• Aucun résultat trouvé

Surface segregation in model symmetric polyolefin diblock copolymer melts

N/A
N/A
Protected

Academic year: 2021

Partager "Surface segregation in model symmetric polyolefin diblock copolymer melts"

Copied!
19
0
0

Texte intégral

(1)

HAL Id: jpa-00248128

https://hal.archives-ouvertes.fr/jpa-00248128

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.

Surface segregation in model symmetric polyolefin diblock copolymer melts

Mohan Sikka, Navjot Singh, Frank Bates, Alamgir Karim, Sushil Satija, Charles Majkrzak

To cite this version:

Mohan Sikka, Navjot Singh, Frank Bates, Alamgir Karim, Sushil Satija, et al.. Surface segregation in

model symmetric polyolefin diblock copolymer melts. Journal de Physique II, EDP Sciences, 1994, 4

(12), pp.2231-2248. �10.1051/jp2:1994258�. �jpa-00248128�

(2)

Classification Physics Abstracts

61.12E 61.40 82.65D 68.90

Surface segregation in model symmetric polyolefin diblock copolymer melts

Mohan Sikka

(I), Navjot Singh (I),

Frank S. Bates

(I), Alamgir

Karim

(2),

Sushil

Satija (2)

and Charles F.

Majkrzak (2)

(')

Department of Chemical

Engineering

and Materials Science,

University

of Minnesota,

Minneapolis~

Minnesota 55455. U-S-A-

(2) National Institute of Standards and Technology~

Gaithersburg,

Maryland, U-S-A- (Receii>ed 5 March 1994, received in final form 5 July J99J, accepted 29 July J994)

Rdsumd. Nous avons dtud16 le comportement de s6gr6gation en surface de couches minces de

polyoldfines

(copolymbres modbles h blocs) contenant une

phase

lamellaire. Les films sont pr6pards par «

spin coating

» sur diff6rents substrats, la microstructure est

analysde

par r6flexion de neutrons, r6flexion de rayon X,

ellipsom6trie, microscopie optique

et

spectroscopie

de masse d'ions secondaires. Chacun des

copolymbres

h blocs est suffisamment

proche

de la transition ordre-d6sordre (TOD) pour que le

profit

de

composition

dans la

phase

lamellaire puisse dtre

approxim6

par une fonction sinusoidale. L'interface avec l'air ou le solide est

toujours

enrichie par le mdme bloc, mdme lorsque I'£nergie de surface du substrat est

plus

61ev£e que celles des deux

blocs du copolymbre. Nous ne sommes pas capables de relier ce rdsultat avec l'enrichissement pr6f6rentiel bass seulement sur la diff£rence de tension de surface

dispersive

entre [es diff6rents blocs. Toutefois, nous remarquons que dans tous les cas, le bloc ayant la plus

petite

conformation est isold en surface. Ceci suggbre qu'un facteur entropique, attribud h

l'asymdtrie

conformation- nelle~

joue

un r61e dans le comportement en

proche

surface des

copolymdres

en milieu fondu.

Abstract. Surface segregation in thin films of model symmetric

polyolefin

diblock

copolymers exhibiting

a lamellar

morphology

has been

investigated.

Films were

spin

coated on a

variety

of

substrates and the

resulting

microstructure

analyzed using

neutron reflection, X-ray reflection,

ellipsometry~

light

microscopy

and secondary ion mass spectrometry. Each of the block copolymers is sufficiently close to the order-disorder transition (CDT) so that the

composition profile

within the lamellar domains can be

represented by

a nearly sinusoidal function. In all cases, the same block was found to enrich both the

polymer/solid

and the

polymerlair

interfaces, even

when the surface energy of the solid substrate

was

higher

than that of either block of the copolymer. These results cannot be reconciled with preferential enrichment based

solely

on the difference in the dispersive surface tension between blocks. In all instances, we find that the

conforrrationally

smaller block is the surface-active one. We propose that an entropic

driving

force, attributable to conformational asymmetry, plays an

important

role in the near-surface

behavior of block

copolymer

melts.

1. Introduction.

Block

copolymers

are useful as

large

surfactant molecules because of their

ability

to straddle

phase

boundaries and interfaces.

Applications

that can be visualized to

exploit

this interfacial

@Les

Editions de

Physique

1994

(3)

activity

include

stabilizing

colloidal

dispersions Ii

and

compatibilizing

the interface between immiscible

homopolymers [2].

The surface

properties

of

polyolefin (saturated hydrocarbon) copolymers

are of

particular

interest to us for several reasons. In

1992, poly(ethylene)

and

poly(propylene) together represented

about 40§b of the total volumetric

production

of

thermoplastic

resins in the United States

[3]. Polyolefins

are

widely

used in

products

where interfacial

properties

are

important,

such as

adhesives, packaging,

lubricants and

coatings.

Also, polyolefins

have been

recognized

as

superior

materials for fundamental studies of

polymer properties,

and a

variety

of

compounds

with controlled molecular architecture have

recently

been

synthesized

and characterized

[4, 5].

An

ability

to control intersurface interactions

using

block

copolymers

is dictated

by

our

understanding

of the

physical

and

thermodynamic

behavior of these materials in the bulk state.

Considerable theoretical and

experimental

efforts have been made in the last twenty years to

categorize

the distinctive microstructure and

ordering

transitions

displayed by

block

copolym-

ers in the melt, as reviewed

recently by

Bates and Fredrickson

[6].

Of

particular

interest in this class of materials are diblock

copolymers, composed

of sequences of distinct monomeric units A and B

covalently

linked at one

point.

The overall

degree

of

polymerization N,

the fraction

f

of block A in the chain, and the A-B segment-segment interaction parameter x, are three

parameters

that control the

phase

behavior of bulk diblock

copolymers.

The

product xN

and

f

are

generally

used to

classify

the bulk

phase

state.

Recently,

Bates and coworkers have used model

polyolefin

diblocks and

homopolymer

blends to

experimentally

demonstrate two facets of

phase

behavior not

anticipated by

conventional mean-field treatment. First, the

existence of

multiple

ordered

phases

in the weak

segregation

limit

[7]

for

polyolefin

diblocks that are

compositionally off-symmetry,

I.e. when

f

is

significantly

different from 0.5

[8, 9].

Second, the correlation of

binary phase

behavior with a parameter e

(see below)

that

provides

a

measure of the conformational asymmetry between the two blocks or two

homopolymers [10-

l2].

In a thin diblock

copolymer

film

(with

film thickness

comparable

to several times the bulk

domain

spacing),

one may expect bulk

thermodynamics

to be modified

by

the presence of

symmetry breaking

interfaces that can influence the bulk

equilibrium configuration.

Most

experimental

work to date in this area has been done on

nearly symmetric ~f

m 0.5 diblocks that form the lamellar

morphology

in the bulk. The first

study,

on

poly(styrene)-poly(methyl-

methacrylate) (PS-PMMA)

diblocks cast on silicon dioxide substrates and annealed above the

glass transition,

indicated a lateral

alignment

of lamellar microdomains

parallel

to the film surfaces for

xN

>

(xN)o~~ [13-16].

Here the order-disorder transition

(ODT)

refers to the

point

in

xN

space above which the

copolymer

would form the lamellar

morphology

from a disordered melt. This lateral orientation has since been observed in other systems

[17-19].

A second effect that has been

reported [17-20]

is the presence of surface induced orientational order at values of

xN

~

~XN )o~~,

where the bulk state

(I.e., essentially infinitely thick)

is

disordered. Another is

that,

for

XN

>

(xN )o~~, affinity

for the same or different blocks at the air and substrate interfaces leads to the

symmetric [17, 18,

2

Ii

or the

asymmetric [13-16, 19]

thin-film

geometries depicted

in

figure

I. In

addition,

the

deposition

of

non-integral

and non-

half-integral

amounts of material

(t)

# dn or

(t)

#

d(n -1/2)

for the

symmetric

and

asymmetric

cases,

respectively,

where

(t)

is the average film thickness and n is a

positive integer]

leads to a surface

topology composed

of islands or holes of thickness d. These defects

were first observed for PS-PMMA diblock films

using light

interference measurements

[22].

Currently,

we are

studying

the effect of reduced

dimensionality

on

phase

transitions and domain size for

off-symmetry ~f

#

0.5) polyolefin

diblocks that exhibit non-lamellar and

multiple

ordered

phases

in the bulk

[8, 9,

23,

24].

Related issues that emerge are whether surface

topology (holes

and

islands)

can be observed for non-lamellar

morphologies

cast in the

(4)

a.

Asymmetric

Film b.

Symmetric

Film

I ~

~

~

~t3 ~~z

~~~~~

~

~

~,

' ~~3

..~ ~+.. ~..~i.'~'"?~. iii? :J~~S.

~

~~z ~' Z ~.

~jj~jm(()j#jjJ%~ ~

,,,,,,,,,,,,,,,,,,,,,

Substrate Substrate

(Il 1/2) d~ 'l£~fi

,

~, ,''~ ,§~

'

l'

"

l'~ l~

/ /

',

/

a ' ~

~ i

j

~

i i '

I j

' ' '

I ' I i I ' ' f

,' ',

,I ,

,_,

,~'

,,

z z

Fig.

I. The two possible thin-film

geometries

for symmetric

~f=

0.5) diblock

copolymers.

d is the lamellar

period

and n is

an integer. (a) Enrichment of different blocks at

polymerlair

and

polymer/substrate

interfaces~ and (b) enrichment of the same block at both interfaces. Each type of film

can have two different

composition profiles,

phase shifted by 180°,

depending

on which block is surface- active.

thin-film

configuration,

and whether the formation of surface structures

depends

on the

majority

component

being

surface-active. We are also interested in

examining

how confine- ment of a structured diblock

copolymer

melt between

parallel plates

influences its microstruc- ture

[25],

since the film will be unable to form holes and islands in such a

configuration.

From the discussion above, three interrelated aspects emerge in

characterizing

the behavior of block

copolymers

in thin films and near interfaces

I)

internal microstructure for a system that may no

longer

be three-dimensional in nature,

it)

lateral

topology

at the free

(air)

surface of a

film,

and

iii)

enrichment at a free surface or a solid interface. The

coupling

between two or

more of these

phenomena

introduces a

degree

of

complexity

that makes thin-film research

challenging,

as will be discussed in a

forthcoming publication [26].

For

example,

we have found that internal microstructure and surface

topology change

with temperature for at least three reasons : thermal

expansivity changes

the relative dimensions of the overall film thickness and the domain

periodicity,

coil dimensions

(hence,

the bulk

periodicity)

themselves

change

with temperature

[27],

and

phase

transitions between ordered

morphologies

can occur

[8, 9].

We also observe that microstructure and

topology

for non-lamellar ordered mor-

phologies,

cast as thin films, are

susceptible

to

dynamic

barriers to true

equilibrium [26].

Examples

of

trapped complex morphologies

in solution cast films are

reported by Hasegawa

and coworkers

[28]

for the

poly(styrene)-poly(isoprene)

system. The observation of

quasi-

(5)

ordered structures and the

suppression

of island formation in

asymmetric ~f

m 0.77 PEP-PEE

thin films have

already

been

reported by

us

recently [24].

In this paper, the [east

complicated

of the three issues discussed above, surface

segregation

in ordered diblock

copolymer melts,

is examined in detail. A recent letter summarized our initial

findings

on this

subject [18]. Compositionally symmetric J

m 0.5

polyolefin

diblocks

were used with molecular

weights

that

placed

them in the ordered state~

forming

the lamellar

morphology

in the bulk, as confirmed

by small-angle

neutron

scattering (SANS)

and

rheology experiments.

Thin films

supported

on

just

one side

by

a solid substrate were studied. Molecular

weights

were chosen to make all the diblocks

relatively

close to the ODT at the measurement

temperature the

composition profiles closely

resemble those

reported

for ordered diblock

specimens

near TODT

l17],

and

proximity

to the ODT ensures that the films achieve an

equilibrium

state. Also, the

samples

were annealed

sufficiently

above the

glass

transition temperature

(T~)

and, where

appropriate

the

crystalline-amorphous

transition

(T~),

of the blocks to eliminate kinetic barriers to

equilibration

associated with the

glassy

and

crystalline

states. It is

pertinent

to stress that an

equilibrated

lamellar

morphology,

with

alignment

of

microdomains in the lateral film

direction,

is

important

in

studying

surface

segregation

in a direct way. This

configuration

ensures

wetting

of both film interfaces

by

the associated

surface-active block [8~

9]

without

distorting

the bulk

morphology

as observed in more

complex

systems, for

example,

where a

wetting

block is a

minority

component

[6].

Also,

limiting

the

study

to model

polyolefins

has enabled us to examine the effect of molecular

architecture and coil size on

polymer

surface

activity

in a controlled way.

2.

Experimental.

2.I MATERIALS. Four model

polyolefins

were chosen for this

study polyethylene (PE),

poly(ethylene-propylene) (PEP), poly(vinylcyclohexane) (PVCH),

and

poly(ethylethylene)

(PEE). The microstructure and relevant

physical properties characterizing

these

polymers

are

listed in table I. The statistical segment

length

b is of relevance to this paper. It is defined

by

the

unperturbed

Gaussian coil radius of

gyration,

R~

=

b(N/6)~'~, (l

where N is the

degree

of

polymerization

defined as the molecular

weight

divided

by

the

monomer mass. This parameter has been determined for PE, PEP, PVCH and PEE in the bulk

state

by small-angle

neutron

scattering [4, 10].

For each of these

polymers,

we

recognize

that

conformational asymmetry effects should be defined in an

N-independent

way. We can

combine the volume of the molecule

(V)

and space

filling (R~)

parameter into a

single

parameter Ii,

~

R( b2

~

V 6 vo ' ~~~

where vo is the volume

occupied by

a statistical segment. Conformational asymmetry between

two

species

is then defined

by

:

p

2

E=

~)) (3)

Table I also lists the value of the parameter

p~,

for each of our

polyolefins

at 25 °C.

The number average molecular

weights, M~,

for the diblock

copolymers

used in this

study

are listed in table II. Also listed are the

isotope

content, the block

copolymer composition

(6)

Table I.

Polyolefin

molecular characteristics.

~°~~°~~~~ ~~~~~ ~'~~~°~~~

g/cm3

~~l~~~

K-1

PE

0.855~.b 7.5~ 8.9b>h

-0.58k

12.I

PEP

(Pclyefllylene-

0.854~ 7.0f

8.l~

-0.58~ 8.I

propylene)

~PC'Yetl1yl

~ ($

o_869~ 6.0f 4.9l 0.2i 3.8

e~~e)

~~~

~2

cH~

PUGH

~CH~-CH~

~Pc'Yvinyl N o_947b.4t 4.0~t 7.ld =0° 4.4

cycloltexwe)

° PE and PEP flso conwin small percen&ges ofrandomly dbtribttted branched repeat ttni~ as reportedin Reference 10. a Reference 29.

bEmapo~tttedfrom above the melt tempenture. CDentity gra&ent column measuremem. d Reference 12.e Reference 30. fReference 31.

i Based on a chemical repeat unit of the homopolymer melt as determined by BANS. h Reference 32. i Reference 33.I Reference 34.

k Reference 35.

Table II.

Polyolefin

dib/ock

copolymers.

A-B 103Mn,

IA

TOD~°C Bulklamellar Ref

Diblock gmol"~

period,

d,

I

PVCH(d8)-PEE

52 0.48 212 248 (25°C) 12, 36

PEP-PEE(d6)

55 0.55 125 341 (25°C) 27, 37

PE(d6)-PEE

26 0.47 182 246 (130°C) 38

PE-PVCH(d8)

15 0.48 238 180 (25°C) 12, 36

PE(d~)-PEP

100 0.50 139 534 (130°C) 39

(7)

f, order-disor4er

transition

To~~,

and the bulk lamellar

spacing

d. These materials were

synthesized

and characterized as described in separate

publications [4, 5, 12].

Methods for

obtaining To~~

and d in the bulk state are also discussed elsewhere

[27, 40].

2.2 SAMPLE PREPARATION. Thin film

samples

were

prepared by spin coating

from solution

onto either 2- or 4-inch flat substrates. Silicon surfaces were

prepared

in a clean room in the

following

way. Polished silicon wafers were cleaned either with

methylene

chloride and methanol or with a 3 : 7

hydrogen peroxide/sulfuric

acid solution at 120 °C. The wafers were then rinsed

thoroughly

with distilled water for several minutes and then etched with a buffered 1: 9

HF/NH~OH

solution to remove the native oxide. After

thorough rinsing

with distilled

water once

again

and

drying

with

high purity nitrogen,

the wafers were

immediately

coated.

Smooth

polystyrene

substrates were

generated by spin coating

this

polymer

from toluene solutions onto silicon wafers followed

by drying

in a vacuum oven ; film

uniformity

and flatness were confirmed

by X-ray

reflection measurements. Silver surfaces were obtained

by evaporating

500

A

of metal onto silicon wafers

using

an E-beam metal evaporator. Polished

optical

quartz disks were cleaned under an ultra-violet

lamp

before

being

coated.

PE-PEE and PE-PEP diblock

copolymers containing

a

crystallizable

block

(PE

melts at

108

°C)

were

spin

coated from hot

o-xylene

solutions onto heated substrates. PVCH-PEE and PE-PVCH diblocks

(PVCH

has a

glass

transition at 140

°C) [4]

were

spin

coated from

o-xylene

solutions at room temperature. PEP-PEE

films,

which are characterized

by glass

transitions

near 56 °C and 20

°C, [5]

were

spin

coated from

low-molecular-weight

alkane or aromatic solutions at room temperature. Film thickness could be controlled with a

precision

of better than 50

A by

a careful choice of

solvent,

solution concentration and

spinning speed.

All films were dried in vacuum for several hours after

coating

to remove residual solvent.

Films made from the

crystallizable polymers

that were used for SIMS measurements were

heated above T~~j~ for PE, but below the ODT for the

copolymer,

and then

quenched

in

liquid nitrogen.

This

methodology

ensured that the

morphology

in the film was

sufficiently

undistorted

by

the

crystallization

process for

depth profiling

at room temperature

using

SIMS.

Neutron reflection measurements on films made from PE-PEE and PE-PEP diblocks were

made in situ at 130 °C in a vacuum

chamber,

to ensure that both blocks were in the

amorphous

state. Measurements on PEP-PEE films were made at 25±2°C since the material is

amorphous

and well

annealed

at

room temperature. Films made from PVCH block

polymers

were annealed for up to 24h at 170 °C and then

quenched

to the

glassy

state at room

temperature where all measurements were made. The

profiles

in these

samples correspond

to those associated with the diblock at the

glass

transition of PVCH.

2.3 INSTRUMENTATION. Neutron reflection

(NR)

measurements were

performed

at the National Institute of Standards and

Technology (NIST) using

the reflectometers

[41]

on beam tube BT7

(with

fixed

wavelength,

A

=

2.35

A)

and beam tube NG7

(with

fixed

wavelength,

A

= 4.I

A). Background scattering

was measured for several values of k

=

2 gr sin 0/A,

A~

~, the incident momentum

perpendicular

to the surface where 0 is the

angle

of

incidence,

and was subtracted from the data when it exceeded

m 10 9b of the total

specular reflectivity.

Measurement of film structure at elevated temperature

using

NR were made in situ

using

a

vacuum

sample

chamber with controlled

heating

surfaces both behind and in front of the film

in order to

optimize temperature uniformity.

Nominal film temperatures

reported

here are

subject

to an error not

exceeding

10 °C at

higher

temperatures. For films that were heated to

higher temperatures

for measurement on the neutron

beam,

a substantial

portion

of the

reflectivity

curve was

repeatedly

monitored until it no

longer changed

with time before

recording

the curves used for

analysis.

Film thickness was measured after

casting using

a

Sopra~

ES4G

spectroscopic ellipsometer

(8)

at the Center for Interfacial

Engineering (CIE), University

of Minnesota or a

Scintagg~ X-ray

reflectometer at

NIST,

with an estimated error not

exceeding

± 10

A. Analysis techniques

to

extract film thickness can be found in the literature for

ellipsometry [42]

and

X-ray reflectivity

[41].

For films with an

incomplete

top

layer,

the measured thickness

(t) corresponds

to an

average

single layer

as seen

by

the

ellipsometer.

Interference

micrographs

of the surface

topology

of films were

acquired using

an

Olympus~ light microscope

in a reflection

configuration. Secondary

ions mass spectrometry

(SIMS)

was

performed

in the static mode

[43]

on some films to characterize the

polymerlair

surface For

this,

a Perkin

Elmer~

PHI

6300 SIMS machine was used with 7 kev Xe+

primary

ions rastered over a 1.5 x 1.5

mm~

area at a beam current of 0,15 nA. The same machine was used in the

dynamic

mode for

depth

profiling

measurements

[14]

on other films with 3 kev

O( primary

ions rastered over a

700 x 700 ~Lm~ area at a beam current of 70 nA.

3. Results.

Figure

2 shows the surface

topology

of thin films of PEP-PEE cast on silicon at 23 °C. The

films are close to 3.35 and 4 bulk lamellar domains

(d~)

in thickness as measured

by

fi§

O

.

.

~

~

~

10~m

Fig.

2. Interference

micrographs

of the surface of films of M~ 55,000 PEP-PEE

spin-coated

on

hydrogen

passivated

silicon. (a)

Sample

with measured

it)

=

385 ± 5

1

4

db.

A

mainly

featureless surface is observed. (b) sample with measured

(t)

=

1421

=

3.35 d~.

Light

areas

corresponding

to

/

it)

=

4d~ show up as islands on a darker

background

that

corresponds

to

it )

= 3d~. d~ is the

bull

lamellar period, estimated as

3411

at 23 °C from small angle neutron scattering [27, 40].

(9)

spectroscopic ellipsometry immediately

after

casting.

The

light

interference

micrographs

were

taken after the films were allowed to anneal at room temperature

(23 °C)

for two weeks. The film close to an

integral

number of

bilayers

in thickness shows a

relatively

featureless surface

whereas the film with measured thickness close to a half

integral

number of

bilayers

shows islands

(light

structures on dark

background)

that are one lamellar domain in

height.

The contrast that makes the surface

topology

visible under a reflection

microscope

arises from the constructive interference of

light

at two different colors for the top and the base of the islands

(lighter

at the top, where

it)

=

4

d~,

and darker at the

base,

where

(t)

=

3

d~).

This result

clearly

demonstrates that the PEP-PEE diblock assumes the

symmetric

thin-film geometry

depicted

in

figure

I with the same block

(A

or

B)

situated at both air and substrate interfaces.

For this geometry~ as discussed in the introduction

section,

the

deposition

of

non-integral

(a)

i

PEP

@

=

U ' ,

4J

qQ ~, '

4J

f~ ,

"

' ',

",

"-,

10~k

,

i~' (b)

Air

si i

~PEE o

I

, ,

', /, i, , ,

, ,

,/,

/, i, i,

i ,

,

, ' , , , 1 ,

, , , ,

'

, , , , ,

, , , ,

' , , , , ,

, , ,

, '

, , , , ,

~ i , ,

° 900 1800

distance from air surface,

A

Fig. 3. Example of fitting neutron reflectivity data to determine the surface-active block of a diblock

copolymer

system. (a) NR data

(open symbols)

for a M~ = 55, 000 PEP-PEE film

spin-coated

on

stripped

silicon. The solid line is the fit corresponding to the optimized profile shown in (b) a~ a solid curve. The dashed line in (al

corresponds

to the

composition profile phase

shifted

by

180°, shown as the dashed

curve in (b). The best fit to the data clearly shows that both film surfaces are enriched by PEE.

(10)

amounts of material

((t)

#

nd~,

where

(t)

is the average film thickness measured

by ellipsometry

and n is an

integer)

would lead to a surface

topology composed

of holes and islands of

height d~,

as observed.

The same result is seen for all our

polyolefin

diblocks cast on etched silicon each of our

copolymers

forms the

symmetric

lamellar geometry where

(t)

=

nd~,

with the same block

enriching

both air and substrate interfaces. This result is further substantiated

by

NR and SIMS

measurements as described below. For these

experiments

we have

prepared,

as far as

possible,

integral

thickness films.

Determination of the surface active block in the diblocks was

accomplished using

neutron

reflectivity (NR)

measurements.

Techniques

for

analysis

of NR data, as well as the

sensitivity

of this

technique

to surface

composition,

have been discussed

by

us and other authors in separate

publications [17,

18, 41,

44].

To

recapitulate,

the NR data can be

satisfactorily

fit

only

when the correct

phase (I.e.,

surface-active

block)

is included in the

composition profile

used to simulate the

reflectivity.

As an

example, figure

3b shows two

composition profiles

that

are used to fit the data for PEP-PEE in

figure

3a ; one

profile

shows an enrichment of PEE

(solid curve)

and the other shows a

depletion

of PEE

(dashed curve)

at the films surfaces. The

corresponding

reflectivities calculated for these

composition profiles

are drawn as the solid and

PVCH-@

b PE

-@

fl

C

ij it

I PE

@

PE

-fl

0 2 3 4 5 6

io~k

,

A-i

Fig.

4.

Experimental

neutron reflectivity curves (open symbols) for representative

polyolefin

diblock

copolymer

films cast on etched silicon, listed in table III. Progressive vertical shifts of 4 orders of

magnitude

have been

applied

to the

plots.

Solid lines are the calculated reflectivities for the optimized concentration

profile,

The surface-active block is outlined above each data set. The temperature at which

each measurement was done is listed in table III.

(11)

Table III, Thin

film

characteristics,

Diblock Subsmte

Temp,°C

102fl

(

d

, n Surface #s

copolymer

block

A-B

PVCH(d8)-PEE tsj

25 4.4/3,8 230 4 PEE 0,96

PEP-PEE(d~) tsi

25 8.1/3.8 348±5 5

PEE(d~)

0.90

'~ 25 '~ 330 3

PEE(d~)

0.87

"

Ag

25 '~ 325 5

PEE(d~)

0.83

"

Quartz

25 '~ 330 5

PEE(d6)

0.87

tsi 130 9.9/3.7 255 12 PEE 0.87

tsj 25 12,I/4A 168 8

PVCH(d8)

0.94

tsi

130 9.9/6.7 535 6 PEP 0.72

tNafive odde removed usinga1:9HF:NII~OH e~h.

dashed curves in

figure

3a. It is clear that

only

the

profile

with PEE

enriching

both air and Si lamellar

period

and the number of

periods

for each film. Also included in table III are results for PEP-PEE films cast on a number of substrates besides etched silicon. These

experiments

were

performed

to see if surface

segregation depends

on the surface energy of the solid substrate. Neutron reflection data for these is not

presented

since

composition profiles

for these films matched the

optimized profile

in

figure

3b and the surface

segregation

is identical to that

observed for etched silicon.

For diblocks where resolution was

adequate,

direct

imaging

of film

profiles

was done

using dynamic

SIMS. In this

technique, primary O(

ions are used to sputter

through

a

specimen

and

secondary

ion intensities recorded as a function of time

(depth)

from the surface. The

interfaces

gives good

agreement between calculated and measured reflectivities

(open

symbols).

The

plots

in

figure

4 show

reflectivity

data

(open symbols)

for films of the other diblock

copolymers

listed in table II cast on etched

silicon, along

with calculated fits for

optimized composition profiles (solid

curves).

Table III summarizes the thin film characteristics extracted from the NR data

analysis

the value of p ~ for each of the blocks at the temperature at which the NR data was taken, the type and volume fraction

#~

of the surface-active block for all the diblock

films,

and the measured

secondary

H, D~ and Si~ ion

yields

for two of the films as a function of

sputtering

time is shown in

figures

5 and 6. Since one of the blocks in each of our

copolymers

is

partially deuterated,

the H~ and D~

signals provide

information about the

composition profile

of both chemical

species

in the thin films. The Si~

signal gives

an estimate of the

point

in time that the substrate is reached after

sputtering through

the full thickness of the

sample. Figure

5 shows the SIMS

profile

for a three

bilayer

PEP-PEE

(d~)

film cast on etched silicon. The presence of

three oscillations in both the D~ and H~ spectra is clear, as is the enrichment of the deuterated

(12)

species (PEE)

at the silicon interface. Instrumental factors

beyond

our control make the

etching

rate non-uniform in the first 100

A

of the film.

Consequently,

the H~ and D~ oscillations are not distinct in this

region

and it is not

possible

to tell

directly

if there is an H~ or a

D~ maximum at the

polymerlair

interface. However, a

comparison

of the

wavelength

of the

oscillations and the thickness of the

film,

as measured

by ellipsometry,

makes it clear that there should be enrichment of D~

,

and therefore

PEE,

at the air interface as well.

Figure

6 shows the SIMS

profile

for a two

bilayer PE(d~)-PEP

film cast on etched silicon. For this

sample, following

the

example

above, H~ maxima are seen at both the air and substrate interfaces

which indicates that PEP is the surface-active

species

for this diblock. Both these results are consistent with our conclusions from NR for the same diblocks.

For each of our diblock films,

qualitative

confirmation of the surface-active

species

at the

polymerlair

interface was also

accomplished using

static SIMS. This

technique

uses a

primary

Xe+ beam at a energy level less than that used for

dynamic profiling.

The

secondary

ions are

collected

only

from about the top 15

A

of the

film,

and the

presence of

mainly

deuterated or

mainly hydrogenated

ionic

species

in the SIMS spectrum indicates which of the blocks enriches the air surface.

~ Si

e

I

B I

°

~

z

g

I

uttering

Fig. 5. SIMS

profile

at 23 °C of a ca.

0001thick

PEP-PEE

(d~)

diblock sample annealed 24 h at

room temperature in vacuum, shown as secondary ion counts versus time. The vertical line at

higher

etching times marks the

position

of the

copolymer/silicon

substrate. The C~

signal

(not shown) remains

constant throughout the whole film, while H~ and D~ signals exhibit pronounced oscillations that

correspond to lamellar domain~ oriented parallel to the film surface.

(13)

H

Si U I

U E

~ O

°

# fl E

~

D

Sputtering Time (sec)

Fig.

6. sIMs

profile

at 23 °C of a ca.

10001thick

PE(d~ )-PEP diblock sample annealed 24 h at

120 °C in vacuum, shown as secondary ion counts i>ersus time. The vertical line at higher etching times

marks the position of the

copolymer/silicon

substrate.

4. Discussion.

Conventionally

surface enrichment in a system with two chemical

species

has been understood in terms of a minimization of interfacial tension at a system

boundary.

For

example~

in work

by

Russell and coworkers

[13-16],

on

symmetric

PS-PMMA

diblocks,

an enrichment of PS at the air surface is

explained

in terms of its lower surface

tension,

and the

segregation

of PMMA at the silicon dioxide substrate is understood in terms of its chemical

affinity

for this material. If

specific

chemical effects are absent, and interfacial tension is associated

only

with

enthalpic (van

der

Waal's) effects,

the block with the

higher

surface energy is

expected

to segregate to the

(high energy)

substrate and the block with the lower surface tension is

expected

to

prefer

the free surface of the film. This follows from a conventional definition of interfacial tension

between a melt and a solid

phase

that relies on the difference between the surface tensions of the two

phases.

An

example

of such an effect can be seen in recent work

by

Krausch and coworkers

[45]

on mixtures of

poly(ethylene-propylene) (hPEP)

and the deuterated

homologue (dPEP)

cast as films on etched silicon surfaces dPEP is found to enrich the

polymerlair

interface, while hPEP is found to be surface-active at the

stripped

Si surface. This behavior can be attributed to the

slightly

lower cohesive energy

density

that characterizes the heavier

isotope [46]

and the fact that the solid surface has a

higher

surface energy than either

polymeric species.

Table IV lists known values of the

dispersive

component of the surface energy

(y~)

of the chemical

species

present in our

diblocks, along

with the substrates used for

spin-

(14)

Table IV.

Surface energies.

Yd,

PEP

30.6b

PEE (20-34)~

PE 35d

PUGH 35t

Hydrogen passivated

44.7b

silicon~

> PEPf

> PEB

Silicon w1tll native

36.5b

oxide

Poly(styrene)

38d

Silver 1500'

a Native oxide removed with a llydrofluoric acid etch.

b Reference 47. CAn experimentally determined value of the stirtace tension for PEE is unavailable in the literature. The lower estimate corresponds to the critical surface tension for closely packed methyl groups (Reference 53). The higher value is estimated using group contributions flom Reference 48, d Reference 49. 'Measured using contact angleswith water and methylene iodide, fReference 47, i Deduced flom isotopic segregation (See text).h Reference 50.

coating.

The

dispersive portion

of the surface energy is the relevant

quantity

in our

analysis

since

polar

and ionic forces are

expected

to be

unimportant

in

simple

saturated

hydrocarbons.

For

hydrogen passivated silicon,

the substrate used in most of our

experiments,

we are

interested in a

quantitative

estimate of y~ as well as a definitive answer to the

question

does etched silicon have a

higher dispersive

surface energy than our

polyolefin

melts ?

Comparison

of measured surface

energies,

we

feel,

is an

inadequate

way to answer this

question

since

techniques

to measure the surface energy of solids and the surface tension of

polymer

melts are~ in

general,

different and

rely

on different definitions of these

quantities [49].

Part of the

answer is

provided by

Krausch and coworkers PEP has a lower surface energy than

stripped

silicon since dPEP and hPEP segregate to the

polymerlair

and

polymer/silicon

interfaces of thin films,

respectively.

We

performed

a similar

experiment using (h~)

PE with a molecular

weight

of 1.8 x

10~ mol/g,

and the deuterated

homologue (d~)

PE. A ca. 4 000

ji

thick film

was

spin-coated

at

high

temperature from a

o-xylene

solution that contained

equal weight

percent of both

polymers.

The film was annealed at 160 °C for 10 h under vacuum and then

analyzed

at 23 °C with

dynamic

SIMS. The

resulting

H~ and D~

profiles,

shown in

figure 7,

(15)

H H

o~

fl

f

~ o

°

Si

# fl

3

D

i

Spunering Time (sec)

Fig.

7. SIMS

profile

at 23 °C of a ca. 4

5001thick

film of PE and the deuterated homologue, dPE cast on

stripped

silicon. The film was annealed at 160 °C for 10 h in vacuum and then

quenched

in

liquid nitrogen

before measurement. The

asymmetric segregation

of dPE to the air surface and hPE to the silicon

substrate is

clearly

seen.

show

clearly

that hPE is

preferred

at the

polymer/silicon

interface and dPE at the

polymerlair

interface. This establishes that

stripped

silicon has a

higher

surface energy than PE as well.

Segregation

based on an

enthalpic

rationale alone cannot

explain

the enrichment of both low energy

(polymerlair)

and

high

energy

(polymer/substrate)

interfaces

by

the same block. A

possible explanation

for our

experimental

observations is that blocks that have different space

filling

characteristics will

experience unequal

conformational

perturbation

in the presence of a

symmetry breaking

interface. This statement is illustrated

graphically

in

figure

8. Block A in a

symmetric

A-B diblock

copolymer

has the same molecular volume as block

B,

a condition that di1be cintrolled

by synthesis stochiometry.

This necessitates that the lamellar domains of size d will be

composed

of A- and B -rich subdomains that are

equal

in lateral size

(d~

=

dB ).

If the

overall

density

remains constant

(which

it does to a close

approximation)

then the I-

dimensional division of space dictates this condition.

However,

the blocks are con-

formationally asymmetric

as

captured

in the different

p

parameter for each chain type

(Eq. (4)).

A

large

value of this parameter

signifies

an

unperturbed

chain that is more

conformationally expansive (I.e. occupies

less space per unit conformational

volume)

than a

chain that has a smaller p value. In this

example, p(

~

p(.

Thus B chains are thicker and shorter than A chains. To preserve

incompressibility,

and the

equality

in subdomain

size,

chains from different sides of an A-A interface will be more intermixed than chains across a B- B interface

(dotted

lines in

Fig. 8).

The creation of a free

interface, therefore,

will occur at

higher

conformational

penalty

at an A-A interface than at a B-B

interface,

and this

entropic

cost should be factored into the calculation of the overall surface free energy.

Here,

we have

(16)

d~

~

d~

~

(17)

degree

of

interpenetration

between blocks from

neighboring

molecules. Such a

compositional

state would be more

susceptible

to conformational

perturbation

than a

strongly segregated

system with

sharp

interfaces. This raises an

interesting

issue. As X or N

(I.e. XN) increase,

and the

degree

of

overlap

between

opposing

interfaces within microdomains decreases

(Fig. 9) [53],

the non-local

entropic driving

force for surface

segregation

should decrease. If a

competing enthalpic

or local

entropic

factor exists, this could lead to an inversion in the

wetting tendency,

which would

provide

a definitive test of our concept.

~

~ 'l# ~~j$

~

Weak segregation Limit (WSL) strong segregation Limit (ssL)

zN

= 10

XN

» 10

Fig.

9. Schematic illu~tration showing decreasing overlap between like chains from

opposing

microdomains in an A-B diblock copolymer as the product XN increases~ and interfacial widths become

narrower [53]. The influence of conformational asymmetry or preferential wetting should therefore

decrease as the SSL is

approached.

~~ is the volume fraction of

siecies

A.

Segregation

based on conformational asymmetry is

supported by

results from all five diblock systems for each case the block with the smaller p value is surface-active near the ODT

(see

Tab.

III). Particularly compelling

are the results from the PVCH diblocks. In one

instance, PE-PVCH,

the PVCH block is surface active~ while in a second,

PVCH-PEE,

it is

not. In both cases, the PVCH block is deuterium labeled while the other

polyolefin

block is

hydrogenous (Here

we note that there is no correlation between

isotope

content and surface

activity

in the results listed in Tab.

II).

A

comparison

of the relative

magnitude

of

enthalpic

and

entropic

components of the surface

free energy can be made

by considering

the PE-PEP system, since the

dispersive

and

conformational characteristics for both blocks are well established

experimentally (Tab.

IV).

Enthalpically

driven

surface-activity

based on a y~ value of ca.

31dynes/cm

for PEP and 35

dynes/cm

for PE should lead to the

(asymmetric) segregation

of PEP to the air and PE to the substrate interface.

Entropy,

on the other

hand,

favors the

(symmetric)

enrichment of the

conformationally

smaller PEP block at both interfaces. Since the latter situation is observed

experimentally~

we conclude that conformational

activity

is dominant for an

~ value of 1.5

(given by

~

=

p (~/p(~p)

and a surface tension difference of 4-5

dynes/cm

between the blocks.

In

closing,

we make a brief remark about local

packing

effects

[52].

While the correlation between conformational asymmetry and surface

segregation

is rather

compelling,

we do not

Références

Documents relatifs

Nevertheless, their convergence has only recently be studied in the only case of a constant function λ: the convergence of a numerical scheme involving a finite volume method for

The main purpose of this paper is to present the structural and chemical properties of TbFe or TbFeCo alloy thin film stacks grown by coevaporation in relation with their

Renormalization group calculations based on the two parameter model pre- dict that the effects of chemical mismatch on the size of diblocks appear only as corrections to scaling

It is shown that the kinetics of growth depends on the initial density of islands (or holes) : at 140 °C, for intermediate annealing times, there is no time-evolution of the

In the previous sections we have investigated the behavior of the symmetric AiBi copolymers in the gel matrix. We have found that the weak preferential adsorption of A monomers

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

Near the order-disorder transition temperature, T - TODT, and at low shear frequencies, the lamellae arrange with unit normal perpendicular to the flow direction and parallel to

T. PHOTOEMISSION FROM THE DECAY OF SURFACE PLAS- MONS IN THIN FILMS OF Al.. - We have measured the yields and energy distributions of electrons generated by the decay of