• Aucun résultat trouvé

Surface interactions in simple electrolytes

N/A
N/A
Protected

Academic year: 2021

Partager "Surface interactions in simple electrolytes"

Copied!
8
0
0

Texte intégral

(1)

HAL Id: jpa-00210767

https://hal.archives-ouvertes.fr/jpa-00210767

Submitted on 1 Jan 1988

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Surface interactions in simple electrolytes

Roland Kjellander, Stjepan Marčelja

To cite this version:

Roland Kjellander, Stjepan Marčelja. Surface interactions in simple electrolytes. Journal de Physique,

1988, 49 (6), pp.1009-1015. �10.1051/jphys:019880049060100900�. �jpa-00210767�

(2)

Surface interactions in simple electrolytes

Roland Kjellander and Stjepan Mar010Delja

Dept. of Applied Mathematics, Research School of Physical Sciences, Australian National University, Canberra, ACT 2601, Australia

(Requ le 2 octobre 1987, accepté le 30 novembre 1987)

Résumé. - Nous présentons des résultats et des concepts

nouveaux

qui proviennent de l’application rigoureuse des méthodes d’équation intégrale de la physique des liquides

au

problème des fluides ioniques qui

sont inhomogènes

au

voisinage de surfaces chargées. Cet article rassemble nos résultats récents obtenus par

l’approximation HNC anisotrope pour les interactions entre surfaces dans les électrolytes

mono-

et divalents.

Par comparaison

avec

la théorie de champ moyen de Poisson-Boltzmann, les différences principales sont

une

contribution attractive

aux

interactions entre surface due

aux

corrélations ion-ion et

une

répulsion causée par les

c0153urs

durs des ions. L’attraction est particulièrement forte dans les électrolytes divalents, la répulsion de

double couche est spectaculairement réduite

ou

transformée

en

attraction. Nous indiquons quelques conséquences de cette contribution

aux

pressions ainsi que son lien

avec

les interactions de Van der Waals. La taille des ions intervient dans la pression surtout pour les faibles distances entre surfaces, et principalement

pour les ions de grands rayons

ou

à forte densité de charge.

Abstract.

2014

We present

some new

results and concepts which have emerged from the rigorous application of

the integral equation methods of liquid state physics to the problem of inhomogeneous ionic fluids in the

vicinity of charged surfaces. The report summarizes

our

recent results for surface interactions in mono- and divalent electrolytes obtained using the anisotropic Hypemetted Chain approximation. Compared to the

Poisson-Boltzmann mean-field theory the major differences

are an

attractive contribution to the surface interaction due to ion-ion correlations and

a

repulsion from the ionic cores. The attraction is particularly strong in divalent electrolytes, where the double-layer repulsion is dramatically weakened

or

turned into

an

attraction. Some consequences of this pressure contribution and its connection to Van der Waals interactions

are

pointed out. The size of the ions affects the pressure predominantly at short surface separations, and then

to

a

substantial extent for large ion radii

or

at high surface charge.

Classification

Physics Abstracts

82.70

1. Introduction.

The introduction of integral equation methods and computer simulations into liquid state physics has greatly advanced the theoretical understanding of

bulk solutions of simple electrolytes. In the last three

years, those methods have been fully applied to the

related problem of electrostatic interactions between

charged surfaces or between colloidal particles in

ionic solutions. The specific problems associated

with accurate treatment of simple inhomogeneous

Coulomb fluids in the vicinity of charged surfaces

have now been overcome, and a limited number of results has been obtained within the primitive model

of electrolyte solutions. The calculations of surface interactions have mainly been performed in Lund (Sweden) by Monte Carlo (MC) simulations [1-3]

and in Canberra (Australia) via the anisotropic Hypernetted Chain (HNC) approximation [4-7].

The accurate theoretical results are of particular

interest in view of the experimental information on

the interaction between charged surfaces immersed in aqueous electrolyte solutions [8] brought by the

introduction of the molecular force measuring ap-

paratus in Canberra some ten years ago. In evaluat-

ing surface interactions, the present calculations represent the first major progress beyond the Pois-

son-Boltzmann (PB) equation, used by Gouy and by

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019880049060100900

(3)

1010

Chapman at the beginning of the century. The excellent reviews by Carnie and Torrie [9] give a good background of various double layer theories.

Primitive model electrolytes consist of charged hard-sphere ions immersed in a continuum dielectric medium. The use of the primitive model for electric double layer systems is a slight generalization of the

usual modelling of such systems in physical or colloid chemistry : point charge ions in a dielectric con-

tinuum. In the conventional treatment

-

the DLVO

theory [10, 11]

-

surface interactions are calculated

on the basis of the PB equation with the addition of

the Van der Waals force. The DLVO theory is still

in widespread practical use, but since it is based on

the mean-field PB treatment of the ion interactions

-

neglecting ion-ion correlations

-

in some circum- stances it gives erroneous predictions. The aniso- tropic HNC and the simulation studies give accurate

results within the primitive model. However, in a description of the inhomogeneous systems on this level it is still not feasible to go beyond the simple

Coulomb and spherically symmetrical core interac-

tions and explicitly include the structure of aqueous solvent. As a result, the interaction of surfaces in aqueous solutions of electrolytes at small separations

remains poorly understood.

In this paper we shall present some of our recent results for the interaction between charged plates

immersed in electrolyte solutions as calculated using

the anisotropic HNC method. For simplicity, we

shall restrict ourselves to systems where the dielectric

constant of the plates is equal to that of the solvent,

i.e. we neglect image charge effects. However, the anisotropic HNC method is capable of handling the image charge interactions without any further

approximations [4, 5], and this topic will be briefly

discussed in the concluding section in conjuction

with Van der Waals forces.

The HNC closure was selected because of its known accuracy in .calculations for particles interact- ing with Coulomb potentials. The accuracy of the calculated density profiles and pressures has been checked by comparison with the MC simulations [3, 4, 7]. While relatively few suitable MC results are

available, in all cases we have found good agreement between the HNC and the simulation results.

2. Theory and method.

The HNC approximation for the pair distribution function 9 ij (r, r’ ) between ionic species i and j is to

set

where /3 = (kB T)- 1, kB is Boltzmann’s constant, T is the temperature, Uij is the pair potential, hij = gij - 1 and Cij is the direct correlation function. The

functions hij and cij are (by definition) related by the

Ornstein-Zernike equation

where nk (r ) is the number density of species k.

These equations are used in the standard HNC

theory for homogeneous systems, in what case n is independent of r and all pair functions are isotropic,

i.e. only depend on r - r’ . For inhomogeneous

systems, like the diffuse double layer, the density

varies from point to point, and the pair functions are anisotropic. The set of equations (1), (2) is no longer complete and has to be supplemented by an equation

for the ion density distribution. In the anisotropic

HNC approximation, as adopted by Kjellander and Mar6elja [4, 5], equations (1) and (2) are used as they stand, and the density distribution is determined

by

where j

and where’ = exp (,B tk i )/A 3 is the activity of species i, JL the chemical potential, Ai the thermal

wavelength, t/J av (r) the average electrostatic poten- tial at r, t/J 00 the potential at infinity and ues the

electrostatic pair potential between the ions. The function ffINc(r) gives the contributions from the

pair correlations of the ions. Incidentally, we remark

that without f HNC we would obtain the density

distribution of the PB approximation (in which case 4’a, should be evaluated in the same approximation).

It is noted that the explicit expression for frNc(r)

given above originates [5] from a general formula for the chemical potential evaluated in the HNC ap-

proximation. Therefore its validity is restricted to that case.

Equations (1)-(3) together with the condition of

electroneutrality form a complete set of equations

that can be solved numerically. For an open system in equilibrium with a bulk salt solution, the mean

chemical potential JL:t from the bulk is inserted in

equation (3) (for details cf. Ref. [7]). We note that

in planar geometry, for instance for a liquid between

two walls, the density only depends on the coordi-

nate perpendicular to the walls, n (r) = n (z) (pro-

vided the ion-wall potential is also a function of

z

only). The pair functions depend on three indepen-

dent coordinates, e.g. the z-coordinate of the two

(4)

particles and their separation. If the two surfaces are equal, which we shall assume, the function n (z) is symmetrical around the midplane.

We solve [5] the problem in practice by dividing

the space between the surfaces into many layers, and evaluating the structure of the equivalent physical

system : a multi-species two-dimensional fluid,

where each species corresponds to one layer of the

actual physical system. Typical calculations may involve some hundred layers, with denser subdivi- sions closer to the surfaces. In the limit of many, very thin layers, one obtains solutions for the

inhomogeneous fluid with the same accuracy as that

normally achieved in the corresponding homo-

geneous fluid calculations. The method has been described in our earlier publications [5, 7]. It suffices to say here that equations (1)-(3) are solved by a

double iterative procedure. Given a starting profile,

for example from the PB theory, equations (1), (2)

are solved by iteration for the pair correlation functions. These correlation functions are inserted in

equation (3) and a new profile is determined itera-

tively. The new profile is then inserted in equa- tions (1), (2) and the whole procedure is repeated

until full self-consistency is attained. The long-range

tails and the discontinuities caused by the Coulomb and the hard core potentials (in both real and Fourier space) are subtracted off and treated analyti- cally.

Once the density profile and the pair distribution functions have been calculated, various ther-

modynamical quantities such as the pressure between the walls and the free energy, as well as the local and the average electrostatic potentials are easily deter-

mined. The internal pressure due to the ions is given by

where ni (0) is the concentration at the midplane

between the surfaces, Pei is the electrostatic force per unit area across the midplane due to the ion-ion interactions (this is non-zero since each ion on one

side affect the ion distribution on the other side) and Pcorc is the pressure component due to core-core

contacts across the midplane. Explicit formulae for these components can be found in reference [7].

The net pressure between the surfaces equals P net = Pint - Pbulk, where P bulk is the pressure of the bulk electrolyte with which the system is in equilib-

rium. The bulk pressure is analogously composed of

an ideal contribution kB T £ npulk and an electrostatic

i

and a core contribution, which should all be calcu- lated within the same approximation as the internal pressure. We accordingly have

Note that in the PB approximation, which neglects pair correlations, the last two brackets in equation (5) are identically zero, and only the first, ideal contribution remains.

3. Density profiles.

The ion density profiles show large deviations from the mean-field results only for rather high surface charge densities. Near the surfaces, the density profile can have a secondary peak indicating layering

due to the crowding of the counterions. This

phenomenon was first found in simulation studies for a single surface [12], and has since been con-

firmed by various integral equation techniques [7, 13]. The anisotropic HNC calculations for two surfaces [7], show that the secondary peak increases

in size at short surface separations.

At low to medium surface charge densities, the

deviations of the profiles from the PB results are

larger for coions than for counterions on a relative

scale, but on an absolute scale the differences are not large. Very good theoretical profiles have been

obtained earlier with a simpler approximation : the

modified PB equation [14, 15]. Comparison of those

results with extensive MC simulations is available in references [12] and [15, 16]. While those earlier workers had the technical framework to examine surface interactions, this apparently has not been attempted.

In contrast to density profiles, the surface-surface interactions are very sensitive to the deviations of the ionic fluid structure from the predictions of the

mean-field theory. Whilst the overall ionic density

may be very close to the mean-field value, the correlation between ions will significantly change the

force balances in the system, which leads to a strong attractive contribution to the surface-surface interac- tion. In cases where the bulk reservoir solution contains dissolved salt, this sensitivity is further enhanced by the fact that the net pressure is the difference between the internal pressure (between

the surfaces) and the external pressure (in the bulk).

4. Surface interactions in monovalent electrolytes.

We begin the examination of surface interactions in the presence of monovalent ions by considering one component electrolytes : solutions containing only

counterions. The system without coions is numeri- cally faster and simpler to evaluate. The results are

also of relevance at low salt concentrations when the surface separation is short. The coions are then

largely excluded from the space between the sur- faces. Under such conditions, the internal pressure between the surfaces in the presence of salt is hardly distinguishable from the pressure in cases where counterions are the only ion species in the system.

The interaction between the surfaces is affected by

(5)

1012

two major contributions which are not present in the PB theory : the attractive effect of electrostatic correlations and the repulsive pressure of the hard

cores. Each of these terms can dominate the other,

and the pressure can be either larger or smaller than the corresponding PB value (Fig. 1). The hard cores

are only important at small surface separations and

Fig. 1. - The pressure

as a

function of separation be-

tween two charged surfaces interacting

across an

electro- lyte phase

as

calculated in the PB and anisotropic HNC

theories. The pressure P/RT, where R is the gas constant and T the temperature, is measured in mol dm-3 [M]. The

surfaces have

a

uniform surface charge density of

one

elementary charge per 0.6 nm2. The electrolyte phase

consists of monovalent counterions with radii 2.3 A. The dielectric constant is 78.358 and the temperature 298.15 K (this also applies to all other cases below unless explicitly

stated otherwise). Throughout this paper the surface

separation is defined

as

the distance between the points of

closest approach of the ion centers to the two surfaces.

Fig. 2.

-

Interaction between the

same

surfaces

as

in

figure 1 for various counterion radii. The pressures

are

calculated within the anisotropic HNC theory and

are

shown relative to the PB values by plotting the ratio

between the HNC and the PB results.

relatively high surface charges. This is illustrated in

figure 2, where at short separations the increase in

repulsion compared to the PB theory is clearly

attributable to the effect of ion radius.

Next, we treat systems with added salt, i,e. in equilibrium with a bulk electrolyte solution (reser- voir). From figures 3 and 4 it can be seen that

Fig. 3. - Interaction (P/RT[M ] ) between two plates

with uniform surface charge density of

one

elementary charge per 0.85 nM2 in the anisotropic HNC and the PB theories. The plates

are

immersed in

a

1.0 M 1:1 electro-

lyte solution which consists of ions with the radius of 2.125 A. The temperature is 298.0 K and the dielectric constant 78.5.

Fig. 4.

-

Some examples of the interaction (relative to

the PB pressure) between two uniformly charged plates in

1:1 electrolyte solutions. The three systems have the following parameters (listed from top to bottom) : 0.1 M electrolyte, 0.6 nm2 per unit surface charge and ion radii 2.3 A ; 0.5 M, 0.714 nm2 and 2.3 A ; 1.0 M, 0.85 nm2 and 2.125 A (same

case as

in Fig. 3) respectively. The symbols

at the end of the 0.5 M

curve

signify points with much

lower relative accuracy than the full

curve.

The net

pressure for these large separations is

a

very small

difference between two large numbers.

(6)

beyond the short separation region where hard core

effects are important, the double-layer repulsion is

much smaller than that expected from the PB

theory. The effect is most dramatic at higher salt

concentrations. In figure 3 we have selected a surface charge of 0.188 C m-2, which for separations above

0.85 nm corresponds to the constant potential example presented by Lozada-Cassou and Hender-

son [17]. The calculated interaction is very different from that of reference [17], which uses a simpler theory, a singlet HNC theory, where the direct correlation function for the inhomogeneous electro- lyte solution is approximated by its bulk values.

For concentrations up to about 0.1 M, the interac- tion at large separations appears as a roughly exponential law, shifted by an approximately con-

stant factor with respect to the PB values. For the 0.1 M case, figure 4, this behaviour is evident at

separations larger than about 2 nm, where the ratio between the HNC and the PB results is nearly

constant. At 0.5 M and 1 M electrolyte concen- trations, the PB predictions are hardly ever valid as

seen from the examples presented in figure 4.

5. Surface interactions in divalent electrolytes.

In case of divalent ions, our currently available

results are restricted to one-component electrolytes.

The reason is the increased difficulty in achieving

sufficient numerical accuracy in calculations of the

rapidly varying anion-cation correlation functions.

While this does not greatly affect the calculation of ion density profiles, the interaction values are not

yet sufficiently reliable ; but work on the improve-

ments is currently in progress. Nevertheless, we can

say that in the case of 1 : 2 and 2 : 2 electrolytes we

still find the attractive double-layer interaction which is the most interesting feature of the results

presented below. It should be noted that excepting

very large surface separations, the one-component electrolyte results are not substantially changed

when millimolar amounts of salt are added.

In figure 5, we begin the examination of the interaction by gradually increasing the surface

charge density from a very low initial value. Com-

pared to the PB theory prediction the repulsion is substantially reduced, and the ion radius which

equals 2.125 A has no significant effect on the

results. On further increasing the surface charge density (Fig. 6) the repulsive force is overcome, and at moderate surface separations the net result is a double-layer attraction. This behaviour of the double

layer interaction at increasing coupling strength has previously been established by MC simulations [1]

and anisotropic HNC calculations [4]. In a separate

publication [18], we have argued that this attractive double-layer interaction is responsible for the very limited swelling of calcium clays in water. They

Fig. 5.

-

Interaction within the anisotropic HNC theory (relative to the PB pressure) between plates with various surface charges. The one-component electrolyte consists

of divalent counterions with radii 2.125 A. The values of the surface charge

are

(from the top to the bottom)

one

elementary charge per 100 nm2, 40 nM2 , 20 nm2, 10 nm2,

4 nM2 and 2.5 nM2 respectively.

Fig. 6. - Same

as

in figure 5, but for

one

elementary charge per 2.5 nm2 (same

as

the last

curve

of Fig. 5), 2 nm2, 1.35 nm2 and 0.6 nm2. Note the change in the separation scale.

remain stable at surface-surface separations between

the clay platelets of approximately 0.6 to 1.4 nm.

Furthermore, the appearance of attractive double

layer forces when changing from monovalent to divalent counterions in some surfactant-water sys- tems can explain the phase behaviour of such systems [19].

But even the reduced double-layer repulsion found

at larger surface separations and lower surface

charges has important practical consequences. In

figure 7, we have presented the interaction at the

constant surface separation of 25 nm as a function of

surface charge. If the values of the interaction

(7)

1014

Fig. 7.

-

The HNC and the PB values for the interaction

(P /RT[mM]) between two uniformly charged plates at

the constant separation of 25

nm as a

function of the

area

per unit surface charge. The

one

component electrolyte

consists of divalent counterions with radii 2.125 A. Note,

for example, that the HNC pressure between two surfaces with

a

surface charge of 12.6 nm2 per unit charge is the

same as

the PB prediction for surfaces with 25.0 nm2 per

unit charge.

measured in some experiments are lower than those

expected from the PB theory, one may be inclined to

incorrectly interpret the deviation as resulting from a

lower surface charge due to ion binding to the

surface (see the example given in the caption of Fig. 7). As we have seen, the decreased repulsion is

a consequence of the ion-ion correlations. Of course, this does not rule out ion adsorption in particular cases, but it means that the amount of

such adsorption will be overestimated if determined

by a comparison based on the PB pressure.

6. Concluding remarks.

Accurate results for surface-surface interactions in

electrolyte solutions have so far been obtained using

the Monte Carlo simulation method and the aniso-

tropic HNC approximation described in this report.

In practice, these numerical methods are currently

limited to calculations of the pressure in systems where the surface separation is not too large. If

one adopts the Debye-Huckel approximation

Cij(rl’ r2) = - Uij(rl’ r2)/kB T for the inho-

mogeneous fluid between the surfaces, the asymp- totic forms of the surface-surface interaction at large separations may be calculated analytically [20].

While this work is still in progress, the results obtained so far are very encouraging for the extreme

surface separation regime which is inacessible to the numerical methods.

Within a wider point of view, the attractive correlation contribution to the surface-surface inter- action discussed in this work is a part of the static

(zero frequency) term of the Van der Waals force.

This fact becomes apparent when one discusses interaction in the presence of electrostatic images,

where consistent results are obtained only when the

zero frequency Lifshitz interaction between the dielectric media is added to the contribution calcu- lated from the electrolyte between the surfaces [20- 23]. The Lifshitz theory only considers the Van der Waals interaction due to correlations in polarization

fluctuations in the dielectrics. When ions are present,

one has to consider the additional electrostatic fluctuation forces due to correlations in charge

fluctuations (ion-ion correlations). Furthermore, the charge and the polarization fluctuations are interde-

pendent, as described by the image charge interac-

tions. To obtain fully consistent theoretical treat- ment, all of these interaction mechanisms have to be included on an equal footing. In the anisotropic

HNC theory this can be done without any other

approximations [5]. In the results [22-23], the static

contribution to the Van der Waals force between the dielectric media is screened by the electrolyte as a

consequence of the cancellation [20-22] of the static

Lifshitz term by a term due to the ion-image charge

interactions.

The need to extend Van der Waals interactions to include the response of the electrolyte was well appreciated some 15 years ago [24]. We have now an

accurate method to actually carry out that pro- gramme for the primitive model electrolytes. What perhaps was not obvious before the present calcu- lations is that the effect of ion correlations can be

extremely dramatic, and qualitatively change the picture based on the classical mean-field theories.

Note added in proo f : The attractive double layer

interactions for divalent ions have recently been

confirmed experimentally by direct measurements of the force between mica surfaces in 0.15 M CaCl2

solution. Both the strength and the range of the attraction agree with HNC calculations for 1:2 elec-

trolytes. (R. M. Pashley, R. Kjellander, S. Marcelja

and J. P. Quirk, in preparation).

(8)

References

[1] GULDBRAND, L., JONSSON, B., WENNERSTRÖM, H.

and LINSE, P., J. Chem. Phys. 80 (1984) 2221.

[2] SVENSSON, B. and JÖNSSON, B., Chem. Phys. Lett.

108 (1984) 580.

[3] BRATKO, D., JÖNSSON, B. and WENNERSTRÖM, H., Chem. Phys. Lett. 128 (1986) 449.

[4] KJELLANDER, R. and MAR010CELJA, S., Chem. Phys.

Lett. 112 (1984) 49 ; 114 (1985) 124E.

[5] KJELLANDER, R. and MAR010CELJA, S., J. Chem. Phys.

82 (1985) 2122.

[6] KJELLANDER, R. and MAR010CELJA, S., J. Phys. Chem.

90 (1986) 1230.

[7] KJELLANDER, R. and MAR010CELJA, S., Chem. Phys.

Lett. 127 (1986) 402.

[8] PASHLEY, R. M. and ISRAELACHIVILI, J. N., J.

Colloid Interface Sci. 101 (1984) 511, and refer-

ences

therein.

[9] CARNIE, S. L. and TORRIE, G. M., Adv. Chem.

Phys. 56 (1984) 141 ;

CARNIE, S. L., Ber. Bunsenges. Phys. Chem. 91 (1987) 262.

[10] DERJAGUIN, B. V. and LANDAU, L., Acta Phys.-

Chim. USSR XIV (1941) 633.

[11] VERWEY, E. J. W. and OVERBEEK, J. Th. G., Theory of the Stability of Lyophobic Colloids.

(Amsterdam, Elsevier) 1948.

[12] TORRIE, G. M. and VALLEAU, J. P., J. Chem. Phys.

73 (1980) 5807.

[13] NIELABA, P. and FORSTMANN, F., Chem. Phys. Lett.

117 (1985) 46 ;

CACCAMO, C., PIZZIMENTI, G. and BLUM, L., J.

Chem. Phys. 84 (1986) 3327 ;

BALLONE, P., PASTORE, G. and TOSI, M. P., J.

Chem. Phys. 85 (1986) 2943.

[14] LEVINE, S. and OUTHWAITE, C. W., J. Chem. Soc.

Faraday Trans. 2 74 (1978) 1670.

[15] OUTHWAITE, C. W., BHUIYAN, L. B. and LEVINE, S., J. Chem. Soc. Faraday Trans. 2 76 (1980) 1388 ;

OUTHWAITE, C. W., BHUIYAN, L. B., ibid. 79 (1983)

707.

[16] TORRIE, G. M., VALLEAU, J. P. and PATEY, G. N., J. Chem. Phys. 76 (1982) 4615.

[17] LOZADA-CASSOU, M. and HENDERSON, D., Chem.

Phys. Lett. 127 (1986) 392.

[18] KJELLANDER, R., MAR010CELJA, S. and QUIRK, J. P., J. Colloid Interface Sci. , in press.

[19] KHAN, A., JÖNSSON, B. and WENNERSTRÖM, H., J.

Phys. Chem. 89 (1985) 5180.

[20] ATTARD, P., MITCHELL, D. J. and NINHAM, B. W., J. Chem. Phys. , in press.

[21] ATTARD, P., KJELLANDER, R. and MITCHELL, D. J., Chem. Phys. Lett. 139 (1987) 219.

[22] ATTARD, P., KJELLANDER, R., MITCHELL, D. J. and JÖNSSON, B., J. Chem. Phys. , in press.

[23] KIELLANDER, R. and MAR010CELJA, S., Chem. Phys.

Lett. 142 (1987) 485.

[24] MAHANTY, J. and NINHAM, B. W., Dispersion Forces

(London, Academic Press) 1976.

Références

Documents relatifs

Ensuite, l’influence du temps de greffage a été mise en exergue par des mesures de mouillabilité (angle de contact avec l’eau, taux de recouvrement, énergie de surface,

Restons calme, respirons gentiment Merci de

Tous les systèmes physiques de l’exposé sont régis par des équations différentielles d’ordre 1 dans leurs espaces des états.. Ils

For ions in a finite lattice with open boundaries, this means we have two type of states (see Fig. 2): states deep in the effective atom-ion potential which would not exist in a

Le composé est repris dans du dichlorométhane, la phase organique est lavée avec de l’eau afin d’extraire les sels, puis séchée sur sulfate de magnésium, filtrée et

Field ion microscope images recorded before and after the field- free heating interval were used to determine the temperature at which restructuring takes place and

-Comparaison d'un spectre de diffusion d'ions lents (lignes pointilldes) avec des spectres obtenus par simulation (lignes solides) pour diverses distances d entre la couche de

Ces interactions font intervenir les champs électriques appliqués simultanément sur la surface des cristaux piézoélectriques qui peuvent interagir entre eux et avec les