• Aucun résultat trouvé

Langmuir monolayers of mesomorphic monomers and polymers

N/A
N/A
Protected

Academic year: 2021

Partager "Langmuir monolayers of mesomorphic monomers and polymers"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: jpa-00209984

https://hal.archives-ouvertes.fr/jpa-00209984

Submitted on 1 Jan 1985

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.

Langmuir monolayers of mesomorphic monomers and polymers

K.A. Suresh, A. Blumstein, F. Rondelez

To cite this version:

K.A. Suresh, A. Blumstein, F. Rondelez. Langmuir monolayers of mesomorphic monomers and poly-

mers. Journal de Physique, 1985, 46 (3), pp.453-460. �10.1051/jphys:01985004603045300�. �jpa-

00209984�

(2)

453

Langmuir monolayers of mesomorphic monomers and polymers

K. A. Suresh (*), A. Blumstein (+) and F. Rondelez

Laboratoire de Physique de la Matière Condensée, Collège de France, 11, Place Marcelin Berthelot, 75231 Paris Cedex 05, France

(Reçu le 5 octobre 1984, accepté le 5 novembre 1984)

Résumé.

2014

Nous avons préparé des monocouches de Langmuir avec des molécules organiques qui forment des

phases cristal liquide en volume. Tous les polymères et monomères utilisés présentent le même type de structure

avec un c0153ur central aromatique rigide contenant plusieurs groupes esters hydrophiles et deux ou plusieurs chaînes aliphatiques hydrophobes. Un étalement bi-dimensionnel satisfaisant est obtenu dans la plupart des cas, avec une

augmentation monotone de la pression superficielle en fonction de la concentration dans la monocouche. Aux

pressions les plus fortes, l’aire minimum correspond à une orientation verticale des molécules avec décollement de certains groupes hydrophiles initialement en contact avec le substrat aqueux. Dans la région de basses pressions,

les données expérimentales tendent à prouver que les molécules sont orientées parallèlement à l’interface.

Abstract.

2014

Monolayers of organic compounds which exhibit liquid crystalline behaviour in their bulk phases have

been formed at an air-water interface. All selected polymers and monomers were of the same basic structure with a rigid aromatic central core containing several hydrophilic ester groups and two or more aliphatic hydrocarbon

chains. Spreading was satisfactory in most cases and the surface pressure was found to increase monotonously with

surface concentration. At the highest pressure attainable the minimum surface area per molecule corresponds to a

vertical orientation, with some of the anchoring ester groups lifted up from the water subphase. In the lower pres-

sure region, there is some evidence that the molecules lie flat on the interface.

J. Physique 46 (1985) 453-460 MARS 1985,

Classification

Physics Abstracts

82.65

-

61.30

-

61.40K - 64.30

1. Introduction.

It has been known for a long time that amphiphilic

molecules such as soaps and phospholipids sponta-

neously spread at the air-water interface to form two- dimensional monolayers [1]. On the other hand, only recently has it been realized that thermotropic liquid crystalline compounds can also be considered as

amphiphiles. Their molecular structure is rather

complex but the existence, on the same molecule, of hydrophilic and hydrophobic groups is generally recognized without great difficulty. As a consequence of this dual character, it is expected that they should spread as Langmuir monolayers. Moreover, since

these molecules are fairly rigid, it should suffice in

principle to have two anchoring points on the water

to insure planar orientation. Unfortunately, the exist- ing monolayer data on such materials are very

scarce. To the best of our knowledge, there exists only

two papers in the literature which describe mono-

layers of calamitic (rod-like) molecules [2, 3] and one

which deals with discotics [4]. Moreover, all these studies have been performed on low molecular weight, monomeric, materials exclusively.

In this paper, we will describe a systematic study of

monomers and polymers bearing hydrophilic ester

groups and exhibiting liquid crystalline phases in the

bulk. We will show how the planar configuration (with

the molecules lying flat on the interface) is in compe- tition with the vertical orientation when the surface concentration is gradually increased. We will also suggest that it is possible to form in this way sweetie phases composed of a single molecular layer, which is

even closer to an ideal two-dimensional system than the freely-suspended two layers smectic films obtained

by Moncton and Pindak [5, 6].

2. Experimental.

The materials used for this study are listed in table I.

They were chosen on the grounds that they exhibit thermotropic mesomorphic behaviours in the bulk or

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01985004603045300

(3)

Table I.

-

Characteristics of the polymers and monomers used in this study. See text for the signification of the

acronyms used for the various compounds. The thermotropic bulk liquid crystalline phase transitions are indicated in column 4. K is for solid, N : nematic, I : isotropic, SA : smectic A, Dc : unspecified discotic phase, DB : rectan- gular disordered columnar discotic phase (Ref. [7]). The transition temperatures are indicated in °C. Parentheses indicate a monotropic transition. The polymer molecular weights have not been indicated here but viscosity mea-

surements in chlorinated hydrocarbon solvents (1, 1, 2, 2 tetrachloroethane, chloroform, m-cresol) have showed them to be in the range 5,000 to 8,000. The minimum areas Ao per molecule (or per monomer in the case of a polymer chain) measured in the monolayer experiments are given in column 6. They were derived from an extrapolation

to zero pressure of the highest surface pressure data points.

(4)

455

are closely related to well-known liquid-crystals. Their

transition temperatures are listed in column 4 of table I. Note that rufigallol hexa n-octanoate (labelle-

das RGH-8) has no nematic phase but two columnar phases (DB and DJ [7] while 4-4’ azoxy a-methyl n-nonyl cinnamate (AMNC) has only a smectic A phase.

All the listed materials are composed of a rigid

aromatic central core surrounded by long, flexible, hydrocarbon chains. They also possess several ester groups which are known to provide good anchoring points to the water surface [1]. With the exception of rufigallol which is an anthraquinone derivative, the

central part is either of the azoxy or the azobenzene type, eventually substituted with methoxy groups in the ortho position. These structures are mostly hydro- phobic and prevent direct solubilization of the spread

molecules into the water subphase. In the case of the polymers, the same monomeric unit is repeated n times (here, n is relatively low, of the order of 10).

Separate studies [8] have shown that polyester

macromolecular chains behave as flexible coils in dilute solution and also in the melt. Indeed their

persistence length, which describes the local chain

rigidity, is found to be no greater than the length of

the monomeric unit. In that respect, these polymers

thus belong to a different class than the typical semi- rigid polymers such as the aromatic polyamides, polyisocyanates or polybenzyl glutamates.

The polymers were synthesized by a condensation reaction between diacids or diacid chlorides and diols

or diphenols [9]. Depending on the synthesis, the mesogenic unit was contained in the di-acid and the flexible spacer in the diol or vice versa. For instance,

PC azoxy-8 and PC azoxy-10 refer to polyesters of

p-azoxy benzoic acid with octane diol and decane diol

respectively. Their full names are poly(4,4’-carboxy azoxy benzene-1, 10-dioxyoctane) and poly(4,4’-car- boxy azoxy benzene-1, 10-dioxydecane). On the other hand, PO azoxy-12 was obtained by condensation of 4-4’ azoxy phenol with tetradecane dioic acid. Its full

name is poly(4,4’-oxyazoxybenzene dodecane dioyl).

Similarly PC azo-8 and PC azo-10 refer to polyesters

of p-azobenzoic acid with octane diol and decane diol respectively while PO azo-10 and PO azo-11

refer to polyesters of 4,4’ azophenol with dodecane and tridecane dioic acids. Their full names are poly(4-4’- carboxy-azoxybenzene-1, 10-dioxy octane or decane)

and poly(4-4’-oxyazobenzene decane dioyl or unde-

cane dioyl) respectively. The molecular weights of

these various polymers, as obtained from intrinsic

viscosity data, are fairly low, typically in between 5,000 and 8,000.

The 4-4’-nonanoyloxy-(2,2’ methoxy) azobenzene (NOMAB) and 4,4’-hydroxy(-2,2’ methoxy) azoben-

zene (HMAB) monomers were prepared by standard

chemical procedures and purified by recrystallization.

Their purity was checked by thin layer chromato- graphy. The rufigallol hexa-n-octanoate (RGH-8)

and the 4,4’ azoxy a-metyl n-nonyl cinnamate (AMNC)

had been synthesized previously by others and were

used as received

Spreading solutions were prepared in the usual way

by dissolving 1 mg of powder in 25 ml of chloroform.

Clear solutions were generally obtained after stirring

for a few days and occasionally heating to 40 OC.

Only in the case of the PC azo-10 and PO azo-10 poly-

mers were we not able to achieve the desired dissolu- tion and these materials were not studied further. It is

possible that the molecular weight of these two poly-

mers was too large to allow good solubilization into chloroform.

Drops (4-10 yl) of these solutions were deposited

onto the free surface of tri-distilled, surfactant-free,

water contained in a cylindrical quartz container.

The monolayer forms spontaneously, following the evaporation of the spreading solvent. Its surface density can be gradually increased by successive deposition up to a state of saturation is finally reached

At this point, the surface density is maximum and additional drops will stay as lenses floating on the

interface. These lenses will slowly disappear by sol-

vent evaporation, leaving on the surface a small flake of solid material which is detectable with the naked eye. On the contrary, when the molecules do not have the correct hydrophilic-hydrophobic balance

for spreading, the monolayer does not form and satu-

ration can never be observed even after deposition

of large amounts of solution. PO azoxy-12 is a good example of such a behaviour.

The surface pressure exerted by the monolayers was

measured by the Wilhelmy hanging plate technique

with a platinum foil (2 x 0.9 x 0.01 cm’) attached to

a force transducer (Model FTAI-1, Sanbord, Waltham, MA) fed into a phase-sensitive amplifier (Hewlett

Packard HP 8805 B). The sensitivity was of the order of 10 gN m-1. The accuracy of the measurements were further limited by a noticeable drift of the order of 0.2 gN m - ’ min - ’. This drift, which was more pronounced at large surface concentrations, was due,

in our opinion, to material adsorption onto the blade.

The temperature of the trough was controlled to

± 0.1 OC by a circulating water bath (Haake, model F-3). In addition to the cylindrical trough (6.24 cm diameter, 0.5 cm height), a rectangular trough (25 x 9.5 x I cm’) equipped with a movable barrier

was also used to perform compression-expansion cycles within the monolayer. The compression rate

was adjusted with a stepping motor to an average value of 0.05 nm2 molecule-1 min-’. The pressure

was never allowed to exceed the equilibrium spreading

pressure (i.e. the pressure reached at saturation in the

drop deposition method) in order to avoid the forma- tion of metastable states in the monolayer.

3. Results.

The surface pressure isotherms obtained by the

drop deposition method at or near room temperature

(5)

456

Fig. 1.

-

Surface pressure isotherm of the polymer poly(4,4’- carboxyazoxybenzene-1, 10-dioxydecane), (PC azoxy-10),

at 20 °C. Substrate is pure water. The solid line extrapolates

to a minimum monomer area A0 ~ 0.25 nM2 at zero pres-

sure.

Fig. 2.

-

Same as figure 1 but for the polymer poly(4,4’-car- boxyazobenzene-1, 10-dioxyoctane), (PC azo-8) at 20°C.

Ao N 0.16 nm2.

Fig. 3.

-

Same as figure 1 but for the polymer poly(4,4’- oxyazobenzene undecane dioyl), (PO azo-11), at 20 °C.

A0 ~ 0.15 nm2.

are shown in figures 1 to 5. The first three curves are

for polymeric materials (PC azoxy-10, PC azo-8 and PO azo-11) while the next two are for the RGH-8 and AMNC monomers. All display similar behaviours.

At very large areas per molecule A > 2.5 nm2 (in the

Fig. 4. - Surface pressure isotherm of rufigallol hexa

n-octanoate (RGH-8) at 40 °C. Substrate is pure water. The solid line extrapolate to a minimum area per molecule

Ao N 0.33 nm2 at zero pressure.

Fig. 5.

-

Surface pressure isotherm of 4-4’ azoxy a-methyl n-nonyl cinnamate (AMNC) at 20°C. Substrate is acidified water (0.01 N HCI, pH

=

2.0). The solid line extrapolates

to a minimum area per molecule Ao N 0.22 nm’ at zero

pressure.

case of polymers, the abscissa scale is calculated per

monomer), the pressure H is low and typically less

than 10- 5 N. m - 1. As the area is further reduced,

the pressure increases gradually and reaches values of about 10-4-10-3 N . m-’ at 0.5 nm’. In this

region the monolayer compressibility, which is the

reciprocal of the slope, is still fairly high. The pressure

variation, however, becomes much more rapid for

even lower areas per molecule. Typically Il changes by several 10- 3 N. m-1 between 0.3 and 0.2 nm2. The

compressibility is then so low that the curve is prac-

tically vertical. Extrapolation of the slope at zero

pressure yields limiting minimum areas per monomer,

Ao, between 0.15 and 0.33 nm’ for all compounds

studied (see column 6 of Table I).

We have also measured the surface pressure iso-

therms for AMNC at several temperatures. The

results are shown in figures 6 and 7. The most

(6)

457

Fig. 6.

-

Same as figure 5 but at three different tempera-

tures (20, 30, 35 °C). The solid lines are just a guide to the

eye.

Fig. 7.

-

Same as figure 6 but the surface pressure isotherms have now been plotted as a function of the surface concen-

tration. Note the very sharp pressure increase around C ~ 0.6 x 10- 3 g. m - 2.

interesting feature is the sharp surface pressure increase observed around 1.4 nm’ for the two higher tempe-

ratures. The change is about 1 x 10- 3 N . m -1 at

30 °C and 2 x 10-3 N. m-1 at 35 °C (see Figs. 6 or 7).

From there down to molecular areas of about 0.5 nm2,

the pressure variations are more gradual, and the slope is fairly similar whatever the temperature is. In this region the compressibility is large again. The

surface pressures measured at even lower areas and for the two temperatures of 30 and 35 OC, have not been plotted in figures 6 and 7 which focus more on

the low pressure regimes. It suffices to say that they

are qualitatively similar to the one plotted in figure 5

at room temperature. Extrapolation of the slopes at

zero pressure yields Ao

=

0.22 nm2 at 20 °C, 0.3 nm2

at 30 °C and 0.5 nm2 at 35 °C.

A word of further comment should be made concern-

ing the spreading conditions for the AMNC mono- mers and also for the azo materials, although for

different reasons. In the case of AMNC, and contrary

to all other experiments, the aqueous subphase has

been acidified, by addition of dilute HCI, down to a pH value of 2.0. This is a routine procedure which is

known to favour spreading in the case of surface- active molecules bearing carboxylic or amine groups.

It is also applicable to some non-ionizable groups such as esters [10]. For a given surface concentration of AMNC we have repeatedly observed that the surface pressures are typically 10 % higher on an acidic subphase, which indeed implies better spreading. In

order to rule out possible hydrolysis effects by the acid, we have performed a separate experiment in

which a chloroform solution of AMNC was stirred with

an equal volume of acidified water during 24 hours.

After decantation, the AMNC solution was spread

onto a pure water subphase. The surface pressure data

were absolutely identical to those of a fresh AMNC solution, never exposed to dilute hydrochloric acid.

This check eliminates the possibility of chemical degra-

dation under the mild conditions of the experiment,

and justifies the data points of figures 5, 6 and 7.

In the case of the materials containing azo linkages,

a different kind of precaution has to be taken. Indeed,

it is well known that azobenzenes undergo trans to cis

isomerization under ultra-violet light excitation in the range 300-380 nm [11]. With PO azo-11, a decrease in the optical absorption spectrum was observed at 350 nm, together with the appearance of a new

shoulder at 400 nm, when the solution was illuminated with alms ultra-violet pulse through a UG-1 Schott filter. Such a photo-isomerization is not desirable

here since there is a considerable conformation

change between the two isomers. As a consequence, all solutions of azo compounds were stored in brown bottles and the corresponding monolayer experiments

were always conducted under dim yellow light condi-

tions.

Last, there are several compounds listed in table I which have only been studied qualitatively and at

room temperature. PC azoxy-8 spreads and shows a

pressure isotherm comparable to that of PC azoxy-10.

Its minimum molecular area is Ao - 0.15 nm2. On

the contrary, no surface pressure could be detected with PO azoxy-12, which therefore should be consi- dered as non-spreading at the air-water interface. As discussed earlier, PC azo-10 and PO azo-10 could not be dissolved in chloroform, making it impossible to

check their spreading behaviour. Finally the two

monomers NOMAB and HMAB are observed to

spread satisfactorily and their surface pressure iso- therms yield a minimum molecular area of about 0.23 nm2 in both cases.

4. Discussion.

4.1 MOLECULAR CONFIGURATION IN THE MONOLAYER.

-

That thermotropic nematic polymers and mono-

mers containing ester groups can be spread as mono- layers at an air-water interface and form stable two- dimensional films, is evidenced from the well-defined surface pressure isotherms observed in our experi-

ments. On the whole, the curves are typical of mono-

(7)

458

layers exhibiting a gaseous phase at large areas and a

condensed phase at smaller areas. By analogy with

the known behaviour for the simpler long-chain fatty acids [1], the gaseous phase corresponds to

a state of weakly-interacting molecules and where both the polar heads and the aliphatic chains are

in contact with water. Similarly, the condensed phase corresponds to a state where the molecules are much

more densely packed and where some of their chemical groupings, generally the methylene units, have been

lifted up from the air-water interface.

4.1.1 Small molecular areas (less than 0.4 nm2).

-

For all the compounds investigated here, the minimum

area per molecule Ao (corresponding to maximum compression of the monolayer) is of the order of 0.15-0.35 nm2. For the azo and azoxy derivatives it is included in an even narrower range of 0.15-0.25 nm2 (see Table I). The value for rufigallol hexa-n-octanoate is higher, of the order of 0.33 nm2, but this is consistent with its bulkier aromatic core.

We can now try and compare these values with esti- mations from molecular models and crystallographic

data in order to gain information on the molecular orientation at the air-water interface. The azo and azoxy cores, not counting the aliphatic chains, are

calculated to be 1.36 nm long, 0.65 nm wide and

0.36 nm thick [12]. Similarly the anthraquinone core

of rufigallol is 1.17 x 0.86 x 0.36 nm3 [13]. Using

these sets of values, it is easy to see that in all cases the minimum molecular area Ao observed in our mono-

layer experiments is not compatible with a planar

orientation. More precisely, if one assumes the ring system to have its long axis parallel to the water sur- face, Ao should be of the order of 0.49 nm2 (0.42 nm2)

for the azo and azoxy benzene (anthraquinone) cores

if the short axis is perpendicular to the interface, and

0.88 nm2 (1.0 nm2) if the short axis is also parallel to

the interface. The only possibility left, and compatible

with our experimental Ao values, is that the ring sys- tem is oriented vertically, with its long and short axis

respectively perpendicular and parallel to the water

surface. In that case the cross-sectional areas should be 0.23 and 0.31 nm2 respectively for the two types of compounds. This is in excellent agreement with our

experimental findings. It is also interesting to note

that polyester polymers very similar in structure to PO azoxy-10 have been recently reported to align homeotropically on untreated microscope glass

slides [14].

Of course this vertical orientation requires that

some of the ester groups get lifted up from the inter- face. Such a possibility has been suggested long ago

by Adam et ale in their study of molecules with two

ethyl ester groups at opposite ends of a long methylene

chain [ 15-a] and also with benzene derivatives contain-

ing two or more hydroxyl or methoxyl groups in the

ring [15-b]. When all the molecules are standing vertical, the benzene derivatives in particular can adopt

a close-packed configuration with the faces of their aromatic rings all in register against the faces of the

neighbouring rings. This is very favourable energetical- ly and may more than compensate for the loss of adhe- sional affinity of the polar ester group with the water.

In this vertical configuration, the aliphatic chains will try to arrange themselves to fill up the void available between the aromatic cores. At not too large compres-

sions, they may form a very thin liquid phase, acting as

a diluent and the rigid cores will gather in lamellar, cylindrical and other arrangements similar to the

lyotropic micellar systems in which water is the diluent [16]. At maximum compression, they will be expelled from the interstitial region and will form two thin pure hydrocarbon layers, one below and one

above the rigid cores. In this configuration, they will

no longer contribute to the molecular area occupied by one monomer in the interfacial plane. This explains why the measured Ao values are independent of the aliphatic chain length, at least in the range of seven to twelve carbon atoms investigated here.

4.1.2 Large molecular areas (more than 1.5 nm2).

-

When the surface occupied per molecule is larger

than the cross-sectional areas in the plane of the ring system, it is probable that the hydrophilic ester groups located at the farthest ends of the azo or azoxy benzene

ring system tend to maintain the molecule flat on the interface. The best evidence for this planar orientation is provided by the AMNC data (Figs. 6 and 7). At

30 and 35 OC, the surface pressure stays very low down

to a molecular area of about 1.4 nm2, and then starts increasing rapidly over a narrow region. We suspect that this sudden compressibility change occurs when

the surface density occupied by the molecules in the

planar configuration becomes of the order of unity.

This interpretation is compatible with the cross-

sectional area deduced from molecular models which is 1.2-1.4 nm2. On further compression, the molecules tilt progressively until they reach a fully vertical

orientation. This molecular rearrangement was also

proposed for the same compound, but at a single temperature, by Dorfler, Kerscher and Sackmann [2].

Here we have the additional information that the pressure jump occurs at a surface area value which is

independent of temperature. This confirms a purely geometrical interpretation. Moreover the pressure

jump gets larger with increasing temperatures. This shows that the new configuration is due to the increased intermolecular cohesion in the tilted orientation rather than to the weakening of the interaction between the ester pinning groups and the water subphase. We are tempted to interpret in the same way the observations of Diep-Quang and Ueberreiter [3] on 4-n-heptyl- phenyl (4’-n-hexanoyloxy) benzoate,

At that time a « rolling-over » liquid collapse mecha-

(8)

459

nism, in which the molecules start to form multilayers

below molecular areas of the orders of 1.0 nm’, was suggested. We think, however, that the observation of a

pressure re-increase around 0.2 nm2 is a strong indica- tion of a progressive molecular tilting towards a

condensed phase where all molecules are vertically

oriented. In the present experiments, we have carefully checked, using the trough equipped with the moving

barrier that the pressure isotherms were always fully

reversible on both sides of the transition region. This is clearly not compatible with the formation of multi-

layers.

It should be technically feasible to follow directly

the progressive tilt of the molecular axis upon com-

pression. Ellipsometric measurements, which give

access to the monolayer thickness, have already been performed at an air-liquid interface [17]. Another possibility would be to build multilayers on a solid reflecting substrate using the horizontal lifting method [18] and then to perform standard interferometric measurements. In this last case, however, it is not absolutely certain that the molecular packing will be fully preserved at all surface coverages during the

transfer onto the solid surface.

4.2 INFLUENCE OF THE HYDROPHILIC GROUPS ON MONOLAYER SPREADING.

-

We have already men-

tioned that the spreading is controlled by a careful

balance between the hydrophobic and hydrophilic

groups on the molecule. The comparison between PC azoxy-10 and PO azoxy-12 strikingly illustrates this

point. Apart from a minor variation by two methylene

units in the flexible spacer, these two azoxy polyester

chains mainly differ by a mere inversion in the sequence of their ether and carbonyl linkages. It appears that the displacement of the carbonyl group one atom away from the benzene ring is sufficient to jump from a spreading (PC azoxy-10) to a non-spreading (PO azoxy-12) behaviour. That the rule is fairly general is

evidenced from the comparison between PC azo-8

and PO azo-11. Here also there is a reversal between the carbonyl and the ether bonds. Although both polymers are observed to spread at the air-water interface, the spreading was found to be much easier

and reproducible in the case of PC azo-8.

Another very legitimate question is related to the

nature of the anchoring points to the water substrate.

We have seen that, with ester groups, all the molecules

flip over to the vertical orientation upon compres- sion. In order to try and keep the molecule flat on the interface at all surface pressures, it would be tempting

either to increase the number of ester groups per molecule or to exchange the esters with more polar

groups (e.g. hydroxyls).

The former approach has been used successfully by one of us in an earlier paper dealing with benzene- hexa-n-alkanoates [4]. These disc-like molecules consist of a single benzene ring to which are connected

six hydrocarbon chains, each with one ester group.

Surface pressure isotherms have evidenced that the aromatic ring lie flat on the water at all compressions

and that the limiting molecular areas correspond to

the cross section of the molecules parallel to the plane

of the ring. Moreover it was possible to show that the

aliphatic chains are parallel to the interface in the

expanded liquid monolayer state and perpendicular

in the condensed state [19]. By comparison, rufigallol hexa-n-octanoate, which has been studied here, has

also six ester groups but its aromatic anthraquinone

core is much bulkier. As a consequence the vertical orientation is still favoured over the parallel orienta-

tion at maximum compression.

The possibility of exchanging esters with more polar groups has been tried long ago by Adam et al., however the results have been deceptive [15-b].

Palmityl resorcinol and stearyl phloroglucinol all

take the vertical orientation despite the presence of

hydroxyl groups situated all around the benzene

ring. It was suggested at that time that «the same

groups which must increase the adhesion to the water also increase the adhesion between adjacent benzene rings standing upright ». This argument may well be very general. It will explain why, so far, all experiments

on monolayers of liquid crystal-forming compounds (including the present ones) have shown evidence for

a vertical reorientation upon compression.

Recently Nakahara and Fukuda [20] have reported

that molecules of 4,4’-bis-stearyl amino azobenzene also get gradually lifted up from the interface on

compression despite the presence of two strongly hydrophilic amide groups. Here also it is possible

that the strong hydrogen-bonding forces between the amide groups belonging to two neighbouring mole-

cules ultimately favours the vertical orientation. It would now be interesting to introduce ortho substi- tuents in the ring system in order to try and decrease this intermolecular interaction and to look if the pla-

nar orientation is preserved at larger compressions

than before.

5. Conclusion.

To conclude, we have prepared Langmuir monolayers

of several polymers and monomers bearing ester

groups and forming liquid crystalline phases in the

bulk. Surface pressure data show that upon compres- sion the molecules tilt from a planar to a vertical

orientation relative to the air-water interface. In the

high pressure region, the molecules have probably

their aromatic ring systems in register. On the other hand, the aliphatic chains do not seem to contri- bute noticeably to the minimum molecular area. This

is suggestive of a smectic-type conformation, involving

a single molecular layer with the aromatic cores on

the inside and the hydrocarbon chains on the outside.

A direct experimental check of this hypothesis is not

easy. However, it may show up in macroscopic pro-

perties such as surface elasticity, shear viscosity, etc.

(9)

460

In the lower pressure regions, the stratified organiza-

tion is probably lost. It is possible that the aromatic

cores gather in a variety of two-dimensional lamellar, cylindrical, etc... aggregates, as in the bulk lyotropic

micellar systems. The hydrocarbon chains will then act as a diluent. Modem x-ray synchrotron tech- niques should allow to test this point in the near

future [21]. Eventually, for the lowest pressures, the

monolayer becomes so dilute that all organization is

lost and the molecules lie flat on the interface with all their hydrophilic groups in direct contact with the water.

Acknowledgments.

We thank P. G. de Gennes for his interest in this study.

RGH-8, and AMNC were gifts of Drs. J. C. Dubois and J. Billard respectively. HMAB and NOMAB were

synthetized specially for us by V. Surendranath. This work was also made possible by a travel grant under the CNRS-CSIR exchange program between the

College de France and the Raman Research Institute

and by a scholarship from the Joliot Foundation for K. A. Suresh.

Note added in proof.

-

After this work was comple- ted, we became aware of experiments by D. Cadenhead

and M. Philips (J. Colloid Interface Sci. 24 (1967) 291)

on fl-estradiol diacetate, a sterol molecule bearing

two polar heads (acetate) at its extremities. The

reported surface pressure isotherms are strikingly

similar to those of AMNC, with a break in the curves

for an area of 0.96 nM2 and a limiting value of 0.38 nm2.

As in the present analysis, this was taken as evidence

of molecular rearrangement from a flat to a vertical molecular orientation, relative to the air-water inter- face. The lifting-up takes place when the close-packed

state corresponding to the flat orientation is reached.

The value of the corresponding area per molecule is also independent of temperature, which confirms the

geometrical interpretation. On the other hand, the

surface pressure at the transition is a much larger

function (increasing) of temperature in the case of AMNC than for p-estradiol di-acetate. This reflects differences in the intermolecular interaction energies

for the two types of compounds.

References

[1] GAINES, G. L., Insoluble Monolayers at Liquid-Gas Interfaces (Wiley, New York) 1966.

[2] DÖRFLER, H. D., KERSCHER, W. and SACKMANN, H., J. Phys. Chem. 251 (1972) 314.

[3] DIEP-QUANG, H. and UEBERREITER, K., Colloid Polymer

Sci. 258 (1980) 1055.

[4] RONDELEZ, F., KOPPEL, D. and SADASHIVA, B. K., J.

Physique 43 (1982) 1371;

BARET, J. F., BOIS, A. and RONDELEZ, F., in preparation.

[5] MONCTON, D. E. and PINDAK, R., Phys. Rev. Lett. 43

(1979) 701.

[6] PINDAK, R., MONCTON, D. E., DAVEY, S. C. and GOODBY, J. W., Phys. Rev. Lett. 46 (1981) 1135.

PINDAK, R., SPRENGER, W. O., BISHOP, D. J., OSHE-

ROFF, D. D. and GOODBY, J. W., Phys. Rev. Lett.

48 (1982) 173.

[7] QUEGUINER, A., ZANN, A., DUBOIS, J. C. and BIL-

LARD, J., Proc. Int. Liq. Cryst. Conf. Bangalore,

December 1979, Ed. S. Chandrasekhar, Heyden,

London (1980) p. 35.

[8] BLUMSTEIN, A., MARET, G. and VILASAGAR, S., Macro-

molecules 14 (1981) 1543.

[9] BLUMSTEIN, A., VILASAGAR, S., PONRATHNAM, S., CLOUGH, S. B. and BLUMSTEIN, R. B., J. Polym.

Sci. Polym. Phys. Ed. 20 (1982) 877.

[10] See Ref. [1] p. 243.

[11] Ross, D. L. and BLANC, J., in Techniques of Chemistry, Photochromism, Vol. 3, ed. by G. H. Brown

(Wiley, New York) 1971.

[12] BROWN, C. J., Acta Crystallogr. 21 (1966) 146.

[13] SEN, S. N., Indian J. Phys. 22 (1948) 347.

[14] BLUMSTEIN, A., SCHMIDT, H. W., THOMAS, O., KHA-

RAS, G. B., BLUMSTEIN, R. B. and RINGSDORF, H., Mol. Cryst. Liq. Cryst. 92 (1984) 271.

[15] a) ADAM, N. K., DANIELLI, J. F. and HARDING, J. B.,

Proc. R. Soc. 147A (1934) 493.

b) ADAM, N. K., Proc. R. Soc. 119A (1928) 628.

[16] We are indebted to P. G. de Gennes for this suggestion.

For a review on lyotropic systems, see for example : PERSHAN, P. S., J. Physique Colloq. 40 (1979)

C3-423.

[17] TENEBRE, L., J. Physique Colloq. 38 (1977) C5-123 ; SMITH, T., J. Opt. Soc. Am. 58 (1968) 1069.

[18] FUKUDA, K., NAKAHARA, H. and KATO, T., J. Colloid Interface Sci. 54 (1976) 430.

[19] RONDELEZ, F., BARET, J. F. and BOIS, A., to be pu- blished.

[20] NAKAHARA, H. and FUKUDA, K., J. Colloid Interface

Sci. 93 (1983) 530.

[21] WINNIK, H. and DONIACH, S., Synchrotron Radiation

Research (Plenum Press, New York) 1980.

Références

Documents relatifs

Using X ray surface diffraction we have investigated crystalline monolayers of short alcohols (1-decanol to 1-tetradecanol) at the air-water interface in the vicinity of the first

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

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

b) Comparison of the interfacial moduli (p) and the bulk moduli (G) is only possible at the same frequency and deformation. Therefore, interpolation between several bulk data

Was aber meinen Rücktritt vom Amte betrifft, so ist der zwingende Grund wirklich der gewesen, daß ich den besonderen Anstrengungen durch zwölf wöchent- liche Vorlesungen und Übungen

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

Abstract: Let IE be a complete ultrametric space, let IK be a perfect complete ultra- metric field and let A be a Banach IK-algebra which is either a full IK-subalgebra of the