• Aucun résultat trouvé

Static and dynamic properties of two-dimensional polymer melts

N/A
N/A
Protected

Academic year: 2021

Partager "Static and dynamic properties of two-dimensional polymer melts"

Copied!
19
0
0

Texte intégral

(1)

HAL Id: jpa-00212422

https://hal.archives-ouvertes.fr/jpa-00212422

Submitted on 1 Jan 1990

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.

Static and dynamic properties of two-dimensional

polymer melts

I. Carmesin, Kurt Kremer

To cite this version:

(2)

Static and

dynamic properties

of two-dimensional

polymer

melts

I. Carmesin

(1)

and Kurt Kremer

(2)

(1)Max

Planck-Institut für

Polymerforschung,

D-6500

Mainz,

F.R.G.

(2)Institut

für

FestkOperforschung,

KFA

Jülich,

D-5170

Jülich,

F.R.G.

(Reçu

le 8 novembre

1989,

accepté

le

7 février

1990)

Abstract. 2014 We

present a detailed

analysis of

the

properties

of dense linear two-dimensional

polymer

chains. The

systems

are simulated

by

the

recently developed

bond fluctuation method. We

investigate

systems

of up to 80 %

density.

The chain

length

and densities considered cover the crossover from the

expanded

single

chain limit to a dense

polymer

melt. The chains

completely

segregate and follow

a power law of the chain extension

R2

>~

N203BD,

N

being

the number of bonds of the chains and 03BD =

1/2.

While for

single

two dimensional chains the Rouse modes are not

eigenmodes they

are

eignemodes

for the chains in a 2-d melt. Relaxation functions and mean square

displacements

display

typical

Rouse-like

behavior,

with the

distinction,

that the

prefactors

for the various relaxation processes differ.

Classification

Physics

Abstracts 61.40-66.10c

1. Introduction.

Most

experimental investigations

of

polymers

consider chains in three dimensions. This is the

most natural

situation,

however,

there are many exact theoretical results for d = 2

[1-3].

Thus,

a two-dimensional

polymer

system

would be an ideal

testing

case for many modern theoretical

concepts,

such as renormalization

group

and conformal invariance methods. Besides this there is not much known about the

dynamics

of such

systems

[4].

Again

a two-dimensional

system

should be ideal to

investigate

the Rouse model

experimentally.

For d = 3

hydrodynamic

interactions

always

dominate the behavior of thé chains in solution. For d =

2,

e.g. chains near a

surface,

this

probably

is not the case since

long

range

hydrodynamic

interactions may be

completely

screened

out. For two dimensional melts the situation is

especially

interesting.

The

topological

interaction is

completely

different

compared

to the d = 3 case.

Following

recent theoretical

investigations

the chains should

collapse

and

segregate

[1].

However,

how does the

shape

of the chains look like? Do

they

appear

as ideal random

walks,

although

the

topological

aspect

of the excluded volume interaction is a dominant

aspect

for d = 2? Recent numerical simulations were not able to discuss this

question

in detail

[6-8].

Since

they

use bond

breaking

simulation methods their

samples

are,polydisperse.

As a

consequence

there is a

mixing

of

long

and short chains

leading

to

an effective

expansion

of the

longer

chains. The classical

papers

of Wall et al.

[9]

were not able

to address the above

questions.

Baumgârtner

[10]

used the

reptation algorithm

to

analyse

the

(3)

916

differences between the ideal random walk structure and the 2-d melt chains. There the

analysis

was confined to the ratio of the radius of

gÿration

and the end to end distance.

Bishop

et aL

[11]

use a Brownian

Dynamics

method in order to

analyse

2-d melts at rather low densities and for

short chains

( N

50,

p =

0.50).

In addition not much is known about the

dynamics

of dense 2-d

systems.

A recent

investigation

of the

Rouse/Zimm

model of 2-d chains

[12],

of course, could not

take into account the

topological

aspects

of the

problem. Experimentally

these

systems

are also

not

completely

out of reach. Modern technics in

producing

various

polymers

could e.g. construct

chains with small side branches

preferring

two different solvents. If these solvents are

immiscable,

such

polymers

would be located at the interface between the two solvents. One can also think of chains between

lipid

monolayers.

Also recent

developments

in the surface

coating

of mica sheets

[13]

may

lead to

stricly

2-d

systems.

Finally,

of course, the classical

approach

of

spreading

chains out of a volatile

good

solvent onto a

liquid

air interface

may

be used. After some

equilibration

time one then can

compress

the film towards a semidilute or dense

layer.

Of course, there one

ltas to take care not to

get

double or multi

layers.

Considering

these

developments,

we think that it is

very

desirable to have a better

under-standing

of such

systems.

Thus we

performed

an extensive numerical

study

of a 2-d

many-chain

system.

To do this we made use of a

recently developed

Monte Carlo

algorithm

[8, 15].

This bond fluctuation method was used to

analyse

the

properties

of an isolated 2-d chain. The way the

moves are constructed also allows to

study

the

dynamics

of 2-d

polymers

[15].

We use this method

to

study

systems

with

up

to 80 %

density

and chain

length

of

up

to N = 100 monomers. A short

preliminary

account of this work is

given

in reference

[16].

Chapter

two

gives

a short

description

of the

algorithm

and the

systems

considered. Then we

describe in some detail the

équilibration

process.

Chapter

three

gives

our results for the

statics,

while

chapter

four

gives

the

dynamics. Finally

chapter

five contains the conclusions and a short outlook.

2. Model and methods.

In order to

investigate

the

dynamical properties

of a 2-d

polymer

system

by

mean of Monte

Carlo,

we have to assure that the

algorithm displays

Rouse

dynamics

for the non reversal random walk

[8].

This random walk assures that consecutive monomers are not allowed to sit on

top

of

each other. Besides this short range

repulsion along

the chains no additional interaction is taken into account. For such a

system

one can use the

complete

set of moves as for chains with full

ex-cluded volume. This then

gives

a

good

and

dependable

check on the static and

dynamic

properties

of an

algorithm.

For

dynamic

Monte Carlo

algorithms

the

typical

requirements

for the moves are,

that

they

create new bond vectors within the chain. Standard lattice methods for d = 2

only

cre-ate new bond vectors at the

ends,

which then diffuse into the chain. This enhances the relaxation times

artificially.

To overcome this

problem

the bond fluctuation method was introduced

recently

[15].

It combines the

advantage

of a lattice

simulation,

which means

e.g.

fast

algorithms

on both

scalar and vector

machines,

with continuum

approaches,

so that new bond vectors are created in

successful moves.

Figure

1 illustrates the method. Each monomer

occupies

four lattice sites. The bond

length 1

is restricted to be

1

13.

In

addition,

a lattice site can never be

occupied

by

more

than

one monomer. For a move, a monomer is selected at

random,

then it

jumps

at random

by

one

lattice distance into one of the four lattice directions. If the new

position

complies

with

both,

the bond

length

restriction and the excluded volume

restriction,

the move is

accepted

and otherwise

(4)

Fig.

1. - Illustration of the bond fluctuation method. Ibe

typical

moves are indicated.

Note that the initial state must not contain crossed bonds. Since the bond

length

constraint is

set to

prevent

crossing

of bonds such a

configuration

would be

preserved

forever. For dense

sys-tems the first blocked

configuration,

which is a situation the

system

cannot reach or

leave,

occurs

at

density exceeding

80% for infinite chains. Since this

configuration

is

ordered,

its

probability

is

vanishing compared

to the other random coil states.

in

the

subsequent investigation

we

only

use densities of

up

to 80%. Since our chains are finite

(N

100)

there will be no

problem

of unresolvable blocked

configurations. (There

is a very small

probability

that a random initial

configuration

of a chain is blocked

by

itself. These

configurations

have a local

density

of one.

The initial conditions

employed

in the

present

work excluded such

configurations automatically).

The simulations were

performed

on a 100 x 100

square

lattice with

periodic

boundary

condi-tions and densities with

20%, 40%, 60%,

80% of the lattice sites

,occupied.

We

analysed

chains with

20

N 100. The N = 100 and 80%

density

system

then constains 20 chains. The

calculations were

performed

on a

Fujitsu

VP

100wector

processor.

typically

100

statistically

in-dependent

systems

of

thé

same kind were simulated in

parallel

The vectorization was set

in

a

way that one monomer in each

system

was moved

simultaneously. By

this

approach,

even

during

the standard simulations

part,

we were able to vectorize the

program

efficiently

[18].

lb start our

systems,

we filled the lattices with ordered

arrays

of chains of

length

N = 100 monomers and

80%

density.

The chains all had a U-like form and were

pairwise

nested into each other.

Fig-ure 2a

gives

a

snapshot

picture

of such a

system

shortly

after the simulation started. One

easily

identifies from this the initial

type

of

configuration. Figure

2b

gives

a

typical snapshot picture

of an almost

equilibrated

structure of the same

sample.

The three different marked chains

give

the

typical

cases. 1Bvo chains have a

fairly

smooth round

surface

while the third is still

strongly

stretched. The stretched one

(which

is the extreme of that

sample)

displays

a rather unfavourable

configuration indicating

that the

system

is not

yet

in

complete

equilibrium. During

the actual

(5)

918

Fig.

2. -

(a)

Snapshot configuration

of a p = 0.8,

N = 100 system a short time after the initial state. The internested structure is

clearly displayed.

(b)

Snapshot

configuration

of the structure of

figure

2a,

where the

system

was almost

equilibrated.

Fig.

3. -

Typical equilibrated

structures for

a)

p =

0.2,

N

= 100;

b)

p =

0.4,

N = 20.

has the same

amplitudes

while in the very

beginning they

were very

asymmetric

due to the initial condition. This

provided

a sensitive test of

equilibration.

The short

chain/low

density

systems

then were made from the N =

100,

p = 0.80

sample

simply

by

cutting

bonds and

eliminating

-

(6)

had started the

systems

from scratch

again.

The

special

initial condition was used to make sure

that no

memory

was

kept during

our simulation and to check the

proposed

segregation

[1].

The initial internested structure is

specific

enough

to follow its

decay

in détail. After

equilibrating

the

systems

we followed the motion of the chains at least

up

to a distance of their own diameter.

Considering

that we

always

ran 100

systems

in

parallel

this

gives

even for

the p

=

0.80,

N = 100

system

20 x 100

independent

diffusion

paths,

which is

enough

to

provide

data of

good

accuracy.

Figure

3

gives

two other

typical

configurations

at

density

p = 0.20 and p = 0.40.

3. Static

properties.

In

polymeric

melts the excluded volume

effect,

which causes the

expansion

of chains in

good

solvent

[2, 19, 20],

is

expected

to be

effectively

screened out

by

the interaction

among

the

differ-ent

polymers.

This is known

theoretically

to hold also for d = 2

[1, 2].

The

major

différence

between d = 2 and d >

2,

however,

is

that,

due to the

topological

constraints in d =

2,

the chains cannot

entangle.

If

they

want to

interpenetrate,

this

only

is

possible by partial alignment

of the chains. This

certainly

causes a reduction of

entropy.

The

consequence

is,

that the chains are

going

to

segregate.

One of the

interesting questions

now

is,

how

strong

is the

segregation depending

on

chainlength

and concentration. The

equilibrium

structure of chain is

supposed

to be like a

Hamiltonian walk.

In order to

investigate

the structure of such a

polymer

melt,

we first check the mean

squared

end to end distance

R2(N)

> and the radius of

gyration

R 2 (N)

> . With rcm

being

the center of mass of a

polymer

and ri the

position

vector of the i-th monomer

they

are

given by

Table 1

gives

the results for the various densities and chain

lengths

considered,

while

figure

4 shows the data for p = 0.40 and

p = 0.80 for several chain

lengths.

For ideal chains one

expects

làble I. -

values for

RG

> and

R2

>

for

the various cases considered. The error bars

given for

R2

> are estimated

from

the

fluctuation

between the 100

statistically independent

systems,

which

(7)

920

Fig.

4. -

Log-plot

of

R2

> and

RG

> vs. N for p = 0.80

(a)

and p = 0.40

(b).

The indicated

slopes

of 2v =1 show that the

expected

behavior is reached

quite clearly.

R2

>

/

4

> = 6 and

R2

>

o:N211, V

=

1/2

[2, 20].

As can be seen from the

data,

for N = 100

compared

to the

expected

ratio

R 2>

is about 10% too small.

Sjrnilar

but

stronger

effects were also found in

[10].

This

certainly

is a

consequence

of the fact that the chains rather

homogenously

fill a little

fluctuating

disk,

while the random walk is a fractal

object displaying

self

similarity

[2].

The

expected

power

law with v =

1/2,

however,

is fulfilled

very

well. From

scaling

for d = 2 one would

expect

R2(N)

>1/2=

Nvf

(N/p-2)

(here

v =

3/4

the

single

isolated chain value

[2]

has to be

used). làldng

the data from table 1 we find that the data for

f (x)

reasonably

well

collapse

on a

single

curve

(besides

N =

20, 25),

however,

the

slope

is not the

expected

one. Ideal

scaling gives

a

slope

of -1/2

for

f (x)

in a

log log

plot,

while here the

slope

is too

large.

This can be

explained

by

the

following:

Scaling

is

only

valid for the limits

x =

N / p-2 ’"

const but N --+ oo, p - 0. Hère we

certainly

are out of this

regime.

Similar effects

were also seen for d = 3

[21].

There is another

interesting

aspect

about how dense the

systems

really

are. For

regions

where the chains

already

interact it makes sense to write

where all the excluded volume effects are taken into the variable

density dependent

persistence

length

p2 p (p)

> 1/2 . Using (2)

we

get land

Ip

as function of

density

as

given

in table II. This decrease of

lp

and

also £,

as also shown

by

figure

5,

certainly

is an effect of

high density. Up

to

(8)

7àble II. - Persistence

length

and mean bond

Iength for

various

densities following

the

definition of

equation

(2).

For p = 0

equation

(2)

requires

Ip

= 00 since there the

exponent

is v

= 3 /4

instead

of

1/2,

as is described

by

the above-mentioned

scaling function.

"

Fig.

5. -

(a)

Plot of the

persistence length

lp

vs.

density

p due to

equation (2). (b)

Plot of the mean bond

length Î

vs.

density

v.

This leads us back to the above discussion of the

scaling

of

R2 ( N )

> and

lîâ (N)

> .

Scaling

assumes that the internal structure of the chains do not

change.

From the

data, however,

we see

that not

only

the

persistence length

varies with

density,

but also the

average

bond

length

itself. The

compression

of the chains is so

strong,

that there is no

space

for a

selfavoiding

walk like

blob,

which is

required

for the

validity

of

density scaling.

Indeed,

if we normalize

R2

> not

only

by

N2"

but also

by

£2

> at least for N = 100 between

p =

40%

and

p =

80%

the

expected slope

(9)

922

chains and the Rouse modes.

First,

we turn to the static

scattering

function of the individual chain

where the

index 1 k

denotes

the

spherical

average over the orientation of k.

Following

the standard

scaling

theories

S(k)

should,

for d = 2 and v

= 4

for the isolated chain

[2], display

the fractal

scattering

law

here e

is the

typical screening length

on which the chain still is

expanded.

From

scaling e

-vl(l-vd) _ -3/2 ,

however as one could see from the data of

R2

>,

R 2

> this

screening

length

does not follow that

power

law for the

high

densities

investigated

hère. Thus we

hardly

can

expect

the occurrence of a

k -4/3

regime

for the

highest

densities considered.

Figure

6a shows

S(k)

for N = 100 and p = 0.20

leading

to 03BE ~

30.

Going

back to

figure

5 this is a reasonable

number,

giving

n =

(l212p/E2) ~- 110

monomers

per

screening·distance.

It means

for p

= 0.20

even the

largest

chains considered

just

interact with each other. This also

explains,

why

there is no distinct

k-2

regime

for p = 0.20 rather than the

typical overshooting

effect of the

single

chain. The

overshooting, coming

from the

fluctuating

ends was discussed

earlier[15]

and can also be found for d = 3. For d =

3,

since the chain can

penetrate

through

itself this effect is much weaker. With

increasing density

this effect is more and more reduced.

Simultaneously,

even the onset of the

k-4/3

region

disappears,

indicating

that e

is

significantly

reduced to about one or two

bond

lengths

for p = 0.80. The

sequence

of

figures

6a-c

nicely displays

this. For p = 0.80 we

give

a

scaling

plot

of

S(k)

vs. k for various N with the normalization that

S(k

=

0, N

=

100)

= 1.

The data

reasonably collapse

onto a

single

curve.

However,

following

the

snapshot

pictures

of the introduction this is somewhat

surprising.

There the cliains seem to

homogeneously

fill

disk-shaped regions.

Such

disks, however,

would

display

the

(d

=

2)

Porod

scattering

of

S(k) -

k-3.

Obviously

the

shape

fluctuations are still

strong

enough

to

prevent

such a

scattering

behavior. In order to see this we have to consider the overall structure of the

system.

With R2

>~

R 2

>N N the

average

density throughout

the chains

approaches

a constant

for distances

Ar »

e.

Thus for d = 2 a

given

chain

only

directly

can interact with a constant number of

neighbors.

For a dense solution of disks

typically

one

would

expect

6

neighbors giving

a

triangular

lattice like short

range

order. If we assume the chains to form

s herical objects

of radius r with

homogeneous

density they

would cover a disk of radius r =

2 12 RG.

Using

this,

we can

estimate the

space

which would be

occupied

by

such chains.

’Paking

the data of table 1 the

chains

typically

would need about 220% for

(N

= 100,

p =

0.80)

to 150% for

(N

=

20,

p =

0.20)

of the

space

available.

However,

this is

simply

impossible.

Therefore,

there must be

strong

fluctuations of the chain

shape.

typically

we find the chains to have between 5 and 7

neighbors averaging

to

6. This leads to a characteristic

packing

pictured

in

figure

7,

as

expected

by

figure

2.

Although

their overall

density

is

relatively

constant there must be

strong

shape

fluctuations. For ideal

ran-dom walks one has

(d

=

2)

For the radius of

gyration

one

typically

gets

a fluctuation reduced

by

a factor of two

[22].

Here we

(10)

Fig.

6. -

(a,b,c) Scattering

function

S(k)

of the individual chains for various densities and N = 100 as

indicated in the

figures.

The data are

averaged

over

randomly

taken orientations of the

scattering

vector k.

(d)’Scaling

plot

S(k,

N)/

(1002/N2)

vs. k for N =

25, 50,

100 and

(11)

924

Fig.

7. -

Illustration of the average local

packing

structure.

which is close to the ideal random walk value.

Although

the chains are

densely packed,

this

suggest

that these fluctuations allow for a

relatively high mobility

of the chains. As we shall

explain

in the

next

chapter

this is the case. These fluctuations which indicate that the chains almost look like ideal

chains,

also

propose

that it

might

be useful to

analyse

the Rouse normal modes of a random walk. For a discrete

system

they

are

given by

[23]

’Iÿpically,

this one dimensional Fourier transform

along

the

sequence

of the chain is done with

periodic

boundary

conditions,

corresponding

to a

cyclic

chain.

By subtracting

the second term we

cut the

ring.

Note that we

only

can

expect

the Rouse modes to be

eigenmodes

for ideal chains. In

principle

this would mean that the monomers

arbitrarily

can

pass

trhough

each other.

However,

we find the Rouse modes are

eigehmodes

of the chains. Within the error bars the cross correla-tions

Xp(t)Xq(t)

>p#q

are zero. Even the

amplitudes

of the modes follow the

prediction

of the Rouse model

As

figure

8 shows

they

follow a common curve for p = 80 and

p = 0.40 for the various chain

lengths.

The

expected slope

is

reproduced

very

well and

only

for small

n/p2

values and p = 0.40

significant

deviations can be found which is consistent with the results of

S(k).

Forp large

the

length

scale is reduced to the

region

where the

self-avoiding

walk behavior starts to dominate. There deviations are

expected

and were also

seen

for d = 3

[24].

This also shows that for

p = 0.80

already

N = 25 reaches the

asymptotic

regime.

In addition the ratio of the

amplitudes

for the

two densities

pictured

is the same as for the

squared

radii of

gyration.

Similar to the fluctuation of the radius of

gyration

we can calculate

6z: =

z

X4

> -

X2

> 2)1/2

/

X2

> . Since

Xp

p p p p

(12)

as

large

as the one for

RG.

This is

essentially

similar to the end to end distance

R2

>. For all densities and chain

lengths

and

p-modes

which follow the

N/p2

power

of the

amplitudes

we find

0.8 b.,

1.0 in very

good

agreement

to the above line of

arguments,

demonstrating

the overall

consistency

of the

picture.

Fig.

8. -

Scaling

of the mode

amplitudes

Xp (N,

p)

> vs.

N/p2

for p =0.80

and p

=0.40. The ratio of

the two sets of data agree with the ratios of

R2

> and

R2

> as shown in table 1 for

N/p2 >

5. The

date

give

the average over the Cartesian coordinates. The absolute value is somewhat smaller than

expected

from ideal chains.

4.

Dynamics.

After

finding

that the static

properties

of dense

polymers

very well

compare

to ideal

chains,

it is

especially interesting

to check the chain

dynamics.

Since the chain cannot

entangle,

the

rep-tation model

certainly

is not an

appropriate description.

On the other hand in order to move

around the chains as well as the monomers have to

get

along

each other. This

puts

strong

con-straints onto the motions of the individual monomers.

However,

how does this affect the overall

dynamical

behavior?

To

investigate

these

questions,

we first

analyse

the mean

square

displacement

of the individual

monomers and the diffusion constants of the individual chains.

Figure

9 shows the data of the diffusion constant D for p = 0.40 and

p =

0.80,

as

given

in

ta-ble III.

Following

the Rouse model we

expect

D =

KBTIN(,

where

kBT

is the

temperature and (

the bead friction. Since there 1 - o

temperature

in our

system

but

only

the

acceptance

rate, we can

(13)

926

Fig.

9. - Diffusion constant 4D = lim

g3(t)

vs. number of monomers N. t->oo

this later in the context of the modes. What is

surprising

is that the diffusion constant seems to

display

a weaker

N-dependence

than the Rouse model

suggests.

Any

crossover to

reptation-like

behavior would cause the inverse

effect,

namely

an

increasingN-dependence. Certainly

we would

need a more extensive

analysis

of a

larger

system.

As is showm

later,

such an acceleration does

not show

up

in the

meansquare

displacements

of the monomers. Thus since for this estimate of D

large

time intervals were

used,

figure

9

probably

shows a residue of the

equilibration.

Thus

Table III. -

Diffusion

constants 4D

for

various N and

density p

= 0.4 and

(14)

we think that the data are consistent with D -

N-1.

Figure

10

gives

examples

for the motion of individual monomers N = 50 and p =

0.40,

0.80.For N = 100 the data show the same behavior.

Fig.

10. - Mean

square

displacement

gl(t)

and

g2(t)

and the diffusion of the

center

of mass of the chain

g3(t)

for p

= 0.80

(a)

and p = 0.40

(b)

and N = 50. The two sets of data for

Yl/ Y2

give the

results for middle

(lower data)

and end monomers

(upper

data).

The difference in the motion between inner and the outer

monomers decreases with

increasing

density.

What is

especially

striking

is that the difference between the mean

square

displacements

gl , 92 of

the total

chain,

given by

and

corresponding

data of gl, 92 for individual inner monomers is much smaller than for d = 3. For d = 3 this effect is

quite

dramatic since this shifts this

visibility

of

reptation

in these

quatities

far

beyond

the

entanglement length

[23-25].

However,

similar to the chain confined to a

straight

tube,

g2-for-inner

monomers saturates much earlier than the mean

square

displacement

gl. Such

an effect was first observed for the chain in the

straight

tube

[26]

and seems to be

typical

for such constrained

geometries.

This

certainly

must be an effect of the

segregation

of the chains and the

confinement due to the

surrounding compared

to the 3-d case. For both cases we find a nice

t112

behavior for gl, 92 up to the

longest

relaxation time of the chain. After that time the diffusion of the whole chain takes over. Table IV

gives

a collection of relaxation times. So far all data seem to

reasonably

agree to the Rouse model. The most crucial test for the

dynamic properties

is the

(15)

928

Thble IV. -

Relaxation times Tp

for the first

three

modes p for

chains at densities

of

p = 0.4 and 0.8.

The error bar

of

the data is around 15% .

rates of the Rouse modes

Xp

as well as the autocorrelation function of the radius of

gyration

squared. Figure

11 shows the relaxation function

for N =

50,

p = 0.40 as an

example

for the first three modes.

They display

a

good

single

exponen-tial behavior for

longer

times,

which enables us to calculate the relaxation times which enables us

to calculate the relaxation times Tp . For the Rouse model one

expects

Figure

12

gives

a

scaling plot

of the relaxation times 7p from table IV.

They

all follow the

single

curves and

roughly

follow a

slope

of

(N/p)2.

From these data as well as from the diffusion

con-stants D we

directly

estimate the normalized bead friction

(IKBT. Using

the data of table II for l and

lp

and table III for D the two friction constants do not agree very well. We find that the relaxation of the individual modes is

considerably

slower than one would

expect

from the diffu-sion. For d = 3 melts it was found

[25]

that

they

agree

very well.

Taking

e.g.

N =

50,

p = 0.8

(here

the statistic is better than for N =

100)

the modes

give

(IKBT

6.4 x

102

while D

gives

1.5 x

102,

giving

a ratio of ~ 4.3. For p

= 40%

this ratio is reduced to a value of less than 3. Thus with

increasing density

this deviation increases as well.

However,

this reflects that the monomers

cannot

interpenetrate

freely,

as

required

by

the Rouse model.

Thus,

although

the Rouse modes

are

eigenmodes

and 91 does not show

significant

deviations from

Rouse,

the diffusion seems to be

governed by

shape

fluctuations,

which are faster

by

a constant factor than the internal

reorgani-zation which would allow the Rouse modes to relax. Similar distinctions in time scale have been

proposed

for star

polymers

[29].

Finally,

let us check the overall relaxation of the radius of

gyration.

The

autocorrélation

function shows a

single

exponential

decay

similar to the modes. A more detailed

inspection

shows that

rouphlv

In order to obtain the

precise

N and p

dependence

much more estensive simulations are needed.

The relaxation times are

given

in table V. Since the fluctuation of

RG

are the

shape

fluctuations

(16)

Fig.

11. -

Examples

of a

typical

mode relaxation

plot

for N =

50, p

= 0.4. The time is

given

in Monte

Carlo

steps

(mcs).

Fig.

12. - Relaxation

times of the first three Rouse modes vs.

N/p

for p = 0.40 and p = 0.80 and different

(17)

930

Thble V. - Longest

relaxation times

of

the auto-correlalion

function

R$(t)Rl(0)-(flÕ(0»)2

»

/(

RG

>2 -

(Rô)2

>)

for

various

chain

lengths

and densifies p = 0.4 and

0.8. The

N2

dependence

is

rougihlyfulfiued.

The error bar

of

the data is

expected

to be between 10% and 15%.

Fig.

13. -

Acceptance

rate

A(S)

of

attempted

moves vs.

density

p. The insert

gives

the normalized devia-tion of the

acceptance

rate

A(S)

with

respect

to the

acceptance

rate  =

A( p

=

0)

of an inner monomer of

a

single

isolated chain.

mode.

Using

again

equation (11)

with p = 1 shows that we arrive at a bead friction which is in

very

good

agreement

to the diffusion

data,

while the modes

give

a different result. This rather

regular

behavior indicates that there are no

precursors

of the

glass

transition

yet

in

agreement

to the mode

analysis.

The

density

dependence

is

directly

seen also from the

acceptance

rate of the moves. For lattice

gases

it is known that due to

back/hopping

corrélations the

acceptance

rate

decays

slower than the increase of relaxation times

suggest

[27].

Figure

13 shows the relative

(18)

power

law with

density

as for

THe;

indicating

that back

hopping

only

enters in the

prefactor

and is dominated

by

the intrachain correlation of monomers

nearby along

the chemical

sequence

of the

polymers.

5. Conclusion.

We

presented

a detailed

study

of the

properties

of a melt of two-dimensional

polymer

chains.

By

the use of the bond fluctuation method we were able to

study

both static and

dynamic properties

of the chains.

For statics we found that the

chains,

although they

cannot cross each

other,

display

typical

random walk behavior for

many uantities

like

R2

>,

RG

>,

S( k)

and the Rouse modes. As function of

density

the

decay

of

R ,

Rb

shows that we are

explonng

the hmit of

high

density

melts. The chains

segregate

comptetety

and the ends seem

only

to ftuctuate on small scales. Nevertheless the overall

shape

fluctuations turn out to be sufficient to result in a

scattering

law

S(k) -V k2

instead of the Porod

scattering.

The

amplitude

of these

shape

fluctuations is of the same order as for

ordinary

random walks. For the lowest

density

considered we

nicely

find the inner

k-4/3

regime

for short internai distances. This is

important

with

respect

to recent ideas of the

e-collapse

transition of

polymers [28].

Considering

the

present

result for these

concepts

the disorder of the defects in the

lattice,

althougt they

are

annealed,

is crucial. Otherwise the

scattering

function of

our

systems

would show a

very

different behavior.

Although

the overall

properties

of the chains

are the same as for ideal Gaussian chains there are some distinct differences with

respect

to the inner structure, such as the uniform

density

in the chains or

probability

of monomers

approaching

each other. The latter

quantities

have been calculated

[1]

and a future

investigation

should be able

to check this.

The most

unexpected

results were found for the

dynamical

properties.

The Rouse model does not consider

any

other interaction betweeen the monomers than the chain

connectivity.

Consid-ering

that this

requires

monomers

crossing

each other

freely,

we found it rather

surprising

that the

2-melt almost

perfectly

reproduces

the Rouse model. There is no

sign

of decrease of

mobility

due

to

topological

constraints. Ibe

only,,

but

significant

deviation is

given

in the time scales.

Although

the mode relaxation as well as the diffusion

give

the same N

dependence

of the relaxation

time,

there is a difference in the

prefactor

of about 4

for p

= 0.8. It shows that diffusion via

shape

fluc-tuations is faster than internal structure relaxation. This

suggests

that the

similarity

to the Rouse model is rather accidental since the 2-d

exponents

for the dense

system

and the related

quantities

exhibit the same

power

law as the

ordinary

random walks. We therefore think that it is

especially

interesting

to

investigate

the chain

dynamics

of a

collapsed

3-d chain. There many

interesting

de-viations

might

occur, while for the

single

collapsed

2-d chain we do not

expect

different behavior

compared

to our

present

findings.

Acknowledgements.

(19)

932

References

[1]

DUPLANTIER

B.,

SALEUR

H.,

NucL

Phys.

B290

(1987)

291 and references

therein ;

DUPLANTIER

B.,

J.

Phys

A19

(1986)

11009.

[2]

de GENNES

P.G.,Scaling

Concepts

in

Polymer

Physics, (Cornell

Univ.

Press, Ithaca,

N.Y.)

1979.

[3]

NIENHUIS

B.,

Phys.

Rev Lett. 49

(1982)

1062.

[4]

DOI

M.,

EDWARDS

S.F.,

The

Theory

of

polymer Dynamics (Clarendon

Press,

Oxford)

1986.

[5]

MANSFIELD

M.,

J. Chem.

Phys.

77

(1982)

1554.

[6]

TUTHILL

G.F.,

J. Chem.

Phys.

90

(1989)

5869.

[7]

REITER

J.,

ZIFFERE

G.,

OLAJ

O.F.,

preprint (1989).

[8]

For a

general

overview about lattice simulations methods for poymers see KREMER

K.,

BINDER

K.,

Comp.

Phys. Rept.

7

(1988)

259.

[9]

WALL

F.T.,

CHIN

J.C.,

J. Chem.

Phys.

66

(1977)

3143;

WALL

F.T.,

SEITZ

W.A.,

J. Chem.

Phys.

67

(1977)

3722.

[10]

BAUMGARTNER A.,

Polymer

23

(1982)

334.

[11]

BISHOP

M.,

CEPERLEY

D.,

FRISCH

H.L.,

KALOS

M.H., J.

Chem

Phys.

75

(1983)

5583.

[12]

MUTHUKUMAR M. J Chem.

Phys.

82

(1985)

5696.

[13]

GRANICK

S.,

private

communication.

[14]

Such a method is the standard method to measure the 2-d

exponents.

see e.g. VILANOVE

R.,

POUPINET

D.,

RONDELEZ

F.,

Macromolecules 21

(1988)

2880.

[15]

CARMESIN

I.,

KREMER

K.,

Macromolecules 21

(1988)

2819.

[16]

CARMESIN

I.,

KREMER

K.,

in

proceedings

of ILL

Workshop

on

Polymer Dynamics,

Proc. in

Physics,

Eds. D.

Richter,

T.

Springer

(Springer Heidelberg)

1988.

[17]

for a detailed discussion of such

ergodicity problems

see also: MADRAS

N.,

SOKAL

A.D., J.

Stat.

Phys.

47

(1987)

573.

[18]

More than 90% of the code ran on the vector unit of the VP100. Besides this the VP100 does not

give

further information

regarding

the

effectivity

of the vector unit

being

used.

[19]

There is an extensive

experimental

and numerical literature devoted to 3-d melts. For simulations see

references

[8,25]

and K. Binder

preprint

(1989).

[20]

FLORY

P.J.,

Principle

of

Polymer Chemistry (Cornell

Univ.

Press, Ithaca,

N.Y.)

1953.

[21]

See e.g. KREMER

K.,

Macromolecules 16

(1983)

1632.

[22]

LYKLEMA J.W.,

KREMER

K., J.

Phys.

A19

(1988)

279.

[23]

KREMER

K.,

GREST

G.S.,

CARMESIN

I.,

Phys.

Rev. Lett. 61

(1988)

566.

[24]

DIAL

M.,

CRABB

S.,

CRABB

C.C.,

KOVAC

J.,

Macromolecules 18

(1985)

2215.

[25]

KREMER

K.,

GREST

G.S.,

preprint J.

Chem.

Phys.,

in press

(1990).

[26]

KREMER

K.,

BINDER

K., J.

Chem.

Phys.

81

(1984)

6381.

[27]

KEHR

K.W,

BINDER

K.,

Application

of the Monte Carlo Methods in Stat.

Physics, Topics

Current

Phys.

36 and references

therein,

Ed. K. Binder

(Springer Verlag, Heidelberg)

1987.

[28]

DUPLANTlER

B.,

Phys.

Rev. A 38

(1988)

3647;

DUPLANTIER

B.,

SALEUR

H.,

Phys.

Rev. Lett. 59

(1987)

539.

Références

Documents relatifs

Molecular docking experiments suggest that chromeno β-lactam s 8a-p may bind to human inducible nitric oxide synthase protein via hydrogen bond and hydrophobic

This work studies the influence of the plasticizer PEG grade on the properties of coating films made with HPMC E19, stearic acid and polyethylene glycol. The addition of PEG enhanced

Computer simulations were done of the mean square displacement (MSD) of tracer particles in colloidal gels formed by diffusion or reaction limited aggregation of hard spheres..

the spin states of the four triangles, the way they are coupled and the way they are coupled to the central Fe atom. Figure 4 shows the energies of the ~250 lowest spin states as

For GB + CaCl2 the texture of static and dynamic cakes are very different (Plate 4): for the cake obtained in dynamic conditions, cells are smaller and the pore walls are no

We also find in agreement with de Gennes that the critical exponents are well described by their mean field values and that the size of the critical region

By Monte Carlo simulation of two lattice models of polymer melts, the enrichment of the chain ends in the vicinity of a wall is observed.. For the bond fluctuation model

/ La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur. Access