• Aucun résultat trouvé

Polymerization of a diacetylenic phospholipid monolayer at the air-water interface

N/A
N/A
Protected

Academic year: 2021

Partager "Polymerization of a diacetylenic phospholipid monolayer at the air-water interface"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: jpa-00247949

https://hal.archives-ouvertes.fr/jpa-00247949

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.

at the air-water interface

L. Bourdieu, D. Chatenay, J. Daillant, D. Luzet

To cite this version:

L. Bourdieu, D. Chatenay, J. Daillant, D. Luzet. Polymerization of a diacetylenic phospholipid monolayer at the air-water interface. Journal de Physique II, EDP Sciences, 1994, 4 (1), pp.37-58.

�10.1051/jp2:1994114�. �jpa-00247949�

(2)

Classification

Physics

Abstracts

68.60 68.10 61.10

Polymerization of

a

diacetylenic phospholipid monolayer at the air-water interface

L. Bourdieu

(I),

D.

Chatenay (I),

J. Daillant (~>*) and D. Luzet

(~)

(~)

Institut Curie, Section de

Physique

et

Chimie,

Laboratoire de

Physico-chimie

des Surfaces et Interfaces

(**),

II rue Pierre et Marie Curie, 75005 Paris, France

(~)

Service de

Physique

de l'Etat

Condens6,

Orme des Merisiers, Centre d'6tudes de

Saclay,

91191 Gif-sur-Yvette Cedex, France

(Recei~~ed

20

July1993,

received in final form 17

September

1993,

accepted

23

September1993)

Abstract.

Monolayers

of a

polymerizable phospholipid

on water have been studied both

before and after

polymerization.

Before

polymerization,

the

phase diagram

is established

by

isotherm measurements and

optical microscopy (epifluorescence

and direct observation between crossed

polarizer

and

analyzer).

This allows us to

bring

into evidence

a coexistence

region

between a condensed and an

expanded phase,

above a

triple point

temperature Tt # 20 °C. The dramatic influence of

impurities

on the size of coexistence domains between the condensed

phase

and the

expanded

one is

clearly

demonstrated, even at a very low concentration of

impurities.

Structural and

morphological

modifications

during

the

polymerization

were

investigated using

X-ray surface scattering

together

with atomic force

microscopy.

Whatever the

polymerization

conditions

(constant

area or constant

pressure), X-ray reflectivity clearly

shows the reorientation of the

diacetylenic

links.

Only

constant area

polymerization

leads to a viscoelastic behavior of

the

film,

as shown

by

talcum decoration. The

topochemical

nature of the

polymerization

of

diacetylenic

groups induces strong constraints on the

monolayers and,

when the

polymerization

is achieved at constant area, leads to the collapse of the films evidenced

by

both

techniques.

1 Introduction.

Langmuir-Blodgett (L.B.)

films have attracted much

attention,

in

particular

with

regard

to their

potential applications [1,

2] in molecular electronics [3],

integrated optics

[4] or as

biolog-

ical sensors

[5]. However,

their

ability

to be used in manufactured devices is limited

by

their

poor mechanica1and

therma1stability,

which can be

improved by cross-linking

the molecules.

The interest of

diacetylenic

groups in this context was demonstrated

by

the

pioneering

work

(*)

to whom correspondence should be

addressed;

Email:

daillant©amoco.saclay.cea.fr.

(**)

Units Assoc16e CNRS 1379.

(3)

of

Wegner

[6]. In addition to the

expected

mechanical

strength, polydiacetylenic Langmuir- Blodgett

films have been revealed to exhibit

fascinating

non-linear

optical properties [7].

Of

course, the exact

properties

and the

quality

of the L-B- films

critically depend

on the

polymer-

ization process at the air-water

interface,

which is conditioned

by

the

thermodynamical

state of the

film,

and in turn modifies its elastic

properties.

The most

important

outcome of

Wegner's

work was the

recognition

of the

topochemica1na-

ture of the

polymerization

reaction which can

only

occur under severe

packing

constraints

[6, 7].

These ideas have been

exemplified recently

in vesicle

forming

solutions of the

diacetylenic

10,12-tricosadiynoic-sn-glycerc-phosphocholine,

where the

polymerization only

occurs in mi- crotubules

consisting

in

strongly

curved

lipidic

sheets at

temperatures

below the

melting

tran- sition

[8-12].

In the

present study,

we have

investigated

the

possibilities

offered

by

the rich

polymorphism

of this

lipid

on the

polymerization

of a film at the air-water interface. Di-

acetylenic lipids (fatty acids, phospholipids,...)

have

already

been studied at the air-water interface

[13-18].

Beside

thermodynamic studies,

earlier works

generally

dealt with either the

structure of L-B- films

[19],

the

comparison

between the molecular

organization

before and

after

polymerization [16, 17],

or the

optical properties,

and in

particular

the color

changes

of the

polymer [20-23].

A more

profound understanding

of

monolayer

structure has been reached

during

the last few years

owing

to the

development

of new

experimental techniques

such as

epifluorescence microscopy [24, 25]

or

microscopy

at the Brewster

angle [26, 27],

Atomic Force

Microscopy (A.F.M. [28-30],

and

grazing

incidence

X-ray scattering [1, 31-38].

Fluorescence

microscopy

which allows one to

image

a

monolayer directly

on water,

X-ray reflectivity

and surface

scattering

which

yield

both the normal structure and the

in-plane morphology

of a

film, complemented by

A-F-M-

experiments

which

give

access to the local

organization

of the

monolayer,

have been used

together

in this

study.

The paper is

organized

as follows. After an

experimental section,

we first establish the com-

plete phase diagram

of the

non-polymerized phospholipid

on

water, by

simultaneous isotherm

measurements,

fluorescence

microscopy

and

optical microscopy

between crossed

polarizer

and

analyzer.

We then

report

the structure

(thickness

and

roughness)

of each observed

phase

as ob- tained

by X-ray reflectivity

measurements.

Finally,

we

study

the

polymerization

process under various

experimental

conditions

by combining A-F-M-, X-ray reflectivity

and surface

scattering

measurements, in order to compare structure and

elasticity

before and after

polymerization.

2

Experimental

section.

2.I ISOTHERM MEASUREMENTS AND FLUORESCENCE MicRoscoPY. The

phospholipid

10,12-tricosadiynoic-sn-glycerc-phosphocholine,

named

DCB,gPC

has been

purchased

from Avanti Polar

Lipids

Inc. It was used without further

purification

and was dissolved in a

chloroform solution at a concentration of 0.6

g/I.

A

microsyringe

was used to

carefully spread adequate

amounts of this solution on water. As the two

diacetylenic

tails are linked

by

a

phosphocholine head,

this

phospholipid

may build a twc-dimensional network

[16]

whereas

diacetylenic fatty

acids can

only

form linear

polymers

in a two-dimensional

crystal.

The isotherm measurements were

performed

in a clean room,

using

a Lauda film

balance,

in which the surface pressure is

directly

measured

by

an induction

dynamometer,

with a sen-

sibility

of o-I

mN/m.

The pressure measurement device was calibrated with a known mass and

by

a known transition in arachidic acid. The balance was filled with

ultrapure water, purified by

a

Millipore Milli-Q

filter

system (specific resistivity

> 18.2

M~.cm; pH

=

S-S).

The

temperature

of the

trough

was

regulated by

a circulation of water with an accuracy of o_i Oc, The

trough

was cleaned with sulfochromic

acid,

acetone, and

multiple rinsing

with

(4)

pure water. The surface pressure isotherms were obtained

through

continuous

compression

at a rate of 0.02-0.05

nm~ /molecule/min.

The films were

polymerized

at the water surface under inert gas

(N2) atmosphere, by

irra- diation with U-V-

light (~

= 254

nm).

The U-V- source was located 5 cm above the surface

and the size of the films was about 10 x 20

cm~; quasi-uniform

irradiation of the film.

Poly-

merizations were

performed

at different pressures; this allowed

keeping

either the pressure or the area constant

during

the

polymerization

process.

Fluorescence

microscopy

was

performed

with a

Riegler

and Kirstein

Langmuir trough,

dis-

posed

under a Reichert

Polyvar metallographic microscope. During

the

observations,

the surface tension was measured with a

Wilhemy plate,

after calibration

using

a known tran- sition

(L2 -LS)

of arachidic acid. The fluorescent

probe, 2-(12-(7-nitrobenz-2-oxa-1,3-diazol-4- yl)amino)-dodecanoyl- I-hexadecanoyl-sn-glycero-3-phosphocholine,

was

purchased

from Molec- ular Probes and dissolved in the solution of

phospholipid

at a ratio of I to 3

il. Optically anisotropic

domains can also be visualized between crossed

polarizer

and

analyzer using

the Reichert

microscope.

The

images

were taken with an Harnamatsu

amplifier

followed

by

a CCD

camera.

In order to characterize

qualitatively

the

in-plane elasticity

of the

monolayers (polymerized

or

not),

we

performed

after each

experiment

a talcum decoration of the

layer;

the way the talcum

deposited

on the

layer

behaves under a

gentle

blow of air

depends

on the character of the

monolayer: liquid-,

solid- or

polymerized-

like.

2.2 ATomic FORCE MicRoscoPY EXPERIMENTS. In order to

perform

A-F-M-

experi-

ments,

monolayers spread

on water were

vertically

transferred onto a silicon wafer. The

(l10)

silicon wafers were

previously

cleaned in toluene in an ultrasonic bath and then in an ozone

atmosphere

in a chamber filled with oxygen under U-V-

light.

The silicon wafers were verti-

ca1ly brought

out of the water so that one

layer

of molecules was

deposited

onto the substrate.

A.F.M. was

performed

with a

Nanoscope

2

microscope

from

Digital

Instruments. The

images

were obtained

using

a

piezoelectric

scan

head,

with a scan range of 135 x 135

~lm~

and a

tip

attached to a cantilever of

spring

constant 0.54

N/m. During

the data

acquisition,

the

repul-

sive force between the

sample

and the

tip

was

kept

constant

by

vertical

displacement

of the

piezoelectric

head.

Typical

forces used were of the order of

10~~

N.

2.3 SURFACE SCATTERING OF X-RAYS.

2. 3. I

Principles

of the measurements.

Grazing

incidence

scattering

of

X-rays

has proven to be a

powerful

tool to

study

the structure of interfaces. In

particular, grazing

incidence

X-ray

diffraction and the so-called

reflectivity technique,

which consists in

measuring

the ra-

tio of reflected to incident

intensities,

have been

recently developed

in this context. Whereas

grazing

incidence

X-ray

diffraction can

yield

a determination of the

in-plane

structure of e.g.

a

Langmuir film,

the structure normal to the interface can be determined

by reflectivity

mea-

surements.

For

X-ray wavelengths (cs

0.I nm

),

the refraction index of matter is

given by

n = I

ill [39, 40],

where

fl

is

proportional

to the linear

absorption

coefficient and to the electron

density

of the mater1al

[40].

For water and the

CuKoi

radiation

(A

= 0.154

urn), fl

= 0.0126 x

10~~

and = 3.56 x

10~~.

One should note that since the

frequency

of the

X-ray

radiation is much

larger

than the atomic transition

frequencies

of the elements

composing

water or

organic materials,

electrons can be considered as free

particles [40, 41].

The accuracy of the electron

density

determination

only depends

on the resolution of the

experiment. Along

the z axis

normal to the interface

(Fig.

I

),

the accuracy of the determination is

approximately 7r/qzm~~

=

(5)

~/(4

sin

0ma~)

ci 0.6 nm, where qzma~ is the

largest

wave-vector transfer measured in the

experiment [34, 42].

We used a linear focus

(parallel

to

y) X-ray

tube.

Along

x

(intersection

of the

plane

of incidence with the

interface),

the coherence

length

is

27r/Aqx

ci 20 ~lm, where

Aqx

=

27r/~

x sin

0;A0x,

and the transverse y coherence

length

is

27r/Aqy

=

~/A0y

ci I nm.

0; is the

angle

of

incidence, A0x cs10~~

rad is the beam

angular width,

and

A0y m10~~

rad

[35].

q

k out

8j

Fig.

I.

Geometry

of the X-ray

experiments.

@I is the

grazing angle

of

incidence,

and @d the

angle

of reflection

or diffusion. q is the wave-vector transfer.

No real interface can be considered as

being perfectly

smooth with

regards

to the

scattering

of

X-rays.

The

implications

of this

point

cannot be

ignored

without

causing

serious misin-

terpretations. However,

as

long

as the

roughnesses

remain

small,

their effect on the reflection coefficient can be treated as a

perturbation

of the case of a

perfect

interface which will be exam- ined first.

By perfect interface,

we mean a

perfectly

smooth interface which can be described

by

a

density profile p(z),

and therefore does not

give

rise to

off-specular

surface

scattering.

Since the real

part

of the index is less than one, total external reflection occurs at

grazing

an-

gles

of incidence below a critical

angle

0c ci

@ (0c

= 2.67 mrad for

water).

For a

single diopter

the reflected

intensity

decreases above 0c

according

to the Fresnel law

(RF (0)

=

(0 /20c )~,

0 »

0c [36] ).

In the case of a

Langmuir film,

the

monolayer

can be

decomposed

into N

chemically

homogeneous layers

which we model as slabs of constant

density

of index n~ located between z~-i and z~

[37].

The

shape

of the

reflectivity

curve results from the interferences between the beams reflected at each interface. The reflection coefficient can be calculated either

exactly, using

iterative methods

(this

is the method used in this

paper),

or, more

conveniently,

within the Born

approximation ignoring multiple

reflection effects. In the latter case

[33]

N

~(Qz)

"

~F(Qz) ~ (ll~+1 ll~)(ll~+1 ll~) C°S(Qz(Zi Zj)) (I)

i,j=0

As stated

above,

no real interface can be considered to be

perfect

at the

X-ray wavelength

scale. The

system

cannot in fact be described

using

a

simple density profile p(z),

and a small

part

of the incident

intensity

is scattered out of the

specular

direction. In the case of

amphiphiles spread

at the air-water

interface,

the

origin

of this

scattering

lies

precisely

in the most

fascinating aspects

of those

systems,

their

phase

transitions and their surface fluctuations: first order

phase

transitions are characterized

by large density inhomogeneities

and surface fluctuations

originate

from the

thermally

excited

capillary

waves

governed by

the

(6)

elastic

properties

of the film. In the first case, the film must be described

by

the

in-plane density

distribution

p(x,y), just

as in the diffraction case, but at a different characteristic scale. In the second case, it is more convenient to consider the actual location

z~(x,y)

of the

previously

defined interfaces.

Using

a linear focus

X-ray

tube at

grazing incidences,

a

satisfactory

resolution is

only

achiev- able in the

plane

of incidence

(x, z),

as revealed

by

the coherence

lengths given

above. The qx

dependence

of the scattered

intensity

is concentrated in the functions

[35]:

where jz is the

component

normal to the

plane

of the wave-vector transfer in medium and

@UAqx

the half-width at half-maximum of the Gaussian resolution function

(Aqx

is of the order of 1.2 x

10~ m~~

at 0

= 30

mrad).

In the case of

height fluctuations, (~j(X)

= e~~z~iz<~~~°~'~J~~~>

[35, 38],

and in the case of

density inhomogeneities (~j(X)

=

(&p~(0)&pj (X)) /p~.

The qz

dependence

of the scattered

intensity

results from the interferences between beams scattered at the different interfaces. This

dependence

is similar to that of the reflected beam

(see Eq. (I))

except that the contrast of the interference

pattern

is modulated

by

the function

Z~j (qx,

qz

).

Within the Born

approximation (the

most accurate

approximation

used here cannot be cast in a

simple analytical

form

[35] ),

the scattered

intensity

writes:

where

ko

"

27r/~.

The

computation

of the scattered

intensity

is

given

in the

appendix

in the case of

height

fluctuations. In the case of islands on

top

of the

film,

the

intensity

can be calculated either

by considering

the

resulting

surface

morphology,

or

by attributing

these islands to an

incomplete

additional

layer [43].

In both cases, the

aggregate-aggregate

correlation

function,

which can be obtained from A.F.M.

images,

is

required

in order to estimate the

intensity.

In our case, this correlation function decreases

rapidly,

and one is left with the average size of the

domains,

which can be estimated from qx scans. The

height

of the domains is

conversely

obtained from qz scans as

explained

in reference

[44].

The accessible

wavelengths correspond

to wave-vectors

q~ =

~~

x

(cased coso~)

=

(~j

x (0~2

od~)

~ ~

where

0;

and

0d

are

respectively

the

angle

of incidence and the

angle

of the detector above the surface. The upper limit is due to the coherence of the beam and is of the order of 20 ~lm.

The limitation at short

wavelengths (a

few tens of

nm)

is

only

due to the

signal-to-background

ratio and

consequently

to the

intensity

of the incident beam

(the

diffuse scattered

intensity

is

approximately

two orders of

magnitude

below the weak reflected

bearn).

An

important

consequence of

equation (2)

is that the scattered

intensity

in the

specular

di- rection cannot be

neglected,

and even overcomes the reflected

intensity

when

elo~~l~f~+l~~))Z~j (q~

=

0,qz)

> I

(typically

for 0; > 30 mrad for a

liquid

surface and ~ = 0.154

nm).

The estimate of the

intensity

scattered at qx = 0 does not

generally

lead to

simple analytical

forms

(an example

is

given

in the

appendix

for a

liquid

surface

[32, 35]).

In any case, the result is not a

simple

Gaussian attenuation as

generally

assumed. It follows that any

attempt

to fit

"reflectivity

curves"

by including only "Debye-Waller"

factors in

equation (I)

to

give

account of the interface

"roughness" necessarily fails,

and leads to

non-physical

determinations. A

realistic estimate of the effect of surface

scattering

must be included in the calculations.

(7)

2.3.2 Atomic force

Jnicroscopy

and

X-ray scattering.

Atomic force

microscopy

and X-

ray measurements have been used

together

in this

study,

and it is worth

comparing briefly

the main characteristics of both methods. The

complementarity

between

X-ray

measurements and atomic force

microscopy experiments

is

striking

at different

levels,

and stems from the fact

that,

whereas A-F-M- measurements are

local, X-ray experiments (see

e.g.

Eq. (2))

are

fundamentally

non-local. An

important

consequence is that

X-ray

results are often more con-

veniently

viewed in Fourier space

(though

the exact result for the scattered

intensity equation (2)

is obtained in terms of

height-height

correlation

functions).

Whereas individual

objects

are

imaged

in one case, statistical information is obtained in the other case and both determi- nations

yield

a

complete

and consistent

understanding

of the system.

Moreover,

whereas the

accuracy achieved in A.F.M.

experiments

is

subject

to the

quality

of the

calibration, X-ray

results are absolute. An accurate calibration is much easier to obtain for

in-plane lengths

than for thicknesses in A-F-M-

experiments owing

to the existence of well

adapted rulings

at any scale of interest. In contrast,

excluding

diffraction

effects,

the

ability

of surface

scattering

of

X-rays

to

yield

valuable

in-plane

information at a scale smaller than 100 nm is

actually

limited

by

the

intensity

of the available sources.

Finally,

as mentioned

above,

different

scattering

sources can be

present

in the

system,

either structural or related to its fluctuations.

X-ray

measurements allow the determination of elastic

parameters

of the film from its

fluctuations, using

the method

developed

in the

appendix.

This method is however

only applicable

if

capillary

waves are the

only

source of

scattering

in the

system.

This can be

directly

checked

by

A.F.M.

experiments

on films

deposited

on a

substrate,

which allow a

separate investigation

of the

morphological part

of the

roughness.

This

method,

used

throughout

this

work,

is

particularly interesting

in

dealing

with

polymerized

films since

one may

expect only

minor structural

changes

upon

deposition

in this case.

2.3.3

Experimental

details. For surface

scattering experiments,

extreme care has to be taken in order to diminish the

background.

To this purpose, we used a

Si(I

II monochromator

giving

a low

divergence(<

0.I

mrad).

The monochromator was followed

by

antidiffusion slits and the

CuKai

radiation

(~

= 0.154

nm)

was selected

by

a 100 ~lm wide slit

just

before the

sample (0.4

m from the

source).

The beam was also limited

by

a 1.25 mm wide vertica1slit and the

analysis

slit was 200 ~lm

wide,

thus

leading

to an

angular

resolution of 8.7 x

10~

sin

0dm~~

Under these

conditions,

the beam was

perfectly

Gaussian with a

background

level

of10~~Io,

that is to say about 0.I count

Is (lo

is the incident beam

intensity).

The

trough

used for

X-ray experiments

is home-built. The surface pressure is measured

by

a

Wilhemy plate

and is

continuously

recorded

during

the

experiment.

The

compression

barrier is made of a

unique

teflon

ribbon,

in order to minimize leaks. In order to avoid surface

vibrations,

the water

layer

is

only

3 mm

deep,

and the most

important point

is that the water level is

kept

constant

by displacement

of an

auxilliary

reservoir

during

the m 12 h

long experiments.

The

reflectivity

curves were recorded

by performing rocking

scans around each

point

to determine the

background.

The best fit was determined as the absolute minimum of the standard error deviation

x~

and the error bars are deduced from

x~

-

x~

+1.

3. Results.

3 .I PHASE DIAGRAM. Isotherms ofthe

non-polymerized phospholipid

recorded at different

temperatures

between IS °C and 40 °C are

presented

in

figure

2. Two different kinds of isotherms were obtained

depending

on the

temperature.

Below a

triple point temperature TT

= 20

°C,

a direct transition from a gaseous state to a condensed state occurs, whereas

above 20 °C a coexistence

plateau

between an

expanded phase

and the condensed

phase

is

(8)

observed. This value of the

triple point temperature

is rather low for this

long

chain

(23

carbon

atoms) phospholipid

and is

comparable

to the one of a

fully

saturated

phosphocholine

like

dipalmitoylphosphocholine (DPPC)

with

only

16 carbon atoms per chain

[45].

Let us

note that the pressure of the coexistence

plateau

is

always slightly higher during

the first

compression.

As

opposed

to saturated

phospholipids,

we did not observe a transition to a solid

phase

at

higher

pressures

[45].

The area per molecule in the condensed

phase,

which does not

depend

on the

temperature

at

high

pressure, is about 0.50

nm~/molecule.

This area per molecule is

slightly higher

than in the condensed

phase

of a

fully

saturated

phosphocholine (like DPPC),

I-e- about 0.45

nm~/molecule [45].

The

collapse

of the

monolayer

occurs at a pressure of 43

mN/m.

The coexistence

region

seems to end up near 40

°C, possibly

at a tricritical

point,

which is

unfortunately

very difficult to locate

precisely.

The

monolayer

is very stable

throughout

the

phase diagram

and can be maintained at a

high

pressure

overnight

without any loss. These isotherms are not very different from those of a saturated

phospholipid

and the main differences

(absence

of a solid

phase,

a

comparatively larger

area per molecule in the condensed

phase

and low

triple point temperature)

can be

directly

attributed to the presence of the

diacetylenic

group, the size of which

(and

the related

kink)

may affect the

packing

of the molecules.

40

-

~

'

)

~~ ~

cH,~.CH,b~c*c-mc-~R~-I-o-

H,

_ CH,~lcHib~c-c-c-Hlcl~~~-O-o

f

H

f pi

(

~~~ ~ ~~ ~

~,~

~~

(

©

l$

lo

~

o

o 5 o.75 1.z5 1.5 75

area/molecule (nm2)

Fig.

2. Isotherms of

diacetylenic phospholipid DCS,9PC (inset)

at different temperatures between IS °C and 40 ° C. The

triple

point temperature is about 20 °C.

Fluorescence

microscopy experiments

were

performed

in order to

complement

the isotherm

measurements. The three above mentioned

phases

were

clearly

shown. At

large

areas, the

homogeneous

gas

phase

is observed. At 2.0 + 0.2

nm~/molecule,

a transition occurs to a

phase

which

depends

on the

temperature.

Below 20

°C,

the

growth

of

angular

domains of the condensed

phase

is observed

(Fig. 3a)

until the gas

phase disappears

at 0.7

nm~/molecule.

Above 20

°C,

circular domains of the

expanded phase

are first observed to grow in the gas

phase (Fig. 3b).

From 1.2

nm~/molecule,

the

monolayer

is in the

homogeneous expanded

(9)

I) aj~)

a)

b)

Fig-

3. Fluorescence

microscopy images

ofthe

monolayer

of

DCB,9PC.

The lateral size of the

images

is 250 ~tm;

a)

coexistence between the gas

phase

and the condensed

phase

below the

triple

point temperature;

b)

coexistence between the gas

phase

and the

expanded

phase above 20

°C; c)

coexistence between the

expanded phase

and the condensed

phase

above 20

°C; (d) inhomogeneous

dilute state of the

monolayer

just after the

deposition

of the solution.

(10)

4

;

S ~ -'

7 )

-G'

~

'>" ~ ~' Y

"'. '"'

'( )

C)

Pi'(,»

<Sl

'W

'1

~

j'

ill O

~ 4~

n

~,

. ' ~~

'

#

~

,

~ ,

.,w

,

,, ~.

-,

Fig.

3.

(contin~ed)

state, until the curved and branched domains of the condensed

phase (ci

50 to loo ~lm

long)

appear on the

plateau (Fig. 3c).

Similar

domiins

have been observed with other

diacetylenic

phospholipids [46].

These

strongly

counterclockwise curved needles grow under

compression.

At

high

pressures the condensed domains

get

closer and

closer,

and fuse

only

at very

high

pressure.

(11)

Highly

contrasted

pictures

of these domains can also be obtained between

crosied polarizer

and

analyzer

without any

probe.

The

striking

contrast of these

images

can be attributed to the

polarizability

of the

diacetylenic

groups. The

L2

LS transition of arachidic acid has been shown

by

the same

method,

but with a poor contrast. It is nevertheless very

surprising

that

a

monolayer

of a saturated

fatty

acid can be visualized

by

this very

simple

method and this fact

has,

to our

knowledge,

never be

pointed

out. We have to

specify

that the

monolayer

is illuminated with a

polarized

conical

light beam,

which is therefore not

exactly polarized

within the

plane

of the

monolayer; nevertheless,

we have checked that the same

images

are obtained when the

monolayer

is illuminated

by

a

parallel

beam. Almost white branches as well

as

completely

black branches can be observed for a

given

orientation of the

polarizer (Fig. 4).

This observation

provides

evidence for

long-range

orientational order in the condensed domains.

Neither the

expanded

nor the gas

phase

can be observed

by

this method. A very

striking point

is the difference in size between the domains observed

by

fluorescence

microscopy

and those observed between crossed

polarizer

and

analyzer

when no

probe

is added

(compare Fig.

3c with

Fig.

4a and note the difference of

magnification ).

In the latter case, domains are

larger

than a few millimeters and a whole domain cannot even be observed within the field of the

microscope objective

of smallest

magnification.

Note that domains observed

by

fluorescence

microscopy

and between crossed

polarizer

and

analyzer

are

identical,

when a

probe

is added.

Such an effect of

impurities

on the domain size of stearic acid has

already

been observed

[47],

but was never as dramatic as in this case.

Finally

the first

compression displays

distinct features.

Right

after

deposition,

a very in-

homogeneous film,

in which bubbles of gas in domains of the

expanded phase,

bubbles of the

expanded phase

in gas, and

large

lamellae of gas or

expanded phase

are observed

(Fig. 3d).

This

inhomogeneous

state of the film is due to a bad

spreading

of the solution of

phospholipids

on water. After the first

compression,

these structures never appear

again.

In many

experiments,

a

bump

is

apparent

at the

beginning

of the coexistence

plateau (Fig.

5).

This

bump

is often observed with

diacetylenic compounds [13, 16-18, 46, 48].

In the

case of the

DCB,9PC,

one can

point

out the

following

facts. The

bump

is

only

observed upon

compression,

whatever the

compression

rate, and never observed

during

the

decompression.

When

only

o-I

lt

of

cholesterol,

which is known to lower the line tension

[49],

is added to a fresh solution which does not exhibit a

bump,

then the

bump

is observed

(Fig. 5c).

If a

bump

is

already present,

there is no variation of its

amplitude

for a concentration in cholesterol

varying

between o-I

lt

to 5

lt. Lastly,

a

bump

is observed after

storing

for one week a solution which did not

initially

exhibit a

bump.

It is therefore obvious

that,

in our case, the presence of the

bump

is

directly imputable

to

impurities.

At least two kinds of

impurities

can be found in

diacetylenic compounds. Firstly,

the

phospholipid

which is used as

purchased

contains some

lyso derivatives,

and

secondly

some

polymerized aggregates

may also be

present

in the

sample.

As these

impurities

should also exist in the other

diacetylenic compounds,

we can assess that the

bump

present at the

beginning

of the coexistence

plateau

of

diacetylenic amphiphiles

is due to the presence of

impurities.

Such a

bump

has also been observed with other

compounds,

e-g- NBD stearic acid

[50, 51].

In

this case, it has been attributed to the

necessity

of

overcoming

the line energy contribution when

nucleating

a domain

[50].

It therefore appears

paradoxical

that the addition of

impurities

leads in our case to the appearance of a

bump. Indeed, usually impurities

are

supposed

to lower line

tension,

thus

eliminating

any transition

lag.

In this

regard,

we

effectively

observe a dramatic increase of the nucleation rate upon addition of

impurities,

as is evidenced

by

the

previously

described

epifluorescence

and

polarized light microscopy images.

This is

fully

consistent with the fact that the line tension is reduced

by

the presence of

impurities

or cholesterol

[50].

The presence of a

bump

does not mean that there is a transition

lag

but must have a more subtle

(12)

/

f ~ /~

~~

~'

W

~

'~~ ~

~

~_~j-~~~'~z *~~~

' '~'

.~~'~

~~~'~~~~

~

~

~.j'~w'~

~(

.

r

'_' '

= ~

?i" ~~' m.

-. ~~-i )

~ £ %~ ~ ., '~

~

%' ~~

~~~ "- _~' Wm" S .~_~~%

~w'Kc i''~ ~ ~ ~ml"~ '~

~~~~/~ ~~$ ~i$~~~~(/.

'

/~

~

,

'jf~

j. j@ W,

~~

~~

iii lji~~~

~ ~' ~

(~_

_~~j@~'

[[~

jw~)_jZ~

~°?~'~~~"

>

'- jjj~s~f$[~[ '~

.~ V ,i~jo~j/[~~

k f /~ ~~~r ~ < ~

~ " ~. i $ ~c"4 >~'~ '"

*~

~j _ww r 'o

'~~ W~ /h~'~'

% ~ i~~ "1 ~

@

if1°S_Tj~

,

S

~Q#,~~~i~, )j#~~f I

'~

/£j

w~~

i~~

~ ° ~'

/ W"~q~ fl~w .-+~ /.

+~?'

g,'. f ~

b)

C)

Fig.

4.

Images

between crossed

polarizer

and analyzer of the monolayer on water; the size of the

images

is 2 mm.

a)

center of a domain; note the dramatic size difference of the domains

compared

to

figure

2c;

b) extremity

of one branch of the

domain; c)

aspect of the

monolayer

close to the

collapse

pressure in the

homogeneous

condensed

phase.

origin. During

the

compression

of the

monolayer,

the

system

is out of

equilibrium

and one

must take into account both nucleation and

growth

of the domains. A more

quantitative study

of this

problem

is

currently

under way.

X-ray reflectivity

measurements have been

performed

in order to

investigate

the structure of the

expanded

and condensed

phases (Fig. 6).

As

explained

in section

2,

the

experimental

(13)

30

~ d

b o

~

'

)

I

j

I

w

~

I

l$

~

o

0

Fig.

5.

Bump

at the

beginning

of the coexistence

plateau.

All the isotherms are measured at 25 °C

and are shifted by 2

mN/m

and 0.04

nm~

for

clarity. a)

isotherm obtained with

a fresh

product, exhibiting

no

bump; b)

isotherm of

DOB,9PG exhibiting

a

bump; c)

and

d)

isotherm obtained with the same solution as in

a),

in presence of

c)

0.li~

cholesterol; (d)

5i~ cholesterol.

o-~

o-~

o-~

-<

o-~

0'~

0'~

~

~~-s

0 10 20 30 40 50 60

6

(mrad)

Fig. 6.

Experimental

reflectivity curves and best fits for the

DCB,9PC monolayer

in the

expanded phase (filled

squares: II

= 5

mN/m)

and at two surface pressures in the condensed

phase (hollow

circles:

II = 18

mN/m

and filled circles: II

= 25

mN/m).

The temperature is 22.5 °C. The parameters of the fits are

given

in table I. Note that both reflectivity curves in the condensed

phase

were best fitted with identical parameters, except the surface tension, which is found to be

equal

to the value measured with the

Wilhemy plate.

(14)

data are

analyzed by fitting

with a model which consists of a stack of

chemically homoge-

neous larnellae. The data for both the

non-polymerized expanded

and condensed films can

be described

by using only

two

lamellae,

one for the

chains,

and one for the heads.

However,

we shall see that this is not

possible

after

polymerization

due to the

high

electron

density

of the

diacetylenic region.

This led us to use an identical model before

polymerization,

in order to

quantify

the structural

changes

induced

by

the reaction. We therefore use four

layers

for the

description

of the film

[17]:

one for the

heads,

two for the

alkyl parts

of the

chains,

and

one for the

diacetylenic region,

each characterized

by

a

length

and an electron

density.

In

addition,

the fluctuations of the film are taken into account as described in the

appendix (Eq.

(A.3)), yielding

an estimate of the

bending rigidity

modulus K. Of course, this more detailed structural

description

leads to

comparatively larger

error bars.

Table I. ParaJneters of the

DCB,9PC monolayer

at 22.5 °C before

polyJnerization

in the condensed and

expanded phases,

and after

polymerization

at constant pressure, at constant area, and on the coexistence

plateau.

p is an electronic

density.

"I,chains

(total)"

is the total

length

of the

chains,

'~l,

up"

the

length

of the upper saturated

region

of the

chains,

"I.

diacetyl."

the

length

of the

diacetylenic region,

"I. heads" the size of the heads and ~f and K the surface tension and the

bending rigidity

modulus.

PV PCh8b*

~~~~ ~n~j) ~~) ij~ i~~ mt~lm)

(*~T)

(+1-0 (+1-0 (+1-0.2) (+1-0 2) (+1-0.2)

~@~~$~

0 9 1.18 1.6 0.5 0.067 2

conden3ed

phase 0 93 05 1. 63 2. 09 0. 3 0. 34 0. 055 40

polynedzanon

at 0 94 3 1.26 2. is 0. 9 0. 4 0. 4 0. 030

con3tant area

polynedzaflon

at o 91 13 07 75 2 0.3 0 2 0 050 40

polvnedeanon

blthe 0.93 2 0.9 2.I 3 0.55 0.17 0.046 25

The fit

parameters

are summarized in table I. In addition to a small difference in the chain

densities,

the condensed

phase

can

mainly

be

distinguished

from the

expanded phase by

the

larger

chain thickness. The

expanded phase

is characterized

by

a total

length

of the chains of about 1.6 nm, which is

comparable

to the thickness of the

aliphatic

medium in the

liquid

expanded phase

of

DPPC,

the chains of which are shorter

by

seven carbon atoms

[34].

The thickness of the condensed

phase (2,1+

0.2

nm)

is much smaller than the

length

of a

fully

(15)

extended

10,12 tricosadiynoic-sn-glycerc-phosphocholine

chain

(ci

2.9

nm).

It should also be noted that better fits are obtained when

slightly larger

values are allowed for the

diacetylenic density

as

compared

to the

alkyl density.

The location of this denser

region

is

exactly

what was

expected

from molecular models. Another most

interesting point

is the

larger

value obtained for the

bending rigidity

modulus in the condensed

phase (ci 40kBT).

A similar result was obtained in an earlier

study

of

phospholipidic compounds,

and could be

expected

from the lower

compressibility

of this

phase [34].

3.2 POLYMERIzATION oN WATER. In order to

investigate

the

possibilities

offered

by

the

polymorphism

on the

polymerization,

the

diacetylenic phospholipid

was

exposed

to U.V. ra- diation both in the

expanded

and in the condensed

phase, keeping

either the pressure or the

area constant, as well as within the coexistence

region.

3. 2, I

Polymerization

in the

expanded phase.

There is no evidence of any

polymerization

in the

expanded phase.

Isotherms before and after exposure to U.V. radiation are

identical,

and talcum decoration shows that the film remains as fluid as the

original expanded

film. This is consistent with the observation that

DCB,9PC

in solution does not

polymerize

at

temperatures

above the chain

melting

transition and is

clearly

related to the

topochemical

nature of the

reaction

[6-12].

30

~

'

)

j

fl I

w

~

@

l$

~

o

0

Fig.

7. Isotherms after

polymerization. a) polymerization

at constant area;

b) polymerization

at constant pressure.

3.2. 2

Polymerization

in the condensed

phase

at constant area. The

monolayer

was first

polymerized

at constant area.

Upon

exposure to U.V.

radiation,

the pressure increases

during

ci 3 min up to 43

mN/m

and then remains constant. It should be noted that if the

trough

is not filled with an inert gas, the pressure

ultimately drops

because of oxidation of the film.

(16)

0

a 40

~

i

b

~ l

g

i j

I O0

~ d

I ,

o

~

«

-5

~

~ ~

'i

4o

.~

io-~

10~~

' O

"'

0 10

0 mrud)

Fig.

8.

Experimental

reflectivity curves obtained on

polymerized

films and best fits.

a) polymeriza-

tion at constant area;

b) polymerization

at constant pressure;

c)

and

d) polymerization

at two

points

on the coexistence plateau. Each curve is

displaced by

x10 for

clarity.

The parameters of the fits are

given

in table I. The initial conditions before

polymerization

are shown in the inset.

We did not notice any difference between

polymers

obtained above and below the

triple point temperature, demonstrating

that

only

the

phase,

I-e- the molecular

organization,

is

important.

The isotherms after

polymerization

do not exhibit any further transition

(Fig. 7a).

Talcum

deposited

onto the

monolayer

moves

slowly

under a

gentle

blow of air and then moves back.

The

monolayer

moreover appears to be very viscous. The

reflectivity

curves obtained on those

monolayers

are

presented

in

figure

8a. The most

striking change

is the

disappearance

of the

large

destructive interference. This can be

directly

attributed to the 25

%

increase of the

diacetylenic

lamella electron

density,

which appears to be the structural

signature

of the

polymerization.

It can be traced back to a reorientation of the

diacetylenic

links in the connected network

(Fig. 9).

Let us also note a

slight

increase of the chain

length

and of the

alkyl

electron

density (Tab. I).

In order to

get complementary information, monolayers

were

transferred onto a silicon wafer before and after

polymerization,

and

imaged by

atomic force

(17)

III

p p

/ /

/ /

P P

d fl

uv

Fig.

9. Schematic of the reorientation of the

diacetylenic

links upon

polymerization.

The reorien- tation has for consequence the increase of the

z-projected density

of the

diacetylenic region,

measured

by X-ray reflectivity (see

Tab. 1).

~

t

» J~

a)

b)

Fig. 10. - A-F.M. images of the

monolayer

onto a silicon wafer. a) before

the size

of

the image is 80

~m

x 80 ~m.

Note

the

branches reminiscent of

those observed

in

b) after the size

Références

Documents relatifs

10.Fire of the Holy Spirit, cancel and completely destroy all the covenants that bring evil things around my spirit, soul, body and family in the Name of Jesus. 11.I cancel and

Swiss Centre for Antibiotic Resistance (ANRESIS) network: Institute for Laboratory Medicine, Cantonal Hospital Aarau; Central Laboratory, Microbiology Section, Cantonal

Microscopic and molecular structures of ω - and γ -gliadin monolayers at the air - water interface were studied under compression by three complementary techniques:

In this work, we present a study of the influence of TMAO and urea on the structure of water using X-ray Raman scattering (XRS) spectroscopy at the oxygen K-edge of aqueous solutions

The information obtained with this short study of the phase structures of brine-oil-ACT mixtures at low ACT concentrations and close to the optimal salinities are in agreement with

Abstract, Grazing incidence X-ray diffraction is performed on a Langmuir monolayer made of pure fluorescent NBD-stearic acid, spread at the free surface of water.. It shows

Organic molecular crystals with bulk electronic excitons of the Frenlcel type, may present also monolayer surface and sub-surface collective excitations (Site Shift

In figure 4 we present a summary of the temperature dependence of the real surface tension, the relative damping coefficient, and the area per molecule for both a