• Aucun résultat trouvé

Hydrodynamics of polymer solutions via two-parameter scaling

N/A
N/A
Protected

Academic year: 2021

Partager "Hydrodynamics of polymer solutions via two-parameter scaling"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00248044

https://hal.archives-ouvertes.fr/jpa-00248044

Submitted on 1 Jan 1994

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.

Hydrodynamics of polymer solutions via two-parameter scaling

Ralph Colby, Michael Rubinstein, Mohamed Daoud

To cite this version:

Ralph Colby, Michael Rubinstein, Mohamed Daoud. Hydrodynamics of polymer solutions via two-parameter scaling. Journal de Physique II, EDP Sciences, 1994, 4 (8), pp.1299-1310.

�10.1051/jp2:1994201�. �jpa-00248044�

(2)

Classification Ph_vsic.s A h.iiiat I-I

83. ION 83.20F 61.25H

Hydrodynamics of polymer solutions via two-parameter scaling

Ralph

H.

Colby ('),

Michael Rubinstein

(')

and Mohamed Daoud

(2)

(1) Imaging Research and Advanced

Development,

Eastman Kodak

Company,

Rochester, New York 14650-2109, U.S.A.

(2) Laboratoire Leon Brillouin (*), C.E.N. Saclay, 91991Gif-sur-Yvette Cedex, France

(Receii'ed10 Januar~ 1994, accepted 5 May 1994)

Abstract. We discuss the temperature and concentration regimes for the

hydrodynamic

properties

of linear

polymer

solutions. New regimes are reported in addition to the known

regimes

for the static conformations of the coils. The new

regimes

appear because of the ex18tence of a second length scale, the tube diameter, which is not proportional to the screening

length.

In the

theta regime, the polymer volume fraction where entanglement becomes important is

~c m la

j/hl'?

N- "~~ where

aj, h, and N are the tube diameter in the melt, the Kuhn length, and the number of Kuhn segments in the chain, respectively. Except for very large N, ~~ is larger than the

overlap concentration ~ *. In the good solvent region. ~ * and ~~ have the same

scaling

for their

dependence

on N, but have different temperature

dependences.

The results are ~ummarized on a

diagram

in the

temperature-concentration plane.

Introduction.

It has been

previously

shown that static

properties

of linear

polymer

solutions can be described

rather

directly using 8caling

for effects of both concentration and temperature

[1, 2].

Presumably,

the latter is

directly

related to the so-called solvent

quality.

A

simple phase diagram

was

suggested

Ii that summarized the static

scaling properties

of solutions of flexible

linear

polymers. Dynamic properties

of these

polymers

were also studied in the

good

solvent

case

[3]

with

scaling

arguments, and the results

compared favorably

with

experiments.

When

generalized

to theta

solvent, however,

these

scaling

ideas based on a

single length

scale do not agree with

experiments [4~6].

This led to the two-parameter

scaling [4]

for theta solvent

dynamics,

which showed that there are two

important length

scales the

screening length [2,

7, 8] and the tube diameter

[9, 10],

which scale

differently

with concentration in theta solvent.

The

resulting

two-parameter

scaling predictions

were shown to compare

favorably

with

experimental

results for

viscosity

and modulus of semidilute theta solutions

[4-6].

(*1 Laboratoire commun C-N-R-S--C-E-A-

(3)

In this paper we

explore

the consequences of two-parameter

scaling

for the full range of temperature and concentration. We find new results for the temperature

dependence

of the tube diameter in the

good

solvent

regime.

Thus in

good

solvents the

screening length

and the tube diameter have identical concentration

scalings

but different temperature

scalings.

Another

consequence of two-parameter

scaling

is that there are

regimes

for both

good

and theta solvent, which are

semidilute,

but

unentangled (because

the tube diameter is

larger

than the

screening length).

The full temperature and concentration

dependences

of criteria for chain

entanglement

will be elucidated.

After a brief review of

scaling

for static

configurations

of chains, we write

general

expressions

for relaxation

times, modulus,

and

viscosity

in terms of the

appropriate length

scales and known parameters such as solvent

viscosity

7~~, Boltzmann's constant

k,

and

absolute temperature T. We next define the various

regimes

on the

temperature-concentration plane

and obtain

scaling

laws for the concentration and temperature

dependences

of the various

length

scales.

Finally,

we tabulate our results for concentration and temperature

dependence

of relaxation

times, modulus,

and

viscosity

in the different

regimes

and compare with

light scattering

and linear

viscoelasticity experiments.

Static

scaling.

In dilute

solution,

the root-mean-square end-to-end distance,

Ro,

of the

polymer obeys

a

scaling

law

[2,

II in terms of the number, N, and

length,

b, of Kuhn segments

(hereafter

called

monomers)

in the chain.

RombN~' (1)

The

Flory

exponent

v is 1/2 in solvents (we

neglect logarithmic

corrections to

scaling)

and

roughly

3/5 in the

good

solvent limit. As concentration is increased, the dilute

regime

ends

when the coils start to

overlap.

The volume fraction of

polymer

at which this occurs is

determined when it reaches the volume fraction inside a

single

coil

[I Ii.

#i*iNb~/R(wN'~~' (2)

Above

overlap,

the semidilute solution is characterized

by

a

screening length [2, 7, 8]

(or blob

size), f, beyond

which

hydrodynamic

interactions and any excluded-volume interactions are screened. The concentration

dependence

of f was determined from a

scaling

argument

[2, 8],

which

requires

f to be

independent

of N.

it Ro(~b/~b *)-

~"~~ ~'- ' ' (3)

Note that at #i = #i

*,

the

screening length equals

the dilute solution coil size

Ro.

Inside a blob any excluded-volume is felt and thus the

screening length

can be related to the number of

monomers in the blob~ g,

by

a relation

analogous

to

equation

(1).

f I

bg

~

(4j

On

length

scales

larger

than

f

the chain

obeys

random walk statistics (a random walk of

N/g blobs).

(R/f

)~ m

N/g (5)

At some concentration #i~,

larger

than ~b

*,

the coils become

entangled.

The

length

scale

describing

this

entanglement

is the tube diameter from

reptation theory [9, 10],

a. The tube

(4)

diameter has been

suggested

to scale with the

following

concentration

dependence [4]

(aj

is the tube diameter in the

melt).

away ~b~'

(6)

The exponent,r has been determined from the assertion that a fixed number of

binary

contacts in a volume

a~

controls

entanglement. Binary

contacts are determined

by

the

screening length

in the

good

solvent

limit,

where we expect a~f and thus x

=

v(3

v

Ii

=

3/4. In solvent the

screening length

reflects ternary contacts

[2, 12]

and a

simple

mean field argument

[4] gives

x

=

2/3. In

good

solvent a and

f

scale with the same concentration

dependence,

but in theta solvent

they

do not, and a and f

actually

intersect

[4]

at

#i =

(aj/b

)~~ However, except for

polymers

of very

high

molecular

weight,

this concentration is below the

overlap

concentration, and we

ignore

such

ultra-high

molecular

weights

in this

paper.

Since a

~ f, an

entanglement

strand of

N~

monomers is a random walk of

N~/g

blobs.

(a/f

)~

i N

~/g (7)

Combining equations

(5) and

(7) provides

a final static relation that will be used in what follows to

replace

all static parameters

by

the three

length

scales f, a, and R.

(Rla

j~

w

N/N~ (8)

Dynamic scaling.

Below the

overlap

concentration (~fi ~ #i *

) polymers

exist as separate coils and

dynamics

are

known to be described

by

the Zimm model

[13]

with full

hydrodynamic

interaction

[10].

The

longest

relaxation time in dilute solution follows Zimm

scaling.

Tz,mm I 1~

~

R(/kT (9)

The modulus for the Zimm model is determined as kT of stored energy per chain.

Gw

#ikT/Nb~ (10)

The

product

of

equations (9)

and

(10) yields

the

viscosity

of a dilute

polymer

solution.

7~ w

GTz~~~

w 7~

~

#iR(/Nb~

I1)

This is the basis of the earlier

Fox-Flory equation

[I

I].

According

to two-parameter

scaling [4]

there should be a

regime

of concentration above

#i * that is semidilute but

unentangled.

This

regime

occurs for f ~R ~ a, and the

physical picture

of the chain is that of an

unentangled

random walk of

N/g

blobs of size f.

Hydrodynamic

interactions dominate up to size f, so the relaxation time of a blob is

determined

by

the Zimm model.

T~ I 7~~

f~/LT. (12)

The

longest

relaxation time is the Rouse time

[10,

14] of the chain of

N/g

blobs.

T~~~,~ -

(N/g

)~ T~

-

(R/f

)~ T~ w 7~

~

R~/fkT. (13)

(5)

The

unentangled

modulus is still determined

by

kT of stored energy per chain

(Eq. (lo)).

The

viscosity

of the

unentangled

semidilute solution is thus obtained from the

product

of

equations (ID)

and

(13).

7~ m 7~

~

#iR~/fNb~ (14)

At still

higher concentrations,

when R > a, the

polymers

become

entangled.

The

physical picture

of the chain is now an

entangled

random walk of blobs. The

dynamics

inside the blob

are as

before,

and the Zimm relaxation time of the blob is

given by equation (12j.

Entanglement

strands of

N~/g

blobs relax

by

Rouse

dynamics (with

relaxation time

T~).

T~ m

(N~/g

j~T~ m

(a/f

)~ T~

m 7~,

a~/fkT. (15)

The

longest

relaxation time is controlled

by reptation

of the chain of

N/N~ entanglement

strands [10].

T~~~ ~ IN IN~)~ T~ m

(Rla

)~T~

w 7~

~

R~la~ fkT

II

6)

The

entangled (plateau)

modulus is determined

by

kT of stored energy per

entanglement

strand

[4].

G

w

kTla~ f (17)

The

viscosity

of the

entangled

semidilute solution is determined from the

product

of

equations (16)

and

(17).

7l m 7l

~

R~la~

f~ (l

8)

We next examine the

scaling

of

lengths

f, a, and R as functions of temperature and

concentration to make

specific predictions

for the various

regimes.

Regimes

on the

temperature-concentration plane.

For

temperature scaling Ii

we make use of the standard reduced variable

[2]

r, defined below

in terms of the excluded volume

[2,

1II, v, such that

r =

0 in a theta solvent

(T

v

=

0) and r

=

I in the

good

solvent limit

(T

~ oJ ; v =

I)

r =

v/b~

I

(T )IT (19)

Previous works

[1, 2]

have demonstrated that in semidilute solutions, r

~ #i means that the

two-body

interactions

(v)

dominate and thus

r ~ ~fi

corresponds

to

good

solvent

scaling.

Conversely,

r ~ ~fi

implies

that

three-body

interactions dominate and therefore theta solvent results govern

scaling.

In other words, the criterion

[1, 2]

~~ ~ ~b

(20)

defines the crossover from theta to

good

solvent

regimes

in the

r-~fi plane.

In the theta solvent

regime equation (I gives

the coil size

(with

v

=

1/2).

RibN"~

r~~fi.

(21)

Equation (3) gives

the

screening length (with

v

= 1/2).

I m b~b ' r ~ ~b

(221

(6)

Equation (6) gives

the tube diameter

(with,t

=

2/3).

a w a ~fi

~'~

r ~ #i

(23)

The

good

solvent

regime

is more

complicated

because of the presence of the thermal

blobs

[2]

that

bring

in an

explicit

temperature

dependence

to these

length

scales. In dilute solution the

physical picture

of the chain is

a random walk up to the thermal blob size

f~

and an excluded~volume walk

beyond f~. Therefore,

there are two

length

scales in the

dilute

good

solvent

regime.

The thermal blob size

[2]

is

only

a function of temperature.

f~

w br ' r >

#i1 (24)

The coil size

[1, 2]

is determined from a

self-avoiding

walk of thermal blobs.

The

overlap

concentration

[1, 2]

is determined from

equation (2).

~ _415

~- 315 ~ > Qi

$

~~~~

The coil size

(Eq. (25))

crosses over the standard theta solvent result

(Eq. (21))

when

r = #i

i,

and also crosses over to the standard result

(Eq. II )

with v =

3/5)

in the

good

solvent

limit

(T

~

).

In semidilute solutions in the

good

solvent

regime,

the chain is

again

a random walk up to the thermal blob size

f~,

and thus

equation (24)

is valid for all concentrations in the

good

solvent

regime.

Excluded volume is effective on

length

scales between

f~

and the

screening length,

f. The

screening length

is obtained

using equations (3), (25),

and

(26) (with

v =

3/5) [1, 2].

f w b~fi ~'~ r "~ r > ~fi

(27)

There are g monomers in a blob of size f.

g ~

iiliTi~'~ (iT/b)~

i ~b ~'~ r~ ~'~ r ~ ~b

(281

Beyond

f the chain is a random walk of swollen blobs.

R

w f

(N/g )"~

w hN "~

~fi

"~

r

"~

r ~ ~fi ~ #i *

(29)

We determine the

scaling

for the tube diameter in the

good

solvent

regime

with a

simple scaling

argument at a fixed

r. Since the tube diameter is

always

controlled

by two-body

interactions

[4],

and the

screening length

crosses over from

two-body

to

three-body

interaction dominance at #i

= r, we

anticipate

a crossover for the tube diameter at #i

= r. For

~fi > r,

equation (23j

holds (it

effectively

uses a mean field count of

binary contactsj.

For

~fi ~ r we know that the

density

of

binary

contacts scales with the

screening length.

a WA

~fi

~'~

~fi ~ r

(30j

We evaluate A

by matching equations (23)

and

(30)

at ~fi = r.

-213 ~ -3'4

(~~)

aj T I T

Thus A

w aj r

"'~, giving

us a

prediction

for the tube diameter in the

good

solvent

regime.

a w aj ~fi ~'~

r

"'~

r ~ ~fi

(32)

JOURNAL DE PHYSIQUE ii T 4 N' 8 AUGUST 1994

49

(7)

We now

identify

six

regimes

in the

temperature-concentration plane,

shown in

figure

I.

Regimes

I and I' are the dilute

good

and dilute theta

regimes, respectively. Regimes

II and

II' are the semidilute

unentangled regimes

in

good

and theta solvent.

Regimes

III and

III' are the

corresponding

semidilute

entangled regimes.

Dilute and semidilute solutions are divided

by

~fi

*, given by equations (2) (for

r ~ ~b, with v

=

1/2)

and

(26),

which is

equivalent

to the concentration where the

screening length

and the coil size are the same.

fit l12 *

* ~

~ ~

(33)

fit 415 T 315 ~ > *

Good and theta

regimes

are

separated by

a critical reduced temperature

II-

~bi

~b ~

~b1

~~~~

~~~

~b

~b>~bi

Unentangled

and

entangled regimes

are divided

by

#i~ that is determined as the concentration where the tube diameter and the coil size are the same.

Matching equations (21)

and

(23),

and

equations (29)

and

(32)

thus

yields

the

entanglement

criterion.

~b~ w

~'~~)~~~

N~ ~'~

~ ~ ~

(a

16 )~'~ N ~'5 ~ ii15 ~

~

> ~b

e

(35j

Equations (33-35j

are the solid curves in

figure

I.

We use the results for f

(Eqs. (22j

and

(27jj,

a

(Eqs. (23j

and

(32))

and R

(Eqs. (21), (25)

and

(29))

to calculate the

viscosity, modulus,

and various relaxation times

predicted

in the

t~ te

III

I

II

III'

,

II'

Fig,

I.- Phase

diagram

in the space of

polymer

volume fraction ~ and reduced temperature

r w (T )/T, for the

example

of

aj/b

= 5 and N loo. Primes denote theta solvent regimes and no

prime

denotes

good

solvent

regimes. Regimes

I and I' are dilute regimes II and II' are semidilute- unentangled

regimes

III and III' are semidilute-entangled.

(8)

dynamic scaling

section in the six

regimes.

The results are summarized in table I. In dilute solution

(regimes

I and

I')

we recover the standard results of the Zimm model

[10, 13]

for

relaxation

time,

modulus and

viscosity.

In

semidilute-unentangled

solution

(regimes

II and

II')

Shiwa

[15]

has

predicted

the theta solvent results and the

good

solvent results in the

good

solvent limit

(with

r

=

I)

but no temperature

dependence.

We have

previously

derived the

entangled

theta solvent results

[4]

and the

good

solvent limit results in the

entangled

case are

known as well

[2, 3]

without the temperature

dependence.

Table I.

Summary of predictions for dynamic scaling.

Reg,me Tz,~~ iTl~, h~ I') TRou~ kTl~, h~ (~) T,~~ kTl~, h~ Gh'liT ~l~,

' N~~ N~' ~ N"~ ~

l' N~'~ N-' ~ N~'~~

II ~ ~'~ r "~ N ~ ''~ r~'~ N ' ~ N~ ~'~ r

"~

ii' ~-' N2~ N-'~ N~2

-9>4 -~'4 -9'4 7n2 ~'

)~

~i ~i12 ~'6 h ~ ~q,4 1<i2 h 2 ~, j~i4 11<12 h

)~

~~~ ~ ~ ~ ~

h ~ aj ~ Uj

~ ~

aj

(') Zimm time of the entire chain for regimes I and I'. Zimm time of a blob (Tz,mm T~) for regimes II, II', III and III'.

(2) Rouse time of the entire chain for regimes II and II'. Rouse time of an entanglement strand (TRou« " T~) for regimes III and III'.

Note that

scaling predicts

a concentration

regime

that is semidilute and

unentangled,

as first

suggested by Graessley [16].

The width of this

semidilute-unentangled (Rouse) regime

is

determined from the ratio of

equations (33)

and

(35).

(aj16

)~'~ N~ "~ r ~ ~b *

~(

m

(aj/b)~'~ N"~° r~'~

#i * ~ r ~ #i

~

(36)

~ (a,/b)~'~ r~"~

r >

#i~

Comparison

with

experiments.

The concentration

dependence

of

viscosity

in

entangled good

solvents has been reviewed

recently [17],

with

7~

#i~°~°~ (see

Tab. III of Ref.

[17]

and Refs,

therein),

in excellent agreement with the

scaling prediction

of

7~ #i '~'~

Experimentally,

the

plateau

modulus in

good

solvent

[17]

scales as G #i ~.~ ~ °~,

again

in excellent agreement with the

scaling theory.

The theta solvent results for

plateau

modulus have also been reviewed

recently [18],

with G

#i~~~°

' in excellent agreement with the

predicted

exponent of 7/3.

Viscosity

data in

entangled

theta solution are scarce, but appear to be consistent with the

scaling prediction

(7~

#i')

with exponent values x =

5.4

[5, 4],

x

=

4.8

[6, 4],

and x

=

4.7

[19]

in

good

agreement with the

expected

exponent 14/3.

There is

experimental

evidence for the

semidilute-unentangled regime

of concentration as well.

Typically a~/b

m

5, implying

that the

good

solvent limit

(r

=

I of

equation (36) predicts

the width of the

semidilute-unentangled regime

to be

roughly

a factor of ten in concentration,

(9)

in reasonable agreement with

experiment [19]

(see

Fig.

8 of Ref.

[19]).

Another group has

reported

that the

entangled scaling

for the

viscosity

in

good

solvent holds down to a concentration of four times the

overlap

concentration

[20]. However,

their estimation of the

overlap

concentration if a factor of three

larger

than the usual one

[21] (the reciprocal

of intrinsic

viscosity).

Thus we believe the

viscosity

data of reference

[20]

deviate from the power law for

entangled

solutions below a concentration of

roughly

twelve times the true

overlap

concentration,

in excellent agreement with

equation (36).

The rather limited extent of the

semidilute-unentangled regime

and the small amounts of data make

quantitative comparison

with exponents

pointless

at present.

Clearly

a detailed

study

of this concentration

regime

is needed.

Notice that a

typical

«

good

» solvent should not

correspond

to the true

good

solvent limit (r

= I means T~

oJ)

and therefore should have four

regimes

of concentration for the

viscosity (I,

II, III and III' in

Fig. I)

as observed

experimentally (see Fig.

8 of Ref.

[19]j.

The temperature

dependence

of osmotic

compressibility [22], plateau

modulus

[5],

and

viscosity [5]

have been measured

by

Adam and coworkers for

polystyrene

solutions in

cyclohexane

from the theta

point

to T

= + 25 K. We now make use of these data to test our

predictions regarding

temperature

dependence.

Scaling predictions

for the osmotic

compressibility [2]

(c dar/dc, where c is concentration and ar is osmotic

pressure)

conclude it is kT per correlation blob

(I,e.,

c d ar/dc m kT f ~). Thus

osmotic

compressibility

measurements can determine the blob size

(within

a

prefactor).

The

prefactor

can be eliminated

by taking

the ratio of osmotic

compressibilities

at theta and at some

(higherj

temperature.

Equations (22j

and

(27)

suggest the

following scaling

function.

(d&T/dC

)@ "3 I T

~ #i

m

#if/b

m

(37j

d&T/dC (T/#i )~~~~ T ~ #i

We therefore

plot #if/b

i>s. r/#i for the data of reference

[22]

in

figure

2. Above

r/#i

of 0,I the data

obey

a power law

[23].

#i

fib

=

0.57 I ° ~~ ~°°~

(38)

~b

These data are in excellent agreement with the

scaling prediction

[I] of

equation (37),

as also noted in reference

[22].

Below

r/#i

of 0, the data crossover to the

expected

theta behavior (#i

fib

m

I)

as the theta temperature is

approached.

We next compare the

plateau

modulus data of reference

[5] (in

the

semidilute-entangled regime)

with our

predictions.

While the osmotic

compressibility changed by

an order of

magnitude

between and + 25

K,

the

plateau

modulus is much less temperature sensitive,

changing by only

20 %. This is consistent with

comparisons

of

plateau

modulus in

good

solvents and theta solvents

[5, 19, 20],

which indicate that the

plateau

modulus does not

depend

on solvent

quality.

The

plateau

modulus data of reference

[5] actually

decrease

weakly

as temperature is increased

(the

data above

r/#i

m 0.I in

Fig.

8 of Ref.

[5] obey

a power law G

r~° '),

whereas we

predict

a weak increase with temperature

(see

Tab. I, G

r"'~

is

predicted).

However, these data are

actually

in

qualitative

agreement with our

theory,

as our

scaling

exponents may not be the

simple

rational powers we have

suggested.

For

example,

the exponent v in

equation

I is known to be 0.588 from renormalization group calculations

[24],

as

opposed

to the

3/5

we have used here.

Using

v

=

0.588 leads to an

expected

value of the

exponent

in

equation (38)

of

0.23,

in even better agreement with

experiment

than Il.

Similarly,

we cannot claim to know other exponents

(such

as the value of

x =

2/3 in

Eq. (6))

with too much

precision.

(10)

e . o

o.7

It

b

ig.

from smotic data [221 using the ~caling form of equation

(37).

Filled

symbols

for M = 3.84 x 10~ and ~ = 0.0 loo. Open symbols are data for

line is linear fit to ata r/~ ~

0.5

(Eq. (38)).

We can extract the apparent temperature

dependence

of the tube diameter

by combining

the

osmotic

compressibility

determination of the

screening length [22] (Fig. 2)

with the

plateau

modulus data

[5]

on the same system,

through equation (17). Equations (23)

and

(32)

suggest the

following scaling

form.

Ii l~

'~

m ~b ~~

alai

m

~~~

ji~ i1 (391

We therefore

plot

#i ~'~

alai

i>s. r/#i for the data of reference

[5]

in

figure

3,

using interpolation

and

extrapolation

of the data in

figure

2 for the

screening length.

Above r/#i of 0, I the data

obey

a power law

[23].

~b~~ala~

1.4

I

° ~~ ~°°~

~b

(40)

Exponents

of

v = 3/5 in

good

solvent and x

=

2/3 in theta solvent led us to expect a weaker temperature

dependence

of the tube diameter

(a

r

"'~, Eq. (39jj. Using

the renormalization group value

[24]

of v

= 0.588, we conclude from

equation (40j

that the

experimental scaling

for the tube diameter in theta solvent is a

#i°~~~°",

in excellent agreement with the exponent 2/3

(the

exponent for Tin

equation (40)

is

v/(3

v

Ii

-J. for

general

v and x). More data are needed at

higher

r/#i on the temperature

dependence

of

plateau

modulus in

near-theta

solutions,

as the data of reference

[5] only

include

r/#i

~ 2.

Given the success of

scaling

in

predicting

the temperature

dependence

of osmotic

compressibility, plateau

modulus,

screening length,

and tube

diameter,

we would

naturally

expect

equally good prediction

of the temperature

dependence

of

viscosity.

However, this is

far from the case. The temperature

dependence

of

viscosity [5]

in

semidilute-entangled

solutions demonstrate some fundamental flaw in our

reasoning.

Our

scaling theory predicts

(11)

2

o

Q ~

2/3

Qi

o

i

iO-2 10"1 io

'C/#

Fig. 3. Temperature dependence of the tube diameter for polystyrene in

cyclohexane

calculated from plateau modulus data [5] (M 6.77 x10~ and ~ = 0.0535) and the

screening length

(obtained from

Fig.

2

by interpolation

and

extrapolation

for the lowest r/~

point) using

the

scaling

form of

equation

(39). Solid line is a linear

regression

fit to data with r/~ ~ 0.07

(Eq.

(40)).

that both the relative

viscosity

and the terminal relaxation time in the

good

solvent

regime

should increase with

increasing

temperature

(7~/7~~~r""~, T~~~~r~'~). Experimentally,

however, both relative

viscosity

and terminal time decrease with temperature

by nearly

a factor of two between and + 25 K

(see Fig.

5 of Ref.

[5]j.

This result is itself a

paradox,

because in

semidilute-entangled

solutions the

good

solvent limit

clearly

has a relative

viscosity

that is

slightly higher

than in theta solvent

[5, 20],

as

predicted by

our

simple scaling theory.

Apparently,

instead,of the monotonic increase in relative

viscosity

that we

expected,

the relative

viscosity

goes

through

a minimum as a function of concentration.

Where is our

reasoning

flawed ? Static

quantities (osmotic compressibility

and

plateau modulus)

allowed us to calculate the two static

length

scales in the

problem (screening length

and tube

diameter).

The temperature and concentration

dependence

of these

length

scales agree very well with our

simple theory

and

they

both seem to crossover from the theta

regime

to the

good

solvent

regime

at the same

point (r/#i

m 0,I, see

Figs.

2 and

3).

Our calculation of

dynamic quantities (terminal

relaxation time and

viscosity) required

these two

length

scales, but also

required assumptions regarding dynamic scaling.

The

hierarchy

of relaxation times

(Eqs. (15)

and

(16)) might

be suspect, but such

reasoning

is at the heart of molecular theories of

polymer dynamics [10],

and these theories seem to have been well-tested.

We also assumed the

hydrodynamic screening length

behaves in the same manner as the

static

screening length.

While it seems as

though

these

lengths

have the same concentration

dependence

in both

good

solvent and theta solvent, it is not

necessarily

true that

they

are identical. In

particular, they

may well crossover with temperature

differently.

In

fact,

other

crossovers between

dynamic

and static

quantities (such

as the coil size and

hydrodynamic

size

in dilute solution as functions of chain

length [25])

are known to be different. However, to

explain

the

pronounced

minimum in the temperature

dependence

of

viscosity,

the

hydrodyn-

amic

screening length

would need to have a nonmonotonic temperature

dependence (I,e.,

a

strong

maximum),

There is no reason to believe such a maximum should exist.

Certainly

data

(12)

in dilute solutions show no

pronounced

differences between the temperature

dependences

of radius of

gyration

and

hydrodynamic

size

[26].

Clearly

more data are needed for the temperature

dependence

of

viscosity

in near-theta solutions. In

particular,

it would be nice to have data in a

high-boiling

theta system, such as

polystyrene

in

dioctylphthalate [27],

which would allow for a much wider temperature range to be covered,

possibly getting deep

into the

good

solvent

regime (well beyond

the minimum in

viscosity

vs.

temperature).

Conclusions.

We have

presented

a two-parameter

scaling theory

for the concentration and temperature

dependences

of

rheological quantities

such as

plateau

modulus, relaxation

time,

and

viscosity.

The presence of two

independent length

scales

(screening length

f and tube diameter

a)

implies

six

regimes

of behavior

(see Fig, I).

Good solvent

scaling applies

at

high

temperature and theta solvent

scaling applies

at temperatures close

enough

to H. At low concentration there

are dilute

regimes (I

and

I')

where R

~ f. At intermediate concentration there are semidilute-

unentangled regimes (II

and

II')

where f ~ R ~ a. At

high

concentration there are semidilute-

entangled regimes (III

and

III')

where R

> a.

The concentration

dependences

of all

predicted rheological quantities

in the semidilute-

entangled regime

appear to be consistent with

experiment. Using

data for osmotic compress-

ibility

and

plateau modulus,

we were able to

verify

that the

expected

temperature

dependences

of our two

length

scales

(f

and

a)

are observed.

However,

the

viscosity

and terminal relaxation time exhibit nonmonotonic temperature

dependences

that are

qualitatively

different from our

simple scaling predictions.

The

origin

of this

discrepancy

is unknown. It is worth

noting

that similar difficulties are observed in other

polymer problems.

An

example

is

gelation,

where static

properties

such as molecular

weight

distribution

[28], gel

fraction

[28], swelling [28, 29]

and modulus

[29, 30]

are

well-understood,

but

dynamic properties

such as

viscosity [30] (and

relaxation

time)

are not.

References

[ii Daoud M., Jannink G., J. Phys. Franc-e 37 (1976) 973.

[2] de Gennes P. G., Scaling Concepts in Polymer

Physics

(Cornell University Press, Ithaca, 1979).

[3] de Gennes P. G., Macromolecules 9 (1976) 587 and 594.

[4] Colby R. H., Rubinstein M., Macromolecules 23 (1990) 2753.

[5] Adam M., Delsanti M., J. Phys. Fiance 45 (1984) 1513.

[6]

Roy-Chowdhury

P., Deuskar V. D., J. Appl. Polym. Sci. 31(1986) 145.

[7] Edwards S. F., Proc.

Phys.

Soc.. 88 (1966) 265.

[8] Daoud M., Cotton J. P., Famoux B., Jannink G., Sarma G., Benoit H.,

Duplessix

R., Picot C., de Gennes P. G., Macromolecules 8 (1975) 804.

[9] de Gennes P. G., J. Chem. Phys. 55 (1971) 572.

[10] Doi M., Edwards S. F., The Theory of Polymer Dynamics (Clarendon Press, Oxford, 1986).

[I ii Flory P. J., Principles of Polymer Chemistry (Comell University Press, Ithaca, 1953).

j12] Famoux B., Baud F., Cotton J. P., Daoud M., Jannink G., Nierlich M., de Gennes P. G., J. Phys.

Franc-e 39 (1978) 77.

[13] Zimm B. H., J. Chem. Phys. 24 (1956) 269.

[14] Rouse P. E., J. Chem. Phys. 21 (1953) 1272.

[15] Shiwa Y., J. Phys. II France 3 (1993) 477.

[16] Graessley W. W., Polymer 21 (1980) 258.

[17] Pearson D. S., Rubb. Chem. Tech. 60 (1987) 439.

(13)

[18] Colby R. H., Rubinstein M., Viovy J. L., Macromolecules 25 (1992) 996.

[19] Colby R. H., Fetters L. J., Funk W. G., Graessley W. W., Macromolecules 24 (1991) 3873.

[20] Adam M., Delsanti M., J. Phw. Fiance 44 (1983) l185.

[21] Graessley W. W., Adi,. Polym. Sci. 16 (1974) 1.

[22]

Stepanek

P.,

Perzynski

R., Delsanti M., Adam M., Macromoleiules 17 (1984) 2340.

[23] While our theory naively expects the crossover to occur at r/~

m I, scaling

actually

cannot be

expected

to

predict

the crossover

point

exactly. Note that, as expected, all quantities cross at

roughly the same point (r/~ m 0.1).

[24] Le Guillou J. C., Zinn-Justin J., Phy.I. Rev. B 21 (1980) 3976.

[25] Weill G., des Cloizeaux J., J. Phy.I. Fiance 40 (1979) 99.

[26] Miyaki Y.. Fujita H., Macromolecules 14 (1981) 742.

[27] Park J. O., Berry G. C., Macromolecules 22 (1989) 3022.

[28] Colby R. H., Rubinstein M.. Gillmor J. R., Mourey T. H., Macromolecule-I 25 (1992) 7180.

[29] Rubinstein M., Colby R. H., Macromolecules 27 (1994) 3184.

[30]

Colby

R. H., Gillmor J. R., Rubinstein M., Phys. Rev. E 48 (1993) 3712.

Références

Documents relatifs

increase of the concentration exponent value with the reduced concentration c/c* and a molecular weight exponent, bigger than 3, which increases with

2014 A theoretical analysis of the order parameter and temperature dependence of the complete set of five independent viscosities of incompressible nematic

The quality of the solvent should have no influence on the viscosity molecular weight exponent, but it must influence strongly the.. concentration

The important feature of the dynamics of semidilute polymer solutions is that the presence of many polymer chains leads to a screening of the hydrodynamic interaction between

For example, in the concentration range at which solvent prevails (C « I) but polymers do overlap, I-e- in semi-dilute solution, a universal behavior is theoretically expected for

In this paper, the temperature variations of mag- netic susceptibility and electrical resistivity, the ther- mal expansion and the pressure effect on the NBel tem-

have a better understanding of the physics of swollen gels than of that of dry rubber, we will calculate the swelling activity parameter for conditions corresponding to the

Mean square end to end distance, screening length and osmotic pressure in the different regions of the phase diagram. Results for three dimensions are