• Aucun résultat trouvé

Self-organization of discogenic molecules at the air-water interface

N/A
N/A
Protected

Academic year: 2021

Partager "Self-organization of discogenic molecules at the air-water interface"

Copied!
12
0
0

Texte intégral

(1)

HAL Id: jpa-00247614

https://hal.archives-ouvertes.fr/jpa-00247614

Submitted on 1 Jan 1992

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.

Self-organization of discogenic molecules at the air-water interface

Nicholas Maliszewskyj, Paul Heiney, J. Kent Blasie, John P. Mccauley Jr., Amos B. Smith Iii

To cite this version:

Nicholas Maliszewskyj, Paul Heiney, J. Kent Blasie, John P. Mccauley Jr., Amos B. Smith Iii. Self-

organization of discogenic molecules at the air-water interface. Journal de Physique II, EDP Sciences,

1992, 2 (1), pp.75-85. �10.1051/jp2:1992114�. �jpa-00247614�

(2)

Classification Physics Abstracts

68.10G 82.65D 61.30

Self-organization of discogenic molecules at the air-water interface

Nicholas C.

Maliszewskyj II),

Paul A.

Heiney (~),

J. Kent Blasie

(~),

John P.

Mccauley,

Jr. (~) and Amos B.

Smith,

III

(2)

(1)

Department

of

Physics

and

Laboratory

for Research on the Structure of Matter,

University

of Pennsylvania,

Philadelphia,

PA 19104, U-S-A-

(2)

Departrnent

of

Chemistry

and Laboratory for Research on the Structure of Matter,

University

of Pennsylvania,

Philadelphia,

PA 19104, U-S-A-

(Received it

September

J99J,

accepted

27

September

J99J)

Abstract. The behavior of several

species

of

discogenic

molecules at the air-water interface has been determined

by measuring

surface pressure vs. molecular area isotherms.

Alkylthiotripheny-

lene derivatives do not appear to form

monolayers

at the interfaces, due to lack of

amphiphilic

character. Two truxene derivatives were found to

display

molecular areas on the order of 65 A2,

most

likely

due to the molecules lying « edge-on » to the water surface.

Hexacyclens generally

lie with the core group

parallel

to the surface of the water with the tail groups

extending

away from the interface ; a higher density

phase

may

correspond

to conformational changes within the

molecule.

1. Introduction.

The

organization

of

amphiphilic

molecules such as

lipids

and soaps into

monolayers

on water surfaces is a well known

phenomenon [I]. Recently,

it has become clear

[2-7]

that

Langmuir monolayers

can also be formed from the

disc-shaped

molecules of which discotic

liquid crystals [8]

are

composed.

Such molecules

typically,

but not

universally,

have flat,

conjugated

cores and 4-8

aliphatic

tails. The bulk

liquid crystalline phases

may be nematic («

ND »)

or

hexagonal-columnar

Dhd »),

the latter structure

consisting

of a

hexagonal

array of columns

of molecules. The column axis is

approximately

normal to the molecular core and the

aliphatic

tails extend more or less

radially

from the column axis.

By

contrast, surface pressure isotherm measurements have shown that several

quite

different conformations are

possible

for

discogenic

molecules at the air-water interface.

For

phthalocyanines [3],

benzene hexa-n-alkanoates

[2, 5],

and

hexacyclens [6, 7],

the

molecular areas

extrapolated

from measurements of

spreading

pressure 17 versus area

A are consistent with a model for a condensed

phase

in which the core lies flat in the

surface,

with the

aliphatic

tails

projecting

away from the surface.

By

contrast, the measured molecular

area of several

alkoxytriphenylene

derivatives is much smaller than the area of such a

(3)

X X la

X,R

m

SCSH,,

16

X,R

-

SC~H,~

lc

X,R

=

SC7H,s

id X ~

COO(CH~)~CH~

R m

OCSH,,

1e

X,R

=

OCSH,,

R

2a R =

COO(CH~),~CH~

2b R

m

COOPhO(CH~),OCH~

R R

R

~~'~~

f~~'h

O

R

~N ~j

a R =

C6H40C,~H~

R~O

b R

"

C6H40C~H~

~~"~ c R

=

C~HS

R d R

«

C17H35

R

~s~/~ eR=C~H~OC,,H~~

~~

~

~/~~

~~~~~~~

~

~~

~

~'

~N ~j ~

j

R

~ 4a.d

Scheme 1.

(4)

molecule

lying flat, implying

that these molecules most

likely

are oriented «

edge-on

» to the air-water

interface,

with at least two tails

penetrating

into the

subphase [4, 7].

Ther-

modynamic

transitions between different conformations may be

possible;

in

particular,

benzene hexa-n-alkanoate isotherms have been

interpreted

both in terms of a structure for which the core lies

parallel

to the interface with the tails close

packed [5],

and in terms of a transition in which the entire molecule tilts from a conformation in which the core lies

parallel

(« flat

»)

to the interface to one in which it sits

perpendicular

edge-on »)

to the interface at

high compression [4].

We have measured surface pressure vs. molecular area isotherms at room

temperature

for three classes of

thermotropic

discotic mesogen as shown in Scheme I

hexacydens,

truxenes, and

triphenylenes.

Thermal data and literature references for these

compounds

are

given

in table I. The

goal

was to further

explore

the role of molecular architecture in

determining

both

liquid-crystal-phase

structure and structure at the air-water interface.

Small

quantities

of solute

(-10~~g)

were

weighed

on a Cahn electrobalance to an

accuracy of 10 ~ g. The

hexacyclens

3a-3f and 4a-4d are

quite hygroscopic [9, 10],

and for

these

compounds

it was necessary to remove water

by heating

the

sample gently

to 55 °C

under

dynamic

vacuum for at least 36 hours before

weighing

to obtain

reproducible

results.

After

weighing,

the solutes were dissolved in 10 ml of HPLC

grade methylene

chloride.

Methylene

chloride is known to be a

good spreading

solvent, immiscible in water and

quite

volatile

(boiling point -40°C) [6].

The

subphase

for H-A measurements was water

purified

in a

Millipore

filtration system. A

commercially

available Lauda film balance was used to determine the surface pressure

by

the movable

barrier/Langmuir

float method to an

accuracy of

0.lmN/m.

All measurements were

performed

at room temperature,

T

=

21 ± 3 °C. After

performing

a

cleaning

sweep of the surface, 0.1ml of the solution was

spread

on the interface and at least 10 min allowed for

evaporation

of the solvent. The film

was then

compressed

at a barrier

speed

of 0.75

cm/min,

so that the duration of an

experiment

was

approximately

30 min. All measurements were made 3 times from different solutions and the

reproducibility

was within 5 fb.

In a

typical

17- A

isotherm,

steep rises in pressure are associated with a pure

phase,

while

approximately

horizontal

regions

are associated with coexistence between two

phases.

There

is more than one convention for

determining

the molecular area from such isotherms. For

example,

the molecular area of 3a is

variously reported

as 160

Ji2 [7]

and

22512 [6],

even

though

the measured isotherms are

quite

similar. The difference lies in whether the molecular

area is defined

by

the inflection

point

of the vertical rise in pressure

corresponding

to the condensed

phase,

or

by

the onset of that vertical rise. It is the former convention to which we adhere in this paper.

As

recently

observed

[11],

17-A isotherms in

Langmuir

films are in

practice

often

performed

away from

thermodynamic equilibrium.

We and others

[2, 4, 5, 7]

have in some

cases observed a local maximum, or «

bump

», in

spreading

pressure vs. molecular area. Such

a feature is an indication that the isotherm is in fact

measuring

a

nonequilibrium

process such

as irreversible second

layer

formation or a

hysteretic

first order transition. With care,

however,

the isotherms are still useful tools for

determining

molecular surface areas close to

monolayer completion.

2. Results.

We will discuss in tum our studies of

triphenylene derivatives,

truxene

derivatives,

and

hexacydens.

Triphenylene

derivatives: The

liquid crystalline properties

of

alkoxy-

and

alkylthio- triphenylene

derivatives have been

extensively

studied

[12-15].

H-A isotherms of

JOURNAL DE PHYSIQUE II T 2, N'l,JANVJER lW2 5

(5)

Table I. Bulk transition temperatures and molecular areas at the water

su~fiace,

derived at the

inflection point ofthe first

state

ofthe

condensed

phase of

the

monolayer. References

are

given for

the appearance

ofthese compounds

in the literature. Molecular areas are measured

from

the

inflection points

in the

first

« state »

of

the condensed

monolayer phase jkom

the isotherms in

figures 1,

2, 4 and 5.

Bulk

Liquid-Crystalline

Phases Reference

(I

K1 ~ I

95 .c

i

la Kl - K2 -

D~d

-1

[27]

17

54 .c xi .c 91 °c

16 K

- H -

D~d

++

[13-15]

17

62c mc 91c

lc K

- D~~ -

[27]

21

72 x ~c 941.c

id

D~d

-

[27]

53

167c

le K - D ++

[4,12]

68

69 °c 122 >c

~~ ~

6~ ~~ 8~ ~~~ 1/C ~~~ 2~C

~~~ ~~~ ~~

2b K ++ D~d +-

ND

- D~d +-

N~

+-1

[17-23]

74

9u .c 117 c 179 °c 284 °c 297 °c

3a K - T

[6], [7], [10]

159

97 °c ml °c

3b K ++ I

[IQ]

168

125 °c

3c

glass

-

[10]

17

157 °c

3d K

+-

[10]

122

92 °c

3e K +- T ++

[10]

176

58 °c 144 °c

3f K +-

l10]

165

v6 °c

4a K

-

[10]

163

71 >c

4b K - 155

25-61 °c

4c K ++

l10]

157

118 °c

4d K +- I

[10

12

42 °c

compounds

la-le are shown in

figure

I. Our isotherm for le agrees within 6 fb with the

previous

determination

[4]

of a 7-9

mN/m

pressure and

70h2

area per molecule at the inflection

point. Compound

ld

gives

a similar molecular area, with a smaller

compressibility

in the condensed

phase.

This

supports

the

hypothesis

of an «

edge

on » arrangement for both

compounds.

The

alkylthiotriphenylenes (la-c)

lack

polarizable

atoms which would serve to

anchor the molecule to the

surface, suggesting

that

they might

be less

likely

to form

(6)

N° DISCOGENIC MOLECULES AT AIR-WATER

~~

Q lall~R-SC,H~~

40

~

R

O ~

O

R

10

~

0

70 60

50

lbll~R=SCJf~,

40 30

_

20

~

10

~ o

Z

E

~~

©$

lcll~R=SC~II~,

20

idX=COOICH~)sCBj

lo

R=OC,H~~

o

iell~RmOC~lI~~

loo 200

Area (A~/molecule)

Fig.

I. 1I A isotherms for several derivatives of

triphenylene.

Note the extremely small molecular

areas for the

alkylthiotriphenylenes

la-c- The low molecular areas and

high

surface pressures indicate

that these

compounds

are not

forming monolayers,

but rather oily

phases

or

microcrystallites

at the surface.

Triphenylenes

id and le are

proposed

to lie

edge-on

to the water surface, with at least two tails

penetrating

into the

subphase.

monolayers

on water.

Indeed,

the very small measured molecular areas for these

compounds

is smaller than the known area for

packing

of

hydrocarbon

chains attached to

hydrophilic

groups

(m18h~), leading

us to conclude that these

alkylthiotriphenylenes

do not form

Langmuir

films

[16].

(7)

Truxene derivatives The

liquid crystalline properties

of the truxene

derivatives,

hexa-n-

alkanoyloxytruxene (2a)

and

hexa-n-alkoxybenzoyloxytruxene (2b),

are also well understood

[17-23].

The

spreading

behavior of these mesogens is shown in

figure

2. The H A

diagrams

all show the formation of

relatively incompressible monolayers

with surface pressures at the inflection

point

on the order of 17.5

mN/m.

The areas measured at the inflection

points

of the isotherms are much smaller than would be

expected

if the molecules were assumed to lie with

the cores flat on the surface

(about

154

h2

for 2a and 380

h2

for 2b). The

presence of

highly electronegative

oxygen atoms to serve as

polar anchoring points

to the water surface and

relatively high

surface pressures

(m

20

mN/m) coupled

with

non-negligible

molecular areas indicate that the molecules are not

being pulled

into the

subphase

and are not so

hydrophobic

as to form

oily droplets

at the surface.

R R

/ '

/ R

b~~

R

2a R

= COO lClI~)uCBj

f

Z E

~

~~

30

20 2b R

= COO

@

O(CIIz)iocBj

io

°

ioo 200

Area (h~/molecule)

Fig.

2. 1I A isotherms for two derivatives of truxene cores. The low molecular areas of 2a and 2b

are reminiscent of those of ld and le and are consistent with a model in which the molecules lie edge-on

to the water surface. The larger area of 2b can be attributed to the presence of

bulky phenyl rings

in the

substituted tails.

The most

plausible explanation

of the observed molecular areas is that, like the benzene and

triphenylene

derivatives discussed

above,

the molecules lie

edge-on

to the interface.

Taking

the bulk core-core distance of 4.5

h

for the

core-core distance in the

Langmuir film,

the

corresponding long

axis of the molecular

projection

on the water surface must be on the order of 15-20

h,

consistent with the transverse dimensions of the molecule. The «

edge-on

»

hypothesis requires

that two

hydrocarbon

tails penetrate the

subphase

to some extent.

Figure

3 shows a

possible

conformation for 2a at the

surface,

and illustrates the manner in which the

(hydrophobic)

tails that

penetrate

the

subphase

most

likely

coil up so as to minimize their surface area

[24].

The cross sectional area of the molecule shown in

figure

3 is consistent with the measured molecular areas from

figure

2.

(8)

Fig. 3. Model of 2a at the air-water interface,

assuming

an edge-on arrangement of the truxene cores at the interface. We

postulate

that the tails

submerged

in the water adopt a coiled structure to reduce the

free energy of the

hydrophobic hydrocarbon

tails in direct contact with water.

Hexacyclen

derivatives : We have studied a

variety

of hexasubstituted

[18]-N~

azacrowns, or

hexacyclens (Scheme1);

H-A isotherms are shown in

figure

4 and 5.

Among

the

hexacyclen compounds

3a-3f and 4a-4d,

only

3a and 3e form

liquid crystalline phases.

The bulk

mesophase

structure of 3a is

unclear,

and has been

variously

described as tubular-

columnar

[25]

or smectic

[10].

The interfacial behavior of this molecule,

however,

is much

more

straightforward,

and we have

duplicated previous [6, 7]

H- A isotherms.

Compounds

3a, 3b, and 3e are all similar to each

other, sharing

the same hexamide core and

benzoyl linkage,

and

differing only

in the

length

of the attached

alkyl

tail. Substitution of

a hexamine core

(amounting only

to

substituting

the

carbonyl

functional group for a

methylene linkage) produces

4a and 4b. All of these

compounds display

isotherms characteristic of a transition between a

low-density, low-pressure

state and a

high-density, high-pressure

state. The molecular areas in the

low-density

state are on the order of 160

l~,

while those for the

high-density

state are either 50 or 100

Ji~.

The transition between

the two states

generally

involves a

non-equilibrium «bump»

as discussed above. The

generally accepted [6, 7] interpretation

of the

low-density

state is that the

macrocycle

core lies flat on the water surface and the

alkyl

tails

project

away from the water surface a space-

filling

model of such an arrangement

yields

a molecular area in agreement with that observed.

(9)

R/)/~Q

~$ l~

N N

~~UN/~~

Q ~ R

RQ

3aRm©OICII~)iic~

3b R m

@ OICII~)sCB~

o

20

lo

3cRm@

11

*

~

E

80

11

60

40 3d R

=

lCIIz)i~CB~

20

o

3eR-©OiC%)ioCl3i

3fR

"1C%)AC'&

100 200 300

Area (A~/molecule)

Fig. 4. 1I A isotherms for derivatives of the hexamide core 3. Note the distinct second phase at low molecular areas for several of these

compounds,

corresponding either to

multilayer

formation or a

more

compact molecular conformation.

(10)

~~

/)R

4*~*©°(~~~a)i1~3

~N N

~§~)

R

40

20 4bRm@OICII~)~CH~

z

fi

E tQ 20

io

o

60

40

4d

20 100

Area

Fig.

5. 1I A isotherms for derivatives of the hexamine core 4. The second phase is clearly manifest in 4a and 4b.

It is conceivable that the

high-density

state

corresponds

to

multilayer formation;

ellipsometry

measurements could

help

resolve this

question. Altematively,

the second state may

quite likely correspond

to a conformational

change

in the

monolayer.

For

example,

there may be more than one conformation of the

macrocyde

core : an «

expanded

» conformation in which the

nitrogen

atoms are

relatively

far away from each

other,

and a «

collapsed

»

conformation in which the core

occupies

much less volume. If we assume a conformation in which a

collapsed macrocycle

lies flat on the water surface and the

alkyl

tails are constrained to extend away from the water

surface, thereby incorporating

the known cross-sectional area per

alkyl

chain of 18

h~,

then the minimum molecular

area is

approximately

110

Ji~,

in

good

agreement with the measured areas for 3a, 3b and 4a. If we assume a linear conformation of the molecule

[10],

in which three chains group

together

on either end of the molecule, we

calculate a molecular area of

m 55

h~,

in reasonable agreement with our results for 3e and 4b.

Completely removing

the

alkyl

tails leaves us with

compounds

3c and 4c.

Compound

4c has

an isotherm with

only

a first state

observable,

whereas 3c has an isotherm with a measurable molecular area of 17

Ji~,

smaller than that for a

single alkyl chain,

and

certainly

too small to support

interpretation

as a

monolayer.

As this

compound

is

weakly

soluble in water, the low

molecular area for 3c most

likely

arises from the molecules

dissolving

into the

subphase.

Finally, removing

the

phenyl ring

and associated oxygen and

leaving

the

alkyl

tail

gives

compounds 3d, 4d,

and 3f.

Compounds

3d and 4d share similar

spreading behavior,

with

(11)

comparable

areas per molecule and absence of a second state. The molecular areas associated with these two

compounds

appear to be consistent with the

hexacyden

core

laying

flat on the

water surface with the

alkyl

tails

extending

away from the

subphase.

As 3f is

structurally

similar to 3d and

4d,

one would expect to see similar

spreading

behavior.

However,

the isotherm measured

clearly displays

the two state

phenomenon

observed in

compounds

such as 3a.

3. Conclusions.

A number of

quite

dissimilar conformations are

possible

for

discogenic

mesogens at the air-

water interface.

Depending

on the details of the molecular structure,

discogenic

molecules

with flat

conjugated

cores

(such

as

benzene, triphenylene,

or truxene

derivatives)

may either

lie flat on the ~vater surface with the tails

extending

away from the

interface,

or

edge-on

with

two tails

penetrating

the

subphase

to some extent. In the

hexacyclen compounds,

which have

quite

flexible cores, the molecules may make a transition between a conformation with the

cores

lying

flat on the water surface and a second state in which either the cores are

collapsed

(I.e. compressed

such that the

projection

of the core on the water is as small as

possible)

and the

alkyl

chains are close

packed,

or in which the molecule is

essentially

linear

(quite

similar to the

edge-on

model for more

rigid compounds).

De Gennes first

suggested

in 1971

[26]

that

elongated

molecules at interfaces could form two-dimensional nematic

phases,

with

short-range positional

order and

quasi-long-range

orientational order.

Discogenic

molecules such as

phthalocyanines

and benzene-hexa-

alkanoates lie with their core groups flat on the water

surface, projecting roughly

circular cross-sections on the water surface. These materials are therefore not

good

candidates for

two-dimensional nematics.

Fatty

acid molecules exhibit tilted

monolayer phases, projecting elliptical

cross-sections on the water

surface,

but in this case the three-dimensional aspect of the molecules becomes more

important, making

them

again inappropriate

models for two-

dimensional nematics. On the other

hand,

molecules such as the

triphenylene

and truxene alkanoates

project elongated

cross sections on the water surface due

solely

to the molecular

shape.

This raises the

interesting possibility

that similar

compounds

may exhibit two-

dimensional

liquid crystal phases.

Future progress in

understanding

the structure of

mesogenic compounds

at the air-water interface will

clearly rely

on structural

probes

such as

high

resolution x-ray diffraction.

Acknowledgements.

We thank R.

Fischetti,

V. Skita, S. Xu and T.

Lubensky

for useful discussions. We thank A. R. McGhie and S. H. J. Idziak for assistance with

calorimetry

measurements. This work

was

supported by

the National Science Foundation Grants DMR MRL 88-19885 and DMR 89-01219.

References

Ii GAINES G. L. Jr., Insoluble Monolayers at Liquid Gas Interfaces (Interscience, New York 1966).

[2] RONDELEz F., KOPPEL D. and SADASHIVA B. K., J.

Phys.

Fiance 43 (1982) I37I.

[3] KALINA D. W. and CRANE S. W., Thin Solid Films 134 (1985) 109 ORTHMANN E. and WEGNER G., Angew. Chem. 98 (1986) 1114.

[4] ALBRECHT O., CUMMING W., KREUDER W., LASCHEWSKY A. and RINGSDORF H., Colloid Polym.

Sci 264 (1986) 659.

(12)

[5] RONDELEz F., BARET J. F. and BoIs A. G., J. Phys. France 48 (1987) 1225.

[6] MALTHETE J., POUPINET D., VILANOVE R. and LEHN J. M., J. Chem. Soc., Chem. Commun.

(1989) 1016.

[7] MERTESDORF C. and RINGSDORF H., Liq. Cryst. 5 (1989) 1757.

[8] CHANDRASEKHAR S., SADASHIVA B. K. and SURESH K., Pramana 9 (1977) 471.

[9] TATARSKY D., BANNERJEE K. and FORD W. T., Chem. Mater. 2 (1990) 138.

j10] IDzIAK S. H. J., MALISzEWSKYJ N. C., HEINEY P. A., MCCAULEY J. P. Jr., SPRENGELER P. A.

and SMITH A. B. III, J. Am. Chem. Soc. l13 (1991) 7666.

[I ii SCHLOSSMAN M. L., SCHWARTZ D. K., PERSHAN P. S., KAWAMOTO E. K., KELLOGG G. J. and LEE S.,

Phys.

Rev. Lett. 66 (1991) 1599.

[12] LEVELUT A. M., J.

Phys.

France 40 (1979) L81 J. Chim. Phys. 80 (1983) 149.

[13] KOHNE V. B., POULES W, and PRAEFCKE K., Chem. Zeit. 108 (1983) l13.

[14] GRAMSBERGEN E. F., HOVING J. H., DE JEU W. H., PRAEFCKE K. and KOHNE B., Liq.

Cryst.

1 (1986) 397.

[15] HEINEY P. A., FONTES E., DE JEU W. H., RIERA A., CARROLL P. and SMITH A. B. III, J.

Phys.

France 50 (1989) 461.

[16] This

interpretation

was

strengthened by

the observation of a

patchy

structure when crossed

polarizers

were used to examine a

deposition

of

density

similar to that on the surface of the

trough

or on the surface of water in a Petri dish,

suggesting

that three-dimensional

microcrystallites

or

oily droplets

were

floating

on the water surface.

[17] DESTRADE C., MALTHETE J., NGUYEN H. T. and GASPAROUX H.,

Phys.

Lett. A 78

(1980)

82.

[18] DESTRADE C., GASPAROUX H., BABEAU A., NGUYEN H. T. and MALTHETE J., Molec. Cryst. Liq.

Cryst.

67 (1981) 693.

[19] NGUYEN H. T., MALTHETE J. and DESTRADE C., J.

Phys.

France 42 (1981) L417.

[20] DESTRADE C., FOUCHER P., MALTHETE J. and NGUYEN H. T., Phys. Lett. 88 (1982) 187.

[21] FONTES E., HEINEY P. A., OHBA M., HASELTINE J. N. and SMITH A. B. III, Phys. Ret,. A37 (1988) 1329.

[22] LEE W. K., WINTNER B. A., FONTES E., HEINEY P. A., OHBA M., HASELTINE J. N. and SMITH A.

B. III, Liq. Cryst. 3 (1989) 87.

[23] WARMERDAM T., FRENKEL D. and ZIJLSTRA R. J. J., Liq. Crysi. 3 (1988) 149.

[24] Figure 3

actually

shows a structure obtained via

potential

energy minimization of an isolated (gas

phase)

molecule, not

including

the effects of the air-water interface or the

surrounding

molecules. The energy was minimized using the conjugate gradient method in the Dreiding force field available on

Polygraf,

a

product

of

Biodesign,

Inc.

[25] LEHN J. M., MALTHETE J. and LEVELUT A. M., J. Chem. Soc.. Chem. Commun. (1985) 1794.

[26] DE GENNES P. G., Farad. Symp. 5 (1971) 16.

[27] Transition temperatures were measured in our laboratory using Differential

Scanning

Calorimetry.

Références

Documents relatifs

L'introduction d'un desordre moleculaire conduit .?i des valeurs des distances OH (D) en bon accord avec celles obtenues par des calculs recents de chimie

Only the public expenditures (subsidies) would decrease. At a water tariff of Rs. 0.6/m3, farmers’ electricity charges are stabilized at the level of “reference situation” but this

width of the heating and cooling transition also varies along the transition line, generally a large temperature hysteresis is observed when the transi- tion widths are also

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

= 2, the Hamiltonian for the Widom model [6] is given by trie same form as (1) except that the summations in the third term on the nght hand side is to be carried ont over

Indeed the lipid cannot undergo a high level of compression and consequently, during the collapse and the formation of multilayers, the slight free space between the two sugar

Kinetic changes in interfacial rheology can be attributed either to the blocking of thiol groups by the N-ethylmaleimide, or more probably to structural and net

We show that, at small laser power, the particle is trapped in on axis con fi guration, similarly to 2 dimensional trapping of a transparent sphere by optical