• Aucun résultat trouvé

Hydrodynamic properties of the smectic A phase formed in colloidal solutions

N/A
N/A
Protected

Academic year: 2021

Partager "Hydrodynamic properties of the smectic A phase formed in colloidal solutions"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00211123

https://hal.archives-ouvertes.fr/jpa-00211123

Submitted on 1 Jan 1989

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.

Hydrodynamic properties of the smectic A phase formed in colloidal solutions

Xin Wen, Robert B. Meyer

To cite this version:

Xin Wen, Robert B. Meyer. Hydrodynamic properties of the smectic A phase formed in colloidal solutions. Journal de Physique, 1989, 50 (19), pp.3043-3054. �10.1051/jphys:0198900500190304300�.

�jpa-00211123�

(2)

Hydrodynamic properties of the smectic A phase formed in

colloidal solutions

Xin Wen and Robert B. Meyer

Martin Fisher School of Physics, Brandeis University, Waltham, MA 02254, U.S.A.

(Reçu le 2 mai 1989, accepté le 2 juin 1989)

Résumé.

2014

Nous étudions les propriétés hydrodynamiques de la phase smectique A formée dans des suspensions de particules colloïdales. Nous avons trouvé

un nouveau

mode diffusif

en

rapport

avec l’écoulement relatif entre le solvant et les particules colloïdales. Dans la direction

perpendiculaire à la couche,

ce

mode est couplé

avec

la perméation entre couches. Le temps de relaxation du mode diffusif dépend à la fois des composantes parallèle et perpendiculaire à la

couche du vecteur d’onde. On trouve que le deuxième

son

est atténué par le cisaillement

hydrodynamique, la perméation et le mouvement visqueux entre le solvant et les particules

colloïdales. Dans cette théorie, le nombre total de modes hydrodynamiques est toujours égal

au

nombre de relations de conservation,

ce

qui est

en

accord

avec

la théorie générale de Martin,

Parodi et Pershan. Nous montrons aussi, par

une

simple estimation d’un système réel, que le second

son

est suramorti par l’effet de cisaillement hydrodynamique, et que le mode de diffusion relative est observable par diffusion quasi élastique de la lumière.

Abstract.

2014

The hydrodynamic properties of the smectic A phase formed in suspensions of

colloidal particles

are

studied. We have found

a new

diffusive mode related to the relative flow between the solvent and colloidal particles. Along the layer normal direction, this mode is

coupled to the permeation between layers. The relaxation time of the diffusive mode depends

on

the wavevector components both along and perpendicular to the layers. The second sound is found to be damped by hydrodynamic shear, permeation, and the viscous motion between the solvent and colloidal particles. The total number of hydrodynamic modes in this theory is always equal to the number of conservation relations, which is consistent with the general theory proposed by Martin, Parodi and Pershan. We also show by simple estimation in

a

real

experimental system that the second sound is overdamped by the hydrodynamic shear effect and the mode of relative diffusion is observable by the quasi-elastic light scattering technique.

Classification

Physics Abstracts

61.30

-

47.35

-

82.70D

1. Introduction.

There is currently great interest [1-6] in the lyotropic smectic A phase formed by rigid particles which interact mainly through steric exclusions. The phase is entropically driven and

has properties substantially different from the thermotropic smectic A phase. Although

several works [2-6] have considered the formation mechanism of the phase, its dynamic properties have not yet been investigated. Experimentally, the lyotropic smectic A phase is

observed in suspensions of colloidal particles, e.g. solutions [7] of tobacco mosaic virus

(TMV). Studying the hydrodynamic properties of these systems is fundamental for our

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198900500190304300

(3)

understanding of the lyotropic smectic A phase as well as the other ordered phases observed

in colloidal solutions.

The general hydrodynamic theory [8] in condensed matter was proposed by Martin, Parodi

and Pershan (MPP) in 1972. In the landmark work, MPP explicitly defined hydrodynamic

modes in a many body system as the few

«

long lived

»

collective modes that decay in times proportional to some power of their wavelengths. MPP also argued that the number of

hydrodynamic modes is always equal to the number of the conserved quantities, or « the independent degrees of freedom

».

In a single-component isotropic liquid, for example, there

are five hydrodynamic modes : a thermal diffusive mode, two diffusive transverse shear

modes, and a pair of propagating sounds. The number of hydrodynamic modes is determined

by the five conserved quantities characterizing the phase, namely, the energy density, the

mass density, and the three components of the momentum density. As the continuous

symmetries of the isotropic liquid are broken, new independent degrees of freedom will appear, resulting in new hydrodynamic modes. For the nematic ordering in which anisotropic particles orient in a preferred direction, the two possible orientational fluctuations represent

two new hydrodynamic modes. The free energy terms associated with the two orientation fluctuations are proportional to the squares of the spatial gradients of the director

distribution. Consequently the related relaxation times are proportional to the square of the

wavelength of the fluctuations.

In a smectic A ordering, the continuous translational symmetry is broken into periodic layers. Due to the symmetry requirement, directors are bound to the layer normal direction.

As a result, the director fluctuations relax in

«

microscopic times

»

which are determined by microscopic binding forces and which remain finite in the long wavelength limit. Clearly, the

director fluctuations in the smectic A phase are no longer hydrodynamic. The extra

«

independent degree of freedom

»

in the smectic A phase is the layer displacement [9, 10]

whose propagation forms a new hydrodynamic mode, the second sound wave. Similar to the director fluctuations in the nematic phase, the fluctuations of the layer displacement are

associated with a free energy increase determined only by the spatial gradient of the. layer displacement. The characteristic time of the fluctuation, namely the period of the second sound, is consequently proportional to the wavelength. For the smectic A phase formed in binary systems, there is an additional hydrodynamic mode associated with the mass

conservation of the second component. One special system of the latter case, the lamellar

phase formed by lipids and water, was studied [11] by Brochard and de Gennes in 1973. In this

work, we attempt to understand the hydrodynamics of the smectic A phase formed by particles suspended in solutions.

A solution of colloidal particles consists of two ingredients : the solvent and the particles.

The particles usually interact with some types of repulsive forces, either coulombic repulsion

or steric exclusion in origin, so that the particles are separated from each other and can be

suspended by thermal agitations. If the particles are charged, the strength of the repulsion

between the particles can be modified by the ionic concentration and the pH value in the solution. When the smectic A order forms, the particles distribute in a stack of layers and

orient preferentially in the layer normal direction. In the example of the TMV solution, a

virus particle [12] is an electrically charged rigid cylinder, 3 000 Â in length and 180 Â in diameter. The virus [13] has a molecular weight of 4 x 107 g/M and a specific volume of

0.73 cc/g. The smectic A phase [7] forms with layer thickness about 3 300 Â at concentration around 175 mg/ml (corresponding to a particle volume fraction of 0.20) in pH 8.5 and 50 mM

borate buffered solution. By simple estimation, one finds the in-layer area fraction [3] of the

particles to be about 0.18, which gives the mean distance between axes of particles within

layers to be about 400 Â, or a little more than two virus diameters.

(4)

Having an idea of the physical structure of real colloidal systems, we should always

remember that it is the conserved quantities that determine the types of the hydrodynamic

modes. Quantities describing the sizes, the physical shape, or the detailed form of interactions of the particles, however, only affect the magnitudes of the relaxation times of the

hydrodynamic modes. Comparing with a single-component smectic A system, we have in the colloidal solutions a new hydrodynamic variable, the local concentration of the solvent, and a

related conservation relation, the mass conservation of the solvent. From the argument of MPP, one would naturally expect an additional hydrodynamic mode in the smectic A phase in

colloidal solutions. Generally speaking, a new hydrodynamic mode is added in two possible

situations : 1) there is one additional diffusive mode ; 2) combining with an existed mode,

there can be a pair of acoustic modes. The addition of the second component may also introduce modes which are not hydrodynamic in nature. For example, the second component in a binary crystal gives rise to non-hydrodynamic optical modes which vibrate at frequencies

determined by the interactions between the two types of atoms. In the following sections, we

will show that the new hydrodynamic mode in the smectic A phase formed in colloidal solutions is related to the relative diffusion between the particles and the solvent. We also

study the effects of solvent on the properties of the other hydrodynamic modes.

2. The dynamic equations.

Consider a smectic A system of colloidal particles which is in equilibrium at temperature T in

a fixed volume. For studying hydrodynamic properties, we define three dynamic variables : 1)

0 (r) the bulk dilation of the solution defined as 0 (r) =1- p (r )/p , in which p (r) is the local

mass density of the solution and p the average of p (r) ; 2) y

=

aulaz the derivative of the

layer displacement u (r) along

z

the layer normal direction ; 3),6 (r) the relative dilation of the local solvent fraction defined as 6 (r) =1- p S(r )/p s, in which p S(r ) is the solvent mass in a

unit volume of solution and p S the average of p S(r).

With the defined variables, the conservations of the total mass and the solvent fraction can be written as :

where v and v’ are the velocities of the total mass and the solvent.

The free energy density [11] consists of the compressional (dilational) elastic energies and

the curvature elastic energy :

where el = (J, £2

=

y, E3

=

8, Cij are the elastic coefficients associated with the three

variables, and K is the splay elastic coefficient. Of the two côntributions to the free energy, the splay elastic term is of higher order in spatial gradients and is therefore a small contribution for long-wavelength fluctuations. As we will see, this term is important only

when wavevectors are parallel to layers.

Three forces [11] can be derived from the free energy expression : the pressure

the restoring force on the layers along the

z

direction

(5)

and the osmotic force on the solvent (the negative gradient of the osmotic pressure)

The acceleration equation, or the momentum conservation, can then be written as :

where Q’ is the viscous stress tensor.

The entropy source consists four dissipative processes :

which include four fluxes : Vv the shear rate tensor, vP - ù the velocity of particles relative to

the layer displacement in the

z

direction, vs - vP the relative velocity of the solvent to the

particles, J the heat flux density, and their respective conjugate forces : a’, g, f, and

E

= -

VT/T. In this work, we neglect any temperature gradient and associated heat- transport processes, since we are primarily interested in the hydrodynamic modes specifically

attributed to the binary lyotropic smectic A phase which is determined [7] mainly by

concentration instead of temperature.

Since fluxes are small for long wavelength hydrodynamic fluctuations, linear relations may be assumed between the forces and the fluxes :

(1) the symmetry of the smectic A phase requires Q’ and Vv to only couple to each other.

At the incompressible limit, the general relations [10] between Q’ and Vv are [11] simplified

as :

where i and j denote x or y, and 17e, 17t, and 17c are the longitudinal, transverse and cross

viscosities ;

(2) along the z direction, the permeation and the relative diffusion are generally coupled,

which is a new physical property of the smectic A ordering formed by particles in colloidal solution. When particles permeate through layers, this also involves a flow of particles relative

to the solvent in the same direction. The restoring force on layers is always opposite to the particle movement :

in which Q22 is the same as the permeation coefficient in an ordinary smectic A phase.

vg, the velocity of particles in the z direction is related to the other velocities by the

conservation of momentum : p vz

=

p S vi + p p vp in which p P

=

p - p S is the mass of particles

in a unit volume of the solution. When an osmotic force is established on the solvent in the

z

direction, either externally or due to thermal fluctuations, the solvent is driven to flow in the

same direction relative to the particles. At the same time, the particles are also dragged by the

solvent to permeate through layers. The force-flux relation of this process is described as :

According to the physical explanations of the two processes, the viscous coefficients

(6)

Qij are all positive. Equations (10) and (11) can be easily transformed into a mathematically

convenient form :

where P22 = Q22, P32 = Q32, P23 = (P Q23 + pSQ22)/pP and P33 = (P Q33 - pSQ32)/pP.

Since the solvent should flow relative to the total fluid in the same direction with

f z, we have P33

>

0 or p Q33

>

p S Q32· As we will see later, results can be simplified to more physically transparent forms, in the limit of weak coupling between the permeation and the

relative diffusion, namely, Q23, Q32 = 0, or P32 = 0 ;

(3) due to the uniaxial symmetry of the smectic A order, we may write in-layer force-flux

relations only in the x direction :

in which L is the viscosity of the relative flow in lateral directions.

So far, we have included six conservation relations : the two mass conservations (Eqs. (1)

and (2)), the three components of the momentum conservation (Eqs. (7)), and the

conservation relation for ù (Eq. (10)). Together with the energy conservation which is related to thermal relaxation, we have a total of seven conservation relations. From the theory [8] of MPP, we expect six hydrodynamic modes besides the ignored thermal relaxation mode.

3. The no dissipation limit.

In this section, the hydrodynamics is studied with the simplification [9] of ignoring all the dissipations. Mathematically, this approximation is equivalent to neglecting the imaginary part in the frequencies of all the hydrodynamic modes. In this way, wa can easily acquire an

overall picture of all the hydrodynamic modes in the system. In the next section, the second

sound and other slow modes will be studied with all the viscous processes. Without

dissipation, we may take P22, P33 and L in equations (10a), (lla) and (12) to infinity and obtain vz

=

ù, and v

=

v’. The viscous shear term in equation (7) may also be neglected. With

v

=

v’, one easily obtains from equations (1) and (2) à

=

6. Since we are interested in modes with amplitudes oc exp (i q . r + iw t ), this equation thus requires either w

=

0 or 5

=

0. The

first case, shown in the next section, is related to the purely damped mode of the relative diffusion between the particles and the solvent. In the rest of this section, we consider modes that satisfy 8

=

8.

Ignoring the small curvature elastic energy, F in equation (3) is now simplified as :

in which C il = C 11 + C 33 + 2 C 13,

>

CÍ2 = Cl2 + C 23, and C 22 = C 22. With this free energy

expression, f can be combined into - Vp in equation (7). So, within the difference in the elastic coefficients, the dynamics has become effectively the same as the smectic A phase [11]

in a single-component system.

With the rotational symmetry around the

z

axis, we may choose q = (q,,, 0, qz).

Equation (7) now reads :

(7)

Equation (16) gives a zero-frequency mode which will be related in the next section to the

hydrodynamic shear motion. Recognizing ÿ

=

a,ù

=

azvz, equation (14) becomes

Using axUx = 9 - aZvz = 8 - 00FF, equation (15) becomes

For nonzero solutions, the determinant of equations (17) and (18) should vanish :

Solving equation (19), one can easily obtain si, the speed of the ordinary acoustic wave, and

s2, the speed of the second sound. Since Cl,, the elasticity of the average density, is much larger than the osmotic compressibilities C22 and C33 (Sect. 5), one can obtain the simplified

results :

,

in which 0 is the angle between q and the z axis. Except for the small angular modulations,

s2 is isotropic and approximately equal to C11/p which is the expression for an isotropic liquid. s2 is exactly the same as the expression [10] for the smectic A phase with one component.

.

To conclude this section, we have six hydrodynamic modes (in addition to the thermal diffusion mode) in the no dissipation limit : two pairs of acoustic modes, and two zero- frequency modes. The velocities of the propagating modes are basically the same as those of a single-component smectic A phase. The additional variable 6 does not bring in new propagating modes. When dissipations are considered, the two zero-frequency modes are to

be related to the hydrodynamic shear motion and relative diffusions.

4. The low frequency modes.

It is shown in section 5 that the oscillation of the bulk dilations is much faster than the oscillation of the layer compressions and the relaxation of the solvent dilation. The first sound

can therefore be decoupled, within a good approximation, from the other hydrodynamic

modes. With this consideration, the first sound may be ignored by imposing 0

=

0, as we

concentrate on the properties of the rest of the four slow modes. We will look for modes

oc exp (i q - r + i w t ) and set qy = 0 without losing generality. Equations (1) and (3) then

become :

where the elastic coefficients C’s are isothermal relative to the bulk dilations of the first sound.

4.1 MODES WITH OBLIQUE WAVEVECTORS. - We start from the acceleration equations

(8)

For the motion in the y direction, one easily derives from equation (26) :

which is associated with the transverse hydrodynamic shear mode that propagates in the

(x, z ) plane and oscillates in the y direction.

In the following derivation of the two relations between y and 8, we will only keep the

q4 terms that are the lowest q power terms when q is parallel or perpendicular to the layers.

Using equations (25) and (22), we substitute p in equation (24) and obtain :

with

To express v, with y and 6, we get rid of vs in equations (10a) and (12) with equations (2) and (22) :

and obtain

Substituting qZ v, in equation (28), we have :

where

a =

p w - i q 2 ~ .

Another relation between y and 8 is obtained by first deleting û - v,, in equations (10a) and (lla) :

then substituting uz - v, in equations (33) and (12) with equations (2) and (22) :

Then we express g and f, by y and 8 with equations (5) and (6) :

where

In the limit of P32

=

0 (see Sect. 2), it is easy to see P1= P2

=

P33-

(9)

Nonzero solutions require the determinant of equations (32) and (35) to be zero, which gives after ignoring the Kq 2 qx term in equation (32) :

We obtain a purely damped mode at low frequency :

which is the frequency of a new type of hydrodynamic mode, the relative diffusion between the colloidal particles and the solvent. At the small P32 limit, this frequency becomes :

The relaxation time is a function of both qx and qz. This is reasonable because the relative flow between the colloidal particles and solvent can be both parallel and perpendicular to the

smectic layers. In contrast, the slip motion in the lamellar phase formed by lipids and water is

confined in the plane of lipid layers, and as a result, the relaxation time [11] only depends on

qx. The hydrodynamic mode of relative diffusion should generally exist in other orderings

formed in colloidal solutions, e.g., the nematic [15] and columnar [16] phases observed in the solutions of tobacco mosaic virus. In the case of the columnar phase, the relative diffusion

perpendicular to the columns should couple with the permeation process between columns.

Neglecting high power q terms, it is easy to obtain the other two roots of equation (37) :

which are the frequencies of the second sound. The real part is the same as the result obtained in the no dissipation limit (Eq. (21)). The damping part consists of three effects : the

hydrodynamic shear, the relative diffusion, and the permeation. In the limit of C23 --> 0, the damping part is reduced to the form of a single-component smectic A phase comprising only

the permeation and the hydrodynamic shear : i 2- 1 (C22 p- 22 1 q2 z + îîp - q2), This may be

expected since C 23 is the coupling coefficient between the layer compressions and the

dilations of the solvent fraction.

4.2 MODES WITH WAVEVECTORS PARALLEL TO LAYERS.

-

When q

=

(q,,, 0, 0), equation (27) gives w 1 = i p - ~ t qx, which describes the transverse shear modes oscillating in the y direction. The relaxation time of this mode is characterized by q, t the transverse viscosity

in the smectic layers. With qz

=

0, Cd34 in equation (40) becomes purely imaginative and the pair of second sounds become diffusive. To find the relaxation times of the diffusive modes,

we write the counterpart of equation (37) by setting qZ

=

0 in equations (32) and (35) :

One immediately obtains the frequency for the relative diffusive mode

(10)

and the frequencies for the coupled modes of hydrodynamic shear and the in-layer undulation

When Klq,, « qlp, the two types of motions become decoupled :

CJ)3 corresponds to the shear mode oscillating in the

z

direction, which is characterized by the

cross viscosity related to the z and lateral directions. w4 is the frequency of the in-layer

undulation mode which oscillates with constant layer distance and constant mass distributions.

The frequencies of the undulation mode and the two hydrodynamic shear modes are of the

same form as [10] that of a single-component smectic A phase.

4.3 MODES WITH WAVEVECTORS PERPENDICULAR TO LAYERS.

-

When wavevectors are

normal to the smectic layers, the hydrodynamic shear motions in the x and y directions become degenerate and give the same relaxation times (Eqs. (25) and (26)) described by the

cross viscosity :

The dispersion relations of the rest of the modes can be obtained similar to our approach

for the in-layer modes by solving the parallel of equation (37) with qx

=

0 in equations (32)

and (35). However, we will choose a more illustrative path by deriving this directly from the

basic dynamic equations. We first simplify the two mass conservation equations with

qx

=

0: dzVz

=

0 and dzV; =03B4 (Eqs. (22) and (2)). The former relation immediately gives

vz

=

0. We then apply these relations to equations (10a) and (11a), take the derivative over z on both sides, express g and f z by y and 5 with equations (5) and (6), and finally arrive at the following equations :

The determinant should vanish for nonzero solutions, which gives

Solving this equation we easily obtain two modes related to the coupled motion of the

permeation [10, 14] and the relative flow between the solvent and particles. If one imposes complete decoupling between the two variables, i.e., C23, P23, P32

=

0, one has the simplified

results :

describing respectively the interlayer permeation and the relative diffusion perpendicular to

the layers.

(11)

5. Experimental considerations.

In this section, we discuss the possibilities of experimental observation of the hydrodynamic

modes studied in this work. Specifically, we estimate the relaxation times and acoustic velocities in the smectic A phase [7] formed in solutions of tobacco mosaic virus. In addition,

the experimental observation is discussed in terms of the quasi-elastic light scattering (QELS) technique, which is commonly used [17] for observing dynamic fluctuations. The « slip

mode » in the lamellar phase, for example, has been observed [18, 19] with this technique. A typical QELS set-up can measure fluctuations of wavelength from a few thousand angstroms

to a few microns and in the time range from about 0.1 ps to seconds.

Due to lack of experimental data, little is known about the values of the elasticities and the viscosities of the smectic A phase in TMV solutions. Exact calculation of these quantities requires the consideration of detailed mechanical and structural properties of the ordering,

which is beyond the scope of the present work. For this reason, we will only make the order- of-magnitude estimation of the frequencies of the hydrodynamic modes. Since the density of a

TMV solution is about 1.05 g cm- 3 for a virus concentration of 0.175 g cm- 3, the velocity of

the first sound in the TMV system should be similar to the acoustic velocity in pure water

(1483 m/s at 20 °C), i.e. in the order of 103 m/s. At the wavelength of a few thousand angstroms, the period of the acoustic oscillations is in the order of 0.1 ns, which is too fast for the QELS technique. The hydrodynamic shear modes in the smectic A phase can couple to

the change in the refractive index when they propagate along and oscillate perpendicular to

the layers. Due to the periodic distribution of the refractive index in the direction

perpendicular to the layers, the shear motion normal to the layers should cause refractive

index fluctuations. The relaxation time of the shear mode in pure water is in the order of ns in

a QELS experiment. Since the shear viscosity q,, in the TMV smectic A phase is higher than

the viscosity of pure water (10- 2 dyn cm- 2 s at 20 °C), the corresponding relaxation time

(Eq. (44)) in the TMV smectic A phase therefore can not be observed by the QELS technique.

The hydrodynamic modes specifically attributed to the smectic A phase formed in colloidal solutions are the second sound and the mode of relative diffusion. The maximum speed of the

former is (C22/p )1/2 (Eqs. (21) and (40)). The relaxation time of the latter is approximately

determined by C33/L in the x direction and by C33/P33 in the z direction (Eqs. (39), (42) and (51)). For simplicity, we only estimate the relaxation times in the z direction. The osmotic

inter-layer compressibility C22 and the isotropic osmotic compressibility C33 may be estimated with the findings [20] of Crandall and Williams who measured osmotic compressibilities of the

colloidal crystals of polystyrene spheres with gravitational effect. They found that the order of

magnitude of the compressibilities is determined mainly by the number density of the crystal

instead of by the detailed form of the interaction between particles. For example, an ordinary

metallic crystal of density 1023 cm- 3 has compressibilities in the order of 1012 dyne cm- 2 ;

whereas for a colloidal crystal of density 1012 cm- 3, the osmotic compressibilities that

Crandall and Williams found are about 1 dyne cm- 2. Compressibilities approximately scale to

the number density. For the smectic A phase in TMV solutions, the number density is in the

order of 1015 cm- 3. We estimate by simple interpolation that C22 and C33 in TMV smectic A phase are in the order of 103 dyne cm- 2. As the mass density of TMV solutions is about 1.05 g cm- 3,

,

the maximum velocity of the second sound (Eqs. (21) and (40)) is then in the range of 0.1 -- 1 m S-I, which is much lower than that of the first sound. In the wavelength

range of a few thousand angstroms to microns, this maximum velocity corresponds to periods

in the order of 1 03BCS, which is within the detection range of the QELS technique. The

observability of the second sound also depends on the damping effects (Eq. (40)) involved in

(12)

the process. According to our previous estimation, the hydrodynamic shear motion

corresponds to a time scale of 0.1 ~ 1 ns, which is much faster than the oscillation of the second sound. The second sound is overdamped by the hydrodynamic shear effect and is

probably difficult to observe.

P33, the viscosity of relative flow along the z direction, may be estimated in analogy to the

calculation [11] of the slip coefficient in the lamellar phase. Ignoring the irregular shape of the

flow channels and the water molecules bound to the virus surfaces, we make two crude approximations for the order-of-magnitude estimation : 1) the viscous flow of water relative

to the TMV particles can be described by the macroscopic viscosity of water TI W ; 2) water

flows along cylindrical channels that are in contact with TMV rods separated by the mean

distance between particles. We have estimated in section 1 the average distance between TMV particles in layers to be about 400 Â. With the second approximation, one easily obtains

with simple geometric analysis the radius of flow channels RS = 142 Â. When considering the

relative flow, we may assume zero velocity for the TMV particules- Since flows are weak at the

long wavelength limit, Poiseuille flow can be assumed in these channels :

in which vi is the velocity of solvent along the z direction, r is the distance to the axis of a flow

channel, and fz is the osmotic force causing the flow. From this expression, it is easy to obtain the mean solvent speed along the z direction: v’ z

=

fz Rs21 (8 q W) and the average solution

velocity along the

z

direction vz

=

(p y? ) vi

=

fz R 2 P Y (8 ~w p ). Assuming small P32 limit (see Sect. 2), equation (lla) gives :

One easily obtains P33

=

2.4 x 1011 dyn s cm- 4 for the TMV smectic A phase. For the wavelengths of a few thousand angstroms, the relaxation time of the mode of relative diffusion is estimated to be in the order of 1 - 10 millisecond, which can easily be measured by

the QELS technique.

6. Conclusions.

Hydrodynamic properties of the smectic A ordering formed in solutions of colloidal particles

have been studied. We have found a new purely damped mode associated with the relative diffusion between colloidal particles and the solvent. The dispersion relations of the mode of relative diffusion as well as the first and second sounds are calculated in a frame that is strictly

consistent with the general theory [8] of hydrodynamics. By simple estimation in a real

experimental system, we have shown that the mode of relative diffusion should be observable

by the quasi-elastic light scattering technique. We have also found that the second sound is

overdamped by the hydrodynamic shear effect and is probably difficult to be observed. It is expected that a similar mode of relative diffusion also exists in the other ordered phases

observed in solutions of colloidal particles.

Acknowledgments.

We thank K. Migler, Dr. R. Hentschke, and M. P. Taylor for illuminating discussions and critical reading of the paper. This work is supported in part by the National Science

Foundation, through Grant No. DMR88-03582, and by the Martin Fisher School of Physics at

Brandeis University.

(13)

References

[1] STROOBANTS A., LEKKERKERKER H. N. W. and FRENKEL D., Phys. Rev. Lett. 57 (1986) 1452 ; STROOBANTS A., LEKKERKERKER H. N. W. and FRENKEL D., Phys. Rev. A 36 (1987) 2929.

[2] MULDER B. M., Phys. Rev. A 35 (1987) 3095.

[3] WEN Xin and MEYER R. B., Phys. Rev. Lett. 59 (1987) 1325.

[4] PONIEWIERSKI A. and HOLYST R., Phys. Rev. Lett. 61 (1988) 2461 ; HOLYST R. and PONIEWIERSKI A., preprint.

[5] SOMOZA A. M. and TARAZONA P., Phys. Rev. Lett. 61 (1988) 2566.

[6] TAYLOR M. P., HENTSCHKE R. and HERZFELD J., Phys. Rev. Lett. 62 (1989) 800.

[7] WEN Xin and MEYER R. B., The observation and measurement of the smectic A ordering in

colloidal solutions of tobacco mosaic virus, to be submitted.

[8] MARTIN P. C., PARODI O. and PERSHAN P. S., Phys. Rev. A 6 (1972) 2401.

[9] DE GENNES P. G., J. Phys., Colloq. France 30 (1969) C4-65.

[10] DE GENNES P. G., The Physics of Liquid Crystals (Oxford University Press) 1974.

[11] BROCHARD F. and DE GENNES P. G., Pramana, Suppl. 1 (1973) 1.

[12] CASPAR D. L. D., Adv. Protein. Chem. 18 (1963) 37.

[13] Thesis of Seth Fraden, Brandeis University (1987).

[14] HELFRICH W., Phys. Rev. Lett. 23 (1969) 372.

[15] OLDENBOURG R., WEN X., MEYER R. B. and CASPAR D. L. D., Phys. Rev. Lett. 61 (1988) 1851.

[16] WEN Xin and MEYER R. B., to be submitted.

[17] BERNE B. J. and PECORA R., Dynamic Light Scattering (John Wiley and Sons, Inc., New York)

1976.

[18] CHAN W. and PERSHAN P. S., Phys. Rev. Lett. 39 (1977) 1368.

[19] NALLET F., Roux D. and PROST J., Phys. Rev. Lett. 62 (1989) 276.

[20] CRANDALL R. S. and WILLIAMS R., Science 198 (1977) 293.

Références

Documents relatifs

gauge are a result of violent phase fluctuations in that gauge leading to the Landau-Peierls destruction of long range order in the smectic phase.. In

tion is condensed (namely the first harmonic p2qo), SA, a smectic A phase in which both the fundamental P qO and the first harmonic p2Qo of the density modulation are

2014 A model based on second rank tensor orientational properties of smectic phases is developed.. The model predicts a second order smectic A-smectic C phase transition

- Experimental phase diagram showing the existence of the three smectic phases : At, A2 and A (from

Considering the above one has to conclude that the smectic A-smectic C phase transition must be due to the appearance of a new order parameter ( X ),..

NAC point as a Lifshitz point [9] implies the first order character of the N-C transition as a result of the tilt fluctuations in the nematic phase; furthermore the N-C

Colloidal crystals are an aqueous suspension of colloidal particles which organize themselves in long-range-ordered crystals so that their description requires both

two-dimensional order (uncorrelated layers) the sam- ple mosaic combined with the two-dimensional fea- tures of the scattering, which consists of rods perpen- dicular to