• Aucun résultat trouvé

Kinetics of interface formation between weakly incompatible polymer blends

N/A
N/A
Protected

Academic year: 2021

Partager "Kinetics of interface formation between weakly incompatible polymer blends"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: jpa-00212487

https://hal.archives-ouvertes.fr/jpa-00212487

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.

Kinetics of interface formation between weakly

incompatible polymer blends

J.L. Harden

To cite this version:

(2)

Kinetics of

interface formation between

weakly

incompatible

polymer

blends

J. L. Harden

Materials

Department, College

of

Engineering, University

of California, Santa Barbara, CA 93106 U.S.A.

(ReceiB’ed

on March 7, 1990,

accepted

on May 11,

1990)

Résumé. 2014 Nous étudions l’évolution

temporelle

de la

composition

de l’interface entre deux

mélanges

de

polymères

faiblement

incompatibles

à l’aide d’une théorie de

champ

moyen. La

dépendance

en temps de

l’épaisseur

de l’interface

présente

deux

régimes

en loi de

puissance;

d’une loi de croissance initiale en

t1/2,

on passe à une loi de croissance en

t1/4

avant de relaxer vers le

profil d’équilibre

de l’interface.

Abstract. 2014 A

mean field model of interface formation between two

weakly incompatible

polymer

blends is used to

investigate

the time

development

of interfacial

composition.

The interface width as a function of time exhibits two power law

growth regimes ;

a

t1/2 growth

law at

early

times crosses over to a

t1/4 growth

law before

relaxing

towards the

equilibrium

interfacial

profile.

Classification

Physics

Abstracts 05.70 - 64.75 - 68.22 1. Introduction.

In the last few years, there has been intense and rather successful

study

of interdiffusion of

compatible polymer couples

both

experimentally

[1]

and

theoretically [2].

However,

there has been somewhat less effort towards an

understanding

of interface formation between

incompatible polymer species.

This case is of

particular

interest since most

polymer couples

are immiscible over a wide range of

temperatures,

including

those

typical

of

polymer

processing

conditions. This

immiscibility

is a direct consequence of the

inherently large

size of

polymer

chains which

strongly

suppresses the

entropy

of

mixing

of molecules but has little effect on the

typically repulsive

and molecular

weight independent enthalpic

interactions between unlike

polymer species [3].

When bulk

incompatible polymer couples

are

brought

into contact and heated above the

glass

transition limited

interpenetration

of unlike chains

occurs,

leading

to an interfacial

region

of finite extent at

long

times. The

relatively large

spatial

extent of such

polymer

interfaces and the slow kinetics of their formation make them ideal model

systems

for the

study

of interface formation in non-critical

binary

mixtures. In

addition,

the kinetics of interface formation in these

systems

may have relevant

technological

implications

to the

engineering

of

polymer alloys

and

polymer welding techniques.

(3)

1778

Very recently,

experimental

techniques

have been

developed

that allow

high

resolution studies of interface formation between

partially incompatible polymer couples

[4-6].

One such

technique, utilizing

an ion beam method based on nuclear reaction

analysis,

allows direct

measurements of interface

composition profiles

between deuterated and

protonated

polymer

blends with very

high spatial

resolution

[4].

Mixtures of deuterated and

protonated

polymers

are

usually

very

weakly incompatible [7], allowing

for the formation of interfacial

regions

of

finite

yet

very

large

extent. Such a

technique

allows for detailed

study

of the time

development

of interface structure in these

partially incompatible

systems.

Recents exper-iments

employing

this method

[8]

indicate that interface

development

between

incompatible

species

is

quite

unlike that of free interdiffusion between

compatible couples

in which interface width grows

asymptotically

as the square root of time

[1].

A theoretical

description

of the

equilibrium properties

of such interfaces was

proposed long

ago

[9, 10].

Until now,

however,

theoretical studies of the

dynamics

of interface formation have focused on

moderately incompatible

mixtures,

which have rather narrow

equilibrium

interfaces

[ 11, 12].

This paper considers the case of interface formation between very

weakly incompatible

polymer couples.

In this case the

equilibrium

interface is many radii of

gyration

in extent, and

experimental

observation of interface

growth

kinetics is

feasible,

allowing

theoretical

predictions

of

dynamical scaling

laws to be tested.

2. Model.

Consider two semi-infinite blocks

composed

of A and B

polymer

mixtures in the

glassy

state

brought

into contact to form an

initially sharp planar

interface. For

simplicity,

attention will

be restricted to

incompressible

mixtures of

equal

molecular

weight

polymers

(NA

=

NB

=

N). Subsequently,

the

system

is heated above the

glass

transition to a

Fig.

1. - Initial interface

configuration.

Two semi-infinite blocks

composed

of

polymer

mixtures

~1

and ~2

from the bulk coexistence curve at the intended

annealing

temperature

To (near

the critical

(4)

temperature

To

near the critical

point

of

de-mixing

of the A-B

system,

tg

TO:5 T,.

The

composition

of the blocks on either side of the interface is chosen from the bulk coexistence curve for mixtures of

polymers

A and B at the intended

annealing

temperature

To,

as shown in

figure

1. Such a choice is made in order to eliminate the

complication

of bulk convective

transport

of

polymer

across the interface.

Upon

heating,

the interface

undergoes

a

healing

process that smears the

sharp

interface to a narrow but finite

region

some fraction of a

radius of

gyration

in extent. This

healing

process is

complicated by

constraints

imposed by

chain folds near the initial interface and is rather sensitive to the initial

density

of chain ends at

the

interface,

a

quantity

that

depends strongly

on the method of

sample preparation.

The

healing

of

sharp

interfaces is a

topic

of active

investigation

that we will not address here

[13,

14]. Rather,

the

analysis

of this paper will consider the time

development

of the interface

as it

expands

from the

pre-healed

initial width

Wo

towards the final

equilibrium

width

Weq,

as sketched in

figure

2.

Fig.

2. - Sketch of the

range of interfacial

growth

to be considered. The

analysis

considers the

development

of interfacial structure from an

initially

broadened

profile ~0(z)

of width

Wo

towards the final mean field

equilibrium

profile ~ eq (z)

of width

Weq ~

N 1/2 a (X / Xc - 1)- 1/2.

We assume that the free energy per unit area of the

system

can be described

by

the well

known

Flory-Huggins

free energy

expression,

extended to account for slow

spatial

variations in

polymer

concentration normal to the interface

[15-17] :

where 0

is the volume fraction of monomers of

species

A, a

is the monomer

size,

and

f[~J(z)]

is the local free energy per

kB

T as

given by

the usual

Flory-Huggins expression

[3, 15] :

(5)

1780

while the last term represents the

enthalpy

of

mixing

as

gauged by

the

Flory

interaction

parameter

X (T).

One obtains the local chemical

potential g

(z)

of monomers of

species

A

by

functional differentiation of the free energy

expression given by equations (1)

and

(2)

with

respect

to

41

(z) : 1£ (z )

oc

d F / d ~ .

This local chemical

potential

drives the

growth

of the interface wi th

time ;

the

system

will redistribute the A

species

in order to relax any non-zero

gradient

of

J.L. The

growth

of the interface may be characterized

by

the time

dependence

of the A

monomer concentration

profile

across the interface. This may be determined from a

one-dimensional

continuity equation :

where

J(z, t )

is the local current of

species

A. Within the context of linear response,

J(z, t )

may be related to the

gradient

of the chemical

potential

via a non-local

transport

coefficient

[ 15] :

where

A (z - z’ )

is a

generalized

non-local

Onsager

coefficient and IL

(z’ )

is the local chemical

potential

per

kB

T. The

non-locality

of A is a consequence of the

large degree

of

connectivity

of the monomers in a

polymer

medium. After

integration by

parts

and omission of the

vanishing boundary

terms,

equations (3)

and

(4) yield

an

equation

of motion for

~ (z, t ) :

:

Equation (5) together

with

boundary

conditions, ~ = ~1

1 at z = + oo

and l/J = l/J 2

at

z = - 00, and a

specified

initial

profile, ~/0(z, 0) = ~o(z),

constitute a

well-posed

initial-boundary

value

problem.

In

principle, given u ( ~ )

and

A,

we may solve

equation (5)

for

0 (z, t ).

In

practice,

the

resulting integro-differential equation

is

strongly

non-linear and not

amenable to

simple

solution. Chemical

potential gradients computed

with the full free energy functional as

given by equations (1)

and

(2)

are non-linear in the monomer concentration

variables.

Furthermore,

recent theoretical work

suggests

that

Onsager

coefficients for

transport

in

spatially inhomogeneous polymer

blends are often a

complicated

functional of

the monomer concentration

profiles [12].

However,

we may

exploit

the weak

incompatibility

of the

interpenetrating

bulk mixtures under consideration to make several

simplifying approximations.

First,

the

Onsager

coefficient may be estimated from calculations of

single

chain

transport

in a melt of

non-interacting

Gaussian chains

[16, 17].

Second,

the chemical

potential

may be linearized around the critical

composition l/J c

=

1/2.

It is convenient to write the

resulting simplified equation

of

motion in terms of the

gradient

of the

concentration, 0

=

a ~ / az.

Assuming A (z - z’)

is

independent of 0 (z)

as discussed above and

using

the linearized chemical

potential,

the

equation

of motion

for 0 (z, t) given by equation (5)

can be recast as a linear

integro-differential

equation

of motion

for gi (z,

t ).

The solution to this

equation

must also

satisfy

boundary

conditions, 03C8 =

0 at z = ± oo, and an initial

condition, 03C8 (z, 0 )

=

dl/J

o/dz.

The new

equation

of motion and

boundary

conditions are

easily

solved in Fourier space. The modes of

(6)

where 03C8(q,

t ),

03A8o{q)

and

A(q)

are the

spatial

fourier transforms of

03C8 (z, t ),

03C8o(z)

and

A (z ), respectively,

and where xc =

y (Tc)

=

2 /N

within the context of

Flory-Huggins

mean

field

theory [3].

The wavevector

dependent Onsager

coefficient for monomer

transport

in

weakly

incompat-ible

polymer

blends may be estimated from the response of a

single

chain in a melt of

non-interacting

Gaussian

polymers

to an

applied spatially periodic

chemical

potential gradient

[16].

The

resulting expression

for

A (q )

has the form

where

Rg ~ N 1/2 a

is the radius of

gyration

of a

polymer

coil in melt

conditions,

L oc N is the

length

of the constraint tube

enclosing

a

typical polymer

coil as created

by

its

entanglements

with

neighboring polymer

chains,

and where

llo

is an effective monomer

diffusion constant

proportional

to the

microscopic

monomer

mobility

go :

Ao - kB Tuo.

In the

small q

limit,

qRg 1,

A

approaches

the

q-independent

tube diffusion constant,

Dt ’"

llo/N,

and chains

respond by ordinary

diffusion. In the

large

q

limit,

qRg »

1,

ll scales

as A (q) ~ q - 2,

reflecting

the

increasing

inhibition of chain response to

perturbations

that vary

rapidly

on the scale of a chain size.

Equations (6)

and

(7) yield

an

explicit

form for

the interfacial relaxation

modes ;

modes

of tb

decay

exponentially

with time constant

Tq

given by

The direct space solution to the

equation

of motion and

boundary

conditions is obtained

by

inverse Fourier transformation

of 03C8 (q,t ) :

with

Tq

as

given

in

equation

(8).

The initial interfacial

profile

is as

yet

unspecified. Actually,

the

precise

form

of CPo(z),

and hence of

03C8o(q),

turns out to be of little conséquence ; any initial

profile

that

smoothly interpolates

between bulk values

of .0

over a distance

corresponding

to the initial extent

Wo

of the interface will

yield qualitatively

similar results. For

convenience,

we obtain

03C8o (q ) by

choosing 00

of the form

[CPo(z) - CPc] ~

tanh

(z/ Wo).

3. Results.

Widths of

experimentally

determined

profiles

are often estimated from the inverse of the

maximum rate of

change

of monomer concentration across the interfacial

region ;

in terms of the difference in bulk values of monomer concentration on either side of the

interface,

A’k = 0 2

- ~ 1,

the width is

approximately given by

In the above

analysis,

this

approximation

is

equivalent

to

estimating

the width

by

the

(7)

1782

Although

not

integrable

in closed

form,

this is

easily

evaluated

numerically by treating

time as a

parameter.

Before

doing

so, it is instructive to consider the

limiting

behavior of

W(t)

at short and

long

times. To facilitate this

analysis,

we assume for the moment that

f/lo(q)

is

approximately

constant and that Xc - X is

negligible.

This

corresponds

to the idealized case of a very

sharp

initial interface

separating

two very

nearly compatible

bulk

phases.

In this case, the

integral

at

early

times is sensitive to the

large q

behavior of the

integrand

which

asymptotically approaches

the Gaussian form

exp ( - x2)

with

x2 ~

tq 2.

Consequently,

the interface grows as

W(t) ~ t1/2

at

early

times. On the other

hand,

the

integral

at late times is dominated

by

the

small q

behavior of the

integrand

which

asymptotically

approaches

the form

exp ( - x4)

with

x4 ~ tq4.

Thus,

the interface grows as

W(t) -

t 1/4 at late times. At

sufficiently

late

times, however,

the

growth

of the interface must

slow as the

equilibrium

with

Weq -

N 1/2 a (X / Xc - 1)- 1/2

is

approached ;

in the late

stages

of

interface

formation, X - X c

is no

longer negligible

and we

expect

the

growth

to

eventually

deviate from the

t 1/4 law.

The linear

analysis

breaks down at this

point ;

in order to

correctly

describe the final

approach

to

equilibrium,

non-linear terms in the chemical

potential

must be retained.

As mentioned

earlier,

the

healing

of a

sharp

initial interface is a

complicated

process that is

outside the scope of this

analysis,

and hence we are limited to consideration of

pre-healed

interfaces.

However,

we

expect

the

general

case of an

initially

diffuse interface to behave in a

qualitatively

similar manner

except

at very

early

times. Profiles

computed numerically

confirm this

expectation.

Consider,

for

example,

the time

development

of an interface

pre-healed to an initial width

Wo

= 0.25

Rg

and with an

equilibrium

width

Weq

= 10

Rg.

The results of numerical

integration

of

equation (11)

in this case are shown in

figure

3.

Following

a

period

of internal redistribution of material within the

initially

diffuse

interface,

interfacial

growth rapidly

falls onto a

t1/2 growth

law. Growth kinetics

subsequently

cross-over to a

t 1 /4 growth

law before

finally relaxing

towards

equilibrium.

The cross-over

region separating

the two power law

regimes

occurs at a time t * that scales with the time

required

for a

polymer

chain to diffuse its own diameter in a melt of identical chains :

t * ~- R2g/ Ds

s where

Fig.

3. -

Log-log plot

of interface width

W(t)

vs. time t for the case of an interface with initial width

J-fô

= 0.25

Rg

and

equilibrium

width

Weq

= 10

Rg.

Width is measured in units of

Rg,

while time is measured in units of t * ~

R g2/Ds,

the time

required

for self-diffusion of a

polymer by

one

(8)

Ds -- kB

T/N 2

is the self diffusion constant of a

polymer

chain with

degree

of

polymerization

N. For a

given

choice of bulk

polymer

mixtures this cross-over time is

universal,

i.e.

independent

of the width and detailed features of the initial interfacial structure. 4. Discussion.

The novel feature of these results is the appearance of two power law

growth regimes.

The

early

time

dynamics

in

particular

is

unique

to macromolecular

systems,

being

a direct

consequence of the

strong

non-locality

of

transport

coefficients on submolecular

length

scales.

At

early

times,

interface width is narrow on the scale of a

polymer

coil,

leading

to

rapidly

varying

chemical

potential

gradients

within the interfacial

region.

As discussed

previously,

response to such

rapidly varying

chemical

potential gradients

is

strongly

inhibited,

leading

to

different

growth

kinetics at

early

times than at late times when the interface is broad and

transport

coefficients are

effectively

local. In contrast, the

application

of this

analysis

to

ordinary

small molecule

binary

mixtures would lead to a

single

exponent

governing growth

kinetics ;

transport

in such

systems

is not

strongly

non-local on any reasonable

length

scale

[18].

It is instructive to contrast the

physics

of interface formation between immiscible

mixtures,

which we have considered in this paper, with that of miscible

systems.

One

expects

the

differences in interfacial

growth

kinetics in these two cases to arise from the differences in the free

energies

that drive these processes. In the case of immiscible

systems,

interfacial

growth

is

entirely

driven

by composition gradients

across the interface. This manifests itself in the

t1/4

growth

behavior at late times. In miscible

systems,

however,

the

enthalpy of mixing

favors

growth

and both the bulk and interfacial free energy terms drive the process. As such an

interface

expands,

however,

composition gradients

weaken and

growth

is driven

increasingly

by

the bulk free energy terms. Hence in miscible

systems,

an

ordinary

diffusion

equation

governs the

growth

of the

interface,

leading

to

growth

that scales

asymptotically

with time as

t 1/2

The results

presented

here are in

qualitative

agreement

with recent

experimental

data on

the kinetics of interface formation between mixtures of

protonated

and deuterated

poly-styrenes

[8].

The measured

profiles

are consistent with the two

growth regime

behavior described above.

However,

further refinement both of

experimental

resolution and of the theoretical model are necessary before detailed

comparisons

can be made. Current

theoretical work is

extending

the model to

asymmetric

mixtures, NA # N B,

the situation most

relevant to current

experiments

and most often encountered in

practice. Forthcoming

are

detailed results of interfacial

composition profiles

for mixtures of

asymmetric

chains,

as well

as a proper

analysis

of the kinetics of the final

approach

to the

equilibrium

interfacial

structure.

Acknowledgements.

This work was initiated

during

a

stay

with the group of Jacob Klein in the

polymer

research

department

of the Weizmann Institute of Science in

Rehovot,

Israel. 1 am

grateful

to the US-Israel Binational Science Foundation and the US

Department

of

Energy

under contract # DE-FG03-87ER45288 for financial

support,

and to the

faculty,

staff,

and students of the

polymer

research

department

for their

gracious hospitality.

In

addition,

1 would like to thank the Institute for Theoretical

Physics

at the

University

of

California,

Santa Barbara for

(9)

1784

kinetics. 1 also benefitted from very

stimulating

discussions with D.

Andelman,

G.

Fredrickson,

P.-G. de

Gennes,

A.

Halperin,

K.

Kawasaki,

S.

Milner,

and P. Pincus.

References

[1]

COMPOSTO R. J., KRAMER E. J. and WHITE D. M., Macromolecules 21

(1988)

2580 and references therein.

[2]

BROCHARD-WYART F., Molecular Conformation and

Dynamics

of Macromolecules in Condensed

Systems,

Ed. M.

Nagasawa (Elsevier, Amsterdam)

1988, pp. 249-256 and references therein; BROCHARD-WYART F., C.R. Hebd. Acad. Sci. Paris 305

(1987)

657.

[3]

DE GENNES P. G.,

Scaling Concepts

in

Polymer Physics (Cornell University

Press, Ithaca,

NY)

1985, Ch. IV.

[4]

CHATURVEDI U. K., STEINER U., ZAK O., KRAUSCH G. and KLEIN J.,

Phys.

Rev. Lett. 63

(1989)

616 ;

CHATURVEDI U. K., STEINER U., ZAK O., KRAUSCH G., SCHATZ G. and KLEIN J.,

Appl. Phys.

Lett. 56 (1990) 1228.

[5]

MILLS P. J., GREEN P. F., PALMSTROM C. J., MAYER J. W. and KRAMER E. J.,

Appl. Phys.

Lett. 45

(1984)

957 ;

JONES R. A. L., KLEIN J. and DONALD A. M., Nature 321

(1986)

161.

[6]

RUSSELL T. P., KARIM A., MANSOUR A. and FELCHER G. P., Macromolecules 21

(1988)

1890.

[7]

BATES F. S., WIGNALL G. D. and KÖHLER W. C.,

Phys.

Rev. Lett. 55

(1985)

2425 ; BATES F. S., WIGNALL G. D.,

Phys.

Rev. Lett. 57 (1986) 1429.

[8]

STEINER U., KRAUSCH G., SCHATZ G. and KLEIN J.,

Phys.

Rev. Lett. 64

(1990)

1119.

[9]

HELFAND E. and TAGAMI Y., J. Chem.

Phys.

56

(1971)

3592.

[10]

JOANNY J. F. and LEIBLER L., J.

Phys.

France 39

(1978)

951 ; LEIBLER L., Macromolecules 15 (1982) 1283.

[11]

DE GENNES P. G., C.R. Hebd. Acad. Sci. Paris

(II)

308

(1989)

13.

[12]

KAWASAKI K. and SEKIMOTO K.,

Physica

143A

(1987)

349 ;

Europhys.

Lett. 4

(1987)

339 ;

Physica

148A

(1988)

361.

[13]

ZHANG H. and WOOL R. P., Macromolecules 22

(1989)

3018.

[14]

BROCHARD F.

(unpublished).

[15]

DE GENNES P. G., J. Chem.

Phys.

72

(1980)

4756.

[16]

PINCUS P., J. Chem.

Phys.

75

(1981)

1996.

[17]

BINDER K., J. Chem.

Phys.

79

(1983)

6387.

[18]

Of course, a

simple

mean field

analysis, although appropriate

for macromolecular systems, is

Références

Documents relatifs

Thus, the library loading mechanism is guaranteed to load all symbols correctly as long as the freetype.dll is in the correct path; thus the windows version of the interface, if

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

We provide asymptotic expressions for the cyclization time, for the eigenvalues and eigenfunctions, and for the mean square displacement at all time and length scales, with

after the partially water saturation method were observed: the initial growth, the rapid 97.. growth and the steady

the heat transfer takes place in the direction of the mould thickness, and the heat fluxes in the other directions can be neglected; (ii) we can assume that the

6,7 The basic rational of the "covalent-assembly" type probes is the in situ formation of the electronic push-pull conjugated backbone of a fluorescent

ing force that resists normal phase separation h present at scales of the order of the radius of the linear chains. Thus, even below the threshold, the elastic restoring force

In a recenttpublication [I] we have used a crossover function which is appropriate to describe the static structure factor S(Q ~ 0), with Q being the scattering vector, for