• Aucun résultat trouvé

A mean-field study of a model system of grafted rods

N/A
N/A
Protected

Academic year: 2021

Partager "A mean-field study of a model system of grafted rods"

Copied!
15
0
0

Texte intégral

(1)

HAL Id: jpa-00212457

https://hal.archives-ouvertes.fr/jpa-00212457

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.

A mean-field study of a model system of grafted rods

Zhen-Gang Wang

To cite this version:

(2)

A

mean-field

study

of

a

model

system

of

grafted

rods

Zhen-Gang

Wang

(*)

Corporate

Research Science Laboratories, Exxon Research and

Engineering Company,

Annan-dale, NJ 08801, U.S.A.

(Reçu

le 4 décembre 1989,

accepté

sous

forme définitive

le 26

février 1990)

Abstract. 2014 We

study

the

phase

behaviour of a model system of

grafted

rods on an

impenetrable

surface,

using

a van der Waals-like mean-field

theory. By accounting

for the

anisotropy

of the

in-plane

excluded volume effects, it is shown that the

in-plane

orientational symmetry can be broken for

adsorption

energy

exceeding

some critical value, in a

density

interval, thus

leading

to a

discontinuity

in

(the

deivative

of)

the fraction of molecules in each orientation. It is further

demonstrated,

using

a van der

Waals-type theory,

that the inclusion of attractive interactions,

when

coupled

with the

anisotropic

excluded volume

effects,

may drive the transition to first order,

in addition to

inducing

the usual

gas-liquid

transition.

1. Introduction.

Molecular

monolayers/films

at

gas-liquid

and

liquid-liquid

interfaces are of

great

current

interest because of their numerous

technological applications,

such as molecular

electron-ics

[1]

]

and

biological

membranes

[2],

and have been the

subject

of extensive studies

[3],

thanks to the advent of a number of

newly developed

surface

techniques

such as

synchrotron

X-ray

diffraction

[4]

and second harmonic

generation [5].

From a

physical point

of

view,

it is

important

to understand

the

collective

phenomena

exhibited

by

these molecular

monolayer

systems

[6-9],

which in turn are

intimately

related to

the various

applications.

Of

particular

interest are the successive

phase

transitions that have

been observed in these

systems

and the molecular conformation in each

phase.

There,

a

key

issue has been the nature of the

liquid-expanded (LE)

to

liquid-condensed (LC)

transition. The richness in the behavior of the molecular

monolayers

arises from the

quasi-two-dimensional nature of these

systems,

i.e.,

from the subtle

coupling

between the translational

degree

of freedom and other « extra »

degrees

of freedom of the molecules normal to the

layer

[10].

A number of statistical mechanical

theories,

including

lattice

packing

and

Ising

Hamiltonian

formulations,

have been

developed

to account for the orientational

ordering

in the

monolayers

[ 1 1 ] .

Clearly

the transition from the LE

phase

to the LC

phase

must be associated with the

(*)

Present Address :

Department

of

Chemistry, University

of California at Los

Angeles,

405

Hilgard

Avenue,

Los

Angeles,

CA 90024, U.S.A. Classification

Physics

Abstracts 64.70M

(3)

coupling

between the translational

degree

of freedom

(density)

and the internal

degrees

of freedom or some

symmetry

order

parameter.

For

fully

flexible

chains,

a lattice model has

been

proposed

and

studied,

using

statistical

thermodynamics

and Monte Carlo simu-lation

[12].

More

recently,

a similar model has been studied

by

Cantor and

Mcllroy[13].

Their

theory,

which enumerates the isomeric states for a

pair

of very short chains

(2-4 units),

better describes the

pair

interaction out of the

grafting

surface,

and is able to

predict

a second

phase

transition,

apart

from the

gas-liquid

one. The appearance of a second

liquid phase

is

also

predicted

in a recent

simple

mean-field calculation

by

Shin et al.

[14].

Most

amphiphile

molecules in the

monolayer

are not

long

giant

molecules,

and therefore cannot be viewed as

completely

flexible. A more realistic treatment would involve full consideration of the molecular

packing [10,

13-16]

and finite stiffness of the molecules.

However,

we feel that a

study

of the

rigid

rod

problem,

which is at the other extreme of the

flexibility

spectrum

[17],

should be

elucidating.

Such a

study

is also

important

in view of the

corresponding

bulk

liquid-crystal

system

which shows

isotropic

to nematic

phase

transition [18,

19].

The

grafted

rods

problem

was

recently

studied

theoretically by Halperin et

al.

[20].

Using

a

second virial

expansion (Onsager theory) [18], they numerically

calculated the order

parameter

as a function of

particle density

in the absence of

adsorption

energy. It was

concluded there that an orientational

phase

transition does not ensue, and

they

attributed this nonexistence of orientational transition to the

already-broken

symmetry

imposed by

the surface.

More

recently,

Chen et al.

[21]

consider the case where

adsorption

energy is included. Via a

simple Onsager

treatment of the excluded

volume,

and

using

symmetry arguments,

they

conclude that

anisotropic

excluded volume

alone,

even with surface

adsorption

energy, is not sufficient to

bring

about an orientational

phase

transition.

They

suggest

that attractive

interactions between the rods are

required

for such a transiton to occur. Since the orientation

of the rods is

intimately coupled

to the

density

in the

grafted

rod

problem,

a

strong

rod-rod

attractive

interaction,

which induces the usual

gas-liquid

transition below the critical

temperature,

necessarily

leads to a discontinuous

change

in the orientational order

parameter.

In reference

[20],

the effect of

anisotropy

of the excluded volume in the azimuthal

angle

cp

is

neglected,

thus the

possibility

of a

symmetry

breaking

in that

degree

of freedom was

precluded

from the calculation. This

possibility

was first

pointed

out

by

Boehm and Martire

[22]

in the

study

of

rigid

rod

adsorption

and was also discussed in the work of Chen et al.

[21]

]

but it was not included in their

phase diagram. Neglect

of the

biaxiality

of the molecules is

equivalent

to

performing

an average of the azimuth

angle

in the

(biaxially)

symmetric phase,

which can be

justified

when there is no

adsorption

energy.

However,

explicit incorporation

of this

angle

becomes

important

when the

adsorption

energy is

sufficiently

strong

because the

symmetry

can be broken in this

degree

of

freedom,

as will be

shown in this paper. This

symmetry

breaking

in turn leads to a discontinuous

change

in the

vertical

angle

or its derivative. This rotational

degree

of freedom is also

expected

to be

partially responsible

for the tilt

phenomena [5, 23]

in some surfactant

monolayer

systems.

Thus it is desirable to take proper account of the

anisotropy

of the excluded volume in the

azimuth direction. In this paper, we conduct a mean-field theoretical

study

of the

grafted

rods

by incorporating

this

degree

of freedom. For the sake of

comparison,

we

adopt

the

Zwanzig

approximation [24] employed

in the work of Chen et al. The role of

in-plane

anisotropy,

adsorption

energy, and attractive interaction between the

rods,

are

explicitly

taken into

consideration.

The paper is

organized

as

following :

in section

2,

we

develop

a van der Waals-like

(4)

volume

part,

we use, instead of scaled

particle theory [27]

which is rather

sophisticated

and

subtle

[28], especially

when

applied

to the

grafted

system,

a

simpler

and more

transparent

theory, namely

the van der Waals

equation

of state for a

multi-component

fluid

[291.

This

general theory

is then

applied

to the

Zwanzig

model in which

only

orientations

along

the

principal

axes are allowed. In section

3,

we

perform

a bifurcation

analysis

to

study

the

possibility

of a

symmetry

breaking

in the

in-plane

directions,

for the case where there is no

attractive interaction between the rods. The

importance

of the

one-body adsorption

energy is demonstrated and its critical value for the

in-plane

symmetry

breaking

is determined. Both

Onsager theory

and the mean-field

theory developed

in section 2 are

employed

to

study

the order

parameters

as a function of the

monolayer density.

In section

4,

we

investigate

the

effects of the attractive

pair

interaction between the

rods,

and show that in addition to the

gas-liquid

transition,

such an attractive interaction can drive the

in-plane

symmetry

breaking

to a

first order

transition,

thus also

resulting

in a discontinuous

change

in the other orientation

angle. Finally,

results of the

theory

are discussed in section 5 in its relevance to

experiments,

and some

concluding

remarks are made.

2. Van der Waals-like mean-field

theory.

We first

develop

the

theory

in a more

general

form

applicable

to any

anisotropic

fluids.

Only

in the end of this section do we

specify

to the

grafted

rods and the

Zwanzig

approximation.

Consider a system of N molecules in a

(d-dimensional)

volume at

temperature

T. For the sake of

discussion,

we assume that these molecules can take m discrete

orientations,

with

Nj, j =

1,

..., m, molecules in each orientation. The continuum limit may be obtained

straightforwardly by letting

m - oo.

Following

Gelbart and Baron

[25]

and Cotter

[26],

we

write the Helmholtz free energy per molecule

F/N

as

where

{3

=

1 /kT

and

QN

is the canonical

partition

function. In

equation (1),

p is the total

particle

density, sj

=

N j / N

is the fraction of molecules in orientation wj,

u (r, wj,

lù k)

is the

(long-ranged)

attractive interaction

potential

between two molecules in states wj and W k, which are

separated

in space

by

r, and

g (r,

’W j,

W k)

is the

pair

distribution function. The

first two terms

represent

the translational and rotational

entropy

of the

rods,

respectively.

The last term in

equation (1)

contains the

short-ranged strongly repulsive potential

U*(r, w j’, w k)

which we take to be hard core interaction.

Then,

the interaction

part

of

equation (1) (i. e.,

the last two

terms)

can be

regarded

as the lowest order

approximation

in the Barker-Henderson

perturbation theory

for

liquids [30].

We have included a

one-body

potential

h (,wj)

which can

be,

e.g., the

adsorption

energy, when a surface is

present.

We now introduce the

following approximation,

following

Gelbart and Baron

[25].

We

(5)

when two molecules with orientation wj and cv k,

respectively,

do not

overlap

and,

otherwise. We note that the

equations (2a)

and

(2b)

are the lowest order term in the

density

expansion

of the

pair

correlation function for a hard-core

fluid,

which takes account of the

two

particles

involved while

neglecting

the effect of the other molecules

[30].

In the same

spirit,

we make the

approximation

for the last term in

equation (1)

that,

where

bjk

is the van der Waals

pair

excluded volume

parameter

and is defined as

with the

prescription

(2a)

and

(2b)

for

g (r,

W j,

’W k).

The

approximations (2) through

(4)

are

nothing

but the van der Waals

approximations

for a

multi-component

fluid with

compositions

si, s2, etc.

[29,

30].

With these

approximations, equation

(1)

now

becomes,

where

The continuum limit can be

easily

obtained from

equation (5)

as

Each term in

equation (5),

or

equation (7),

has a

transparent

physical interpretation :

the first term is the translational

entropy,

the second term is the

entropy

loss due to the excluded

volume,

the third term

represents

the orientational

entropy,

and the last two terms arise from the two

body

attractive interaction and one

body

external

potential, respectively.

Notice that in the absence of the last two terms,

equation (7),

when

expanded

to the linear order in p, reduces to the

Onsager theory [18]

for

isotropic-nematic

transition in

liquid-crystals.

Thus

equation (7)

is

quantitatively

correct to order p, since it

yields

the correct second virial coefficient. It should be

kept

in

mind, however,

that our estimate of

ajk

in

equation (5) (or

equivalently

of

a(w,

w’)

in

equation (7)) by using equation (2),

is a rather crude

(6)

theory

still

captures

some of thé

important qualitative

features of the

system,

such as its

symmetry

properties.

Equation

(5),

or

(7)

gives

the free energy as a functional of the distribution

sj(s(w)).

The

equilibrium

values for

these sj ’s

are obtained

by minimizing f

with

respect

to sj ’s

at

given

density

p and

temperature T,

for the set of

parameters

characterizing

the

system.

This is

achieved

by setting

These

equilibrium

Si*’s,

when substituted back to

equation

(5),

yield

the

equilibrium

Helmholtz free energy

f*.

Hereafter we will ommit the *

symbol

for the sake of notational

simplicity,

with the

understanding

that the

quantities

to be discussed are

already

their

equilibrium

values.

Using

standard

thermodynamic

relations,

the

(osmotic)

pressure lI and chemical

potential

» can be derived as

and

If there is

two-phase

coexistence,

the pressure and chemical

potential

of

phase

a, must

equal

the pressure and chemical

potential

of the

coexisting phase

8,

i.e.,

Equations (5), (8), (9)

and

(10) completely

determine the

equilibrium

distribution

sj’s

and the

thermodynamics

of the

system. Next,

we

specify

to the discrete model to be studied in what follows.

We use the

Zwanzig approximation [24],

where the molecules are assumed to take

orientations

along

the

principal

axes, which has been

employed

in the

study

of bulk

isotropic-nematic transition of

liquid-crystals.

Such a model is

clearly

the

simplest

of

all,

and at the

same time retains most of the

phenomenologies. Comparisons

with

experimental

data even

yields quantitative

agreement

to certain extent

[31]. Although

when

applied

to the

grafted

rod

[31]

problem

some subtleties remain

(see

discussion

below),

the model is

expected

to catch many of the

important

features of the

problem,

especially

the

anisotropy

in the excluded volume between the various

orientations,

the

breaking

of the

symmetry

in the direction

perpendicular

to the surface and the role of

adsorption

energy. The

theory

developed

above,

however,

does not limit us to this

particular

model,

and a

study by using

a continuum model is

straightforward.

We assume the molecules to be

rectangular

parallelipipeds [21]

]

of

length f

and width

d,

with one end

always

anchored on the

impenetrable

surface,

which can take five

orientations : the directions of z

(perpendicular

to the

surface),

x, - x, y

and - y (in-plane).

The distinction between the orientations x and - x, and

likewise y

and - y, does not

play

a

crucial role in this

particular

model,

where the

in-plane

conformations have the molecules

lying

down

completely

on the surface. It is

essential,

however,

in real

systems

and when focus

is in the

in-plane

orientation itself. Our main concern in this paper is to

study

the

phenomenology

when

symmetry

breaking

is allowed in the

in-plane

direction and the

coupling

between the

in-plane

and vertical orientations as a function of the

density.

The

(7)

of the

grafted

rod

problem

and is absent in the bulk

counterpart.

For this purpose, we

ignore

the subtle differences

afore-mentioned,

and our crude model suffices.

(We

treat x and

- x orientations as the same,

except

for an

entropy

contribution of In 2 associated with two

orientations as

opposed

to

one.)

The excluded volume

(area)

matrix

bjk

between various orientations can be

easily

calculated to be

In order to

explicitly distinguish

between the

in-plane (horizontal)

and

out-of-plane

(vertical)

orientations and to allow the

possibility

of a

symmetry

breaking

in the horizontal

directions,

we define the order

parameters .0

and 6 such

that,

Sz = 1 - l/J, Sx = l/J (1 + u) /2,

sy = 0

(1 - u) /2.

Clearly, 0 =

S x + SY’ which is the total fraction of horizontal

molecules,

whereas a nonzero

signifies

a

symmetry

breaking

in the horizontal orientations. With these

order

parameters,

equation (5)

becomes,

where

B(0, o-)

is defined as

p =

pMo/2

is the reduced

density

with

Mo

=

4 d2, u

=

azzlMo,

and we have used reduced

parameters

Bjk

and,

Ajk

defined as

Using (11),

the

Bjk’S

are found to be

with r =

l /d

being

the

aspect

ratio. The

A jk’S

will be

specified

in section 4.

From

equation (8),

the

equations determining

the

equilibrium

values

of 0

and

(8)

3. Bifurcation

analysis

and the critical

adsorption

energy.

Equations (16a)

and

(16b)

determine the values

of ~

and o- as functions of the

temperature

T and

density

p. Because of the

larger

excluded volumes

experienced by

the

in-plane

molecules,

0,

which is the fraction of

in-plane

molecules,

is

expected

to decrease as the

density

is increased. As

p --> 1, ~2013>0.

When

p --> 0, ~

approaches

a finite value

~0

determined

by

the

adsorption

energy. Thus in both of these

limits,

the

density

of

in-plane

molecules

p ~

vanishes. On the other

hand,

a nonzero solution of 6 is

possible only

when the

density

of

in-plane

molecules is

sufficiently large.

In

fact,

such a transition is

expected

to take

place

at a surface coverage of

in-plane

rods of the order

1/ l2 [32].

Therefore it follows that

only

when the

adsorption

energy is

sufficiently large

do we

expect

a

symmetry

breaking

in the

in-plane

orientations. In the

following,

we use bifurcation

analysis

to determine this critical

hc

in the absence of attractive interaction.

Although

it is desirable to discuss the behavior of the

system

in terms of the total

density

(or

surface

coverage)

of rods p and the

density (or

surface

coverage)

of

in-plane

rods Pin, the latter is related to p

and 0

simply

by

P in =

P Q .

We

prefer

to use p and the fraction

of in-plane

rods 0,

based on consideration of

mathematical

convenience, since

decreases

monotonically

with

increasing density

and is thus more amenable to the

subsequent analysis.

Clearly u

= 0 is

always

a solution of

equation

(16b).

To locate the

point

where a nonzero

solution becomes

possible,

we

expand equation (16b)

to linear order in a,

where the

order o, 2term

is

identically

zero

by

symmetry.

Bifurcation takes

place

when the coefficient of a

vanishes,

which

yields,

On the other

hand,

from

equation

(16a),

for

The emergence of a nonzero is thus tantamount to the existence of a solution

(p-,

0 )

to the simultaneous

equations

(18)

and

(19)

in the

physical

domain

p

E

(0,

1 )

and

l/J

e

(o, Q o). (In

the absence of attractive

interactions, 0

is

monotonically decreasing

with

p,

so Q o

is the maximum value

that qb

can

attain.)

Substituting

p/(1 2013

p B )

from

equation (18)

into

equation (19),

we obtain an

equation

for

0 only,

or,

where the definition of G

(0 )

is obvious from

equation

(20).

Equation

(20)

or

equation (21)

is thus the determinant condition for the emergence

of

’nonzero 0-.

The function

G ( Q)

is

(9)

Fig.

1. -

Curve 1

gives

the maximum value

of 0 (the

fraction of

in-plane molecules),

i.e.,

r6o

as a function of the

adsorption

energy

6h.

Curve 2 is the

graphical representation

of the relation

G (Q ) + Bh = 0 (Eq. (21 )).

Curve 3 shows the limit of

using

the

Onsager expansion by setting

p = 1 ; thus the

region

below Curve 3 cannot be accounted for

by

the

Onsager expansion.

where

G min

is the minimum value of G in the interval

(0, Q 0).

The

critical Oc

and hence

G min

can be found

by letting

ag/aQ =

0,

from which we

obtain,

and

The critical

adsorption energy 6 hc

is thus seen to be

or

For the

rectangular parallelipipeds

with

aspect

ratio r,

equation (25)

is

simply,

When - Bh > - BhB c’ equation (20)

has two solutions

0

1

and ~ b corresponding

to the

emerging

and

vanishing, respectively,

of the nonzero 6. In this case,

then,

an ordered

phase

in the

in-plane

orientation exists

between 0

1 and 02,

or

Fi

and

p 2,

the latter to be

computed

from

equation (19).

That

both .0

1

and 2

and hence

p

1 and

P2,

are within the

physical

domain, i.e.,

0 0

>

(~)

1,

~

2)

>

0,

when -

3 h

> - {3h c’

is ensured

by equation (20),

since the

first term there is

always positive.

Most of the above

arguments

apply equally

well to the

Onsager theory

where

only

the second virial coefficient is retained. In

particular,

it can be shown that

equations (20)-(25)

remain

unchanged.

However,

since the

expansion

is

only

up to order

fi,

within the range

p E

(0, 1 ),

~

cannot decrease below some

minimum ~min

which is obtained from

(10)

equation (19)

may be cut out of the

physical density

range in the

Onsager

approximation,

for

sufficiently large - 8 h (see Fig. 1).

We have carried out a numerical calculation

of 0

and a as functions of the

density

p at T = 300 K. For

aspect

ratio r =

10, -l3hc

=

22/9

+ In

(2/9)

from

equation (24) ;

thus

when - 8 h

=

2,

an ordered

phase

is

expected

to exist in a

density

« window ».

Figure

2 shows

the

0-j5-

and a-fl

plots

for this set

of parameters.

The full curves are the results from

using

our van der Waals

type

mean-field

theory,

and the dash curves are results from

using

the

Onsager

theory.

Both theories show the existence of an ordered

phase

in an intermediate

density

range. It can be seen

that,

associated with the transition in the order

parameter

o-,

0

has discontinuous

changes

in its derivatives at both boundaries.

Analysis

of the

pressure-density

77-pr

relation and the chemical

potential

suggests

that the transition at both boundaries is second since there is no van der Waals

loop.

Fig.

2. - Order

parameters r6 ,

the fraction

of in-plane

molecules, and a, the

in-plane

symmetry order parameter

(see

the lines below

Eq. ( 11 )

for their

definition)

as functions of the reduced

density

p for r = 10 and

6h

= - 2, in absence of attractive interactions.

(-)

from van der Waals

theory ;

(- - -)

from

Onsager approximation.

4. Effects of the attractive interactions.

It is well known that attractive interactions between

particles

induce a

gas-liquid phase

transition below the critical

temperature.

For

simple liquids,

such a

transition,

as well as the critical

point,

can be studied within the framework of various mean-field

theories,

such as van

der Waals

theory

and

Bragg-Williams approximation [33].

For more

complex

systems,

e.g. a

monolayer

of adsorbed

long

surfactant or

polymer

molecules,

the mean-field

theory

becomes

more involved because of the extra

degree

of freedom

(rotation,

for

example)

and of

possible

inhomogeneity

in the

density.

For flexible surfactant or

polymer

molecules,

usually

a

self-consistent

procedure [6,

12(a)]

is

required.

For the

simple

model considered

here, however,

we can use the mean-field

theory

developed

in section 2 of the paper.

In the discussion that

follows,

we use the letter G for the gas

phase

and L for the

liquid

phase.

To make distinction between a

symmetric

(with

respect

to the

in-plane orientations)

and a

symmetry-broken phase,

we attach the letter S for the former and B for thé latter. Thus

in the full parameter space, four different

phases

are

possible,

at least

mathematically.

These

(11)

S/B

boundaries and

p G

and

PL

the

density

for the

coexisting

gas and

liquid phase,

respectively,

then

mathematically,

the

following possibilities

exist,

with their

corresponding

sequence of

phase

transitions :

In

practice,

however,

not all of the above

possibilities

are realizable for

physically

reasonable

parameters.

The

analysis

in the case where attractive interactions are

included,

proceeds

in much the

same manner as the last section. We first consider the situation where the interaction is

independent

of the orientations. In this case,

A,,

=

A xx

=

A xy

=

0,

and

the ~ - P

relation

and the values

of p

1 and p 2

remain

unchanged.

However,

if the attractive interaction is

strong

enough,

then a first order transition from a gas

phase

to a

liquid phase

takes

place.

As in any

mean-field

theories,

a Maxwell construction is

required

in this case to eliminate the

thermodynamically

unstable

region

and locate the values of the densities at coexistence

[33],

j5-G

and

pL’

Figure

3 shows the results for T = 300

K, r

=

10,

u = - 1 000 K and

h = - 600 K. This

corresponds

to the second

possibility

mentioned above.

Figure

3 shows

that,

associated with the discontinuous

change

in the

density

at the

transition,

the order

parameters 0

and a also

change discontinuously.

The

jump

in

(

at the

gas-liquid

transition has been

pointed

out in the work of Chen et al.

Figure

4 shows another situation for the same

temperature

and

aspect

ratio but with h = - 450 K and u = - 850

K ;

it

corresponds

to

possibility (1).

In both of these cases, the transition between the

symmetric

and

symmetry-broken

phases

remains second

order,

except

when it is overtaken

by

the

gas-liquid

transition,

in which case it becomes first order.

Fig.

3.

- Isotherms ~ -

II, 6 - II and p - II for r = 10, B h = - 2, u = - 1 000 K and T = 300 K.

There is no

anisotropy

in the attractions. The

regions

I, II and III

correspond respectively,

to the GS,

(12)

Fig.

4. -

Isotherms r6 -

II, 6 - II and p - II for r = 10, {3h = - 1.5, u = - 850 K and T = 300 K.

Again

there is no attraction

anisotropy. Regions

I, II, III and IV are

respectively

the GS, GB, GS, and LS

phases.

It is of interest to

investigate

the effects of

anisotropic

attractions. To this

end,

we make the

following

ad hoc choices for the

Ajk’S:

Figure

5 shows the results for T = 300 K and the same

aspect

ratio as the other two. An obvious feature in these

results,

is

that,

besides the

gas-liquid

transition,

the transition GB to

GS becomes

discontinuous,

as

opposed

to the

isotropic

case. In this case we have the

following

sequence : GS --> GB ==> GS =>

LS,

where we have two discontinuous

changes

(denoted by

the =>

symbol)

in the

density

and the order

parameters.

Fig.

5.

- Isotherms r6 -

II, o- - II and p - II for r = 10,

{3h

= - 2, u = - 1 000 K, T = 300 K and

(13)

Discussions of the results of

figures

3,

4 and 5

depend

on the

validity

of

using

a

density-independent

ajk

or

a(w,

w’) (see

Eq. (5)

or

Eq.

(7))

which can be

put

into doubt. The

phenomenology pointed

out,

however,

can be

expected

on

physical

considerations,

and

should not be affected

by

the drawbacks of the

approximations

we make.

Results in this section

suggest

that,

attractive interactions can induce a

gas-liquid

transition,

which in turn leads to discontinuous

changes

in the orientational order

parameters.

Moreover,

when

anisotropy

in the attractive interactions are

allowed,

it is

possible

to have another first order transition. This latter transition is

mainly

orientational in nature. At this

juncture,

it is

interesting

to compare the behavior studied here and the LE to LC

phase

transition observed in

typical

surfactant

monolayer

systems :

both situations can exhibit two

successive first order

transitions,

one of which is associated

mainly

with conformational

changes.

In this

study

it results from a

symmetry

breaking coupled

with orientation

dependent

attractive

interactions,

whereas in the LE to LC

transition,

the effect of conformational

packing

is

suggested

to be

important [10, 13-16]

in addition to the orientational

change.

5. Discussion.

In this paper, we

study

a model

system

of

grafted

rods

using

a mean-field

theory. Emphasis

is

placed

on the

importance

of the orientational

degree

of freedom in the azimuth

angle

in

determining

the

phase

behavior of the

system.

This

complements

earlier studies which have

not taken this

degree

of freedom into consideration. It is shown that when the

adsorption

energy is

large enough,

a

symmetry

breaking

transition in the afore-mentioned orientation can take

place,

which

necessarily

leads to a discontinuous

change

(second

order)

in the other

orientation

angle.

Furthermore,

it is shown that attractive

interactions,

can lead to a

jump

discontinuity

in the order

parameters,

which was

pointed

out

by

Chen et al.

[21],

and in some

cases

(anisotropic interactions)

may also cause another first order transition besides the

gas-liquid

transition.

It is

interesting

to note that the

possibility

of an orientational

phase

transition for

grafted

rods when

adsorption

energy

(which

favors the flat

orientations)

is

included,

was

suggested

in

the work of

Halperin et

al.

[20].

They

first attributed this

possibility

to the

weakening

of the

alignment by

the

impenetrable

surface. In a more recent

discussion, however,

these authors

point

out that such surface

potential

have the effect of

constraining

the rods to lie

parallel

to

the

surface,

and hence

causing

a

symmetry

breaking

in the

in-plane

orientations at some

appropriate

densities.

Although

the

potential proposed

in reference

[32]

is

slightly

different in form than the

adsorption

energy

employed

here,

the effect of such a

potential

in the two

studies is the same,

namely

to favor the

in-plane

orientations. In the

preliminary study

presented

in this paper, we find no

alignment phase

transition if the

symmetry

breaking

in the

azimuth directions is

neglected,

thus

confirming

the

symmetry

argument

given by

Chen et al. Our

study employs

a discrete

model,

where

only

three orientations are allowed. While this

kind of

approximation

is

commonplace

in

studying

bulk

systems,

a few comments should be made in

applying

it to the

grafted

system.

Such an

approximation

makes sense

only

if the

fraction of molecules in each direction is

comparable

to each other. This is

possible

in the

grafted

rod

problem

if there is

(1)

a

strong

surface attraction that favors the

flat-lying

orientation or

(2)

other

energetic

factors,

e.g.

bending

energy, that

strongly

confine the

molecules to orient at some

prefered 0-angle.

The second

possibility

arises,

for

example,

in

systems

of

fatty

acid molecules at water-air

interface,

where the -COOH group, with the

hydrophilic

part

immersed in water, orients the

-CH2-

tails. This is

probably

one of the factors

in the tilt transition observed in these

systems.

Another

experimental

realization,

of which the

theory

presented

in this paper is

expected

to be a more faithful

representation,

is to use

rigid

(14)

observe the

breaking

and

restoring

of the

in-plane

symmetry.

A more direct test of the

predictions

given

in this paper can be done

by performing

Monte Carlo simulations of the

same model

[35].

Since the

general

mean-field

theory developed

in this paper is not limited to

the discrete

approximation,

the continuum can also be studied

[36, 37].

Acknowledgments.

1 thank T. A.

Witten,

A.

Halperin

and M. W. Kim for

helpful

discussions,

and W. M. Gelbart for

sending

me a copy of their work

(Ref. [21]) prior

to

publication.

1 am

grateful

to W. M. Gelbart for his critical

reading

of the

manuscript

and some valuable

suggestions.

References

[1]

LARKINS G. L., THOMPSON E. D., DEEN M. J., BURKHART C. W. and LANDO J. B., IEEE Trans.

Magn.

19

(1983)

980.

[2]

PHILIPS M. C. and CHAPMAN D., Biochim.

Biophys.

Acta 163

(1986)

301.

[3]

SWALEN J. D., ANDRADE D. L., CHANDROSS E. A., GAROFF S., ISRAELACHVILI J., MCCARTHY T. J., MURRAY R., PEASE R. F., RABOLT J. F., WYNNE K. J. and YU H.,

Langmuir

3

(1987)

932.

[4]

KJAER K., ALS-NIELSEN J., HELM C. A., LAXHUBER L. A. and MÖHWALD H.,

Phys.

Rev. Lett. 58

(1987)

2224 ;

BARTON S. W., FLOM E. B., RICE S. A., LIN B., PENG J. B., KETTERSON J. B. and DUTTA P., J.

Chem.

Phys.

89

(1988)

2257.

[5]

RASING Th., SHEN Y. R., KIM M. W. and GRUBB S.,

Phys.

Rev. Lett. 55

(1985)

2903.

[6]

WIEGEL F. W. and KOX A. J., Adv. Chem.

Phys.

41

(1980)

195.

[7]

BELL G. M., COMBS L. L. and DUNNE L. J., Chem. Rev. 81

(1981)

15.

[8]

GAINES G. L., Insoluble

Monolayers

at

Liquid-Gas Interfaces (Wiley,

New

York)

1966.

[9]

ADAMSON A. W.,

Physical Chemistry of Surfaces (Wiley,

New

York)

1982.

[10]

RICE S. A., to be

published.

[11]

See Refs.

[6, 7]

for a review of the theoretical efforts.

[12]

(a)

WANG Z.-G. and RICE S. A., J. Chem.

Phys.

88

(1988)

1290;

(b)

HARRIS J. and RICE S. A., ibid. 88

(1988)

1298.

[13]

CANTOR R. S. and MCILROY P. M., J. Chem.

Phys.

90

(1989)

4423 and 4431.

[14]

SHIN S., WANG Z.-G. and RICE S. A., J. Chem.

Phys.

92

(1990)

1427.

[15]

GELBART W. M., J. Chem.

Phys.

87 3597

(1986) ;

BEN-SHAUL A., SZLEIFER I. and GELBART W. M.,

Physics of Amphiphile Layers,

Eds. J.

Meunier,

D.

Langevin

and N. Boccara

(Springer, Berlin)

1987.

[16]

DILL K. A. and CANTOR R. S., Macromolecules 17

(1984)

380.

[17]

KHOKHLOV A. R. and SEMENOV A. N.,

Physica

A 112

(1982)

605.

[18]

ONSAGER L., Ann. N. Y. Acad. Sci. 51

(1949)

627.

[19]

DE GENNES P. G.,

Physics of Liquid Crystals (Oxford University

Press,

London)

1974.

[20]

HALPERIN A., ALEXANDER S. and SCHECHTER I., J. Chem.

Phys.

86

(1987)

6550.

[21]

CHEN Z.-Y., TABOLT J., GELBART W. M. and BEN-SHAUL A.,

Phys.

Rev. Lett. 61

(1988)

1376.

[22]

BOEHM R. E. and MARTIRE D. E., J. Chem.

Phys.

67

(1977)

1061.

[23]

SAFRAN S., ROBBINS M. O. and GAROFF S.,

Phys.

Rev. A 33

(1986)

2186.

[24]

ZWANZIG R. W., J. Chem.

Phys.

39

(1963)

1714.

[25]

GELBART W. M. and BARON B. A., J. Chem.

Phys.

66

(1977)

207.

[26]

COTTER M. A., J. Chem.

Phys.

66

(1977)

4710.

[27]

REISS H., FRISCH H. L. and LEBOWITZ J. L., J. Chem.

Phys.

31

(1959)

369.

[28]

COTTER M. A.,

Phys.

Rev. A 10

(1973)

625.

[29]

HIRSHFELDER J. O., CURTISS C. F. and BIRD R. B., Molecular

Theory of

Gases and

Liquids

(15)

[30] MCQUARRIE

D. A., Statistical Mechanics

(Harper

& Row, New

York)

1976.

[31]

TUMANYAN N. P., and SOKOLOVA E. P., Zhur. Fiz. Khim.

(Russ.

J.

Phys. Chem.)

58

(1984)

2444 ;

ibid., 58

(1984)

2448 ;

SOKOLOVA E. P. and TUMANYAN N. P., ibid., 61

(1987)

1349.

[32]

HALPERIN A., ALEXANDER S. and SCHECHTER I., J. Chem.

Phys.

91

(1989)

1383.

[33]

REICHL L. E., A Modern Course in Statistical

Physics (University

of Texas,

Austin)

1980.

[34]

KIM M. W.,

private

communication.

[35]

KRAMER D., BEN-SHAUL A., CHEN Z.-Y., TALBOT J. and GELBART W. M.,

unpublished.

[36]

In a continuum model the various

phase

transitions are

expected

to be less

pronounced

than in a

discrete model of

only

a few states.

[37]

A more serious drawback of the mean-field

theory applied

to two-dimensional systems is the

neglect

of fluctuations. The treatment of such fluctuations is

beyond

the scope of this paper;

see however: KOSTERLITZ J. M. and THOULESS D. J., J.

Phys.

C 6

(1973)

1181 ;

Références

Documents relatifs

Simulated normalized optical power absorption spectra (a) and average field enhancement spectra (b) for Au nanorod arrays with different lengths on a Si or ITO

Si l'on considère que la transparence phonographique de l'italien se vérifie tout particulièrement dans la morphologie du verbe (Dressler et al. 2003), qui

Thus, to be able to build a complete model of the respiratory system and to include a second parameter to the monitoring process of the lungs, we introduce in this paper

We describe athermal packings and their observed geometric phase transitions using equilibrium statistical mechanics and develop a fully microscopic, mean-field theory of the

conditions prevent the director field from being uniform, as it is the case in hybrid cells or in capillaries with ra- dial anchoring, then at equilibrium uniaxial order is de-

tuitively, the propagation of chaos means that the particles become independent so that we have a Poincaré inequality with a constant which does not depend on the dimension.. In

Magnetic after-effect(MAFi) is a very interesting property, from which one may ob- tain valuable informations of micraprocesses in ferromagnetic materials since MAE is generated 3

- Final configurations of binary alloy at low-temperature 0.08To with (a) Lennard-Jones potential, (b) Lennard-Jones potential supplemented by orientational three body