• Aucun résultat trouvé

Micellar shape distribution in the isotropic phase near a prolate-to-oblate nematic phase transition

N/A
N/A
Protected

Academic year: 2021

Partager "Micellar shape distribution in the isotropic phase near a prolate-to-oblate nematic phase transition"

Copied!
7
0
0

Texte intégral

(1)

HAL Id: jpa-00210285

https://hal.archives-ouvertes.fr/jpa-00210285

Submitted on 1 Jan 1986

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.

Micellar shape distribution in the isotropic phase near a prolate-to-oblate nematic phase transition

C. Rosenblatt

To cite this version:

C. Rosenblatt. Micellar shape distribution in the isotropic phase near a prolate-to-oblate nematic phase transition. Journal de Physique, 1986, 47 (6), pp.1097-1102. �10.1051/jphys:019860047060109700�.

�jpa-00210285�

(2)

Micellar shape distribution in the isotropic phase near a prolate-to-oblate

nematic phase transition

C. Rosenblatt

Francis Bitter National Magnet Laboratory,

Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, U.S.A.

(Reçu le 6 janvier 1986, accepté le 17 février 1986)

Résumé. 2014 Nous avons mesuré le coefficient de Cotton-Mouton en fonction de la température et de la concen-

tration dans la phase isotrope du système ternaire laurate de potassium (KL), 1-décanol, D2O. Les résultats sont analysés dans le cadre d’un modèle où la solution est un mélange de cylindres (tiges) et de disques, les interactions entre cylindres étant de nature entropique. On obtient ainsi les fractions relatives de cylindres et de disques de part et d’autre du point de coexistence entre la phase isotrope « à tiges » et la phase nématique « à disques ».

Ces résultats complètent les données existantes de diffraction des rayons X obtenues dans la phase nématique.

Abstract

2014

The Cotton-Mouton coefficient was measured as a function of temperature and concentration in the isotropic phase of the ternary system potassium laurate (KL), 1-decanol, and D2O. Based on a micellar picture

of mixed rods and disks, and in the context of an entropy model for hard cylinders, the data indicate the relative fraction of rods and disks as the KL concentration is adjusted in the vicinity of the isotropic

2014

rod-like nematic

2014

disk-like nematic coexistence point on the concentration-temperature phase diagram. This data complements X-ray scattering results in the nematic phase.

Classification

Physics Abstracts

61.30G - 64.70M - 82.70K

As a function of temperature and amphiphile concentration, Yu and Saupe demonstrated the exis- tence of a biaxial micellar nematic (NBX) phase in the ternary system potassium laurate (KL), 1-decanol

and D20 [1]. At a fixed weight fraction of decanol

they found that a biaxial phase exists between two uniaxial nematic phases, viz., a cylindrical prolate

nematic (Nc) phase at high KL concentrations and a

disk-like oblate nematic (ND) at low concentrations.

In principle the three ordered phases can coexist

with the high temperature isotropic phase (I) at an

isolated multicritical (o Landau ») point, such that

the ordered-to-isotropic phase transition becomes second order at this point [2-4]. In addition, other topologies for the general lyotropic phase diagram

have also been discussed For example, only the isotropic and two uniaxial phases can coexist at a

Landau point with an Nc - ND transition at lower temperatures [6], or an intermediate biaxial phase

can appear with a finite linei of first order NBx - I tran-

sitions along the concentration axis [6].

Subsequent to the discovery of the biaxial phase

a number of related experiments were performed to

better understand both the macroscopic properties

of this phase and the structure of the constituent

micelles. Optical birefringence measurements, for

example, have been interpreted in terms of the biaxial order parameter [7-9]. In addition, both light scattering [10] and magnetically-induced birefringence experi-

ments [11] have been performed near the ND - NBx

transition to extract the order parameter exponent

and susceptibility exponent y, with apparently con- flicting results : in one case classical behaviour was.

observed [10], whereas in the other (with much higher

temperature resolution) critical behaviour was seen [11]. In the microscopic realm a number of investi- gations of the micellar configuration are also being aggressively pursued A dozen years ago Alben and Shih [3] and Alben [2] pointed out that a uniaxial-

biaxial nematic phase transition can take place in a system composed of long (uniaxial) rods and flat disks.

Although this model was and remains a popular interpretation of the micellar configuration in biaxial lyotropic systems, recent X-ray and neutron data have

also been interpreted [12, 13] in terms of inherently

biaxial micelles preferentially fluctuating about a given axis; the nature of the fluctuations then define whether a macroscopic oblate, biaxial, or prolate phase would obtain. Although such measurements have been unable to distinguish between the models,

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019860047060109700

(3)

1098

they have nevertheless yielded effective average mi- cellar dimensions in the nematic phases. In an early experiment Hendrikx et al. [14] scanned both tem-

perature at a fixed concentration and concentration at a fixed temperature, extracting average occupation numbers, diameters, and lengths (thicknesses) for the

rods (disks) in the two uniaxial phases. The results indicate a remarkably concentration-independent and temperature-independent effective shape within the entire range of each uniaxial phase, although no report was made of the shape changes very close to the uniaxial-uniaxial transition. (The authors report

no intermediate NBx phase, perhaps owing to the precise region of the ternary phase diagram in which

the experiments were performed) Unfortunately,

within the context of the mixed rod-disk (Alben) model, the analysis of reference [14] precludes a

determination of the relative concentration of disks in the rod-like phase, and rods in the disk-like phase.

This deficiency arises because of orientational averag-

ing in the azimuthal direction of the minority micellar species. Unless the X-ray experiment were performed

at very large q, for example, one would expect that

orientationally-averaged rods in the ND phase would

appear similar to the more ubiquitous disks. (This

can easily be understood by noting that a disk pos-

sesses a uniform in-plane density. An infinitesimally

thin rod, on the other hand, exhibits a 1 /r density

distribution when orientationally averaged about

one axis. This distribution is equivalent to a uniform background disk plus a cusp diverging at r = 0. The background term is identical to the real disk, whereas

the cusp can be properly resolved only if large scat- tering vectors are used)

More recently, Figueiredo-Neto et ale reported [12]

X-ray results for the experiments performed at a

fixed concentration while scanning temperature

through all three nematic phases. Analysing their

results in terms of the biaxial micelle model (although

unable to discount the possibility of the mixed rod-

disk model), they concluded that effective local micellar order over distances ç 500 A remains similar in all three phases near the biaxial region, and

that within each of the two uniaxial phases the globally averaged effective micellar dimensions are nearly temperature-independent. Here ç is the biaxial cor-

relation length. Thus in both experiments relatively

constant and well-defined average dimensions were

obtained within each uniaxial phase. A question now arises : using complementary techniques, is it possible

to investigate the distribution of micellar shapes at a given point on the phase diagram ? High field optical techniques [15, 16] have been shown to be useful in several micellar structural studies, and thus may

provide insight into the shape distribution in the

isotropic phase. To this end I performed a magneti- cally-induced birefringence experiment as a function

of sample composition and temperature in the iso-

tropic phase in the region above the Nc phase, close

to the ND crossover. My central result is a rapid

evolution of the quantity 0 with composition, where

4Y * (T - T*) C, C is the Cotton-Mouton coefficient,

and T* is the concentration-dependent supercooling

limit of the uniform isotropic phase. These results are

analysed in terms of an entropy model for hard right-

circular cylinders [17] in the mixed rod-disk picture,

and indicate the presence of a significant fraction of disks above the rod-like nematic phase, increasing

as the concentration W is reduced Unfortunately,

in terms of the inherently biaxial picture [12, 13], an analysis was not possible owing to local biaxial

correlations of the micelles.

Potassium laurate was obtained commercially from

Pfaltz and Bauer, Inc. and recrystallized twice from

absolute ethanol. Both 1-decanol (99 % purity) and D20 (99.8 %) were obtained from Aldrich Chemical

Company and used without further purification.

Samples were prepared to contain a fixed weight

percent 6.238+0-00’ of decanol, closely corresponding

to the compositions used to obtain the phase diagram

in reference [1]. The weight percent W of KL was varied from 25.475 % to 26.895 %, and the concen-

tration of D20 was adjusted accordingly. Samples

were contained in a 0.2 cm pathlength stoppered

cuvette housed in a temperature-controlled oven.

The oven was in turn situated in an 11.2 T Bitter magnet possessing a transverse optical port Details of the oven and birefringence apparatus are given

elsewhere [118]. The field H was swept from zero to 10 T in 30 s and the induced birefringence An was

computer recorded Over the duration of the sweep temperature control was better than 3 mK.

Well above T* the birefringence was found to be

linear in H2 and the Cotton-Mouton coefficient

C(W,1) == ð 4.n/ðH2IH=O was determined by a linear least-squares fit to the data. Closer to T * (T - T*

400 mK) small deviations from linearity were observed;

here the initial slope ð An/ðH2IH=O was determined

from a least-squares fit to a polynomial quadratic in

Fig. 1.

-

Extrapolated supercooling temperature T* of the isotropic phase as a function of weight percent potassium

laurate.

(4)

H2. Data were taken over a temperature range

T* T P,,: + 2 K, where Tg is the onset of a

small biphasic region, and in all cases C -1 was found to be linear in temperature, indicating a mean field susceptibility exponent y = 1. At all concentrations studied T,* - T* was found to be a few hundred

millikelvins ; this is unlike the case of the binary system caesium perfluorooctanoate (CsPFO) and H20, which

can exhibit values of r: - T* ;g 20 mK for low weight fractions of CsPFO [15]. In addition, the

transition remained discontinuous to fields of 10 T, contrary to the results of Saupe et al. [9, 11] who

observed no discontinuity in An vs. temperature at the Nc - I phase transition in the presence of a 1 T

aligning field.

Since the behaviour in the isotropic phase is apparently mean field, for weak ordering the Cotton- Mouton coefficient is given by [19]

where As and AX are the volume dielectric and

magnetic susceptibility anisotropies for fully saturated order, n = (8)1/2, and ao is the coefficient of the

quadratic term in the Landau free energy expansion.

The two fitted parameters T*(W) and O(W) were

obtained from the C(W, T) data using a logarithmic fitting procedure. Fitted values of T* vs. W are shown in figure I and the quantity cP[ = AB Ax/(9nao)] is plotted in figure 2. As is readily apparent 0 varies

Fig 2.

-

0[ =- ð An/ðH2IH=O] vs. weight percent potassium

laurate.

smoothly over the entire concentration range, with the fastest change occurring at lower concentrations.

To interpret the data in figure 2 we need a model

for each factor in equation (1), such as ao, åe, etc.

We deal summarily with n by noting that it is virtually

constant over the concentration range studied Next

we consider ao by invoking a model for the entropic

part of the free energy, similar to that calcu14ted by Onsager [17] but taken only to é>(Tr t.J2), where Q is

the nematic order parameter. Consider an assembly

of rigid, right-circular cylindrical micelles exhibiting only repulsive excluded volume interactions and,

for the moment, monodisperse in aggregation number

s ; we thus are dealing with a transition from the

isotropic phase to only a uniform Nc phase or uniform ND phase. In the context of a Landau model - Tr ao t.J2

represents to é>(Tr t.J2) an orientational entropy

change 8 per unit volume between the la = 0 and t.J =F 0 states. As argued earlier [15] this entropy has

a direct orientational component 80 and a transla- tionally-coupled orientational component 80r. For

weak ordering we assume Onsager’s orientational distribution function f(0) =( rx/4n sinh a) cosh ( a cos 0) [17], where rx4 oc Tr t.J2 for small a. Then to the level

of the second virial coefficient we can obtain [15] the entropy 8 to é>(Tr t.J2) and thus obtain

ao[ = - 8/Tr t.J2] :

where NA is the amphiphile concentration of the

sample, kB the Boltzmann’s constant, d the micelle diameter, and I its thickness (or length). Note that the first term represents 80 and the second term ST.

Later it will be argued that for an ensemble of all disks

or all rods of similar size, ao is nearly identical in both

cases. The function 0 [ct: Eq. (1)] will thus depend

almost exclusively on the shape dependences of the

factor As Ax.

We consider finally the relationships between Ax

and As and the shape of the micelle. Such an analysis requires a precise knowledge of the molecular confi-

gurations within the micelle, still a subject of much investigation. The amphiphile conformations within

spherical micelles, bilayers, and infinitely long rods

have been extensively investigated theoretically [20-24], although the problem of finite, anisometric micelles has yet to be tackled (Recent experimental work by

Charvolin and Hendrikx [25] on oblate micelles will

hopefully stimulate such calculations.) Thus a gross simplification will be introduced : we consider right-

circular cylinders and neglect «end caps » for the rods and « rims » for the disks [26]. Moreover, we

assume that for disks the amphiphile chains lie

perpendicular to the disk face and for rods they lie

in a plane normal to the micellar symmetry axis;

more about this later. The discussion will now be

specialized to disks. Since the amphiphiles within the

micelle are all parallel, Ax is obtained by summing

the individual magnetic susceptibility anisotropies.

Thus the susceptibility anisotropy per micelle scales

as s åXm, where AXm is the molecular susceptibility

anisotropy, and the susceptibility anisotropy per

volume AX scales as NA AXm. For As we can make a

(5)

1100

similar argument, although an important difference

should be noted In this case the summation of the molecular polarizibility anisotropies Op is not straight- forward, although a Lorentz term can be used to approximate the local field corrections. It turns out that for sufficiently small birefringence these cor-

rections are negligible and the optical dielectric anisotropy per micelle scales as s dp; the volume

averaged AR as seen in a birefringence experiment,

thus scales as N A ð.p. Gathering together all these factors, we find from equation (1) that for disks above the I - ND transition,

For rods the story is a bit different. Here the molecules lie in a plane normal to the symmetry axis with no

net orientation in the plane. Along the rod axis the micellar susceptibility scales as sxl ; in the normal

plane, however, the susceptibility scales as -1 s(xl + X II).

It thus turns out that the micellar susceptibility ani- sotropy scales as 2 s AX. and the volume anisotropy Ax scales as 2 Np LBXm’ a value one-half that of the disks. Similar geometric arguments apply to A8, and we find that for rods A8 scales as 2 NA Ap, again

one-half the value of disks. Thus by substituting the

rod-like values for Ax and A8 into equation (1), we

find that the functional form for Qrod = 1/4 0 disk above

the I - Nc transition, such that the appropriate rod-like

values of 1, d, and s are used in equation (3). The experiment reported herein is based upon this result,

that is, the measured values of tProd and tPdisk differ (primarily) because of their respective forms for

A8 Ax. If a mixture of rods and disks were present,

an average A8 Ax would obtain. The measured value of 0 would thus correspond to the number average

of 45rod and 45di.,,, and thus figure 2 reflects the relative concentration of the two micellar species.

Hendrikx et al. have used X-ray scattering [14] to

deduce the average micellar dimensions and occu-

pation number as a function of temperature and concentration. They found for disks I - 23 A, d - 64A

and s - 215; for rods 1 - 63 A, d - 36 Å and s ~ 185.

Although these measurements were performed at a slightly different set of temperatures and concentra- tions than my experiments, they probably represent reasonable values for use in equation (3). Upon substituting these values we find the 1- and d-dependent

factor in the denominator to be virtually identical

for both rods and disks and of order 1/3. Gelbart

et al. have also suggested [27] that a growth in rod-like micelles occurs in the Nc phase below the I -Nc

transition as the micelles attempt to minimize the loss in orientational entropy attendant upon align-

ment ; thus in the isotropic phase one might find even

slightly smaller micellar dimensions with a con- comitant reduced contribution from ST. The important point is that given the large experimental variation

of 4Y with W, to first order the actual values of tfJrod

should be approximately 4 (P disk where the Ts are

dominated primarily by their respective forms for AeA/, rather than the precise shape contribution

to ST.

Figure 2 shows the evolution of 4Y with W in the

region primarily above the Nc phase, where the Nc - ND transition occurs in the vicinity W = 25.2 %.

This is to be compared with the data of Hendrikx

et al. [14], which indicate concentration-independent

average micellar dimensions on either side of the concentration-driven Nc - ND transition in the nematic

phases; only within the range A W - ± 0.1 % of the

critical concentration (at constant temperature) does

a change in dimensions occur in their data. As men-

tioned earlier, the micellar shape distribution cannot be extracted in a low q X-ray scattering experiment.

In such an experiment the disks would appear washed out due to orientational averaging about the uniaxial

direction, and thus the statistics of the rod-disk mixture

are relatively inaccessible. Only at higher q can the

X-ray form factors be distinguished Figure 2, on the

other hand, shows a slow evolution of the quantity 0

with W. Although the analysis for ST, and therefore 4Y, is formally applicable only to an assembly of either

all rods or all disks, the virial corrections arising from ST are somewhat smaller than the experimental

variation in the quantity 0. Thus since (AS ex)dlsk = 4(As AX)rod for comparably-sized micelles, disks can

be experimentally distinguished from rods by an

increase in 0. In the lower concentration region,

for example, the data clearly indicate an increasing

fraction of disks in the mixed rod-disk picture; thus

even above the Nc phase, the fraction of disks can be

quite substantial, and grows with decreasing W.

Although 0 is expected to vary by a factor of four when going from all rods to all disks, the experimental

variation (cf. Fig. 2) is at least this over a more limited concentration range. Moreover, the curve does not

seem to be leveling-off at low W. There are numerous

possible sources for this discrepancy. For example, although ST was formally included in equation (3)

it was not deemed very important Yet this term can ultimately be responsible for a 20-30 % difference

between 0,.d and 45di,,,. In addition, the aggregation

number s of the micelles may be growing as W decrea-

ses. This hypothesis, however, is at odds with the data of Hendrikx et al. [14], who found the average dimen- sions within each nematic phase to be concentration-

independent In fact, in light of their results and those

presented herein, the shape distribution is probably bimodal, peaking about given positive and negative aspect ratios with relatively few spherical micelles.

This result is expected since rods and disks both

require smaller surface areas per head group [26]

than do spheres. Another potential difficulty in 0

(6)

arises from the neglect of « rims » and « caps ». On a

purely geometric basis these end corrections can account for nearly 50 % of the micellar volume, although the volume fractions turn out to be com-

parable for both rods and disks; it is therefore unlikely

that end corrections can account for a very large experimental variation in 4Y, although these cor-

rections can certainly add to an overall discrepancy.

A more significant source of potential problems is probably the molecular conformation within the micelles. Szleifer et ale have calculated [21] bond orientationaj order parameters Sk for spheres, infinite rods, and bilayers using representative bond energies Eb. They found that the Sk’s are largest for bilayers (by as much as a factor of 3, depending upon chain

length and Eb), indicating that the quantity Ag Ax

for disks could actually be considerably greater than four times that of rods, although it is very sensitive to the particulars of the model. Gruen, for example, performed a similar calculation [23], where he seg- mented the aggregate into onion-like shells and defined A., as the volume-weighted average of the

mean area of the shells. If Aav were comparable for all

three shapes, he found Sk relatively insensitive to

geometry. On the other hand, if the core radius were

held constant, results similar to those of reference [21]

were obtained Recent experimental work [25, 27-31]

on several micellar systems to determine bond order parameters has been helpful, although more extensive

data is necessary for a fully quantitative analysis of

the data presented herein. Charvolin and Hendrikx,

for example, have shown [25] that the shape of the Sk vs. k curve (where k is the carbon number along the chain) is similar in the lamellar, hexagonal, and

oblate nematic micellar phases of the KL system, and are all of the same order of magnitude, although

factors of two or three are crucial for a full analysis

of my results. Nevertheless, I’ve been able to obtain

semiquantitative results for the shape distribution in the rod-disk model using the simple conformation

picture outlined earlier.

The above discussion was based upon a model of

primarily rods crossing over into primarily disks.

As mentioned earlier, however, another picture has

been proposed in which inherently biaxial micelles

generate the three ordered phases [12, 13]. This is a particularly appealing model since experimentally

T* vs. W is relatively flat; in the mixed rod-disk

picture one might expect a more rapid variation in T*

as the shape distribution changes. Since local corre-

lations of the micelles play a fundamental role in such

a model, and the statistical mechanics have yet to be worked out, the data in figure 2 cannot yet be discussed in terms of such a picture. It is hoped that

the results of references [12] and [13] and those pre- sented herein will provide sufficient impetus for the

necessary calculations.

To summarize, Cotton-Mouton data taken in the

isotropic phase above the rod-like nematic phase

near the Nc - ND crossover indicate an evolution in the mixture of rods and disks in the rod-disk picture, complementing average size data obtained by X-ray scattering. At this time, unfortunately, the data cannot

be analysed in terms of the inherently biaxial micelle model.

Acknowledgments.

I wish to thank Dr. P. Photinos for his critical com- ments on the manuscript. This work was supported by the National Science Foundation, Division of

Materials Research under contract number DMR- 8211416 to the Francis Bitter National Magnet Laboratory.

References

[1] Yu, L. J. and SAUPE, A., Phys. Rev. Lett. 45 (1980) 1000.

[2] ALBEN, R., Phys. Rev. Lett. 30 (1973) 778.

[3] SHIH, C. S. and ALBEN, R., J. Chem. Phys. 57 (1972)

3055.

[4] PRIEST, R. G. and LUBENSKY, T. C., Phys. Rev. B 13 (1982) 4159.

[5] GRAMSBERGEN, E. F., LONGA, L. and DE JEU, W. H., Phys. Rep. (in press).

[6] ALLENDER, D. W., LEE, M. A. and HAFIZ, N., Mol.

Cryst. Liq. Cryst. 124 (1985) 45.

[7] GALERNE, Y. and MARCEROU, J. P., Phys. Rev. Lett.

51 (1983) 2109.

[8] GALERNE, Y. and MARCEROU, J. P., J. Physique 46 (1985) 589.

[9] SAUPE, A., BOONBRAHM, P. and Yu, L. J., J. Chim.

Phys. 80 (1983) 7.

[10] LACERDA-SANTOS, M. B., GALERNE, Y. and DURAND, G., Phys. Rev. Lett. 53 (1984) 787.

[11] BOONBRAHM, P. and SAUPE, A., J. Chem. Phys. 81 (1984) 2076.

[12] FIGUEIREDO-NETO, A. M., GALERNE, Y., LEVELUT,

A. M. and LIMBERT, L., J. Physique Lett. 46 (1985) L-499.

[13] HENDRIKX, Y., CHARVOLIN, J. and RAWISO, M., Phys.

Rev. B (to be published).

[14] HENDRIKX, Y., CHARVOLIN, J., RAWISO, M., LIEBERT, L.

and HOLMES, M. C., J. Phys. Chem. 87 (1983) 3992.

[15] ROSENBLATT, C., Phys. Rev. A 32 (1985) 1924.

(7)

1102

[16] ROSENBLATT, C., J. Physique Lett. 46 (1985) L-1191.

[17] ONSAGER, L., Ann. NY Acad. Sci. 51 (1984) 627.

[18] ROSENBLATT, C., J. Physique 45 (1984) 1087.

[19] STINSON, T. W., LITSTER, J. D. and CLARK, N. A.,

J. Physique Colloq. 33 (1972) C1-69.

[20] BEN-SHAUL, A., SZLEIFER, I. and GELBART, W. M., J. Chem. Phys. 83 (1985) 3597 ; Proc. Nat. Acad.

Sci. 81 (1984) 4601.

[21] SZLEIFER, I., BEN-SHAUL, A. and GELBART, W. M., J. Chem. Phys. 83 (1985) 3612.

[22] DILL, K. A. and FLORY, P. J., Proc. Nat. Acad. Sci.

77 (1980) 3115 ; 78 (1981) 676.

[23] GRUEN, D. W. R., J. Phys. Chem. 89 (1985) 146 ;

89 (1985) 153.

[24] OWENSON, B. and PRATT, L. R., J. Phys. Chem. 88 (1984) 2905.

[25] CHARVOLIN, J. and HENDRIKX, Y., Nuclear Magnetic

Resonance of Liquid Crystals, J. W. Emsley ed.

(P. Reidel) 1985, p. 449.

[26] MCMULLEN, W. E., BEN-SHAUL, A. and GELBART,

W. M., J. Colloid. Int. Sci. 98 (1984) 523.

[27] GELBART, W. M., MCMULLEN, W. E. and BEN-SHAUL, A., J. Physique 46 (1985) 1127.

[28] MELY, M., CHARVOLIN, J. and KELLER, P., Chem. Phys.

Lipids 15 (1975) 161.

[29] WALDERHAUG, H., SÖDERMAN, O. and STILBS, P. J.,

J. Phys. Chem. 88 (1984) 1655.

[30] CHARVOLIN, J., J. Chim. Phys. 80 (1983) 15.

[31] SEELIG, J. and SEELIG, A., Rev. Biophys. 13 (1980) 19.

[32] SEELIG, A. and SEELIG, J., Biochem. 13 (1974) 4839.

Références

Documents relatifs

Effect of micellar flexibility on the isotropic-nematic phase transition in solutions of linear

2014 We present an ultrasonic investigation of the nematic-isotropic phase transition in p-azoxyanisole (PAA). Our measurements were performed at several frequencies

on elastic interaction, recalling that its average energy is proportional to 1 /r per unit length of straight and parallel edge dislocations. Simple considerations

Transport currents flow in a layer of thick- ness A near the surface ; out of this layer the vortex system remains uniform.. Let us consider the inhomogeneous mixed phase in

In the framework of density functional theory 共DFT兲 all these studies showed that the interfacial tension of hard rods is minimal in the case of parallel anchoring, i.e., when

In the literature, several attempts to solve the boundary problem by means of hard walls have been reported. 26,30 In these studies, standard periodic images were used in two

2014 The influence of permanent dipoles on the nematic isotropic transition is studied by adding a simple polar interaction to the Maier-Saupe model of nematic

Precise measurements of the Kerr effect [6, 9] and magnetic birefringence [6, 7] in the isotropic phase of. some nematic substances have been