This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
= \drNp(r)P(rN), 2l\ + 2ZTJ) = fj((/>)• This is a consequence of a more general fact that solutions of (46) - Weierstrass Elliptic Functions - are doubly periodic functions in the complex plane [49]. Periods are connected with the invariants through the formulas a
(9)
A
where p (r) is the density operator A
p(r)=JT8(r-ri).
(10)
Note that p o (r)= N/V = p 0 for a homogeneous fluid (Eq. (8)). In an ideal gas, different particles are uncorrelated [UN(rN)=0] and, in the absence of an external potential, p/vW(r") » po". It seems then appropriate to define the ^-particle distribution function g{n\rn) by
Statistical Mechanics of Fluid Interfaces
gN("\V) = pN("\r")/]\9{r,),
5
(11)
where |~J p(r ; )= p o ' ! , for a homogeneous fluid. The particle distribution 1=1
functions defined in this way are a measure of the extent to which the structure of the fluid deviates from complete randomness. In particular, the pair distribution function gA(2)(ri,r2) plays a central role in the theory of liquids. In a homogeneous and isotropic fluid, g7v<2)(i"i,r2) depends only upon the relative distance between particles r 12 =|ri - r2|: gN(2)(rbr2)=g(rn). (12) The quantity g{rn) is called the radial distribution function of the fluid and is directly measurable by radiation-scattering experiments. Moreover, in systems for which the total potential energy UN is pair-wise additive, all thermodynamic properties of the uniform fluid can be expressed as a function of g(r]2) [5,8]. The distribution functions that we have considered have been defined in a closed system with a fixed number of particles N. However, we can generalize these ideas to open systems characterized by fixed values of V, T and chemical potential p,. In this case, the probability of observing n particles in dr\...drn at the space point r^, irrespective of the value of N, is p(n)(r")=X
PNWpK
(13)
where PN is the probability that an open system contains N molecules, namely
p =
* i(^]v
(14)
S in this equation is the grand partition function
*-%*?"•
<15)
and z is the activity z=A"3exp(Pn),
(16)
6
Vicente Talanquer
where A=(/22(3 /2mn )1/2 is the de Broglie thermal wavelength and h is Planck's constant. Thus, the ^-particle density in the grand canonical ensemble p(n) can be expressed as
P ( "V)= I^^—Jrfr"'-»)exp[-P(£/A,+ ^)].
(17)
Note that this function is normalized such that
W"V)
= <^>,
(18)
where <..> represents an ensemble average. The distribution functions in the grand canonical g(n)(r ") can be defined in the same way as we did for a closed system (See Eq. (11)). In particular, the pair distribution function is given by g(ri,r 2 )= g(2)(r ")= p(2)(r")/po(ri)po(r2), and the total pair correlation function h(r\,r2) is then defined as A(r,,r 2 )=g(r J ,r 2 )-l.
(19)
This correlation function /z(r1;r2) is a measure of the total influence of a particle at i"i on a second particle at r2 and vanishes in the limit r12=|ri - r2| -^ <x>. Based on ideas developed by Ornstein and Zernike, the total correlation function can be separated into direct and indirect parts. The direct part is given by a function c(i"i,r2) called the direct correlation function (Ornstein-Zernike correlation function), which is defined by the integral equation h(r,, r 2 ) = c(r,, r2) + \dr' c(r,, r3) po(r3) h(r3, r2).
(20)
The direct correlation function c(ri,r2) tends to be a simpler function than the distribution function g(r\,r2) and therefore can be more reliably approximated when working with inhomogeneous fluids. For a system of N particles with a total potential energy UN composed of pair potentials <j)(r,y),
UN(r")=J2^)>
( 21 )
the equilibrium particle densities p(n)(r") are coupled together by a set of (Af-1) equations. This can be demonstrated by differentiating Eq. (17) with respect to
Statistical Mechanics of Fluid Interfaces
7
the position of one particle in the set of n particles, for example with respect to r, [2]. If we divide the potential energy in Eq. (21) in three classes of terms u
» (rA')= E
(22)
the differentiation of Eq. (17) leads to
-kBIWi p W (r ")= p(n)(r ") fV, Fex,(r,) + ] T V.-XO) i=2 {
]
+l
+ \drn+ip "' \r" )V^(ru+,).
(23)
This set of equations is called the Yvon-Born-Green (YBG) hierarchy, and can be used to derive expressions for the distribution and correlation functions of an inhomogeneous fluid. For example, the first equation in the hierarchy can be expressed as -kBT Vp o (r,)= p o (r,) V V,JT\) + \dr2 p (2) (r b r 2 )V(Kr 12 )
(24)
or -kBT Vlnpo(r,)= V Vext{r,) + \dr2 g(\rhr2)Po(r2)V§{rn),
(25)
given that p(2)(ri,r2) = ^(2)(r1,r2)po(ri)po(r2). This equation can be solved to obtain the equilibrium density profile p o (r) for an inhomogeneous fluid given some means of evaluating the pair distribution function g (i"i,r2). For an ideal gas in which (t>(r12)=0, this equation leads to the Boltzmann distribution in the presence of an external field Po(r)=p o exp[-PfUr)],
(26)
where p o is the uniform density of the ideal gas in a field-free environment. In general, the YBG hierarchy is useful to derive the thermodynamic properties of a system only if a closure relationship between p (r") and p (r"+ ) is available. 2.2. Fluid Interfaces The previous formalism can be applied to the study of fluid surfaces, such as a liquid-vapor or a liquid-liquid interface. In such systems, the evaluation of
8
Vicente Talanquer
the single-particle densities for the different species p,(r) present in the fluid serves as the basis for the determination of a variety of interfacial properties, such as the interfacial tensions and surface adsorptions. In this section, we consider the case of a single monoatomic fluid with a free liquid-vapor interface to illustrate the application of the theory. 2.2.1. Planar interfaces Let us first assume that the fluid has a planar interface of area A parallel to the x-y plane and a total volume V. Although in the absence of an external field (e.g. gravity) one can expect the interface to be spherical, we may suppose that the liquid phase takes the shape of a sphere of infinite radius and thus zero curvature. The density profile for the system will then be a function of a single variable po(z), and should satisfy the boundary conditions po(z) -^ p o ' when z -^ - oo, and po(z) -^ pov when z -> + co , where p o ' and pov are the bulk densities of the coexisting liquid and vapor, respectively. If one assumes that the potential energy of the fluid can be expressed as a sum of pairwise intermolecular potentials ^(r^), the structure of the interfacial profile can be obtained using the first YBG Eq. (25), which for a planar interface reduces to
(27)
This equation can be solved by iteration to obtain the density profile po(z) given a suitable expression for the distribution function g(2\ru,z\,Z2). A natural approximation is to replace this quantity by the radial distribution function of a uniform fluid g(pm; ri2) evaluated at some mean value pm of the local densities at z\ and z2. Results obtained in this way for the Lennard-Jones fluid tend to agree well with those generated by computer simulations [15,16]. The grand potential Q of this system can be decomposed into a bulk and a surface contribution Q = -PV+yA,
(28)
where P is the pressure of the coexisting phases and y is the surface tension. The surface free energy y can also be expressed in terms of the integral of the difference between the normal and the tangential components of the pressure tensor across the interface [2,10]:
Statistical Mechanics of Fluid Interfaces
CO
9
00
Y= \dz\pN(z)-pr(z)]=
\dz[P-pT{z)l
(29)
~<x>
-co
where pN and pT are the normal and tangential components of the pressure. For a fluid in hydrostatic equilibrium in the absence of an external field, pN (z) is a constant equal to the equilibrium pressure P. Eq. (29) is usually called the mechanical definition of the surface tension. For a fluid with a configurational energy composed of pair potentials <j)(ry), the normal and tangential components of the pressure tensor for a planar surface are given by [2,10] 2
,
Z
pN(z)= kBTPo(z) - I \drn -n^hl 1 rn
r. 2
2
pT(z)= kBTPo(z) - I \drnXn
(2)
fdap dr
J
[r-ar, 2 , r+(l-a)r 12 ],
(30)
o
I
+y d
^ ^2) f j a P (2)[r-arl2, r+(l-a)r12]. (31) rn drn {
4
In a homogeneous fluid the pair density function p(2) is only a function of r)2, and these expressions reduce to P= kBTp0-^[drn^^Mp^(ru), 6
(32)
~ Ay 12
which is often called the virial equation for the pressure. The combination of Eqs. (29-31) leads to the Kirkwood-Buff formula for the surface tension oo
Y=iJ«fe, 1
-<*
9
(X
9
Jr, 2 ^^) '2"^ p ( 2 ) (r l .r 2 ), drn
)
(33)
rn
where p(2)(n>r2) can be expressed in terms of the pair distribution function g(2)(r,,r2) as p (2) (r b r 2 ) = g(2)(r,,r2) po(r,)po(r2). Given the equilibrium density profile po(z), which can be generated using Eq. (27), this relationship can be used to calculate the surface tension of a planar interface in a self-consistent manner [17]. More general expressions for the surface free energy of a fluid interface will be discussed in section 3.
10
Vicente Talanquer
2.2.2. Curved interfaces In comparison to planar fluid interfaces, the statistical mechanics of curved interfaces was not formally developed until recent years [18-20]. This work has been fueled by the realization that the curvature of interfaces determines the physical behavior of a variety of systems such as micellar solutions, fluid membranes, microemulsions, lyotropic liquid crystals, and interfacial films [9]. The free energy of a surface element in a curved interface is not only characterized by its area, but also by its curvature. In particular, for small curvatures the surface tension can be expanded in terms of the mean curvature H=C,+C2 and Gaussian curvature K=C[C2, where C, and C2 are the principal curvatures of the interface [18]: JS(H,K)
= J-KC0H+
V2KH2+KGK.
(34)
In this equation y represents the surface tension of the planar interface, Co is the spontaneous curvature, K is the bending rigidity, and KG is the saddle splay module. This expression for the surface free energy per unit area reduces to that derived by Helfrich under the assumption that the bending properties of the interface are those of a continuum thin elastic plate [21]. For a spherical interface of radius R and curvatures H=-2/R and K=\/R2, Eq. (34) can be rewritten in terms of the inverse radius yc(R) = y + 2K C 0 ^ j
+ (2K+K G ) ( I j
.
(35)
The coefficient of the second term in this equation is usually rewritten as KC 0 = -yS, where 8 is the Tolman length. Microscopic expressions for Co, K, and KG can be derived from statistical mechanics by calculating the change in the free energy (partition function) of the interface under an infinitesimal coordinate transformation that changes both the surface area and the curvature of the system but leaves the volume unchanged. This transformation demands the selection of a particular geometry for the system and a choice for the location of the interface. For a spherical surface with a radius R determined by the location of the Gibbs (equimolar) dividing surface, the calculation for systems with pair-wise intermolecular potentials leads to the Kirkwood-Buff formula Eq. (33) for the surface tension of the planar interface, and to the following expressions for the spontaneous curvature and rigidity constants [18]:
Statistical Mechanics of Fluid Interfaces
9
GO
(X
11
9
2 2)
KC0 = i f «fe, J Jr l 2 ^^) ' 2 ~ ' (z l + z 2 )p';)(r,,r 2 ),
(36)
GO
2K + K G = I Jdfe, /
* 12 ^U)J_[(xf 2 -zf 2 )pW(r 1 ,r 2 ) drn
rl2
-4(4-f-)p2(r,,r 2 )].
(37)
These relationships are derived assuming that the pair density of the spherical interface p(t2)(r,,r2) can be expanded around the pair density of the flat interface p^r,,!*,) to second order in the curvature : Pi
2)
(r,,r2) = p^r,,!-,) + ^jp ( j 2 , ) (r 1> r 2 )+ ( I j P « (r,, r2)
(38)
As can be seen from Eqs. (36) and (37), the deformation of a spherical interface does not lead to separate expressions for K and KG. In order to generate these independent relationships it is necessary to analyze both a spherical and a cylindrical interface [18]. Further approximations for the pair density of the planar interface similar to those described in the previous section can then be used to calculate the spontaneous curvature (or related Tolman length), and the rigidity constants for the fluid interface. 3. DENSITY FUNCTIONAL THEORY A significant proportion of the work developed in the last 25 years regarding the statistical mechanics of fluid interfaces is based on the application of density functional theory (DFT) [10-11,14]. Density functional methods are predicated on the idea that the free energy of an inhomogeneous fluid can be expressed as a functional of the distribution functions, in particular the singleparticle density p(r). This free energy functional has two basic properties: a) it is at an extremum when the distribution functions are those of the equilibrium state, and b) the value of the functional at the extremum corresponds to the equilibrium value of a thermodynamic potential such as the Helmholtz free energy F or the grand potential Q. The density functional approach is particularly useful because one can rely upon methods of calculus of variations to derive all the relevant thermodynamic properties of an inhomogeneous fluid [5]. Derivatives of the free energy
12
Vicente Talanquer
functional determine the equilibrium density profiles and the equilibrium distribution functions that describe the microscopic structure of the system. DFT also allows for easier identification of approximate functionals that can be used to derive the properties of a wide variety of non-uniform fluids. The density functional theory of inhomogeneous systems was pioneered by Hohenberg and Kohn [22] and Mermin [23] for a quantum mechanical electron fluid. The classical version of the theory was originally developed by Lebowitz and Percus [24] among others and, in its present form, by Ebner and Saam [25] and Yang et al [26]. The general formalism has proved extremely useful to study interfacial phenomena in a variety of systems and has become the preferred path to derive both formally exact and approximate results for correlation functions and thermodynamic functions. In this section we review the basic formalism of DFT, the common approximations that have been used in the study of fluid interfaces, and some of the most recent developments and applications of the theory. 3.1. Formalism In an open system of volume V at a given temperature T and chemical potential \x, the equilibrium distribution of particles po(r) will be determined by the nature of the external potential Vexl(r) (see Eqs. (4) and (9)). In particular one can expect po(i") to be uniquely determined by the local potential «(r)=n - VUr).
(39)
Under these conditions, the grand potential Q should become a functional of w(r), Q [«]. If we consider an infinitesimal change SVexl(r) of the external potential, a straightforward statistical mechanics calculation yields the following result for the change in the grand potential [6, 10-11]: 5Q = n [w(r)+8 «(r)] - Q [w(r)] = - J dr p(r) w(r).
(40)
Thus, the single-particle density po(r) is given by the functional derivative of Q with respect to w(r), po(r)=-8n[«]/8«(r).
(41)
One can prove that not only is po(r) a functional of w(r), but also Vex,(r) is uniquely determined by po(r) (only one external potential can determine a specified equilibrium density) for a given interaction potential energy UN, at a fixed p. and T [10]. In general, the local density is a more convenient variable
Statistical Mechanics of Fluid Interfaces
13
than u(r). This is particularly true in the study of phase transitions, in which inhomogeneities can arise in the absence of external potentials. Thus, the density functional approach focuses on functionals of p(r) rather than w(r). A convenient functional of p(r) can be constructed from a generalized Legendre transformation of Q [u]: Q[u]=F[p]- \drp(r)u{r),
(42)
where F[p] is a unique functional of the local density. F[p] is called the Helmholtz free energy functional, since the total Helmholtz free energy is given by F=F[po]+ \drp0Vext(r)-
(43)
F[p] is an intrinsic property of the system of particles and its structure can be derived without any reference to an external perturbation. For a fixed value of the local potential w(r), the equilibrium local density po(r) minimizes the functional in Eq. (42) over all density functions that can be associated with the potential Vext(r) [10-11]. We can express this variational principle as follows: ^ { F [ p ] - /rfrp(r)«(r)} ||i=|io(r)= 0 ,
(44)
which leads to the fundamental equation for the equilibrium density functional theory of inhomogeneous fluids: &F[p],
n= vur) + JTT U w = vUr) + LUPO ; r],
(45)
where
tUp;r]=™
(46)
8p(r) is called the intrinsic chemical potential. Eq. (45) expresses the constancy of the chemical potential u. throughout the inhomogeneous system and, given some means of determining F[p], it is an explicit equation for the equilibrium density
14
Vicente Talanquer
po(r) (density profile) for any external potential. For example, for an ideal gas of non-interacting atoms the free energy is given by [2,10] Fdp]= P "' \dr p(r){ln[AJp(r)]-l},
(47)
and Eq. (45) yields the familiar Boltzmann distribution A3po(r)=exp{p[n -Vexl{r)}} = exp[p w(r)].
(48)
For many-body interacting systems, important non-local contributions to the free energy need to be taken into account and only approximations to F[p] tend to be known. The effects of interactions between particles in the system can be incorporated by introducing a set of direct correlation functions [10-11]. For this purpose, let us divide the functional F[p] into an ideal and an excess part in the form ^[p]=^[p]+^[p],
(49)
where the ideal part is given by Eq. (47). If Eq. (45) and Eq. (49) are now combined, the equilibrium density can be expressed as AJpo(r) = exp[(3 M(r) + c(1)(r)],
(50)
where !>f [p]
is called the single-particle direct correlation function and is itself a functional of the density. The quantity -(3 'c(1)(r) is a measure of the contributions to the intrinsic chemical potential that arise from interactions between particles. c(1)(r) is the first term of a hierarchy of correlation functions that can be generated from the excess free energy functional Fex[p]. The usual two-body direct (Ornstein-Zernike) correlation function can be viewed as the second functional derivative of Fex [10]: §2F [p]
^-"w^o^-
(52)
Statistical Mechanics of Fluid Interfaces
15
A useful expression for this correlation function can then be derived by differentiating c(1)(ri) in Eq. (50) with respect to po(r2) 2 6 r r c< >(ri>r2)= ( i ~ 2 ) r
Po( i)
PMr,) = 6(r,-r 2 ) P 0 ( r 2) Po(ri)
8
|
(36^(r.) Spo(r2)
(53)
Further differentiation of the free energy functional yields three-body, four-body, and so on, direct correlation functions. This hierarchy of correlation functions can be used to derive integrodifferential equations for the equilibrium density. For example, we can transform Eq. (50) by taking the gradient on each side of this relationship [10]. Using Eq. (52) we then obtain Vln Po (r,)=-(3VF M ,(r 1 ) + \dr2 c(2)(r1;r2)Vpo(r2),
(54)
known as the Lovett-Mou-Buff-Wertheim equation, which in principle can be solved for po(r) given some means of evaluating or approximating the OrnsteinZernike correlation function c<2)(ri,r2). For a fluid in which the particles interact via pairwise potentials, this relationship is equivalent to the first member of the Yvon-Born-Green hierarchy (Eq. (24)). A similar hierarchy of correlation functions can be obtained by functional differentiation of the grand potential Q [w] with respect to w(r). The first term of the hierarchy is simply the average single-particle density as shown in Eq. (41). The second derivative yields the density-density correlation function G(r!,r2), which is proportional to the static linear density response function [10]:
^r^-p-'-^ff-L
(55)
8w(r,)5w(r2) G ( r 1 , r 2 ) = [ 3 - ' ^ l ^ = <[p(r,)-p 0 (r 1 )][ P (r2) - po(r2)]>. 6w(r2)
(56)
This density fluctuation function is a measure of the correlation of fluctuations in numbers of particles at ri and r2 and is related to the usual pair-wise distribution function p(2)(ri,r2) via G(r,,r 2 )= p(2)(r,,r2) - po(r,)po(r2) + po(r,) 8 (r, - r2).
(57)
For a bulk fluid of uniform density po, this last relationship can be expressed as G(ri2)=po2(g-(r12)-l) + po8 (r,2), where g(r,2) is the radial distribution function.
16
Vicente Talanquer
Thus, the Fourier transform of G(r]2) is directly related to the static structure factor of the bulk liquid by G(k)=poS(k) [5, 8], 3.2. Approximate functionals Practical applications of density functional theory to realistic physical systems require some approximation for the free energy functional F[p]. Once this is established, Eqs. (44) and (45) provide the recipe for a variational calculation of the equilibrium density and the total free energy at a given T, u,, and Vex,(r). Direct correlation functions and distributions functions can then be obtained by further functional differentiation. As in any variational calculation, the success achieved depends on the skill with which the approximate F[p] is constructed. However, it is important to realize that there is no reason to expect that the calculated properties correspond to the exact solution of any particular Hamiltonian. Useful approximations for the free energy functional F[p] can be obtained only in certain limiting cases. In this section, we describe the most common approximations developed for the study of fluid interfaces. 3.2.1. Square gradient approximations For systems in which the density p(r) exhibits long wavelength variations such that Vplp « l/<;, where £, is the typical range of density correlations in the bulk fluid, the functional F[p] can be expressed as a function of p and its spatial derivatives [10, 27] F[p] = \drAtfr),
V p, V V p ...],
(58)
where J[p] is the free energy density. Assuming that the gradient is small, J[p] can be expanded as a series of density gradients F[p] = \dr [/o(p(r)) + / 2 (p(r)) ( Vp(r))2 + O(Vp) 4 ].
(59)
The coefficients /(p(r)) in this series are functions of scalar p(r) and thus are invariant under rotations in fluid systems. Symmetry arguments can then been applied to eliminate certain terms in the expansion [27]. /,(p( r )) m Eq. (59) is the Helmholtz free energy of a uniform fluid of density p, so that truncating the expansion after the first term leads to the local density approximation (LDA). In this limit, the free energy of the non-uniform phase is approximated by evaluating the free energy of the uniform phase at a density equal to the local density of the non-uniform phase. The first correction
Statistical Mechanics of Fluid Interfaces
17
to this approximation is quadratic in the gradient and the resulting functional is the base of the "square gradient" approximation. The structure of the coefficient / 2 (p(r)) can be determined only by imposing additional requirements on F[p]. For example, one can expect Eq. (59) to be consistent with linear response theory [10], which results in the expression f2(p) = (12(3)-' \drurl22c(2\p;r]2),
(60)
where c(2)(p;r12) is the two-body direct correlation function of the uniform fluid. Thus, given some means of calculating this correlation function, fo(p) and/j(p) can be derived, and the functional in Eq. (59) can be used to fully derive the properties of an inhomogeneous fluid. In particular, from the variational principle in Eq. (44) it follows that the equilibrium density po(r) in the square gradient approximation should satisfy the differential equation H= VUr) +/oXp o (r))-/2'(Po(r))| Vp o (r)| 2 -2/ 2 (p 0 (r)) V2p0(r),
(61)
where the prime denotes differentiation with respect to p0. Although technically the square-gradient approximation should be valid only in the case of very slowly varying density profiles, such as for a two-phase interface near the critical point, the theory has often been applied to cases in which the profile varies rapidly [2,10]. Its generalization to fluid mixtures is straightforward and has been useful in analyzing interfacial tensions, adsorption, and wetting phenomena at fluid-fluid interfaces [12,13]. Simple perturbation theory using a mixture of hard spheres as a reference system can be used to calculate the direct correlation functions Cy2) for species i,j in the bulk mixtures, and thus calculate the expansion coefficients/i(p,,p/) [28]. To illustrate the application of the square-gradient approximation to determine the interfacial profile and interfacial tension of a fluid-fluid interface, let us consider the simple case of a symmetric, incompressible mixture of two components A and B near the critical point and in the absence of any external potential [6, 27]. Thanks to the incompressibility condition, p /) (r)+p B (r)=p, where p is the packing density, this system is characterized by only one independent local density. For a system of two coexisting phases with a common planar interface parallel to the x-y plane, the density profile PA(Z) varies only in the direction perpendicular to the interface. The grand potential functional in the square gradient approximation can then be expressed as WPA] = F[pA] - u* \dx pA=A \dz [fo(pA) - u*p, +/2(pA) (^f dz
],
(62)
18
Vicente Talanquer
where A is the interfacial area and u* is an effective chemical potential for the incompressible mixture. In the vicinity of the critical point, the grand potential energy density of the bulk fluid, G>(PA), may be approximated as a quartic form of the particle density o>(P/()=/o(P^) - V*PA = © b + -
(PA -
p/)2
(PA -
p/)2
(63)
where pAa and pj1 correspond to the bulk densities of the coexisting phases, cob is their associated grand potential energy density, and C is a constant. This approximate expression captures the essential features of the free energy close to the critical point and will simplify our calculations. The equilibrium density profile can be obtained by minimizing the grand potential functional in Eq. (62) with respect to variations of PA(Z), under the following boundary conditions: PA(Z) ~$ PA" when z -> - oo, and p^(z) -^ p/ when z -> + co . The minimization yields the following differential equation (see Eq. (61))
2/2(p,)^ dz
+
^ 2 = dz az
^
. apA
(64)
By multiplying both sides by dpA/dz, this equation can be integrated to yield d fiiPA) -jdz
2
= co(P/0 - «)6= Aco(p/,),
(65)
which can be used to determine the interfacial density profile. In particular, if we assume that f2 is a constant independent of the density, the resulting profile is given by
Px(*)=
P
] ,^ + 1 , P\ / n = ^ + P [ : ) + ^-p:)tanh(z/2O,(66) l + exp(—z/0 l + exp(z/0 2V ' 2y '
where ^=(2/2/C)1/2 Ip^-p^^l"1 defines the interfacial width and diverges at the critical point. This quantity is identical to the correlation length of the density fluctuations as obtained from the Ornstein-Zernike form of the structure factor. Although the square gradient approximation is only valid for density variations
Statistical Mechanics of Fluid Interfaces
19
that take place over scales which are large when compared to molecular sizes, the hyperbolic tangent profile described by Eq. (66) tends to be a reasonable approximation even relatively far from the critical point. The interfacial tension of the system y can be evaluated based on the thermodynamic relationship for the grand potential Q = -PV + yA. The pressure of the bulk phases can be expressed as P = -Q b / V= -co b, and the interfacial tension is then given by
A
A
Inserting the result of the density functional minimization, Eq. (65), in the functional for the grand potential, Eq. (62), one obtains a simple expression for the interfacial tension of our system
y = 2 Jdzf(pA)\^\2. J
2
(68)
dz
— no
This particular relationship is consistent with the exact Triezenberg-Zwangzig formula for the surface tension for a planar surface [2]
y = UBT Idz ^ 4
J
1 dz,
\dr2*¥± [xl +y>) c^\rur2),
(69)
dz,
which reduces to Eq. (68) when c (2) (ri,r 2 )= c(2)(r12) is an isotropic function of r12. The expression for the surface tension in Eq. (68) can be further simplified by using pA rather than z as the integration variable
y = 2 fdpAJf2(PA)Au(PA)=j2^c(piiA-P:f.
(70)
The second equality in this equation results from using the approximation for the free energy in Eq. (63), assuming that^ is a constant independent of the density. The minimization of the grand potential functional for this system in spherical geometry can be used as the basis to analyze the effect of curvature on the surface free energy of this system. Comparison of an expansion of the minimized functional to second order in the curvature with the interfacial energy in Eq. (35) yields expressions for the spontaneous curvature Co, the Tolman
20
Vicente Talanquer
length 8, and the rigidity constants K and KG. In particular, it can be shown that the Tolman length in this approximation is given by [29]:
S=--fW2(P>|^H2. 'yJ
(71)
dz
These results represent what is probably the simplest microscopic treatment of a fluid-fluid interface, which is similar to the approach introduced by van der Waals at the end of the nineteenth century [1]. As we have seen, the implementation of this type of approach to the study of interfacial problems implicitly requires some approximation of the structure of fo(p) and/ 2 (p) in the two-phase coexistence region. It is at this stage that mean-field expressions for the free energy are normally introduced to simplify the treatment and assure some sort of analytic continuation of relevant thermodynamic functions for the multiphase system. Similar assumptions are inherent in most approximate theories of liquid surfaces. 3.2.2. van der Waals perturbation theories For systems in which the potential energy of interaction f/(ri, ...rN) is pairwise additive,
t/(r,,...rN)= ijjrfrrfr2<|)(r sr2)p(r )(p(r 2 )-6(r - r ^ ) ,
(72)
the grand potential is a functional of the interaction potential §(t\,xi)=§(\r\-x-^. In this case one can demonstrate that for fixed Tand w(r) [10], 8fi
8
=
&F
o
1 (2),
.
= —P ( r u r ?)
2
, ,
(73)
where p(2)(ri,r2) is the usual pair-wise distribution function. If we now consider a reference fluid at the same temperature and density p(r), for which the pairwise potential is ((>,.(i"i,r2), and integrate Eq. (73) along a one-parameter integration path such that
Statistical Mechanics of Fluid Interfaces
*Ip]=*V[p] + -f^dafdrjdr^i^r^^r,).
21
(75)
Fr[p] in this expression corresponds to the intrinsic Helmholtz free energy of the reference fluid in which the pairwise potential is ((>,.( i"i,i"2) and the density is p(r). p<2)(())a;ri,r2) is the distribution function for a system with the same local density but with an interaction potential §a. As a increases from 0 to 1, one needs to assume the presence of a variable external potential that imposes such a density as ())„varies. The parametric form of the free energy functional in Eq. (75) is normally used as the basis to develop perturbation theories of inhomogeneous fluids. By choosing suitable expressions for Fr[p] and p(2)(())a;r1,r2) of the inhomogeneous fluid, approximate free energy functionals can be derived from Eq. (75). For example, if we assume that for the system of interest p(2)(()>a;ri,r2) can be approximated by the pairwise distribution function of the reference fluid p(2)(ri,r2), and the reference free energy can be expressed in the local density approximation (LDA), Eq. (75) reduces to
F[p] = pr/ r (p(r)) + l~j dr,j dr29{2\rvr2)$p{rx,r2).
(76)
Based on perturbation theory of uniform fluids, it is common to set p(2)(r,,r2) « p(r{]p(r2)g,(pm;rn),
(77)
where gr(pm',r\2) is the radial distribution function of the uniform reference fluid evaluated at some mean value pm of the local densities at i-| and r2. Since the free energy and radial distribution of the uniform hard-sphere system are readily available, popular perturbation schemes, such as those based in the BakerHenderson (BH) [30] and the Weeks-Chandler-Andersen (WCA) [31] prescriptions, use hard spheres as the reference fluid while the attractive part of the pairwise potential §att is treated as the perturbation potential. An even simpler but common approximation is to take p (2)(n>i"2)«p(ri)p(r2), which completely ignores correlations. Using hardspheres as the reference fluid, the resulting functional
F[p] = \drfhs(p(r)) + ^Jdr, Jjr 2 p(r,)p(r 2 H a/ ,(r l2 )
(78)
is the basis of the van der Waals theory of non-uniform fluids [2,10]. In this case, the equilibrium density profile is normally calculated using iteration techniques
22
Vicente Talanquer
to solve the Euler-Lagrange equation that results from the application of the variational principle from Eq. (44), H= U r i ) + Hfa(po(r0)+ \dr2p{r2)§att{rX2).
(79)
Prescriptions iorfhs such as that of Carnahan-Starling [32] are commonly used to derive analytic expressions for the hard-sphere chemical potential \ihs=dfhs/dp, and have been extended [33] to investigate the interfacial properties of multicomponent systems. Lennard-Jones and Yukawa potentials have normally served as the basis for modeling the effect of the perturbation potential in these types of systems [5]. The direct correlation function c(2)(r!,r2) for this theory can be calculated from the second functional derivative of Eq. (78). The result of the functional differentiation yields
c(2)(r,,r2) = - p 4 ^
5
(ri"r2) ~ P W'12),
(80)
dp where the structure of the first term on the right-hand side of this equation is a direct consequence of the LDA. As we can see, in this approximation the hardsphere direct correlation function is replaced by a delta function at the origin and the correlation length for hard-sphere repulsion is zero. Although the simple van der Waals perturbation theory for inhomogeneous systems has provided insight into a wide variety of phenomena, its mean-field nature restricts its application to problems in which long wavelength fluctuations do not play a central role [10-11]. Moreover, LDA is useful for the study of long-range correlations and of problems where density oscillations do not occur. The theory does not include short-range correlations and generates incorrect descriptions of the density at sharp boundaries, such as for fluids in contact with hard-wall or fluid-fluid interfaces in highly immiscible mixtures. This latter shortcoming has been addressed with the introduction of reference functional involving smoothed or "weighted" densities. 3.2.3. Weighted-density approximations Most schemes for improving the van der Waals perturbation approach accept the basic separation of the free energy functional into a hard-sphere reference part Fhs[p], and an attractive force part Fa,,[p], which is normally treated in a mean-field fashion: F[p]= Ffa[p]+Fa,,[p]. Many of these approaches then focus on generating an approximated form for Ffe[p] that goes beyond a local density approximation.
Statistical Mechanics of Fluid Interfaces
23
In the weighted-density approximations (WDA) it is assumed that the hard-sphere functional can be expressed as the sum of ideal and excess parts, Fhs[p] = Fid[p] + Fex[p]. The excess free energy is then taken to be a function of a smoothed density p(r) constructed as an average of the true local density p(r) over a local volume that is determined by the range of interatomic forces. In particular, the excess free energy Fex[p] is approximated as a local function of p(r), a coarse-grained variable that contains smoothed out information on the local variations of the density Fex[p]=\drp(r)VMrn
(81)
The function ^^(p^)) in this equation represents the excess free energy density, and is commonly calculated based on the excess free energy of a homogeneous hard-sphere fluid at a density p(r). Thus, most versions of WDA differ only in the form of the smoothing or weight functions used to calculate the coarsegrained density p(r). One of the simplest recipes for the evaluation of p(r) is based on the use of a weight function w0 that is independent of the local density and smoothes the profile uniformly over a local region of the size of the excluded volume of a single particle 4na3/3 [34]. The weighted density is then calculated as p(r)=jdr'w o (|r-r'|)p(r'),
(82)
where w0 is taken to be proportional to the Heaviside step function 0 , wo=T^TG(a-r). 4-Kcr
(83)
The structure of this function assures that p(r) reduces to the homogeneous value p in the uniform fluid. This particular form of the weight function leads to a hard-sphere direct correlation function that has the correct qualitative features, including a discontinuity at r=a. The application of this type of WDA, with the mean-field approximation for FaU, has provided realistic descriptions for a variety of phenomena, from drying transitions at solid interfaces [35], to the presence of density oscillations at fluid-fluid interfaces in highly immiscible mixtures [36], the adsorption of surfactant molecules at interfaces [37], and the characteristic structure of micelles, membranes, and vesicles in amphiphilic systems [38].
24
Vicente Talanquer
More sophisticated versions of the WDA introduce normalized densitydependent weight functions w(|r-r'|;p(r)), chosen in such a way that the second functional derivative of the free energy leads to an accurate direct correlation function c (2 \|r-r'|;p) for the bulk hard-sphere fluid over a wide range of densities (given by the accurate Percus-Yevick approximation) [39]. In this approach the weighted density is given implicitly by p(r)= Jdr'w(|r-r'|;p(r))p(r).
(84)
The solution of this equation is simplified by using the additional assumption that w(|r-r' |;p(r)), and hence c (|r-r"|;p), can be expanded in a virial series w(|r-r'|;p(r))= wo(|r-r'|) + w,(|r-r'|)p + w 2 (|r-r'|)p 2 + ...,
(85)
which reduces to the simpler theory described before when the expansion is truncated at zero order. The use of a truncated power-series expansion of the weight function has proved to be a useful and versatile approximation scheme to the study of a variety of inhomogeneous systems. It has been widely used to investigate wetting transitions in simple liquids in contact with hard walls [13], and the structure and phase equilibria of confined fluids [40]. The theory has also been successfully extended to study the interfacial properties of multi-component systems [41]. 3.2.4. Geometry-based approaches In 1989, Rosenfeld introduced novel ideas for deriving a density functional theory for hard-sphere mixtures [42]. His approach is based on the geometrical properties of the particles and is called fundamental measure theory (FMT). Present-day formulations of Rosenfeld's ideas [43,44] have given a major boost to recent density functional studies of a wide variety of inhomogeneous fluids. FMT rests on the assumption that the excess Helmholtz free energy FM[p] can be described as a functional of weighted densities obtained by averaging the local density with geometric measures of the particles. The theory is based on a geometrical representation of the particle shapes using different characteristic measures: volume, surface, integral mean curvature, Euler characteristic of the particles. These functions are used to obtain a set of weighted densities n,lr) = \df
Wa (r-r')p(r'),
(86)
Statistical Mechanics of Fluid Interfaces
25
with weight functions w^r-r') that have the range of the molecular radius, R, and are constructed to recover the Mayer / function Xr)=exP[-P
(87) (88) (89)
where 0(r) is the Heaviside function and 8(r) is the Dirac delta function. An alternative deconvolution of the Mayer / function, suggested by Kierlik and Rosinberg, avoids vector-like weight functions but introduces instead weights containing first and second derivatives of the Dirac delta function [45]. Although both formulations lead to an equivalent functional for hard-sphere systems, Rosenfeld's selection is more general in scope. The excess free energy functional in FMT is given by Fex[p] = kBT \dr O M r ) ] ,
(90)
where the excess free energy density for the hard-sphere fluid is usually written in the form <$>hs= Oz+cP^+Oj with
^2-nvl.nv2j
(91) (92)
l-«3
r? — 3« 2 n ,»n , O3 = ^ ^ - ^ . 24vr(l-« 3 ) 2
(93)
In these equations the weighted functions nvl and nv2 are vectors and nV| • nv2 is a dot product. In the limit of a homogeneous fluid, the two vector weighted densities vanish, and the Helmholtz free energy O becomes identical to that derived from the Percus-Yevick theory. The scalar-weighted densities coincide
26
Vicente Talanquer
with the variables of the scaled-particle theory for a uniform hard-sphere fluid: no=po, n\=Rp0, n2=4nR2p, and «3= 4izR3p/ 3 (packing fraction). The form of the functional
njj) = Z \dr w'a(r-r")p,(r),
(94)
with weight functions w'J^r-r^) given by Eqs. (91-93) in the case of spherical particles with a radius /?,-. This type of approach has been used recently to study effective depletion interactions for: non-additive mixtures, polydisperse particles, mixtures of non-spherical particles, and systems with soft interactions [46]. 3.3. Relevant developments and applications Density functional theory (DFT) of statistical mechanic has been applied to the study of a wide variety of phenomena involving fluid interfaces. The theory presented in the previous sections has been generalized from traditional systems of hard spheres to fluids with soft interactions, from mono-atomic to molecular systems, from pure substances to mixtures, and from systems of particles interacting via simple potentials to those exhibiting complex anisotropic interactions. Not only has the complexity of the systems explored increased, but also the range of phenomena being studied has multiplied: adsorbed fluids, thin films, fluids in porous media, fluid membranes. Excellent reviews of much of this work have been presented elsewhere [11-14]. This section will highlight some relevant developments in the treatment of fluids with anisotropic interactions and in the study of wetting phenomena.
Statistical Mechanics of Fluid Interfaces
27
3.3.1. Fluids with anisotropic interactions In general, the molecules forming a liquid have non-spherical shapes and anisotropic pair potentials. Therefore, in contrast to simple fluids, they possess not only positional but also orientational degrees of freedom. The analysis of their interfacial properties thus implies a higher level of complexity than that required for systems with spherically symmetric interactions. Progress has been made in several directions. The study of the interfacial properties of molecular fluids composed of small dipolar particles has been simplified by the use of idealized models for polar molecules in which a point dipole-dipole interaction is superimposed on a spherically symmetric potential. An example is that of the Stockmayer fluid in which a Lennard-Jones potential is used to model the long-range van der Waals interactions. DFT perturbation methods have been successfully applied to investigate the properties of these types of systems [47,48]. For the anisotropic fluid the free energy can now be expressed as a functional of the density of particles at point r, which have an orientation ©=(9,(|)) in a spatially fixed reference system, p(r,co). Based on Eq. (75), the Helmholtz free energy functional of the dipolar fluid can be written as
*Ip] = ^[Pl + - f da. Tc?rlc?r2c?tolc?co2g(2)(r1,r2,co],co2,a)p(r],to1)p(r2,co2)(|) (r,,r 2 ),
(95)
where g(2)(i"i,r2,c0i,eo2,a) is the pair distribution function of an inhomogeneous fluid with a pair-wise intermolecular potential parameterized by a: ^«(i"i,r2,(0i,C02,a.)= <j>re/r1,r2,a),,co2)+ o u ^ r , , ^ , © , , ^ ) .
(96)
The basic interfacial properties of the liquid-vapor interface in planar and spherical geometries have been calculated using this functional in the so-called modified mean-field approximation where g<2)(ri,r2,G)i,C02,a) is given by the lowdensity limit g (2) (ri,r 2 ,co,,co 2 ,a)= exp(-p^).
(97)
The Helmholtz free energy functional can then be approximated by: F[p] =F,.[p] + -^- lcft-,(Ji-7(io>1drco,p(r1,co,)p(r7,a)2)e 2 J
reJ
\~e
p
.
(98)
28
Vicente Talanquer
Results from this theory for the density and orientational profiles indicate that molecules in Stockmayer dipolar fluids are aligned parallel to the interface on the liquid side and perpendicular to it in the vapor side at a liquid-vapor interface [47, 48]. Closely related DFT approaches have been used to explore the interfacial properties of more complex systems such as surfactant solutions [49] and diverse amphiphilic structures [50-51]. The main limitation of these types of models for molecular fluids is that they ignore shape effects and thus the angle dependence of the short-range interactions. A relatively easy way to take such effects into account is with models in which the molecule is represented as a set of discrete interaction sites that are commonly located at the sites of the atomic nuclei. The total potential energy of the system is then obtained as the sum of spherically symmetric interaction-site potentials. DFT has been extended using this interaction site formalism to the cases of systems composed of polyatomic species [52-53], such as polymeric fluids. In this generalized theory, the free energy may be expressed as a functional of the density of site type a at point r, p«(r) [54],
p/Tpa(r)] = Zj*pa(r)ln/-Q(r) a
- {... { [ F K / i i v ) ] s{n\{ra})+ (3Fc,[Pa(r)].
(99)
The first term in this equation represents the free energy of an ideal mixture of isolated atoms (sites). The second term accounts for the decrease in entropy of the ideal gas due to intramolecular couplings. This contribution depends on the structure of the intramolecular correlation function s^"\{ra}) for the polyatomic molecule. Fex[pa(r)] is the excess free energy, which in a perturbation approach can be written as
3F«[p]= Jrfrp(r)T«(p(r)) + § £ / * , / ^ P ^ K O ^ f e ) ,
(100)
a,(3
where the first term normally represents the excess free energy of a hard-sphere reference system, and the second term is the contribution from the long-range attractive interactions between atomic sites in a mean-field approximation. The density profile for the polyatomic molecule p(r) is given by P(r) = X Pa(r). a
(101)
Statistical Mechanics of Fluid Interfaces
29
The minimization of the free energy functional in Eq. (99) over the atomic distribution functions pjr) and the auxiliary functions fjj) allows us to obtain the equilibrium site densities
p«(r) =fa(r) J... jirUg/Xrp)] *{#l)({rp}),
(102)
with / f l (r)=exp[P(M o --^)]. 8 P,,
(103)
These equations can be solved in a self-consistent manner to derive the density profile for the polyatomic system at a given temperature 7 and chemical potential The interaction site formalism has been successfully applied to the study of interfacial properties of a variety of systems. Polymers are nonrigid structures that can be modeled as a sequence of freely jointed hard-spheres. Such models have been used to explore the behavior of polymeric fluids near hard walls and in confined geometries [55-59]. Models of fused spheres have also been used to investigate the properties of amphiphilic systems [37-38,60-61]. In combination with theoretical approaches based on Wertheim's theory of associating fluids [62-63], DFT is now robust enough to deal with the effects of molecular association in complex solutions. Density functionals for anisotropic hard bodies have been proposed to model liquid crystals and colloid suspensions of rod-like macromolecules [6465]. Recently, FMT has been extended to the study of the bulk and interfacial properties of colloidal mixtures of hard rods and hard spheres [66-67]. 3.3.2. Wetting at fluid interfaces An important application of DFT has been to the study of the structure and tension of the interfaces in the equilibrium of three phases [12-13, 68];for example, in the coexistence of two liquid phases and a vapor at a triple point in a binary mixture. Depending on the value of the interfacial tensions between the different phases in the system, one of the liquid phases (a) could form a drop at the interface between liquid (3 and the vapor (v). Besides the droplet there will be generally a film of liquid a of microscopic thickness at the interface. For such partial wetting to occur, the interfacial tensions should satisfy the condition
30
Vicente Talanquer
S = Jpv-{j a/9 + Y«v)<0,
(104) where S is the spreading coefficient. For different values of temperature, pressure or chemical composition, liquid a could form a macroscopic film and spread over liquid (3, separating it from the vapor. Total wetting occurs when Antonow's rule is satisfied S = y/iv-(jafi + yav) = o.
(105)
Among the various theoretical techniques that can be used to describe inhomogeneous fluids, only a few are capable of exhibiting the growth of a macroscopic wetting film. In particular, various approaches based on the density functional theory have been very successful in describing the evolution from states of partial to total wetting in systems with a fluid-solid or a fluid-fluid interface [12, 13]. DFT has revealed, for example, that at the transition from partial to complete wetting, the wetting film becomes of macroscopic thickness in a singular manner. The type of singularity is connected with the order of the wetting transition. If the thickness grows discontinuously, the transition is of first order and the spreading coefficient displays a discontinuous first derivative. If the film thickness diverges in a continuous manner, the wetting transition is said to be critical and S exhibits a weak singularity. The application of density functional theory to the study of wetting phenomena in binary mixtures has shown that the order of the wetting transition is strongly influenced by the strength, range, and form of the intermolecular interactions. In particular, it is expected that in fluids with long-range van der Waals forces the transition will be of first order. The nature of the transition is also affected by the strength of the interactions. Ordinarily, a wetting transition will take place if the triple-line is approached from the less volatile liquid and the wetting films are rich in more volatile component. Although most of the basic research on density functional theory as applied to wetting phenomena was developed almost twenty years ago, this is still a vibrant area of research. For example, new approaches based on FMT have been recently developed to study adsorption and wetting by fluids confined in porous media [69], and novel phenomena such as entropic wetting have been identified in phase separating binary fluid mixtures near a hard wall [70].
Statistical Mechanics of Fluid Interfaces
31
REFERENCES [I] B. Widom, Physica A, 263 (1999) 500. [2] J. S. Rowlinson and B. Widom, Molecular Theory of Capillarity, Oxford University Press, New York, NY, 1989. [3] D. Henderson (ed.), Fundamentals of Inhomogeneous Fluids, Marcel Dekker, New York, NY, 1992. [4] C. Caccamo, J. -P. Hansen, and G. Stell (eds.), New Approaches to Problems in Liquid State Theory, Kluwer, Dordrecht, 1999. [5] J. - P . Hansen and I. R. McDonald, Theory of Simple Liquids, Academic Press, San Diego, CA, 1990. [6] J. B. Barrat and J. -P. Hansen, Basic Concepts for Simple and Complex Liquids, Cambridge University Press, Cambridge, 2003. [7] D. Chandler, Introduction to Modern Statistical Mechanics, Oxford University Press, New York, NY, 1987. [8] D. A. McQuarrie, Statistical Mechanics, Univesrity Science Books, Sausalito, CA, 2000. [9] S. A. Safran, Statistical Thermodynamics of Surfaces, Interfaces, and Membranes, Westview Press, Boulder, CO, 2003. [10] R. Evans, Adv. Phys., 28 (1979) 143. [II] R. Evans, in Fundamentals of Inhomogeneous Fluids (D. Henderson, ed.), Marcel Dekker, New York, NY, 1992. [12] D. E. Sullivan and M. M. Telo da Gama, in Fluid Interfacial Phenomena (C. A. Croxton, ed.), Wiley, New York, NY, 1986. [13] S. Dietrich, in Phase Transitions and Critical Phenomena (C. Domb and J. L. Lebowitz, eds.), Academic Press, New York, NY, Vol. 12, 1988. [14] D. W. Oxtoby, Annu. Rev. Mater. Res., 32 (2002) 39. [15] S. Texvaerd, J. Chem. Phys. 64 (1976) 2863. [16] G. A. Chapela, G. S. Saville, S. M. Thompson, and J. S. Rowlinson, J. Chem. Soc. Faraday Trans. II, 73 (1977) 1133. [17] J. Fisher and M. Mathfessell, Phys. Rev. A, 22 (1980) 2836. [18] E. M. Blokhuis and D. Bedeaux, Physica A, 184 (1992) 42. [19] E. M. Blokhuis and D. Bedeaux, Heterogeneous Chem. Rev., 1 (1994) 55. [20] A. Robledo and C. Varea, Physica A, 220 (1995) 33. [21] W. Helfrich, Z. Naturforsch, 28c (1973) 693. [22] P. Hohenberg and W. Kohn, Phys. Rev., 136 (1964) B864. [23] D. Mermin, Phys. Rev., 137 (1964) A1441. [24] J. L. Lebowitz and J. K. Percus, J. Math. Phys., 4 (1963) 116. [25] W. F. Saam and C. Ebner, Phys. Rev. A, 15 (1977) 2566. [26] A. J. M. Yang, P. D. Fleming, and J. H. Gibbs, J. Chem. Phys., 64 (1976) 3732. [27] J. W. Cahn and J. E. Hilliard, J. Chem. Phys., 2 1958) 258. [28] M. M. Telo da Gama and R. Evans, Mol. Phys. 48 (1983) 229;48 (1983) 251. [29] S. M. Oversteegen and E. M. Blokhuis, J. Chem. Phys., 112 (2000) 2980. [30] J. A. Barker and D. Henderson, Rev. Mod. Phys. 48 (1976) 587. [31] J. D. Weeks, D. Chandler, and H. C. Andersen, J. Chem. Phys., 54 (1971) 5237. [32] N. F. Carnahan and K. E. Starling, J. Chem. Phys., 51 (1969) 635. [33] G. A. Mansoori, N. F. Carnahan, K. E. Starling, and T. W. Leland, J. Chem. Phys., 54 (1971) 1523. [34] M. Johnson and S. Nordholm, J. Chem. Phys. 75 (1981) 1953. [35] P. Tarazona and R. Evans, Mol. Phys., 52 (1984) 847.
32
Vicente Talanquer
[36] I. Napari, A. Laaksonen, V. Talanquer, and D. W. Oxtoby, J. Chem. Phys. 110 (1999) 5906. [37] T. A. Cherepanova and V. Stekolnikov, Mol. Phys., 82 (1994) 125. [38] V. Talanquer and D. W. Oxtoby, J. Chem. Phys., 113 (2000) 7013. [39] P. Tarazona, Phys. Rev. A, 31 (1985) 2672. Erratum, Phys. Rev A, 32 (1985) 3148. [40] P. C. Ball and R. Evans, J. Chem. Phys., 89 (1988) 4412;Mol. Phys., 63 (1988) 159. [41] Z. Tan, U. M. B. Marconi, F. van Swol, and K. E. Gubbins, J. Chem. Phys., 90( 1989) 3704. [42] Y. Rosenfeld, Phys. Rev. Lett. 63 (1989) 980. [43] Y. Rosenfeld, M. Schmidt, H. Lowen, and P. Tarazona, J. Phys.: Condens. Matter, 8 (1996) L577. [44] P. Tarazona, and Y. Rosenfeld, Phys. Rev. E, 55 (1997) R4873. [45] E. Kierlik and M. L. Rosinberg, Phys. Rev. A, 42 (1990) 3382. [46] H. Lowen, J. Phys.: Condens. Matter, 14 (2002) 11897. [47] P. Frodl and S. Dietrich, Phys. Rev. A, 45 (1992) 7330. [48] B. Groh and S. Dietrich, in New Approaches to Problems in Liquid State Theory (C. Caccamo, J. - P . Hansen, and G. Stell, eds.), Kluwer, Dordrecht, 1999. [49] M. M. Telo da Gama and K. E. Gubbins, Mol. Phys., 59 (1986) 227. [50] A. M. Somoza, E. Chacon, L. Mederos, and P. Tarazona, J. Phys.: Condens. Matter, 7 (1995)5753. [51] C. Guerra, A. M. Somoza, and M. M. telo da Gama, J. Chem. Phys., 109 (1998) 1152. [52] D. Chandler, J. D. McCoy, and S. J. Singer, J. Chem. Phys, 85 (1986) 5971;85 (1986) 5977. [53] J. D. McCoy, S. J. Singer, and D. Chandler, J. Chem. Phys., 87 (1987) 4853. [54] T. A. Cherepanova and V. A. Stekolnikov, Chem. Phy., 154 (1991) 41. [55] C. W. Woodward, J. Chem. Phys, 94 (1991) 3183. [56] E. Kierlik and M. L. Rosinberg, J. Chem. Phys. 100 (1994) 1716. [57] J. B. Hooper, J. D. McCoy, and J. G. Curro, J. Chem. Phys, 112 (2000) 3090. [58] J. B. Hooper, M. T. Pileggi, J. D. McCoy, J. G. Curro, and J. D. Weinhold, J. Chem. Phys, 112(2000)3094. [59] J. B. Hooper, J. D. McCoy, J. G. Curro, and F. van Swol, J. Chem. Phys, 113 (2000) 2021. [60] I. Napari, A. Laaksonen, and R. Strey, J. Chem. Phys, 113 (7013) 2000. [61] V. Talanquer and D. W. Oxtoby, J. Chem. Phys, 118 (2003) 872. [62] E. A. Miiller and K. E. Gubbins, Ind. Eng. Chem. Res. 40 (2001) 2193. [63] F. J. Bias, E. Martin del Rio, E. de Miguel, and G. Jackson, Mol. Phys, 99 (2001) 1851. [64] A. M. Somoza and P. Tarazona, J. Chem. Phys, 91 (1989) 517. [65] R. Holyst and A. Poniewierski, Phys. Rev. A, 39 (1989) 2742. [66] M. Schmidt, Phys. Rev. E, 63 (2001) 050201 (R). [67] J. M. Brader, A. Esztermann, and M. Schmidt, Phys. Rev. E, 66 (2002) 031401. [68] J. W. Cahn, J. Chem. Phys, 66 (1977) 3667. [69] P. P. F. Wissels, M. Schmidt, and H. Lowen, Phys. Rev. E, 68 (2003) 061404. [70] J. M. Brader, R. Evans, M. Schmidt, and H. Lowen, J. Phys.: Condens. Matter, 14 (2002) LI.
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 2
The structure of fluid interfaces determined by neutron scattering J. Penfold ISIS Pulsed Neutron Facility, CCLRC, Rutherford Appleton Laboratory, Chilton, Didcot, OXON, 0X11 OQX, UK 1. INTRODUCTION The structure of fluid interfaces, the air-liquid and liquid-liquid interfaces, and the nature of surfactant and polymer adsorption at such interfaces is central to our understanding of the nature of a wide range of colloidal dispersions, and a wide variety of important surface and interfacial phenomena. These include the stability and manipulation of colloidal dispersions, emulsion stability, surface adsorption, and the generation of functionalized surfaces and novel surface and solution structures. A key aspect of these broad areas is the widespread industrial, technological, domestic and medical applications; and which include household products, personal care products, lubrication, adhesion, coatings, paints, fuel oils, enhanced oil recovery, sensors, many modern food products, encapsulation, drug delivery vehicles, bio-sensors, and bio-functionalised and bio-compatible surfaces. In recent years a wide range of sophisticated techniques have been developed for the study of the nature of surface and interfaces which extend and complement the classical techniques such as surface tension; and include optical reflectometry, ellipsometry [1,2], the optical techniques of sum frequency and second harmonic generation [3,4], surface plasmon resonance [5], surface sensitive FTIR [6], atomic force microscopy (AFM) and related surface force probes [7-10], and the use of spin probes [11]. The use of the scattering techniques associated with x-rays and neutrons have emerged as particularly important and powerful probes of surface and interfacial structure, and include x-ray [12] and neutron reflectivity [13], and the
34
J. Penfold
corresponding bulk techniques of small angle x-ray scattering, SAXS [14], and small angle neutron scattering, SANS, [15]. The neutron scattering based techniques of reflectivity and SANS are particularly powerful for the organic systems encountered in soft matter, in that hydrogen / deuterium isotopic substitution can be used manipulate the system, and make specific components or fragments of components effectively visible or invisible by refractive index matching. In this chapter the techniques of neutron reflectivity and SANS will be introduced, and it will be shown how they are applied to determine the structure of fluid surfaces and interfaces, and of surfactants and polymers adsorbed at such interfaces. This will be illustrated by a range of recent examples that focus on a variety of air-solution interfaces that are of direct relevance to the liquidliquid interface, and directly on the liquid-liquid interface. It will be shown how neutron reflectivity can be used to determine surfactant adsorption, with examples of nonionic surfactants, and how this can be extended to binary and multi-component mixtures. Neutron reflectivity and SANS will be used to contrast the adsorption of ionic / nonionic surfactant mixtures at the air-water and oil-water interfaces. Related to this it will be shown how SANS has been applied to the study of surfactant mixing at the oil-water interface in microemulsions. It will be shown how neutron reflectivity, in combination with D / H isotopic substitution, can be used to determine the detailed structure of a surface surfactant adsorbed layer. This will be extended to the structure of binary surfactant mixtures and to refine the description of the solvent profile at such interfaces. The effect of the progressive change from a nonionic surfactant to alcohol and to an alkane on the packing with an ionic surfactant at the airsolution interface is discussed in the context of understanding the oil-water interface. Finally the role of alcohols and co-surfactants on the nature of the fluid-fluid lamellar phase dispersion will be discussed, and the nature of the surface in equilibrium with such dispersions will be described.
2. NEUTRON SCATTERING TECHNIQUES In recent years neutron reflectivity has emerged as a powerful technique for the study of surfaces and interfaces [14], and has been extensively applied to the study of surfactant [16] and polymer [17] adsorption at interfaces. The particular feature that makes the technique so powerful and particularly surface sensitive is the different scattering powers of hydrogen and deuterium (hydrogen has a negative scattering length of -0.374 x 10"'2 cm and deuterium a larger positive value of 0.6764 x 10"12 cm). Hence D / H isotopic substitution can be used in organic systems to label particular components or fragments by refractive index matching, whilst leaving the chemistry essentially unaltered. The same approach
The Structure of Fluid Interfaces Determined by Neutron Scattering
35
can be used in the bulk scattering technique of SANS to make it surface sensitive, and this has been successfully applied to the study of meso-phase structure [18], adsorption onto colloidal particles [19] and emulsion droplets [20]. 2.1 Neutron reflectivity The specular reflectivity of neutrons, like the analogous light or x-ray reflectivity, from a surface or interface provides information about the neutron refractive index gradient or distribution in the surface region and in a direction orthogonal to the plane. This can often be simply related to a composition or concentration profile in the direction orthogonal to the surface, to provide directly information about adsorption and the structure of the adsorbed layer. The technique and its broad applications have been extensively reviewed elsewhere [14, 16,21,22]. In the simplest case of the interface between two bulk media the reflectivity is related to the refractive index difference across the interface, and is described by Fresnel's Law [23]. The refractive index, n, is related to the neutron scattering length density, such that, «2=1-—p
(1)
71
where X is the neutron wavelength and p is the scattering length density given by,
where b, is the scattering length and n, is the number density of species i. The refractive index is not a convenient term to use in neutron reflectivity as it is ~1.0 (in contrast to light), and the scattering length density is a more relevant term. For most materials the neutron scattering length, b, is positive (hydrogen is one of the exceptions), and so n is generally < 1.0. This highlights another difference compared to light; in general total external reflection occurs as most materials have a refractive index less than air. So, for example, at the air / D2O interface there will be total reflection from the interface below the critical glancing angle of incidence. However, the information about the surface / interface structure occurs at grazing angles beyond the critical angle. In the simple case of a monolayer (thin layer) of uniform composition and density at the interface between two bulk media (say air and a substrate) the neutron reflectivity can be given exactly as [23],
36
J. Penfold
R _ R 2 = r °' + ^ + 2r°/n COS(2Pl ) l + r02,r122+2r01rI2cos(2pl)
(3)
where the subscripts 0, 1, 2 refer to the air phase, the monolayer, and the substrate, p is the optical path length, p]=qiT1, %\ is the monolayer thickness, and q, is the neutron wave vector in the monolayer and normal to the interface. The reflectivity is usually expressed in terms of the wave vector transfer, K, where, 47Tsin90 K
=
= 2^0
(4)
A,
and 90 is the grazing angle of incidence. The Fresnel coefficients, r,j, are then, f — v
rv=^—^
(5)
r
j+r,
The critical wave-vector transfer (below which total reflection takes place) is, K 2 .,=4^=167t( P / -p,) such that the Fresnel coefficients can be expressed in terms of K and Kci, /
^
2
2 ^
K - K. (K
(6)
2-K
CI)
L
^
~ K +K
For a simple sharp interface between air and the substrate (Fresnels law) Eq. (3) becomes, 2
I6n
/ . A
\2
^ = K,| = — T ( P )
,o. 8
()
and for a simple monolayer, R = —J-[(p, " P0)2 + (p2 - P. )2 + 2 (Pi - Po)(p2 - Pi )COSKX]
(9)
The Structure of Fluid Interfaces Determined by Neutron Scattering
37
Although only approximate, it is sufficient to describe the thin monolayer observed in surfactant absorption, and this is the essence of the measurement of surfactant adsorption. For a 0.088 mole fraction D2O / H2O mixture, null reflecting water, nrw, the scattering length density is zero (a refractive index of unity), and there will be no specular reflectivity at the interface between that and air (in Eq. (8) Ap is zero). Specular reflection will occur if there is a monolayer of deuterated material at the interface, and Eq. (9) will become, ^ r i ? = 2p 2 (l-cosKT) = 4&Vsin— V ; 16TT 2
(10)
Hence treating the adsorbed layer as a thin film of uniform density the adsorbed amount, T, can be determined from the thickness, x, and the scattering length density, p, such that [16],
r = -£L = ^ Nab
(ii)
NaA
where A is the area / molecule, and Na is Avogadro's number. A similar expression to Eq. (10) can be obtained if it is assumed that the monolayer has a Gaussian distribution rather than a uniform form, to give [16], 4
K
~—jR =
2
2
7tCT p
2
K
/
^ ^ ( e x p - K 2a
2\
/1ON (12)
and r=(nap m ) l/2 /2N a , where a is the 1/e width and pm the scattering length density at the maximum of the distribution. In practice either functional form can be used. It is straightforward to extend equation 11 to take into account the adsorption of a multi-component mixture [16], such that for a binary mixture equation 11 becomes,
where the subscripts now refer to the two different components. Beyond the simple thin surfactant monolayer the reflectivity can be interpreted in terms of the internal structure of the layer, and can be used to determine thicker layers and more complex surface structures, and this can be done in two different ways. The first of these uses the optical matrix method
38
J. Penfold
[24] developed for thin optical films, and relies on a model of the surface structure being described by a series or stack of thin layers, This assumes that in optical terms an application of Maxwell's equations and the relationship between the electric vectors in successive layer leads to a characteristic matrix per layer, such that,
c
-U- H
<14)
For n layers the matrices for each layer are multiplied together, [C]=[Ci][C2]...[Cn+i], to give the resultant reflectivity from the elements of the final 2x2 matrix such that, R=cc* / aa*. In practice recurrence relationships between the Fresnel coefficients in the successive layers can be used to provide an efficient calculation. Furthermore, following the approach of Nevot and Croce [25], a Gaussian roughness or diffuse profile can be included at each interface in the stack [26], such that, r} = r, exp-O.SK^K^CT 2
(15)
The second method uses the kinematic approximation in which the reflectivity is given as [27], 167i 2 h , J
R = —T-\P(K)
16n2 „., ,2 =—rP(K)
(16)
where P ( K ) is the one-dimensional Fourier transform of the scattering length density profile normal to that interface, p(z), P ( K ) = £op(z)exp(/Kz)
(17)
and p (K)1S the corresponding Fourier transform of the derivative, p'( z )- For a uniform layer or a Gaussian distribution for a thin layer on a substrate, this approach will also give Eq. (10) and (12). The other commonly used and relevant distributions which can be used to describe the solvent or substrate distribution at the interface are the tanh distribution [28] given by,
p(z) = ^fl±tanhf^jl
(18)
The Structure of Fluid Interfaces Determined by Neutron Scattering
39
where £, is the width parameter, z0 the distance between the centre of the solvent distribution and the interfacial layer, such that [28],
16TT 2
K
J
°{
2
)
{
2 J
and the reflection from a surface roughened by capillary waves is given by, 4
^ - r i ? ( K ) = p^exp(-a 2 K 2 )
(20)
and where a is the capillary wave amplitude. Eq. (16) is only approximate, but corrections have been derived [29] which for thin layers essentially remove that inaccuracy. The use of the partial structure factors [28] provides a basis for determining the surface structure from the reflectivity in terms of the contributions from the different components in the layer. Substituting Eq. (2) into Eq. (16) gives,
R~WbX+Y,2b,bA
(2i)
where hn are the self partial structure factors given by, hn=\h\2
(22)
and the hy are the cross partial structure factors given by,
^. = Re[«,«;]
(23)
The «,(K)are the one-dimensional Fourier transform of n.(z) the average number density profile of group i in the direction normal to the surface (see equation 17). From the shift theorem in Fourier transforms [30] we can express the cross terms in the following form, h,J{K) = ±[h,hJf
COS(/K5 V )
(24)
40
h,l(K) = ±[hihJfsm(iK5IJ)
J. Penfold
(25)
where the ± arises from the uncertainty in the phase and Sy is the distance between the centres of the distributions associated with hj, h>. The two forms (Eq. (24) and (25)) arise from the combination of Fourier transforms that are entirely real or imaginary (arising from even distributions) or where one is real and the other is imaginary (arising from a combination of an odd and an even distribution). The use of the partial structure factors leads to a way of analyzing the reflectivity data from a series of different isotopic combinations, and this approach has been extensively used to determine structure in surfactant and related systems [16]. 2.2 Small angle neutron scattering, SANS Small angle scattering is scattering in the forward direction, and is related to the scattering properties of the material at small scattering vectors, K, where ic is defined as K = —sin — , and 9 is the scattering angle. For cold neutrons A, 2 (wavelengths greater than or of order 4 A) the K range accessible is typically in the range 10~3 to 0.5 A"1. Hence SANS can provide information about structure and organization in dispersions on the length scale of 20 to 1000 A. As with neutron reflectivity, H/D isotopic substitution provides the opportunity to manipulate 'contrast' within the dispersion, and this chemical and spatial resolution makes it a powerful probe of the structure of self-assembling systems, colloids, and adsorbed layers. The relevant quantity measured (the scattered intensity I(K) as a function of K) is related directly to the differential scattering cross-section (a) per unit solid angle (Q), which can be written as a Fourier transform of the scattering length density distribution [15],
[£)4=Wjp(r)cxp(zKr)dr \
(26)
where p(r) is the scattering length density distribution (which contains the structural information), N is the number of scatterers, the integral is over the sample volume, and the ( ) denotes an average over all orientations with respect to K. For a dilute solution of discreet particles (for example, micelles, colloids, emulsion droplets) in solution the scattering can be written as [15].
The Structure of Fluid Interfaces Determined by Neutron Scattering
41
where the form factor which contains the information about the particle shape. The form factor, F(K), is written as, F(K)=
J(p(r)-p,)exp(iKr)
(28)
vp
and ps is the scattering length density of the solvent and the integral is over the particle volume (vp) and p(r) describes the scattering length density distribution within the particle. For concentrated interacting particles, where inter-particle as well as intraparticle correlations are important, Eq. (27) becomes [31],
[ ^
N[s(K)(F(K)fK+A(K)]
A(K) = ( | F ( K ) | 2 ) K - ( F ( K ) ) 2 K
(29) (30)
and S(K) is the inter-particle structure factor which is the Fourier transform of the real space inter-particle radial distribution function, g(r). A(K) is a coherent disorder term which is switched on by the interactions and disappears in dilute systems where S(K) is of order 1.0. For mono-disperse spheres Eq. (29) reduces to,
{J±^=NS{K)P{K)
(31)
andP(K) = |F(K)| 2 . For mono-disperse particles with a centro-symmetrical distribution of scattering lengths, (F2(K))K=(F(K))>FS2(K)
and for a uniform sphere of radius R,
(32)
42
J. Penfold
F(K) = V{p-Px)F0(KR) = V{p-pt)3jl{KR)/KR
(33)
where _/, (KR) is a first order Bessel function. For a core and shell radial geometry then, Fs(K) = Vt(pi-p2)F0(KRl) + V2(p2-ps)F0(KR2)
(34)
4 where V.t =—nR?, p , 2 is the scattering length density of the core and shell, and R], R2 the radius of the inner core and the outer dimension. Effects of polydispersity can be included straightforwardly [31], and standard analytic expressions for the form factors of different geometries (for example, ellipsoids, cylinders, and discs) exist [15]. Various analytical expressions for different structure factors, S(K), for hard spheres [32], repulsive Yukawa [33], and attractive interactions [34], exist. Manipulating the refractive index distribution or 'contrast' by D/H isotopic substitution, as with neutron reflectivity, is a central feature of exploiting SANS to study colloidal systems. Measurements in D2O will highlight micelles, colloids, and emulsion droplets compared to the background solvent [15]. For a core-shell structure, either for micelles [31], or for an adsorbed layer on a colloid particle [19], or at the oil-water interface in colloids, or in micro-emulsions [35-37], deuterium labeling the core and solvent will, for example, highlight the shell (or adsorbed layer). Analysis of such data, using Eq. (34) to obtain a thickness and a scattering length density, will allow, from the known scattering length density of the adsorbate and its molecular volume, the adsorbed amount to be determined from, r=
'(P'-PJ
(35)
NaV(p,-pa)
where d is the layer thickness, ps, pa, p2 are the solvent, adsorbate and layer scattering length densities, V the adsorbate molecular volume, and N a Avogadro's number. In a multi-component system, for example in a mixed micelle [38], or a mixed surface layer [39], deuterium labeling can be exploited to determine compositions (in a way analogous to the use of neutron reflectivity for mixed surface layers, see Eq. (13)). In dilute solution and in the limit of low K Eq. (31) can be simplified as [15],
The Structure of Fluid Interfaces Determined by Neutron Scattering
43
where N is the particle number density, V the 'dry' volume, pp the particle scattering length density. For a binary mixture, a ratio of the measured scattered intensity for two different isotopically labeled combinations (subscript i) will provide an estimate of the volume fraction of one of the components. This has been exploited to measure micelle compositions [38, 39], and the composition of absorbed layers on colloidal particles [19], at the oil-water interface in emulsion droplets [20], and in micro-emulsions [35-37]. In the application to micellar systems [15], it has also been shown that the limitation of dilute systems is not a strong requirement, and providing S(K) is not 'contrast' dependent the same approach can be applied to more concentrated systems. 3. SURFACTANT SYSTEMS 3.1 Surfactant adsorption at the air-solution interface For a deuterium labeled surfactant in nrw the specular reflectivity arises only from the adsorbed surface monolayer. Treating the monolayer as a layer of uniform density or as a Gaussian distribution enables the adsorbed amount to be estimated straightforwardly (using Eqs. (10-12)). The limiting value as K tends 4
to zero of a plot of In
-R
verses K" (from Eq. (12)) is determined by Y,
\_\6n J whereas the slope is directly related to the monolayer thickness, a. This is shown in Fig. la for the nonionic surfactant Ci2E8 (dC|2hE8) for the concentrations 9 x 10"6 and 9 x 10"5 M, and the corresponding adsorption isotherm from 10'6 to 10"3 is shown in Fig. lb [40]. This approach has been used to measure the adsorption isotherm for a wide range of surfactants, and Fig. 2 shows the isotherms for C|2E2 to C|2Ei2 [41]. A particular feature is the ability to make measurements over a wide concentration range, from dilute solutions to concentrations well in excess of the cmc [39], and to study multi-component mixtures [42]. Even at concentrations up to the cmc the technique has made important conmtributions. Careful comparison with surface tension measurements have demonstrated the role of divalent counterion impurities [43, 44], established the correct pre-factor in the Gibbs equation for Gemini surfactants [45], and highlighted the need for careful analysis of surface tension data and the role of effects such as wetting / dewetting [46] in interpreting surface tension data. It also provides an absolute measure of adsorption, in circumstances where the Gibbs isotherm cannot be used, such as measuring dynamic surface excess in an overflowing cylinder [1].
44
J. Penfold
Fig. 1. (a) Ln-^—^R vs K2 for dC| 2 hE 8 in nrw and ( • ) 9 x 10-5 M and (o) 9 xlO-6 M, the 16/r solid lines are calculated curves using Eq. (12) , (b) Adsorption isotherm for C12E8, adsorbed amount, F, (xlO" mol cm" ) as a function of surfactant concentration (M).
Fig. 2. Adsorption isotherm for C12E3, C12E6, C12E8, and C12E12 (symbols are shown in the figure)
The ability to make measurements of the surface excess above the cmc is an important feature of neutron reflectivity, and in combination with Gibbs isotherm measurements the surfactant activity above the cmc can be established; as demonstrated for C,4TAB [47] and SDS [44]. 3.2 Adsorption of surfactant mixtures The technique is particularly powerful, not only for concentrations above the cmc, but also for studying the adsorption of multi-component mixtures; and a range of different mixtures have been studied [16]. Fig. 3 shows the adsorption
The Structure of Fluid Interfaces Determined by Neutron Scattering
45
Fig. 3. Adsorption isotherm for equimolar mixture of C12E3/CnEg, (•»•) total adsorption, ( • ) C12E3, and (•••) Ci 2 E 8 .
for the binary nonionic surfactant mixture of an equi-molar solution of C|2E3 / C12E8 [48]. Measurements in nrw for the three isotopically labeled combinations, dC,2hE3 / dC,2hE8, dCi2hE3 / hC12hE3, hC,2hE3 / dC,2hE8, and using Eq. (8) enables the total adsorption and the amount of each component to be determined unambiguously. The combination dC)2hE3 / dC12hE8 provides a measure of the total adsorption, whereas the combinations dC12hE3 / hCi2hE8, and hC]2hE3 / dCi2hE8, provide a direct measure of the C]2E3 and C12E8 adsorption respectively. The isotherm in Fig. 3 shows the evolution of the surface composition from below the cmc to some two orders of magnitude above the cmc. The adsorption shows a strong preference for the more surface active component, C]2E3 (see also Fig. 2), and the total adsorption and composition tend to a constant value at concentrations well in excess of the cmc of the mixture. The abrupt change in the adsorption pattern at about 5 x 10"5 M is associated with the mixed cmc. Such behaviour is predicted theoretically [49], and is associated with the change in distribution of the two surfactants between the solution and the surface due to the onset of micellisation, reflecting the change in the monomer concentration at micellisation. This is also consistent with the predictions of Regular Solution theory, RST [50], except that due to
46
J. Penfold
Fig. 4. Nonionic surface and solution composition (mole % CnE,,) as a function of EO chain length for 3 x 10"3 M 90/10 Ci2En / SDS, where n is a distribution from 0 to 12 with a solution average of 5 (symbols are as shown in the figure)
differences in surface pressure of the pure components and packing constraints the surface remains relatively rich in Ci2E3. An extreme example of the application of the technique to the study of surfactant mixing is the study of SDS / Ci2E5 mixtures at concentrations greater than the cmc, which mimic the behaviour of commercial detergent formulations [42]. The surface adsorption for the SDS / Ci2E5 mixture was measured at the air-water interface for a pure SDS / Ci2E5 and for a mixture of SDS with nonionic surfactants from C12OH to Ci2E|2 (but with an average ethylene oxide chain length of 5 and a distribution that mimics the commercially impure nonionic surfactants). Fig. 4 shows the surface distribution of the different nonionic components compared to the solution composition. The average ethylene oxide chain length is of order 3.5 compared to the solution average of 5. There is a strong preferential adsorption of the shorter ethylene oxide chain length surfactants at the interface, and this accounts for the significantly enhanced total adsorption compared to the pure SDS / C|2E5 mixture. 3.3 Adsorption at the oil-water interface There have been relatively few reported studies using neutron reflectivity to probe adsorption at the liquid-liquid interface [51]. This is due to the difficulty in reliably establishing a thin upper liquid layer that is sufficiently transparent to cold neutrons. In contrast, the greater flux of x-rays on synchrotron sources has led to extensive x-ray studies of the liquid-liquid interface. Recent developments [53] suggest that this is now resolved for neutron reflectivity, but few systematic studies have yet to emerge. In the
The Structure of Fluid Interfaces Determined by Neutron Scattering
47
Fig. 5. Scattered Intensity (plotted as Q I(Q) vs Q2 ) for 6.4 vol % d/h-hexadecane in D2O emulsion with h-SDS in the concentration range 1.25 to 20 mM (Q is equivalent to K in the text)
meantime SANS has been used to investigate surfactant mixing at the oil-water interface by studying emulsion droplets [20] and micro-emulsions [35-37]. In the pioneering study of its kind, Staples et al [20] used SANS to investigate the nature of surfactant mixing at the oil-water interface for the mixture SDS / Ci2E6 at the hexadecane-solution interface. Hexadecane-in-water emulsions, < 0.2 (im diameter, were prepared with the hexadecane index matched to D2O (93.7 volume % d-hexadecane). The scattering curves then arise only from the hydrogeneous surfactant adsorbed at the interface, and this is shown in Fig. 5 for SDS over a wide range of concentrations. Surfactant can be added to the emulsion to change the adsorbed amount and the composition (in the case of mixed surfactants) whilst maintaining the emulsion stability. In Fig. 5 the scattering in the concentration range 2 to 12 mM is consistent with a thin adsorbed layer of surfactant (~ 13 A thick) with increasing coverage. Above 12 mM the form of the scattering changes markedly and the scattering is consistent with the adsorbed layer and free micelles. For the surfactant mixture SDS / C|2E6 it was possible to determine the composition of the mixed surfactant layer (at the oil-water interface) and of the micelles in equilibrium with the emulsion. Fig. 6 shows a comparison for a 70:30 mole ratio SDS / Ci2E6 mixture of the composition at the oil-water interface, the composition of the corresponding air-water interface [39], and the composition of the mixed micelles [38, 39]. Previous extensive neutron reflectivity [39], and SANS [38, 39] measurements on the SDS / Ci2E6 mixtures have shown that the mixing is broadly consistent with RST, but that an interaction parameter, p, that varies with solution composition is required (for example, for the micelles P is 2.6 for a 70:30 composition, and -2.0 for a 50:50 composition). In contrast, the behaviour at the oil-water interface and in the micelles that are in equilibrium
48
J. Penfold
Fig. 6. Comparsion of oil-water, air-water interface, and micelle compositions (mole % SDS) for a 70/30 SDS / C12E6 mixture as a function of surfactant concentration (M), symbols as shown in the figure..
with the emulsion droplets is not consistent with RST. This was rationalized in terms of the solubilisation and hence partitioning into the hexadecane phase, and of the solubilisation of hexadecane into the micellar phase. Eastoe et at [35-37] have applied the same approach in SANS to investigate the nature of surfactant mixing in water-in-heptane micro-emulsions [36], studying DDAB / SDS and DDAB / CI2E5 [36], and DDAB / C12E3 to C12E23 [37] mixtures. For the DDAB / SDS and DDAB / C,2E5 [35, 36] they showed how the micro-emulsion phase behaviour and properties are a direct consequence of the preferred film curvature, which is manipulated by the surfactant mixing at the interface. In the DDAB / Ci2En mixtures departures from ideal mixing is observed at the micro-emulsion interface. It was found that the longer ethylene oxide chain length surfactants partition more strongly into the DDAB layer than the short ethylene oxide chain length surfactants, and that this has an impact on the microemulsion phase behaviour. This is a result of the shorter ethylene oxide chain length surfactants partitioning more strongly into the oil phase. As with the emulsion droplet study of Staples et al [20], partitioning effects of the nonionic surfactant are shown to strongly influence the overall behaviour. 4. SURFACE STRUCTURE In recent years the partial structure factor analysis approach has been used in neutron reflectivity to determine the structure of surfactant monolayers at the air-water interface [16], (see Eqs. (21-25)). This has been applied to a number of
The Structure of Fluid Interfaces Determined by Neutron Scattering
49
different surfactants [16], and surfactant mixtures [53-56], where deuterium isotopic labeling has been used to probe varying degrees of structural detail. 4.1 Structure of surfactant monolayers Almost the simplest isotopic labeling scheme that can be used to determine the structure of the surfactant monolayer adsorbed at the air-water interface is to label separately the alkyl chain, the headgroup, and the solvent. For this labeling scheme Eq. (21) becomes,
R
=
l
-^[bX + b2hhhh + bX, + 2bcbhhch + 2bJ,what + 2 W J
(37)
where the subscripts c, h and w refer to the alkyl chain, headgroup, and solvent respectively. To determine the six partial structure factors (Eq. (37)) requires six reflectivity measurements with different bj values, obtained by either deuterium labeling the different components, and by suitable D/H isotopic substitution to make a particular component equal to zero. The three self terms, hcc, hhh and h ^ , provide information about the distributions of the alkyl chain, the headgroups and the solvent at the interface. The alkyl chain and headgroups have been shown to be well-described to good approximation by a Gaussian distribution [28], whereas the solvent distribution has been described using a tanh function [28], see Eqs. (12) and (19). The cross partial structure factors describe the relative positions of the different components at the interface, 8ch, 5CW, and 5hw; where 5ch is described by Eq. (24) and 5CW, 5hw by Eq. (25). The structure is generally over-determined, as 5c,,,-5cA + 5fe),
(38)
the phase uncertainty in Eqs. (24) and (25) can be deduced at the air-water interface as the hydrophobic portion of the surfactant molecule will be adjacent to the air phase, and stoichiometry dictates that the area/molecule for the headgroup and the alkyl chain are identical. From this, a convenient way to present the structure is in the form of number density or volume fraction distributions, calculated from the number density distributions deduced above and from the known molecular volumes. Fig. 7 shows the structure for the cationic surfactant, C16TAB, determined at this level of isotopic labeling [28]. Included in the figure for comparison are the computer simulation results of Bocker et al [58], and similar agreement with the neutron reflectivity results were also obtained by Tarek et al [59]. The neutron reflectivity results show a significant overlap between the alkyl chain, headgroup, and solvent distributions, and a substantial contribution to the widths of the distributions
50
J. Penfold
Fig. 7. Number density distribution determined from neutron reflectivity (dashed line) for alkyl chains (a), headgroup (b) and (c) solvent for C^TAB, solids lines are from computer simulation (see text).
from capillary waves [60]. This is characteristic of a disordered surface layer. More detailed labeling schemes (deuterium labeling fragments of the alkyl chain) have enabled the conformation of the alkyl chain to be deduced and the capillary wave contribution to the width to be deduced [28]. Similar labeling schemes have been applied to the monoaikyl ethoxylate nonionic surfactants [16], where partial labeling of the alkyl chain and of the ethoxylate chain has enabled the conformation of both chains to be deduced. A notable feature of the monoaikyl ethoxylate surfactants is that there is increasing overlap between the alkyl chains and the ethoxylate chains with increasing ethoxylate chain length [41]. Detailed labeling of the alkyl chains reveals an interesting difference in the conformation of the alkyl chain of the Cn cationic (Ci2TAB) and the C,2 nonionic (C12E3), see Fig. 8.
Fig. 8. Comparison of chain conformation of C12TAB and C12E3, determined from partial labeling of the alkyl chain.
The Structure of Fluid Interfaces Determined by Neutron Scattering
51
Fig. 8 shows the mean alkyl chain conformation, from the headgroup towards the terminal methyl group. For the cationics the alkyl chain tilts to a more oblique orientation towards the outer part of the chain, whereas the reverse is true for the nonionic surfactant: illustrating the role of the headgroup geometry in influencing the alkyl chain conformation. Whereas the distribution of the individually labeled surfactant components (the alkyl chain, and headgroups) are well described by Gaussian functions [16], in good agreement with computer simulation, the solvent distributions are not universally well described. The tanh distribution is adequate for the ionic surfactants [28], and is reasonably close to computer simulation [58, 59], see Fig. 7, but is not an entirely satisfactory description for the nonionic ethoxylated surfactants. A different strategy was adopted to improve upon the tanh distribution [60]. It is assumed that below a critical depth, zc, solvent fills all the remaining space not occupied by surfactant, <|>,(z) = 1.0-<|>e(z)-<|>A(z)
(39)
and for z>zc the solvent distribution decays as a Gaussian of width, ^,
fv(z) = f (zjexp[-(z-zt.) /^2]
(40)
From the <|)s(z) distribution ns(z) can be estimated, and by Fourier transformation of ns'(z) hww(K) can be calculated. This was shown to provide an improved description for the nonionic surfactants Ci2E6 and C]2E8 [61]. 4.2 Structure of mixed surfactant monolayers Recent neutron reflectivity studies on the adsorption of surfactant mixtures at interfaces [56, 57, 62, 63] have demonstrated the importance of determining not only adsorbed amounts, but also the structure of the adsorbed layer. In Regular Solution Theory [50], the departure from ideal mixing is characterized by a single interaction parameter, P; with the central assumption that the excess entropy of mixing is zero. Recent studies [62, 63] have shown that this assumption is not always applicable, and that correlations between structure and composition in mixed systems suggest that structural changes and especially changes in hydration are important. In conjunction with neutron reflectivity, similar isotopic labeling schemes to those used to study pure surfactant monolayers (as described in the previous section) have been used to study a number of surfactant mixtures. In the context of the focus of this chapter, we highlight briefly two systematic studies of the structure of the mixed nonionic surfactant monolayers of Cj2E3 / C|2E8 and
52
J. P enfold
CioE6 / Ci4E6 [56, 57], where the two systems contrast the effects of the differing ethylene oxide and alkyl chain lengths on the molecular packing at the interface. Comparison with the structure of the pure component nonionic surfactant monolayers reveal the effect of mixing on the conformation of the two components in the mixed monolayer. Isotopic labeling of the alkyl chains, ethylene oxide chains, and solvent in the two binary mixtures enables the distribution of each component in the mixture to be determined (in a way analogous to the use of Eq. (37) for a single component monolayer), and compared with the structure derived for the pure component monolayers. Fig. 9 shows the volume fraction distribution for the C[2E3 / Ci2E8 and C|0E6 / C!4E6 mixtures at the air-water interface. For the Ci2E3 / Ci2E8 mixture the alkyl chain distributions coincide, whereas the centre of the Eg distribution is further from the alkyl chain than the E3 distribution (see Fig. 9a). In contrast, for the C10E6 / C|4E6 mixture the alkyl and ethylene oxide chain distribution are coincident (see Fig. 9b). For the C|2E3 / Ci2E8 mixture the frustration caused by packing the E3 and E8 headgroups results in a change of the surface structure of each component compared to the pure component monolayers. The alkyl chain distributions of both surfactants are more extended. The capillary wave broadened distributions are often not particularly sensitive to changes in conformation, but measurements of the whole chain labeled and of a labeled fragment of the chain (in this case the outer C6 of the chain) provide a sensitive measure of changes in conformation. The E3 headgroup is less hydrated and the Eg headgroup more hydrated and less extended in the mixed monolayer compared to the pure monolayers; and this is illustrated in Fig. 10a. In essence, the C,2E8 structure adjusts to more closely accommodate the structure of the Ci2E3 on mixing. In contrast, the packing of the two different alkyl chain
Fig. 9. Volume fraction distributions for (a) 5 x 10"5M 30 / 70 mole ratio C12E3 / C12E8, and (b) 4.2 x 10"4 M 97.3 / 2.7 C t0 E 6 / Ci4E<;, symbols as shown in the figure.
The Structure of Fluid Interfaces Determined by Neutron Scattering
53
Fig. 10. Volume fraction distributions for (a) Ci2E3 / Ci 2 E 8 (as figure 9a) compared to pure C|2Eg monolayer and (b) C10E6 / C14E6 (as figure 9b) compared to pure C10E6 monolayer (symbols as shown in the figure).
lengths in the Ci0E6 / Ci4E6 mixture also results in structural changes compared to the pure monolayers, but in a different way to the Ci2E3 / C12E8 mixture. The relative positions of both the alkyl and ethylene oxide chains are altered on mixing; with the largest impact on the Ci0E6. For the Ci0E6 in the C10E6 / C14E6 mixture the molecule is partially dehydrated on mixing, the alkyl-ethylene oxide chain overlap decreases, and the alkyl chain is more extended (see Fig. 10b). 4.2 Towards fluid-fluid interfaces Understanding the interaction of an oil (alkane) with a surfactant monolayer is important in the context of a number of situations. The enhanced penetration of short chain alkanes into the surfactant alkyl chain region compared to larger alkanes is associated with the stability of oil-in-water compared to water-in-oil micro-emulsions, to destabilize aqueous foams, and in the context of the structure of the oil-water interface. On a surfactant monolayer the addition of alkane results either in spreading to form a film of macroscopic thickness for long chain alkanes, or the formation of lenses in equilibrium with a mixed alkane / surfactant monolayer. In a series of systematic measurements, Fletcher et al [64] have used neutron reflectivity to probe the composition of such mixed monolayers, and to determine the structure for the surfactants, C,6TAB, C14TAB, C12TAB and C12E5 with dodecane [65-67]. The resulting mixed monolayer structures are compared with mixtures where the alkane is replaced by a progressively more soluble co-surfactant, for SDS / dodecanol [44], and Ci6TAB / C!2E6 [54]. Partial deuterium labeling of the surfactant alkyl chain [65, 67] provides an enhanced sensitivity to the location of the alkane solubilised in to the surfactant monolayer. For C]4TAB / dodecane the alkane
54
J. P enfold
distribution coincides with the distribution of the outer C6 of the C ]6 alkyl chain, and was completely located within the hydrophobic region of the monolayer, see Fig. 1 la. In contrast, the structure of the C|2E5 / dodecane mixed monolayer [65] is similar to that for dodecanol / surfactant mixtures, notably SDS / dodecanol [44], where the oil penetrates deeper into the surfactant monolayer, see Fig. 11 b,c. This is attributed to a more favourable interaction between the dodecane and the ethylene oxide chain of the C|2E5. It also results in a greater capacity for solubilisation of the alkane, and a greater impact on the conformation of the alkyl chain. Dodecanol adopts a location within the surfactant monolayer which is intermediate between that adopted by dodecane and a second surfactant, see Fig. lie. For the mixture Ci6TAB / Ci2E6 [54], the distribution of the two surfactant components at the interface are essentially coincident, see Fig. 11 d. With increasing solubility the co-surfactant shows a progressive shift towards the solvent plane within the mixed monolayer.
Fig. 11 .Schematic representation of the surface structure for (a) C^TAB / dodecane, (b) C12E5 / dodecane, (c) SDS / dodecanol, and (d) C|6TAB / Ci2E6
The Structure of Fluid Interfaces Determined by Neutron Scattering
55
5. FLUID INTERFACES Lamellar phase dispersions are a commonly found solution micro-structure in a range of colloidal systems and in a variety of applications; and as such represent an important class of fluid interface. The micro-structure and accompanying phase behaviour is particularly sensitive to the co-adsorption of co-surfactants, alcohols and alkanes, which manipulate the fluidity of the surfactant monolayer, and alter the spontaneous curvature. In lamellar phase dispersions, for example, the addition of co-surfactants such as model perfume molecules like benzyl alcohol and phenyl ethanol, result in a reduction in the L a / Lp phase transition temperature. The increased flexibility in such mixtures results in a different response to shear, and can have a significant impact on formulation and processing. 5.1 Adsorption of co-surfactants Penfold et al [69, 70] have used neutron reflectivity and deuterium labeling to investigate the co-adsorption of benzyl alcohol and phenyl ethanol into a Ci6TAB surfactant monolayer. Deuterium labeling of the C16TAB or the aromatic alcohol allows the composition of the mixed monolayer to be determined. It is found that the phenyl ethanol, which is more surface active than benzyl alcohol, competes more effectively with the C16TAB for the interface. Deuterium labeling of the aromatic alcohol and partial deuterium labeling of the surfactant alkyl chain, and solvent allows the structure of the mixed monolayer to be determined. It is shown that the addition of the aromatic alcohol alters the conformation of the C^TAB, and draws it closer to the aqueous subphase. Importantly it enables a measure of the location of the aromatic alcohol within the surface layer to be obtained. It is found the centre of the alcohol distribution is located within the surfactant monolayer coincident with the distribution of the C6 of the Ci6TAB alkyl chain adjacent to the headgroup. The more hydrophobic phenyl ethanol is located slightly further from the C]6TAB headgroup than the benzyl alcohol, more completely in the hydrophobic alkyl chain region, see fig. 12. Furthermore it has a slightly greater impact upon the conformation of the alkyl chain. This provides direct evidence on the role of the co-surfactant in enhancing membrane flexibility and in the reduction in the L a / Lp phase transition temperature. 5.2 Lamellar - co-surfactant dispersions Lamellar phase dispersions of di-alkyl chain cationic surfactants are extensively used in a variety of product formulations, fabric conditioners, hair shampoos and shower gels, where the formulations are usually mixtures of the di-alkyl chain cationic surfactants with nonionic surfactants. In terms of their
56
J. Penfold
Fig. 12. Schematic representation of the structure of the mixed C^TAB and benzyl alcohol monolayer, showing the relative positions of the Ci6TAB and the benzyl alcohol.
product formulation and functionality the bulk and surface adsorption behaviour are important. Very little is known about the surface behaviour in such mixtures, and especially in concentration regimes where the surface is in equilibrium with a concentrated dispersion micro-structure. Penfold et al [72], in a recent study have shown how neutron reflectivity can be used to study such mixtures at the air-water interface. For the mixture DHDAB (dihexadecyl dimethyl ammonium bromide) and Ci2E6 it was shown that at a solution concentration ~ 10°M (»cmc) and in electrolyte that there was a marked departure from ideal mixing at the interface (see Fig. 13), which cannot be accounted for by existing theories of non-ideal mixing, such as RST. In the absence of electrolyte and for solutions richer in cationic than 40 mole % cationic the surface is almost entirely cationic; and for solutions richer in nonionic the surface composition changes towards one dominated by the nonionic surfactant. In electrolyte, for solution compositions in the region of 40 to 70 mole % nonionic the surface composition is invariant with solution composition, but otherwise shows a trend similar to that in the absence of electrolyte. Measurements of the structure of the mixed surface layer do not show any dramatic changes from the structure of the pure component monolayers, and cannot account for these unexpected surface compositions. To understand the behaviour it is necessary to consider the surface adsorption within the context of the solution behaviour and the microstructure. The region of constant composition observed in the presence of electrolyte coincides with a region of coexistence of mixed micelles and lamellae in the solution that is in equilibrium with the surface. For solutions rich in nonionic surfactant the micro-structure is predominantly micellar, and for
The Structure of Fluid Interfaces Determined by Neutron Scattering
57
Fig. 13. Surface composition (mole % Ci2E6) for 1.3 x 10 3M DHDAB / C| 2 E 6 , in D2O, 0.05M KBr, and 0.1M KBr (symbols as in the figure)
solutions rich in cationic it is predominantly lamellar. These results provide direct evidence of the way in which the bulk phase behavior is mediating the monomer concentration and composition in the solution phase, and hence the surface adsorption behaviour. This highlights in particular the need to characterize both the surface and the solution (which is in equilibrium with the surface) behaviour, in order to understand such systems. REFERENCES [I] S Manning-Benson, S R W Parker, C D Bain, J Penfold, Langmuir 14 (1998) 990 [2] S Manning-Benson, C D Bain, R C Darke, J Coll Int Sci 189 (197) 109 [3] R M Corn, D A Higgins, Chem Rev 94 (1994) 107 [4] G R Bell, C D Bain, R N Ward, J Chem Soc, Faraday Trans 92 (1996) 915 [5] B Schoeler, E Poptoshev, F Caruso, Macromol 36 (2003) 5258 [6] R A Dluhy, D G Cornell, in D R Scheung (Ed) Fourier Transform infrared spectroscopy in colloid and interface science, ACS symposium series 447 (1991) 192 [7] S Manne, H E Gaub, Science 220 (1995) 1450 [8] S Mawie, J P Cleveland, H E Gaub, G D Stucky, D K hausma, Langmuir 10 (1994) 4409 [9] H N Partick, G S Warr, S Manne, I A Aksay, Langmuir 13 (1997) 4349 [10] J Klein, J Chem Soc, Faraday Trans 79 (1980) 99 II1] H Voiel, H A Scheidt, D Huster, Biophys J 85 (2003) 1691 [12] T Russell, Mat Sci Rep 5 (1990) 171
58
[13] [14] [15] [16] [17]
J. Penfold
J Penfold, R K Thomas, J Phys : Condens Matt 2 (1990) 1369 A Guinier, G Fournet, 'Small angle x-ray scattering', J Wiley (1955) J Penfold in 'Encyclopedia of surface and colloid science', Marcel Dekker (2000) 3653 J R Lu, R K Thomas, J Penfold, Adv Coll Int Sci 84 (2000) 143 R A L Jones, R W Richards, 'Polymers at surfaces and interfaces', Camb Uni Press (1999) [18] J Penfold et al, in ACS symposium series 861 on 'Mesoscale phenomena in fluid systems' (2003) 96 [19] P G Cummins, E J Staples, J Penfold, J Phys Chem 94 (1990) 179 [20] E Staples, J Penfold, I Tucker, J Phys Chem B 104 (2000) 606 [21] J Penfold, Current Opin in Coll and Int Sci, 7 (2002) 139 [22] J Penfold, and R K Thomas,Current Opin in Coll and Int Sci, 1 (1995) 23 [23] M Born, E Wolf, Principles of Optics, Pergammon (1970) [24] O S Heavens, Optical properties of thin films, Dover (1955) [25] LNevot, P Croce, Phys Appl 15 (1980 761 [26] J Als-Neilsen, Z Phys B 61 (1985) 411 [27] T L Crowley, E M Lee, E A Simister, R K Thomas, J Penfold, A R Rennie, Coll Surf 52 (1990) 85 [28] J R Lu, Z X Li, J Smallwood, R K Thomas, J Penfold, J Phys Chem 99 (1995) 8223 [29] E A Simister, E M Lee, R K Thomas, J Penfold, Macromol Rep A29 (1992) 155 [30] R N Bracewell, The Fourier transform and its applications, 2nd Ed, McGraw-Hill (1978) [31] J B Hayter, J Penfold, Coll Polym Sci 261 (1983) 1072 [32] N W Ashcroft, J Lekner, Phys Rev 145 (1966) 83 [33] J B Hayter, J Penfold, Mol Phys 42 (1981) 109; J B Hayter, J P Hansen, Mol Phys 42 (1982)651 [34] R J Baxter, J Chem Phys 49 (1988) 2770 [35] A Bumajdad, J Eastoe, S Nave, D C Steytler, R K Heenan, I Grillo, Langmuir 19 (2003) 2560 [36] A Bumajdad, J Eastoe, P Griffiths, D C Steytler, R K Heenan, J R Lu, P Timmins, Langmuir 15(1999)5271 [37] A Bumajdad, J Eastoe, R K Heenan, Langmuir 19 (2003) 7219 [38] J Penfold, E Staples, L Thompson, I Tucker, J Hines, R K Thomas, J R Lu, N Warren, J Phys Chem B 103 (1999) 5204 [39] J Penfold, E Staples, L Thompson, I Tucker, J Hines, R K Thomas, J R Lu, Langmuir 11 (1995)2496 [40] J R Lu, Z X Li, R K Thomas, E J Staples, L Thompson, I Tucker, J Penfold, J Phys Chem 98 (1994) 6559 [41] J R Lu, T J Su, Z X Li, R K Thomas, E J Staples, I Tucker, J Penfold, J Phys Chem B 1010(1997) 10332 [42] E J Staples, L Thompson, I Tucker, J Penfold, Langmuir 10 (1994) 4136 [43] S W An, J R Lu, R K Thomas, J Penfold, Langmuir 12 (1996) 446 [44] J R Lu, A Morrocco, T J Su, R K Thomas, J Penfold, J Coll Int Sci 158 (1993) 303 [45] Z X Li, R K Thomas, J Penfold, Langmuir 15 (1999) 4392 [46] E A Simister, R K Thomas, J Penfold, R Aveyard, B P Binks, P Cooper, P D I Fletcher, J R Lu, A Sokolowski, J Phys Chem 96 (1992) 1383 [47] J R Lu, E A Simister, R K Thomas, J Penfold, J Phys Chem 97 (1993) 13907 [48] J Penfold, E Staples, L Thompson, I Tucker, Coll Surf 102 (1995) 127 [49] Y F Nikas, S Pruvvada, D Blankschtein. Langmuir 8 (1992) 2680 [50] P M Holland, Coll Surf A 19(1986) 171
The Structure of Fluid Interfaces Determined by Neutron Scattering
59
[51] B J Clifton, T Cosgrove, R M Richardson, Physica B 248 (1998) 289; J Bower, A Zarbakhsh, J R P Webster, Langmuir 17 (2001) 140 [52] M L Schlossmann, A M Tikkonov, ACS symposium series 861, 'Mesoscale phenomena in fluid systems'(2003) 81 [53] A Zarbakhsh, J R P Webster, J Bowers, unpublished results. [54] J Penfold, E Staples, P Cummins, I Tucker, R K Thomas, E A Simister, J R Lu, J Chem Soc, Faraday Trans 92 (196) 1773 [55] J Penfold, E Staples, L Thompson, 1 Tucker, Coll Surf 102 (1995) 127 [56] J Penfold, E Staples, I Tucker, R K Thomas, J Coll Int Sci 201 (1998) 223 [57] J Penfold, E Staples, I Tucker, R K Thomas, R Woodling, C C Dong, J Coll Int Sci 262 (2003)235 [58] J Bocker, M Schlenhrich, P Bopp, J Brinkmann, J Phys Chem 96 (1992) 9915 [59] M Tarek, D J Tobias, M L Klein, J Phys Chem 99 (1995) 1393 [60] A Braslau, M Deutsch, P S Pershan, A H Weiss, J Als-Nielsen, J Bohr, Phys Rev Lett 54 (1985) 114 [61] J Penfold, R K Thomas, PCCP 4 (2002) 2648 [62] J D Hines, P R Garrett, G R Rennie, R K Thomas, J Penfold, J Phys Chem B 101 (1997) 7121 [63] J D Hines, R K Thomas, P R Garrett, G R Rennie, J Penfold, J Phys Chem B 102 (1998) 5834 [64] B P Binks, D Crichton, P D I Fletcher, J R MacNab, Z X Li, R K Thomas, J Penfold, Coll Surf 146 (1999) 299 [65] J R Lu, R K Thomas, B P Binks, P D I Fletcher, J Penfold, J Phys Chem 99 (1995) 4131 [66] J R Lu, R K Thomas, R Aveyard, B P Binks, P Cooper, P D I Fletcher, A Sokolowski, J Penfold, J Phys Chem 96 (1992) 10971 [67] J R Lu, Z X Li, R K Thomas, B P Binks, D Crichton, P D I Fletcher, J R MacNAb, J Penfold, J Phys Chem B 102 (1998) 5785 [68] J Penfold, E Staples, I Tucker, L Soubiran, A K Lodi, L Thompson, R K Thomas, Langmuir 14(1998)2139 [60] J Penfold, E Staples, I Tucker, L Soubiran, R K Thomas, J Coll Int Sci 247 (2002) 397 [70] J Penfold, E Staples, I Tucker, R K Thomas, Langmuir 20 (2004) 1269
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 3
Interfacial rheology of adsorbed layers R. Miller" and V.B. Fainermanb a
Max-Planck-Institut fur Kolloid- und Grenzflachenforschung, Am Muhlenber^ 1, 14424 Golm, Germany
b
Medical Physicochemical Centre, Donetsk Medical University, 16 Ilych Avenue, Donetsk 83003, Ukraine
The dynamic properties of interfacial layers are of increasing interest for many technologies. An important aspect of the non-equilibrium interfacial behaviour is the two-dimensional rheology. Studies of the interfacial response to shear and dilational perturbations are now a wide spread topic and a large variety of experimental technique exists, partially even in form of commercial instruments. Shear experiments, suitable for interfacial layers formed by polymers or mixtures of polymers and surfactant, yield only qualitative information on the interfacial structure. In contrast, the dilational rheology, studied for adsorption layers of surfactants, polymers and also mixtures of polymers and surfactants, has a direct interrelation to the mechanisms of the adsorption dynamics and the thermodynamic state. The presented overview is meant as an introduction into the interfacial rheology of surfactant and protein layers. Some theoretical aspects are discussed, specific for the studied systems. Selected experimental techniques for shear and dilational perturbations of liquid interfaces are given and experimental data presented as demonstration how the experiments work and which output information can be expected.
62
R. Miller and V.B. Fainerman
1. INTRODUCTION Most technologies with liquids are highly dynamic and hence require good knowledge of the mechanical behaviour of the involved liquid interfaces and on the possibilities how to modify the properties by adsorption layers. Examples for the impact of dilational and shear mechanical properties are food processing [1, 2, 3, 4, 5, 6], coating processes [7], flotation [8] or enhanced oil recovery [9, 10]. Also phenomena in nature, such as at the atmosphere-ocean interface [11, 12,13], are strongly influenced by rheological parameters. The formation and stabilisation of liquid disperse systems like foams or emulsions, is mainly governed by surface active compounds. However, at present quantitative relationships between interfacial rheology and the stability of foams, for example, do not exist. The behaviour of foams and emulsions depends on various specific parameters. Discussions are based on elementary processes, such as stability of thin liquid films, coalescence of two drops or bubbles, break-up processes of drops in flow fields, motion of bubbles and drops in a solution, etc. The relevance of dilational properties and exchange of matter mechanisms as well as shear rheological parameters, to phenomena such as foaming and emulsification has already been emphasised and is explained by the rapid expansion of surfaces during the generation of foams or emulsions [14, 15, 16]. The rheology of wet foams has been analysed by Wasan et al. [17] and a relation was derived [18] connecting the foam dilational viscosity with the surface dilational elasticity. In contrast, the long time stabilisation of foams and emulsions requires suitable properties of the interfacial layers against small perturbations. A qualitative correlation between foam stability and effective dilational elasticity or dynamic surface tension, respectively, was found for some selected surfactant systems [19,20,21,22,23]. Although it is broadly accepted that knowledge of surface rheology is essential for many technological processes only recently systematic studies of these properties were started. The reason is that there are different surface rheological parameters which reflect the various aspects of the complex behaviour of liquids and their interfaces and it is not trivial to predict which of these parameters allow the best understanding and control of a technological problem. The number of commercial instruments for the determination of interfacial rheological parameters increased over the last years, so that studies are no longer restricted to original set-ups in very few laboratories. These professional developments, which complement each other in many respects, will help to overcome the experimental limits.
Interfacial Rheology of Adsorbed Layers
63
2. THEORETICAL BACKGROUND OF SURFACE RHEOLOGY In contrast to bulk phases the interface of a liquid may be easily expanded and compressed. Both kinds of motion - dilation and shear - can therefore be easily achieved. The reactions observed and the information we will get depend on the kind of deformation or stress. The oscillation of a bubble leads to dilation and compression of the interface. Assuming soluble surfactants in the bulk phase leads to an exchange of matter between the bulk and the bubble surface when the frequency becomes comparable to the time of the adsorption/desorption process, whereas an oscillation with another frequency may lead to a displacement within the interface, i.e. a lateral displacement of some molecules. The structure of the adsorption layer is the result of the interaction between all molecules and ions in the range of the interface. These interactions may in the simplest case be the interaction between the solvent molecules. In the presence of other molecules (surface active ones) the structure occurring depends on the interaction between these adsorbed molecules as well as the interaction between them and the solvent. In more complicated cases macromolecules or mixtures of different surface active components can form surface layers. The structure of such interfacial layers may take a long time and can change significantly with the components dominating in different time scales. Interfacial aggregation can occur accompanied by rearrangements, conformational changes and interactions between adsorbed molecules. A simple projection of 3D rheology to a two dimensional surface layer was critically discussed first by Boussinesq [24, 25]. He interrelated the hydrodynamic equations of two phases discontinuously and formulated a 2D rheology for the interfacial layer. The mathematics of 2D rheology was given in a comprehensive way on the basis of the stress tensor y (consisting of four elements) by Krotov [26], and partly in [27]. Under dilation or compression interfacial layers behave viscoelastic, and the main parameters are the dilational elasticity E and viscosity r|d: s(ia>) = E(
(1)
At infinite frequencies co —> oo the limiting dilation elasticity Eo E
»=^T' dlnF
(2)
64
R. Miller and V.B. Fainerman
while the dilational viscosity vanishes. Under shear deformation interfacial layers show also viscoelastic behaviour with the viscosity r)s and elasticity Gs as the two main parameters. More details on the quantities are given below in connection to the respective experimental procedures. 3. STUDIES OF DILATIONAL RHEOLOGY OF INTERFACIAL LAYERS The two types of relaxation techniques, based on harmonic and transient area changes, respectively, will be described in this paragraph from the theoretical point of view. It will be shown, that the exchange of matter functions are generally applicable to both types of relaxations [28]. 3.1.
Theoretical background
3.1.1. Relaxations induced by harmonic interfacial disturbances Investigations of the damping of capillary waves at interfaces is the classic version of all surface rheology methods [29]. Harmonic waves are generated and the response of the system is measured in terms of a relative damping of the propagated wave [30]. Detailed descriptions of capillary wave techniques including specific theories are given in many reviews, the most recent one by Noskov et al. [31]. Recent work has been focused on theoretical modifications for this technique [32, 33, 34, 35, 36, 37] as well as on experimental improvements and new equipment [38, 39, 40, 41, 42, 43, 44, 45]. One of the more recently developed methods to investigate the surface relaxation of soluble adsorption layers due to harmonic disturbances is the oscillating bubble method. The technique involves the generation of radial oscillations of a gas bubble at the top of a capillary immersed into the solution under study. The first set-up was described by Lunkenheimer and Kretzschmar [46] and Wantke et al. [47] followed by new designed apparatuses using modern electronic technique to register the pressure changes inside bubbles or drops [48, 49,50,51,52,53]. The transport problem involved in isotropic harmonic area deformations of surface layers is based on the diffusional flux at the interface and reads
!d(rA) = D ac a t A
dt
dx
Here A(t) is the change of the surface area with time. The transport of the surfactant molecules in the bulk is given by Fick's diffusion law,
Interfacial Rheology of Adsorbed Layers
- =D - ^ .
65
(4)
For anisotropic area changes, as it is generally the case in Langmuir trough experiments, the lateral transport of adsorbed molecules has additionally to be considered [54]. The solution to the problem given by Eqs. (4) with the boundary condition (3) has the general form [55], c(x,t) - c0 +aexp(px)exp(icot),
(5)
where c0 is the surfactant concentration in the bulk. The boundary condition (3) can be rearranged to
dc/dX
i ^ I = -f1+ D dlnA
^
V
(6)
(dT/dc)(5c/9t)J
With the definition of the dilational elasticity modulus E in the form E _^dlnT
dlnTdlnA the following relationship results
E = — ^ f l +D ^ - * ^ V dlnl\
(8)
dr(dc/dt)J
Introduction of c(x,t) from Eq. (5) leads to an expression for E(ia>) [47],
E(i03) = E0
^ Vico +^/2co D
with
(9)
66
R. Miller and V.B. Fainerman
The parameter coD is the so-called characteristic frequency which is related to the diffusion relaxation time xd via
At liquid/liquid interfaces an additional complication arises because the surfactant is usually soluble in both adjacent phases. Thus, the exchange of matter takes place with both adjacent bulk phases. The result can be obtained in an analogous way [56]. For surfactant solutions above the critical micelle concentration, CMC, Lucassen [57] derived the respective solution. The method of capillary waves damping is a useful method only for small values of the dilational elasticity and at sufficiently high frequencies. There is, however, a second method, the longitudinal wave technique, which works at low frequencies and complement the capillary wave method. The theoiy for longitudinal waves was developed by van den Tempel and van de Riet [58], and Lucassen and van den Tempel [59]. The theory of pulsating bubbles in surfactant solutions is also well developed. Wantke et al. [47] were the first who derived a hydrodynamic equation for the pressure balance in the present system. Combined with a model which describes the exchange of matter at the oscillating bubble, equivalent to Eq. (9), the dilational elasticity and the exchange of matter mechanism can be obtained. The theory was later generalised by Johnson and Stebe [60] for relaxation mechanisms other than diffusion. A complete revision of the theory for the oscillating drop and bubble method was presented very recently by Kovalchuk et al. [52, 53, 61, 62]. In many experiments the data measured at high oscillation frequencies should reach a limiting value. In absence of any relaxation processes this limit is given by Eo = d n / d l n F , however, the measured values at higher surfactant concentrations are much smaller, even by some orders of magnitude. The same is true for the molecular exchange parameter, i.e. the characteristic frequency of diffusional relaxation coD. Also this parameter exceeds the experimental values by orders of magnitude [63, 64]. A recently proposed explanation for these discrepancies is an intrinsic compressibility of interfacial layers which is a consequence of finite repulsion forces acting between adsorbed molecules when
Interfacial Rheology of Adsorbed Layers
67
the monolayer is close to saturation [65, 66, 67]. The introduction of this intrinsic compressibility coefficient does not change the shape of the surface pressure isotherm significantly, however, the high frequency elasticities are much better described by this model and the resulting diffusion relaxation time is of the right order of magnitude. For capillary waves and a diffusion controlled adsorption a solution of the same physical problem was derived earlier by Lucassen and Van den Tempel [55, 59]. The real and imaginary parts of the complex surface elasticity modulus E(f,c) = E r (f,c)+iE ; (f,c) is a function of frequency f and the surfactant concentration c
E r (f,c) = E0
1+
^
7,Ei(f,c)
= E0
^
T
(12)
with C, = ^/coD /47tf . According to the definition of the parameters Eo and coD can be found from the surface equation of state and adsorption isotherm. For the Frumkin adsorption model
bc = ^ ^ e x p ( - 2 a 9 )
(13)
— = -ln(l-e)-a92 V RT '
(14)
we obtain [4]
0
and
RTY_e__2aQ^ Q(i-e J
68
R. Miller and V.B. Fainerman
Here the monolayer coverage is defined by 6 = FQ, R is the gas constant, T is the temperature, b is the adsorption constant, a is the (Frumkin) interaction constant, and Q. = \IYX is the molar area of the surfactant at saturation. In this model at saturation (9—»1) the values Eo and coD increase to infinity. As mentioned above, the experiments however show, that the high-frequency elasticities predicted for higher concentrations l/b
n = n o (i- e n),
(17)
where Q o is the molar area determined by extrapolation of the surface pressure FT to zero, and s is the two-dimensional intrinsic compressibility of the adsorbed molecules [68, 69, 70]. At high surface pressures, F « Ta we now do not get an infinitely high value for Eo but
E -^L~i dlnF
(18)
8
As discussed in [66] the generalisation of the equation of state for the Frumkin isotherm (13), Eq. (17) and the Gibbs adsorption equation yield nQ0(i^n/2)=ln(l_9)a92
(19)
while the two-dimensional solution concept for a changing partial molar area of the solute and solvent in Eq. (17) yields [71]
nQ0Mn)_ln(ie)ae2
(2Q)
Interfacial Rheology of Adsorbed Layers
69
Alternatively, for a two-dimensional solution we can assume a constant partial molar area of the solvent and a changing partial molar area of the solute, for which we obtain under the restriction sFI « 1: ^ RT
= -ln(l-e)-a62 v /
(21)
These three models were compared to experimental data and it turned out that they describe the experiments much better than any earlier models [66]. 3.1.2. Relaxations induced by transient interfacial disturbances Besides methods based on harmonic disturbances of the equilibrium adsorption layer to generate relaxation processes methods exist where the relaxation processes are induce by arbitrary area changes. Experimental methodologies are, for example, the Langmuir trough technique [54, 72], the elastic ring [73], or the modified pendent drop experiment [74, 75]. The whole theoretical treatment of the derivation of interfacial response functions has recently been summarised in [56]. As the result for a diffusion-controlled exchange of matter the following functions for a trapezoidal area change, the most general and advantageous transient perturbation function [28], are obtained: Ay, (t) = ^-[exp(2co D t)erfc(72co D t) -1]+ j g ^ _
AY2(t) = AY,(t)-AY 1 (t-t 1 ),
; at0
at t, < t < t2,
(22)
(23)
AY3(t) = A y 2 ( t ) - A Y 1 ( t - t 1 - t 2 ) J at t, + t2 < t < 2 t, + t2,
(24)
AY4(t) = AY 3 (t)-AY,(t-2t 1 -t 2 ),
(25)
, •
,
a t t > 2 t , + t2> • .
,,
Here, the relative area change is denoted by a =
dlnA
1 , ,, AA.
= —ln(l ). t, dt t, Ao and t2 are the characteristic times of the trapezoidal perturbation. In [56] the response functions also to other types of area deformations are summarised.
70
R. Miller and V.B. Fainerman
Thus, for a surfactant mixtures of m compounds, the response function Ay(t) for a square pulse of duration t| reads m
AA r Ay,(t) = E0—-X=Lexp(2{BiDt)erfc(72eoiDt) A i=i r T
at 0 < t < t,,
(26)
Ay2(t) = AY 1 (t)-AY,(t-t 2 )
att>t,,
(27)
with the following characteristic frequencies for each of the components
At the liquid/liquid interfaces the surface tension response function Ay(t) has the same form, except the characteristic frequency takes into account the peculiarity that the surfactant can be soluble in both adjacent phases, i.e. the distribution coefficient has to be introduced here [56]. 3.2.
Experimental methods There are many experimental techniques for studying interfacial relaxations of soluble adsorption layers. Except for wave damping techniques, these methods have been developed and only used by single research groups. Up to now, no commercial set-up exists and relaxation experiments are not therefore very common. New developments in this field, especially the transient methods, are simple to construct and to handle data and will probably increase the number of investigators studying the dynamic and mechanical properties of adsorption layers. In this section, wave damping and other harmonic methods as well as transient relaxation techniques will be described. 3.2.1. Relaxation methods for Harmonic Interfacial Disturbances As mentioned above, the classical relaxation techniques are the methods of wave damping. If a wave is generated by mechanical or other means, it propagates along the surface and is damped by hydrodynamic and surface mechanical properties [76]. The constructions differ for instruments to instrument essentially in the way how the waves are generated, and how their propagation and damping is determined [34, 77, 78]. A recent review summarises the main types of wave damping methods [79].
Interfacial Rheology of Adsorbed Layers
71
Fig. 1. Experimental set-up of the oscillating barrier method.
For low frequencies, the oscillating barrier method is suitable, as discussed for example in [80, 81, 82]. Due to hydrodynamic limitations of the method the frequency range is limited to less than one Hz, however, this type of relaxation experiments can be performed with any Langmuir balance, as shown in Fig. 1. For much higher frequencies the method of capillary wave damping was proposed [58, 83, 84]. This technique is suitable in a frequency range from 25 Hz to 4 kHz. There are various designs developed in the past, but so far there is no commercial instrument available for this wave technique. In the example given in Fig. 2, the capillary waves are excited by a sinusoidal voltage applied to a thin metallic blade placed above the liquid surface. The reference platinum electrode is immersed into the solution. The initial amplitude of the transverse waves is controlled by the distance between blade and surface. A laser beam is reflected from the liquid surface and directed to a position-sensitive photo detector. The alternating current caused by oscillations of the laser beam on the photo detector is amplified and compared in phase with the signal from the electric generator. Simultaneous measurements of the laser beam position and the corresponding changes of amplitude and phase of the electric signal allow the determination of the damping coefficient and the length of the capillary waves. The longitudinal wave damping technique complements the oscillating barrier and transversal wave damping methods and fills the frequency gap between them. The existence of this type of surface waves was predicted theoretically and also demonstrated experimentally by Lucassen [85, 86, 87].
72
R. Miller and V.B. Fainerman
Fig. 2. Wave damping setup with electro-capillary wave generator and optical detection.
Fig. 3 Longitudinal wave setup.
Interfacial Rheology of Adsorbed Layers
73
To generate longitudinal surface waves a Langmuir trough can be applied in almost the same way as in the oscillating barrier method, and only the construction of the barrier is different (cf. Fig. 3). Instead of a Teflon barrier we a Teflon frame with a metal wire inside is used. The wire with a diameter of 50 um is wetted by the liquid and moves parallel to the liquid surface to generate the longitudinal surface waves. The wave characteristics are determined by means of transverse surface waves excited perpendicular to the wire at a fixed frequency by the electro-capillary method [87]. Propagation of longitudinal waves leads to oscillations of the surface concentration and thereby to oscillations of the surface tension. In turn, this induces oscillations of the length of transverse waves. Beside the capillary wave techniques, methods based on oscillating drops and bubbles exist. The experiment designed by Lunkenheimer [46, 47] belongs with the first surface dilational elasticity measurements with this type of experiments. The principle of presently running oscillating bubble experiments is shown in Fig. 4 [88].A narrow capillary is immersed into the solution under study and a small air bubble is formed at its tip and generated to harmonic oscillations via a piezodrive. A sensor reads the pressure changes caused by the dilations and expansions of the adsorption layer. From the harmonic surface tension response the dilational elasticity and the exchange of matter can be calculated. The experimental limit for the oscillation frequency depends on many parameters of the measuring cell and capillary geometiy. However, one can say that oscillations up to 100 Hz are easily feasible. Wantke et al. [89] haven even shown that with a small liquid volume in the cell and a sufficiently narrow capillary, frequencies of 450 Hz can be reached.
Fig. 4. Oscillating bubble set-up.
74
R. Miller and V.B. Fainerman
Recently, the theoretical background, established first by Wantke et al. [47, 63, 64], was further elaborated and the new idea of an intrinsic compressibility of packed surface layers incorporated [65, 66, 67]. For relaxation experiments at slow oscillations, the bubble and drop profile analysis tensiometry is suitable. The first experiment of this type was performed by Miller et al. [74]. This method, originally developed in the group of Neumann [90, 91] to measure contact angles and surface tensions, can be successfully applied to relaxation studies by generating harmonic drop volume changes, which in turn induce respective compressions and expansions of the drop or bubble surface. In contrast to the oscillating spherical bubbles or drops, this experiment is limited to frequencies below 1 Hz. If oscillations of larger frequencies are generated the surface tension response can mimic a dilational rheology even for pure water [92]. 3.2.2. Relaxation methods for Transient Interfacial Disturbances The elastic circular ring method as one of the first possibilities for studies of surface relaxation processes in soluble adsorption layers caused by transient perturbations was developed by Loglio et al. [28, 73]. The main feature of the method is an elastic ring, which confines the surface area in the sample vessel and substitutes the traditional barrier on a Langmuir trough. By changing the shape of the ring, which is immersed into the solution, small area changes of the solution surface can be applied. The surface tension response after such deformations is registered via force measurements using a Wilhelmy plate. Step and ramp type, square pulse and trapezoidal area deformations are possible. The construction ensures that area changes are almost isotropic.
Fig. 5. Video enhanced pendent drop method modified for relaxation experiments.
Interfacial Rheology of Adsorbed Layers
75
The classical pendent drop method allows, as described above for slow harmonic oscillations, is advantageous for transient relaxation studies, because definite area changes of the drop surface area are easy performed and controlled [206, 207]. A dosing system consisting of an accurate syringe (cf. Fig. 5) is used to form a drop with a definite volume and change it by small increments according to the experimental protocol. In this way transient relaxation experiments with any area perturbation are feasible. Both oscillating drop methods can be applied to liquid/gas as well as liquid/liquid interfaces and are easily temperature controlled. The same is true for the drop pressure technique. A pressure transducer, in contact with the liquid that forms the drop, can register the change of the pressure inside the drop (or bubble) and therefore, the interfacial tension is determined as a function of time [93, 94, 48, 95] developed different drop pressure experiments. A typical set-up of a drop pressure experiment is shown in Fig. 6, designed for studies at a liquid/liquid interface. The same experiment is possible with an extra cell of a drop profile instrument. In this case, the video image can be used to determine the size of the drop precisely. The fast development of electronic devices, such as high accuracy pressure transducers, dosing systems, high speed computers, made this type of experiments a very powerful tool [96]. There are more transient relaxation methods, such as the one developed by Kokelaar et al. [97], or the various modifications of the overflowing cylinder discussed [98, 99], which however, will be less frequently applied than the commercially available methods.
Fig. 6. Principle of a drop pressure experiment according to Passarone et al. [94],
76
R. Miller and V.B. Fainerman
3.3.
Surface elasticity and viscosity of surfactant layers For diffusion-controlled adsorption the complex surface elasticity modulus can be predicted by the theory of Lucassen and van den Tempel [55, 59]. From Eq. (12) together with (10) we can obtain
(29)
(30)
Due to their physical meaning, the parameters given by Eqs. (29) and (30) should be independent of frequency, although both are functions of the frequency. This represents a simple criterion for applicability of the Lucassen/van den Tempel model [67]. In Figs. 7 and 8 the parameters dF/dc and Eo for dodecyl dimethyl phosphine oxide solutions are given as determined by the oscillating bubble method [64]. The parameters are almost constant over the frequency range between 10 and 400 Hz for the smallest concentrations, however, are not constant for higher concentrations. Thus, the diffusion mechanism derived by the Lucassen/van den Tempel is applicable only for the small concentrations of dodecyl dimethyl phosphine oxides. Wantke and Fruhner postulated a non-equilibrium between the monolayer and the subsurface (slow adsorption kinetic) as well as an intrinsic surface viscosity [64], in order to obtain a satisfactory agreement with experimental data for dodecyl dimethyl phosphine oxides at all concentrations. Above we have discussed the concept of an intrinsic compressibility, as given by the Eqs. (17) to (20). In Figs. (9) and (10) the experimental data from the same paper [64] are shown together with the theoretical curves calculated according to the Frumkin model with zero compressibility (curves 1) and with a final intrinsic compressibility (curves 2-4). The values of the model parameters are given in detail in [67]. It is seen from Figs. 9 that the Frumkin isotherm gives a slope for ln(dF/dc) as a function of ln(c) which is much larger than the experimental one The assumption of a two-dimensional compressibility gives slopes close to the experiment curves (2-4) and coincides much better with the experimental data. Equivalent results are obtained for the elasticity Eo in Fig. 8. Thus, by an appropriate choice of the parameter e the theoretical curves agree quite well with the experimental data, while in contrast the Frumkin isotherm is not in a good agreement with the experimental data.
Interfacial Rheology of Adsorbed Layers
11
Fig. 7. Eo determined for dodecyl dimethyl phosphine oxide solutions measured by the oscillating bubble method [64]: 3 10"5 M (A), 10"4 M ( • ) , 1.5 10'4 M (T), 2.5 10'4 M (O), 3 10~4 M ( • ) , lines are averaged values in the given frequency interval.
Fig. 8. d r / d c for dodecyl dimethyl phosphine oxide solutions measured by the oscillating bubble method [64], used symbols are equivalent to Fig. 7.
78
R. Miller and V.B. Fainerman
Fig. 9. Average values of Eo as functions of concentration for dodecyl dimethyl phosphine oxide solutions obtained from Fig. 7 (symbols), theoretical dependencies are calculated for £ = 0 (curve 1) and different compressibility E = 0.007 m/mN (curve 2), 0.01 m/mN (curve 3).
Fig. 10. Average values of dF/dc as functions of concentration for dodecyl dimethyl phosphine oxide solutions obtained from Fig. 7 (symbols), theoretical dependencies equivalent to Fig. 9.
Interfacial Rheology of Adsorbed Layers
79
3.4.
Surface elasticity and viscosity of protein layers For many applications, for example in food technology, adsorption layers of proteins and mixtures with surfactants are important. Knowledge of the rheological behaviour is important for example for the formation and stability of foams and emulsions [100, 101]. Mixtures of proteins and surfactants are systems which can be tailored such that respective properties can be established. It must be emphasised, however, that there are very few systematic studies on this subject so far. In a recent review Bos and van Vliet summarised studies on mixed protein/surfactant adsorption layers [102[. As examples, some measurements of a recent work by Kragel et al. [103] will be given here. The Figs. 11 and 12 show the frequency dependence of elasticity E and viscosity r|co = rj(27tf) = r\{2nlx) for mixed SDS/p-lactoglobulin ((3-LG) adsorption layers. The measurements were performed with the drop profile analysis technique (PAT1 from SINTERFACE Technologies, Berlin) as described in [ 104]. As one can see, the elasticity and the viscosity of the mixed layers are only slightly dependent on the deformation frequency f = 1/x. This small effect is due to the presence of proteins. With increasing SDS contents the absolute values of the rheological parameters drop down and approach zero, as it is expected for pure SDS solutions at the given concentrations and frequencies of the applied oscillations. This demonstrates the suitability of the rheological studies for understanding the composition of mixed adsorption layers.
Fig. 11. Dilational elasticity E of a 10 mol/1 |3-LG solution at the water/air interface mixed with different amounts of SDS: cSDS = 0 ( • ) , 10"4 mol/1 (O), 4.10"" mol/1 (A), 10'3 mol/1 (O).
80
R. Miller and V.B. Fainerman
Fig. 12. Dilational viscosity (r\ co) of a 10" mol/I P-LG solution at the water/air interface mixed with different amounts of SDS: CSDS = 0 ( • ) , 10'4mol/l (O), 2 10'4mol/l ( • ) , 4 10" 4 mol/l(A), 10" 3 mol/l(O).
4. STUDIES ON INTERFACIAL SHEAR RHEOLOGY Interfacial shear rheology of interfacial layers is a useful methodology for the discussion of the structure of adsorbed or spread layers at liquid/gas and liquid/liquid interfaces. Numerous experimental techniques have been employed for the measurement of the two main parameters, the interfacial shear viscosity and elasticity. Two different types of surface viscometers are typically used which are designed such that surface tension gradients during the generation of surface flow are negligible, the determination of surface velocity profiles (indirect methods) and the determination of torsion stress values (direct methods). In contrast to the dilational rheology, the shear rheological parameters of interfacial layers are in no direct relationship to their thermodynamic characteristics. 4.1.
Experimental methods There are quite a number of experimental set-ups for measuring the shear rheology, because the measured interfacial parameters are very valuable for many technological processes. We give here only a brief overview and discuss in some more detail the torsion surface shear rheometry, practicable to many users due to available commercial devices. There are some recent summaries of shear rheology studies [102, 105, 113].
Interfacial Rheology of Adsorbed Layers
81
4.1.1. Indirect surface shear experiments The surface flow methods are designed in a way analogous to the Hagen-Poiseuille law for measuring the bulk viscosity of liquids. It is based on the determination of the flow rate of a film through a narrow canal or slit under a two-dimensional pressure difference All. The first canal surface viscometers were already proposed in the Thirties [106, 107]. The surface viscosity r\s is calculated from the rate of film flow Q through the canal of width a and length L via the equation n =^ L . 12LQ
(29)
Harkins and Kirkwood [108] included a correction term for the drag of the underlying viscous liquid having a bulk viscosity r\ =
ATV_a n 12LQ n
This equation is valid only for narrow canals with smooth, parallel walls and under the assumptions that there is no slip of the film along the walls. The present canal surface viscometers, called deep-channel surface viscometers, go back to the work of Davies [109] and Mannheimer and Schechter [110, 111, 112]. The design of such a deep-channel surface viscometer consists of two concentric, stationary vertical cylinders and a rotating flat-bottomed dish containing the liquid (cf. [113]). The cylinders are placed such that only a small gap to the bottom is left. The dish rotates at a known angular velocity, shearing the fluid between the channel walls, and enhancing the effect on the interface. The centreline surface motion within the channel is monitored by a very small PTFE particle floating on the fluid interface, from which the interfacial shear viscosity i\s can be calculated [114]. There are more of these indirect methods which will not be further analysed here. We rather go into some more detail for the direct methods, for which some devises are commercially available. 4.1.2. Direct surface shear experiments The direct determination of surface shear viscosity is mainly performed by measuring the damping of an oscillating torque pendulum touching the interface is one of the oldest methods in surface rheology[115]. The first knifeedge surface viscometer of Brown et al. [116] consists of a knife-edge bob
82
R. Miller and V.B. Fainerman
suspended from a torsion wire such that the circular knife just touches the surface of a solution contained in a cylindrical vessel. The measuring vessel is forced to rotate, and the torsional stress on the knife-edge is measured in order to determine the surface shear viscosity. In the seventies Mannheimer & Burton [117] improved the theoretical analysis of the typical "knife edge" torsional viscometer. If a ring or disk is forced to oscillate in the interface, the rheological parameters are better defined, particularly when the interface is bounded by a concentric outer ring [118]. The equation of motion for free [119, 120] or forced oscillations [120] was solved. 4.1.3. AutomatedInterfacial Torsion Shear Rheometers An instrument for measuring the interfacial shear rheology at very low shear-rates was described by Kragel et al. [121]. This equipment does not disturb the structure during the measuring procedure because the deflection angle is very small, of the order of 1 to 5 degrees only. The principle of the rheometer is based on a ring with a sharp edge hanging at a torsion wire. When applying an impulsive torque by an instantaneous movement of the torsion head the pendulum performs damped oscillations with the damping factor (3 and the circular frequency co. Such studies provide simultaneously information on the surface shear coefficient of viscosity and the surface shear modulus of rigidity from a single experiment. The schematic set-up of the rheometer (ISR1 from SINTERFACE Technologies, Germany, is shown in Fig. 13. The active part (stepper motor, gearing, motor controller) comprises a drive for the deflection, at which the torsion wire with a circular measuring body is fixed. The body touches the liquid surface or is situated at the liquid/liquid interface. The angular position of the body is registered by means of a laser and a position-sensitive photo sensor with an accuracy of about ±0.01°. The whole instrument is computer-driven and the software controls all calibrations and the measurements and acquires and analyses the experimental data. The application of this torsion pendulum apparatus and its sensitivity and accuracy has been demonstrated by different model experiments [122, 27]. A second commercial instrument is the Camtel CIR 100 interfacial shear rheometer (Camtel, Royston, UK). It is an oscillating stress controlled instrument, where a virtually frictionless suspension drive, similar to a galvanometer movement, is used [123]. As measuring body a Pt/Ir ring is attached to this drive and a high resolution sensor monitors the displacement of the ring. The ring imparts an oscillation stress to the sample and interfacial viscosity and elasticity are automatically computed from amplitude ratios and phase angle. The instrument can work in a normalised resonance mode and in a controlled stress mode.
Interfacial Rheology of Adsorbed Layers
83
Fig. 13 Automated interfacial shear rheometer ISR1
The former mode is used for measurements near or above the natural frequency of the instrument (> 2Hz) while the controlled stress mode allows measurements in the frequency range between 0.001 - 20 Hz [124]. 4.2.
Examples of interfacial adsorption layers Surface shear studies of interfacial layers modified by usual surfactants have an extremely low shear rheology so that only very special, highly sensitive methods are suitable, such as designed by Petkov et al. [125]. For insoluble monolayers, the shear viscosity is significantly changing at the end of the coexistence region. Beyond this region the shear viscosity increases steeply [126]. For example, DMPE monolayers show an abrupt increase to very high viscosity values after the transition point from the liquid condensed (LC) to the solid-like state at a surface pressure of about n = 21mN/m (cf. Fig. 14). For DPPC a similar behaviour is observed. Although a transition from the LC-phase to a solid-like film at about IT = 32 mN/m. The comparison of the shear rheology results for DPPC with the dilational rheological behaviour of the same monolayer [ 126] has shown that both shear and dilational rheology correlate with the morphology of the monolayer as observed by optical methods, such as Brewster angle microscopy [127, 128].
84
Fig. 14
R. Miller and V.B. Fainerman
Surface shear viscosity r|s of spread monolayers of DPPC (•) and DMPE (•) as function of surface pressure IT [126] Protein interfacial layers are the most frequently studied systems in shear rheology. Extensive studies were published by Dickinson and co-workers [129, 130, 131]. Mackie et al. [132] performed recent studies of the adsorption behaviour of pure and mixed P-casein and P-lactoglobulin at the air/water interface using shear and dilatational interfacial rheological techniques. The shear data are obtained using the CIR 100 instrument in the normalized resonance mode. The obtained data are used to explain the different adsorption behaviour of the two proteins. The adsorption layer formed by P-casein (luM) was so weak as to be almost immeasurable while the layer formed by P-lactoglobulin (luM) gave quite high values of the shear elasticity and rose to a maximum of about 5 mN/m at 50 minutes and then decreased (cf. Fig. 15). A 1:1 mixture of the two proteins at a total protein concentration of 1 uM shows a gradual slow rise to 1 mN/m after 3 hours. The interfacial dilatational elasticity shows a similar trend. Extensive studies of shear viscosity of proteins were performed recently by Martin et al. [133] using a torsion shear rheometer with rotating dish. Fig. 16 gives the shear elasticity of ovalbumin and p-lactoglobulin. The two studied proteins show different behaviour which is also reflected by the strain that can be applied to the protein layer before it fractures and by the way a steady state stress is reached.
Interfacial Rheology of Adsorbed Layers
Fig. 15 Shear elasticity Gs as a function of time for /?-casein (1), /?-lactoglobulin (2), and a mixture (3) each at a total protein concentration of 10" mol/1 [132]
Fig. 16 Apparent surface shear viscosity (mN s/m) of ovalbumin (0.1 g/L, • ) and P-lactoglobulin (0.1 g/L, O) measured at pH 6.7 [133]
85
86
R. Miller and V.B. Fainerman
The authors discuss that the fracture strain of an adsorbed ovalbumin layer is much larger than that for (3-lactoglobulin. The differences induce that the mechanical properties of an interfacial layer vary from protein to protein. The shear elasticity appears to be less suitable to characterize adsorbed protein layers rather than full stress-strain curves. 5. SUMMARY The present manuscript gives an overview of the 2D-rheology of interfacial layers at liquid/fluid interfaces. Some recent theoretical models are discussed as well as the experimental methodologies available as laboratory set-ups and commercial instruments. We discuss here also experimental examples for various systems, containing surfactants, proteins and mixed systems. The history of surface rheology is very young. Only in the sixties the Dutch school around van den Tempel und Lucassen [87] introduced this group of parameters. Since then intensive work was made in order to understand the various phenomena observed in experiments. However, there are still a large number of open questions even for single surfactant solutions and further systematic work is required. REFERENCES [I] [2]
[3] [4]
[5] [6] [7] [8]
[9] [10] [II]
E. Dickinson, J.A. Hunt and D.G. Dalgleish, Food Hydrocolloids, 4 (1991) 403 E. Dickinson, in C. Gallegos (Ed.) Interactions in Protein-Stabilized Emulsions. Progress and Trends in Rheology IV; Proc of the Fourth European Rheology Conference, Sevilla, (1994) 227 D.C. Clark, P.J. Wilde, D.R. Wilson and R. Wiistneck, Food Hydrocolloids, 6 (1992) 173. Food Emulsions and Foams: Interfaces, Interactions and Stability, E. Dickinson and J.M. Rodriguez Patino (Eds.), Special Publication No. 227, Royal Society of Chemistry, 1999 Food Colloids 2000 - Fundamentals of Formulation, E. Dickinson and R. Miller (Eds.), Special Publication No. 258, Royal Society of Chemistry 2001 Food Colloids, Biopolymers and Materials", E. Dickinson and T. van Vliet (Eds.), Special Publ. No. 284, Royal Society of Chemistry, 2003 H. Fruhner, K.-D. Wantke, J. Kragel and G. Kretzschmar, J. Inf. Rec. Mats. 22 (1994)29 S.S. Dukhin, G. Kretzschmar and R. Miller, Dynamics of Adsorption at Liquid Interfaces, in D. Mobius and R. Miller (Eds.) , Studies of Interface Science, Vol. 1, Elsevier Publishers, Amsterdam, 1995 R.B. Dorshow and R.L. Swofford, J. Appl. Phys. 65 (1989) 3756, Colloids & Surfaces, 43(1990) 133 A.R. Kovscek, H.Wong and C.J. Radke, AIChE Journal, 39(1993) 1072 G. Loglio, U. Tesei, G. Mori, R. Cini and F. Pantani, Nuovo Cimento 8 (1985) 704
Interfacial Rheology of Adsorbed Layers
[12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46]
87
G. Loglio, N. Degli-Innocenti, U. Tesei, A.M. Stortini and R. Cini, Ann. Chim. (Rome), 79(1989)571 C. Oppo, S. Bellandi, N. Degli Innocenti, A.M. Stortini, G. Loglio, E. Schiavuta and R. Cini, Marine Chemistry. 63(3-4):235-253, 1999 P.R. Garrett and P. Joos, J. Chem. Soc. Faraday Trans. 1, 69 (1976) 2161 E.H. Lucassen-Reynders and K.A. Kuijpers, Colloids Surfaces, 65 (1992) 175 B.S. Murray, B. Cattin, E. Schuler and Z.O. Sonmez, Langmuir, 18(2002)9476-9484. D.T. Wasan, A.D. Nikolov, L.A. Lobo, K. Koczo and D.A. Edwards, Progr. Surface Sci., 39(1992) 119 D.A. Edwards, H. Brenner and D.T. Wasan, Interfacial Transport Processes and Rheology, Butterworth-Heineman Publishers, 1991 K. Malysa, Adv. Colloid Interface Sci., 40 (1992) 37 P.R. Garrett and P.R. Moore, J. Colloid Interface Sci., 159 (1993) 214 D. Langevin, Adv Colloid Interface Sci., 88(2000)209 P.J. Wilde, Current Opinion in Colloid Interface Science. 5(2000)1 76. C. Stubenrauch and R. Miller, J. Phys. Chem., in press J. Boussinesq, Ann. Chim. phys. (Ser.) , 29 (1913) 349 J. Boussinesq, Acad. Sci., Paris, 156(1913) 1124 V. V. Krotov, A. I. Rusanov, Physicochemical Hydrodynamics of Capillary Systems, Imperial College Press, 1999 R. Miller, R. Wustneck, J. Kragel and G. Kretzschmar, Colloids Surfaces A, 111(1996)75 G. Loglio, U. Tesei, R. Miller and R. Cini, Colloids Surfaces, 61 (1991) 219 M. van den Tempel and E.H. Lucassen-Reynders, Adv. Colloid Interface Sci., 18 (1983)281 E.H. Lucassen-Reynders, J. Lucassen, P.R. Garrett, D. Giles and F. Hollway, E.D. Goddard (Ed.), in Adv. in Chemical Series, 144 (1975) 272 B.A. Noskov, A.V. Akentiev, A. Yu. Bilibin, I.M. Zorin, R. Miller, Adv. Colloid Interface Sci., 104 (2003) 245 J.C. Earnshaw and CJ. Hughes, Langmuir, 7 (1991) 2419 M. Hennenberg, X.-L. Chu, A. Sanfeld and M.G. Velarde, J. Colloid Interface Sci., 150(1992)7 Q. Jiang, Y.C. Chiew and J.E. Valentini, Langmuir, 8 (1992) 2747 B.A. Noskov, G. Loglio, Colloids Surf. A 143 (1998) 167. D.M.A. Buzza, J.L. Jones, T.C.B. McLeish, R.W. Richards, J. Chem. Phys. 109 (1998)5008 M. Hennenberg, S. Slavtchev, B. Weyssow, J.-C. Legros, J. Colloid Interface Sci., 230(2000)216. K. Sakai, H. Kikuchi and K. Takagi, Rev. Sci. Instrum., 63 (1992) 5377 J.C. Earnshaw, E. McCoo, Langmuir, 11 (1995) 1087 B.A. Noskov, D.O. Grigoriev and R. Miller, Langmuir, 13(1997)295 F. Monroy, F. Ortega, R.G. Rubio, Phys. Rev. E 58 (1998) 7629 B. A. Noskov, A. V. Akentiev, G. Loglio and R. Miller, J. Phys. Chem., 104 (2000) 7923 F. Monroy, S. Rivillon, F. Ortega, R.G. Rubio, J. Chem. Phys. 115 (2001) 530 F. Monroy, M.G. Munoz, J.E.F. Rubio, F. Ortega, R.G. Rubio, J. Phys. Chem. B 106 (2002) 5636 B.A. Noskov, A.V. Akentiev and R. Miller, J. Colloid Interface Sci., 255 (2002) 417 K. Lunkenheimer and G. Kretzschmar, Z. Phys. Chem. (Leipzig), 256 (1975) 593
88
[47] [48] [49] [50] [51] [52]
[53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75]
[76] [77] [78]
R. Miller and V.B. Fainerman
K.-D. Wantke, R. Miller and K. Lunkenheimer, Z. Phys. Chem. (Leipzig), 261 (1980) 1177 C.A. MacLeod and C.J. Radke, J. Colloid Interface Sci., 160 (1993) 435 C.H. Chang and E.I. Franses, J. Colloid Interface Sci., 164 (1994) 107 Fruhner H, Wantke KD. Colloid Polymer Sci., 274(1996)576 K.D. Wantke and H. Fruhner, J. Colloid Interface Sci., 237(2001)185 V.I. Kovalchuk, J. Kragel, E.V. Aksenenko, G. Loglio and L. Liggieri, Oscillating bubble and drop techniques, in "Novel Methods to Study Interfacial Layers", Studies in Interface Science, Vol. 11, D. Mobius and R. Miller (Eds.), Elsevier, Amsterdam, 2001, p. 485 V.I. Kovalchuk, J. Kragel, A.V. Makievski, G. Loglio, F. Ravera, L. Liggieri and R. Miller, J. Colloid Interface Sci., 252 (2002) 433 D.S. Dimitrov, I. Panaiotov, P. Richmond and L. Ter-Minassian-Saraga, J. Colloid Interface Sci., 65 (1978) 483 Lucassen J. and M. van den Tempel, Chem. Eng. Sci., 27 (1972) 1283 R. Miller, G. Loglio; U. Tesei, and K.-H. Schano, Adv. Colloid Interface Sci., 37 (1991)73 J. Lucassen, Faraday Discussion Chem. Soc, 59 (1976) 76 M. van den Tempel and R.P. van de Riet, J. Chem. Phys., 42(1965) 2769 J. Lucassen and M. van den Tempel, J. Colloid Interface Sci., 41(1972)491 D.O. Johnson and K.J. Stebe, J. Colloid Interface Sci., 168 (1994) 21 V.I. Kovalchuk, E.K. Zholkovskij, J. Kragel, R. Miller, V.B. Fainerman, R. Wustneck, G. Loglio and S.S. Dukhin, J. Colloid Interface Sci. 224(2000)245 V.I. Kovalchuk, J. Kragel, R. Miller, V.B. Fainerman, N.M. Kovalchuk, E.K. Zholkovskij, R. Wustneck, and S.S. Dukhin, J. Colloid Interface Sci., 232(2000) 25 K-D. Wantke, H. Fruhner, J. Fang and K. Lunkenheimer, J. Colloid Interface Sci., 208(1998)34. K-D. Wantke and H. Fruhner, J. Colloid Interface Sci., 237(2001)185 V.B. Fainerman, R. Miller and V.I. Kovalchuk, Langmuir, 18 (2002) 7748 V.B. Fainerman, R. Miller and V.I. Kovalchuk, J. Phys. Chem. 107 (2003) 6119 V.I. Kovalchuk, G. Loglio, V.B. Fainerman and R. Miller, J. Colloid Interface Sci., 270 (2004) 475 U. Gehlert, D. Vollhardt, G. Brezesinski and H. Mohwald, Langmuir, 12(1996)4892 G. Brezesinski, E. Scalas, B. Struth, H. Mohwald, F. Bringezu, U. Gehlert, G. Weidemann and D. Vollhardt, J. Phys. Chem., 99(1995)8758. U. Gehlert and D. Vollhardt, Langmuir, 1 8(2002)688 V.B. Fainerman, E.H. Lucassen-Reynders, Adv. Colloid Interface Sci., 96 (2002) 295 G. Kretzschmar and K. Konig, J. Signal AM, 9 (1981) 203 G. Loglio, U. Tesei, and R. Cini, Rev. Sci. Instrum., 59 (1988) 2045 R. Miller, R. Sedev, K.-H. Schano, C. Ng and A.W. Neumann, Colloids &Surfaces, 69(1993)209 G. Loglio, P. Pandolfini, R. Miller, A.V. Makievski, F. Ravera, M. Ferrari and L. Liggieri, in "Novel Methods to Study Interfacial Layers", Studies in Interface Science, Vol. 11,D. Mobius and R. Miller (Eds.), Elsevier, Amsterdam, 2001, p. 439 F.C. Goodrich, J. Phys. Chem., 66 (1962) 1858 T. Yasunaga and M. Sasaki, W.J. Gettins and E. Wyn-Jones (Eds.) in Techniques and Application of Fast Reactions in Solutions, D. Reidel Publ. Comp., (1979) 579 D. Wielebinski and G.H. Findenegg, Progr. Colloid Polymer Sci., 77 (1988) 100
Interfacial Rheology of Adsorbed Layers
[79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91]
[92] [93] [94] [95] [96]
[97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] [113] [114] [115]
89
B.A. Noskov, A.V. Akentiev, A. Yu. Bilibin, I.M. Zorin, R. Miller, Adv. Colloid Interface Sci., 104 (2003) 245 J. Lucassen and D. Giles, J. Chem. Soc. Faraday Trans. 1, 71 (1975) 217 J. Lucassen and G.T. Barnes, J. Chem. Soc. Faraday Trans. 1, 68 (1972) 2129. G. Kretzschmar, Progr. Colloid Polymer Sci., 77 (1988) 72 R.S. Hansen and J.A.Mann, J. Appl. Phys., 35 (1964) 152 E.H. Lucassen-Reynders, J. Lucassen, Adv. Colloid Interface Sci. 2 (1969) 347. J. Lucassen, Trans. Faraday Soc, 64 (1968) 2221 J. Lucassen, Trans. Faraday Soc, 64 (1968) 2230 B. Carroll, Adv. Colloid Interface Sci., 107 (2004) 1 H.Fruhner and K.Dd. Wantke, Colloids Surfaces A, 114 (1996) 53 K.-D. Wantke, H. Fruhner, J. Ortegren, Colloids Surfaces A, 221 (2003) 185 P. Cheng, D. Li, L. Boruvka, Y. Rotenberg and A.W. Neumann, Colloids & Surfaces, 43(1990) 151 P. Chen, D.Y. Kwok, R.M. Prokop, O.I. del Rio, S.S. Susnar and A.W. Neumann, in "Studies in Interface Science", D. Mobius and R. Miller (Eds.), Vol. 6, Elsevier, Amsterdam, 1998, pp. 61 M.E. Leser, S. Acquistapace, A. Cagna, A.V. Makievski and R. Miller, poster presented at the ECIS, Florence, 2003 L. Liggieri, F. Ravera and A. Passerone, J. Colloid Interface Sci., 140 (1990) 436 A. Passerone, L. Liggieri, N. Rando, F. Ravera and E. Ricci, J. Colloid Interface Sci., 146(1991) 152 R. Nagarajan and D.T. Wasan, J. Colloid Interface Sci., 159 (1993) 164 L. Liggieri and F. Ravera, monograph in "Drops and Bubbles in Interfacial Science", in "Studies in Interface Science", D. Mobius and R. Miller (Eds.), Vol. 6, Elsevier, Amsterdam, 1998, p. 239 A.J.J. Kokelaar, A. Prins and M. de Gee, J. Colloid Interface Sci., 146 (1991) 507 D.J.M. Bergink-Martens, H.J. Bos, A. Prins and B.C. Schulte, J. Colloid Interface Sci., 138(1990) 1 A. Prins, Chem.-Ing.-Tech., 64 (1992) 73 D. Langevin, Adv. Colloid Interface Sci., 88(2000)209 P.J. Wilde, Current Opinion in Colloid Interface Sci., 5 (2000) 176 M.A. Bos, T. van Vliet, Adv. Colloid Interface Sci., 91 2001 437 J. Kragel, M. O'Neill, A.V. Makievski M. Michel, M.E. Leser and R. Miller, Colloids Surfaces B, 31 (2003) 107 A.V. Makievski, M. O'Neill und R. Miller, Labor Praxis, 25 (2001) 66 R. Miller, V.B. Fainerman, J. Kragel and G. Loglio, Current Opinion in Colloid Interface Sci., 2(1997)578 R.J. Myers and W.D. Harkins, J. Chem. Phys., 5 (1937) 601 D.G. Dervichian and M. Joly, J. Phys. Radium, 10 (1939) 375 W.D. Harkins and J.G. Kirkwood, J. Chem. Phys., 6 (1938) 53 J.T. Davies, Proc 2nd Int. Congr. Surf. Act, 1 (1957) 220 R.J. Mannheimer and R.S. Schechter, J. Colloid Interface Sci., 32 (1970) 195 R.J. Mannheimer and R.S. Schechter, J. Colloid Interface Sci., 32 (1970) 212 R.J. Mannheimer and R.S. Schechter, J. Colloid Interface Sci., 32 (1970) 225 R. Miller, R. Wustneck, J. Kragel and G. Kretzschmar, Colloids Surfaces A, 111(1996)75 D.T. Wasan, L. Gupta and M.K. Vora, AIChE J., 17(1971) 1287 I. Langmuir, Science, 84 (1936) 378.
90
[116] [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] [127] [128] [129] [130] [131] [132] [133]
R. Miller and V.B. Fainerman
A.G. Brown, W.C. Thuman and J.W. McBain, J. Colloid Sci., 8 (1953) 491 R.J. Mannheimer and R.A. Burton, J. Colloid Interface Sci., 32 (1970) 73 J. Ross, J.. Phys. Chem., 62(1957)531 M. Joly, Kollloidzschr. 89 (1939) 26 N.W. Tschoegl, Kolloid-Z., 181 (1962) 19 J. Kragel, S. Siegel, R. Miller, M. Born and K.-H. Schano, Colloids Surfaces A, 91 (1994) 169 J. Kragel, R. Wiistneck, D.C. Clark, P.J. Wilde and R. Miller, Colloids Surfaces A, 98(1995) 127 C. Moules, Principles of Interfacial Rheology & Measurement with CIR 100; Camtel Ltd, Royston, UK, 1998. M. Sherriff and B. Warburton, Polymer 1 5(1974)253 J.T. Petkov, K.D. Danov andN.D. Denkov, Langmuir, 12(1996)2650 J. Kragel, G. Kretzschmar, J.B. Li, G. Loglio, R. Miller and H. Mohwald, Thin Solid Films, 284 285(1996)361 G. Weidemann and D. Vollhardt, Colloids & Surfaces A, 100(1995)187 G. Weidemann and D. Vollhardt, Thin Solid Films, 264(1995)94 E. Dickinson, B.S. Murray and G. Stainsby, J. Colloid Interface Sci., 106 (1985) 259 E. Dickinson, B.S. Murray and G. Stainsby, J.Chem. Soc. Faraday Trans., 84 (1988) 871 E. Dickinson, S.E. Rolfe and D.G. Dalgleish, Int. J. Biol. Macromol., 12 (1990) 189 A.R. Mackie,A.P. Gunning, M.J. Ridout, P.J., Wilde and V.J. Morris, Langmuir, 17(2001)6593 A. Martin, M. Bos, M.C. Stuart and T. van Vliet, Langmuir, 18(2002)1238
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 4
Electric properties of oil/water interfaces A. G. Volkov" and V. S. Markin" department of Chemistry, Oakwood College, 7000 Adventist Blvd., Huntsville, AL 35896, USA b
Department of Anesthesiology, University of Texas, Southwestern Medical Center, Dallas, TX 75390-9068, USA 1. INTERFACIAL POTENTIALS 1.1. Boundary potential difference An electric potential difference is established at the oil/water interface, which is called the boundary potential difference or, briefly, the interfacial potential [1-4]. Discussions of the nature of the potential drop at the interface between two immiscible electrolyte solutions (ITIES) dates back to the beginning of the century. Interfacial potentials were first classified by Lange and Miscenko [5].
Fig. 1. The interfacial potentials at the interface between two immiscible liquids a and p, as well as between these phases and a vacuum.
92
A.G. Volkov and V.S. Markin
The potential \\ia produced by the surface charge of phase a and measured in vacuum in the vicinity of the surface is called the external potential or the Volta potential. The potential tya produced in the bulk phase a and measured from infinity in a vacuum is called the internal potential or the Galvani potential. These two potentials differ by the surface potential %a. Table 1 illustrates the thermodynamic feasibility of experimental measurement of boundary potentials. The difference of Galvani potentials Aa$ = (j)a - ^ is of primary interest in the study of the interface between two phases a and p. This difference is determined by the distribution of charged and dipolar particles near the interface. The dipole component of the interfacial potential difference Aap(j>j is sometimes considered as a separate term to obtain the quantity Aap<|) - Aap<|>j caused by the distribution of ions. This separation is useful in cases where the dipole potential drop does not depend on the total interfacial potential difference produced by ion distribution. One may, therefore, assume that changes in the interfacial potential difference are determined only by the ionic contribution. However, it is uncertain if the interfacial potential difference can be separated into two independent components caused by the ionic and dipole contributions so that interpretation of results is to some extent approximate. Table 1 Thermodynamic feasibility of measuring interfacial potentials in immiscible liquid phases a andp Boundary potentials 1. External potential of one phase \\J 2. Volta potential Aa^\\i 3. Surface potential of one of the liquids % 4. Difference between surface potentials
Feasibility of measurement can be measured can be measured cannot be measured cannot be measured, but can be determined when an additional non-thermodynamic hypothesis is used can be measured
5. Change of Aapx potential during any process (adsorption, variation of temperature, pressure, concentration, etc.) cannot be measured 6. Internal potential § 7. Galvani potential, or difference between cannot be measured two inner potentials Aa^ can be measured 8. Change of Galvani potential Aap4> during any process (adsorption, variation of temperature, pressure, concentration, etc.) can be measured 9. Potential difference in the cell between two phases of the same composition E=2AaB(|)i
Electric Properties of Oil/Water Interfaces
93
The equilibrium potentials at a reversible or non-polarizable interface are classified into three types depending on which species pass between phases: 1. Potentials which arise in an equilibrium system where all the ions are capable of passing from one phase to another are called distribution potentials. 2. If the phases in contact contain ions, some of which are capable of crossing the boundary while others are not, Donnan potentials or membrane potentials are generated. The transfer of ions may be prevented by a semipermeable membrane or by chemical adsorption to ion exchangers. If ions of only one kind can cross the interface, then the membrane potential difference is called the Nernst potential. If two immiscible electrolyte solutions exchange electrons and are in equilibrium with respect to electrons, then a redox potential is established at the interface. Interfacial potentials at polarizable boundaries can be produced by adsorption of charged or dipolar species at the interface or by the charging of the interface from an external source. In an ideally polarizable interface, the potential is an additional independent variable characterizing the state of the boundary and is specified by the external potential source. Electrostatic rather than electrochemical equilibrium is maintained at an ideally polarizable interface, analogous to a charged capacitor. 1.2. Standard potentials and standard Gibbs free energies Suppose that phases a and (3 are in contact and contain ions i\ with charge numbers zj and activity a{. It follows from the condition of electrochemical equilibrium that
ztF
ztF
a\
where p. is the chemical potential and F = 96500 C mol"'. The difference between the standard chemical potentials of the ion / in the two phases ^-^=A;;^(O
(2)
represents the energy of resolvation or the standard Gibbs free energy AapG°tr(/) of the transfer of the ion / from the phase a to the phase (3. According to the standard electrochemical nomenclature the distribution or partition coefficient of the ion / is:
94
A.G. Volkov and V.S. Markin
r»=apz*m
(3)
Kl
and the standard potential of the distribution of a specific ion between two immiscible liquid phases is:
. = _*&»S*L z,F
z,F
According to Eq. (3), it is evident that p«/n=_L_
(5)
The interfacial potential difference can be represented in the form
AJN> = A £ < | > ? + ^ l n ^ = ^ l i i P?*^ ztF
a,
z,F
.
(6)
a,
The equilibrium ratios of the activities of the ions i in two phases is described by the expression
f = ^ f =exP[-§(A;> - A;;^)]= ?r « P ( - § ^ ) ,
(7)
where y; is the activity coefficient and c is the concentration. The standard distribution potential of a given ion represents the interfacial potential difference which would be established at the interface if the equilibrium activities of this ion in the two phases were identical. The partition coefficient of the ions Pa/Pj corresponds to the equilibrium ratio of the activities of the ions in the two phases in the absence of an interfacial potential difference. Values of A^" of interest for the study of ITIES have been compiled for certain solvent systems (Figure 2), and representative values are listed in Table 2. If volumes of phases a and (3 are equal, the electroneutrality condition can be written as
5>,c, = 0-
(8)
Electric Properties of Oil/Water Interfaces
95
The relation between the interfacial potential Aap(|) and the concentration of ions in solution is described by Eqns. (7) and (8). If there are n types of ions in the system, then the number of equations is n + 2 and the number of variables is In + 1, so that the number of independent variables in the system is n - 1. In other words, we can specify the concentration of all species of ions, except one, in one phase. Then the concentration of the remaining species of ions in that phase, the concentrations of all ions in the other phase, and the interfacial potential are automatically determined. Apart from the activities of individual ions, it is convenient to use the concept of the mean activity of a 1:1 electrolyte BA: a
BA = 4aBaA
•
(9)
The mean activity coefficient of the electrolyte is then given by IBA = ^IBIA
•
0°)
The partition coefficient of 1:1 electrolyte can be introduced by analogy: r
BA
~~ V
B
A
VL i I
It is easy to see that the electrolyte partition coefficient P^0 is the ratio of the mean activities of this electrolyte in the bulk phase:
p;i" = a;jdlA.
d2)
This ratio, in contrast to the analogous ratio for individual ions (Eq. 7), is independent of the interfacial potential. It is also independent of the presence of other solute ions. 1.3. Distribution potentials Consider the simplest case where the interface is non-polarizable and both phases a and P contain a binary 1:1 electrolyte B+A": a,
B+A" | B+A", (3
(13)
The interfacial potential can be readily found from Eq. (6) in cojunction with the electroneutrality condition (8):
96
A.G. Volkov and V.S. Markin
Fig. 2. The standard ion distribution potentials at the nitrobenzene/water (a), 1,2dichloroethane/water (b), and dichloromethane/water (c) interfaces versus the ion radius. Solid lines were calculated by Markin and Volkov [2].
97
Electric Properties of Oil/Water Interfaces
Table 2 Standard ion transfer potentials A o
Nitrojenzene H+ 337 T • + 398 358 Na* K+ 252 Rb* 206 Cs + 161 289 Ag + Tl+ 176 Mg 2+ 370 Ca 2+ 354 Sr+ 348 Ba 2+ 328 284 NH 4 + acetylcholine + 49 choline 80 Me 4 N + 30 EUN* -67 Pr 4 N + -170 9 Me 4 P + Me 2 V 2+ -15 Et 2 V 2+ -43 Pr 2 V 2+ -58 -250 Bu 4 N + Pe 4 N + -408 Hex 4 N + -472 -372 Ph 4 As + -454 Cl" -316 Br" -295 -195 r CN" -393 -253 NO3" Pi' 47 -83 C1O4" -75 IO4-121 BF4" 2,4-DNPh" -76 372 Ph4B" -270 CIO3" BrO 3 ' -340 NO," Cr,O 7 2 " MnO 4 " -16 C1O2" -350 -189 SOT
1,2-Dichloroethane
1,1-Dichloroethane
Oichloromethane
580 493
Chloroform Propylene carbonate 518 247 154 55 -10 -73 195 114
490 477 445,246 364 459 269
299 307 299
182 44 -91
181 112 -23
195 44
-114 -134 -228
-226 -360 -494 -364 -581 -481 -400 -274
-121
-230 -377 -455
-321
-330 -60 -178 -170 -190 364 -330 -390 -410 -320 -410 -264
-283
-440 -317
-481 -408 -273
-330 -210
-373 -580 -412 -310 -142 -373
-290 -69 -221
-190 -190 -240
-62 -31
373 -310 -370 -370 -310 -370 -73
98
A.G. Volkov and V.S. Markin
A;tf+A^ +RTlti±C 2
2F
(14)
The first term reflects the individual characteristics of the solute ions in the standard state, while the second depends on the activity coefficients and concentration effects. However, in practice, the activity coefficients of cations and anions can be assumed to be approximately equal, y^ ~ yaB and y^ «y p s , so that a simplified expression for the interfacial potential can be written: A"^0 + Aa(K° 2
J?T
Pa/P
IF
Pa/P '
In this case the interfacial potential difference is equal to the arithmetic mean of the standard distribution potentials of individual ions. The potential difference between two phases with a binary electrolyte in both phases is called the distribution potential. In this approximation the distribution potential does not depend on the concentration of the electrolyte BA or the interface structure, but is determined only by the thermodynamic properties of the bulk phases. This has been confirmed by experimental studies, although under certain conditions deviations have been observed. At low concentrations of the salt BA, the interfacial potential measurements can be affected by HCO3" if they are performed in air rather than in an inert gas. Distribution potentials also depend on the concentration of ionic solutes when more than two ionic species are present. The structure of the interphase between two immiscible electrolyte solutions is determined by the bulk properties of the contacting phases. When a monolayer of surfactant that is practically insoluble in both phases is adsorbed at the interface, the surfactant should not alter the interfacial potential determined from Eq. (14). However, due to surface blocking the thermodynamic equilibrium can be affected; and as a result the measured potential difference will be a nonequilibrium quantity that differs from the true distribution potential. One may consider the general case of an asymmetric electrolyte Bk*A'~ partitioning between phases a and |3:
a,Bk;4~ |^ + 4'-,p
(16)
In this case the interfacial potential determined by the distribution of electrolytes is described by the expression
Electric Properties of Oil/Water Interfaces
99
*AB.
Fig. 3. The distribution potential of AB2 electrolyte depends on concentration in the nonaqueous phase.
vw-,A-tf+
RT ^
(17)
For an asymmetric electrolyte, the partition potential depends on the salt concentration. The second term on the right-hand side of Eq. (17) is given by the following expression according to the Debye-Huckel theory:
where Ja is the ionic strength of the solution in the phase a, N a is Avogadro's number, e0 is the electric constant, and ea is the dielectric constant of the phase a. In terms of the Debye approximation, the distribution potential is independent of concentration only for a symmetrical electrolyte, where zA = -ZBIn Figure 3 the dependence of the distribution potential on concentration for a 2:1 electrolyte, for example MgCl2, is presented. For this illustration we
100
A.G. Volkov and V.S. Markin
chose s a = 10 and sp = 80, which corresponds to the common experimental system of water/1,2-dichloroethane. In this situation a hydrophilic electrolyte MgCl2 is dissolved in a two-phase system so that the term Jp/28p/2 can be neglected. As one can see from Fig. 3, the distribution potential of a nonsymmetrical electrolyte strongly depends on concentration. 1.4. Incomplete dissociation of the salt We will next describe how incomplete dissociation of the electrolyte can affect the interfacial potential distribution (Fig. 4). Suppose that the reaction B+ + A'
o
BA
(19)
occurs in phase a, where KgA denotes the dissociation constant of the BA molecule. This means that in equilibrium the equality aaBa°
= K"BAaaBA
(20)
is fulfilled. The dissociation constant KgA is related to the chemical potentials of the corresponding species by the equation
RTinKaBA=^-vT+-vT-,
(21)
Fig. 4. The partition of incompletely dissociated binary electrolyte A'B in a system of two immiscible liquids.
Electric Properties of Oil/Water Interfaces
101
Similar equations can be written for dissociation of the electrolyte in the phase (3. The undissociated BA molecules are distributed between the phases and the partition coefficient Ps"/p is defined by the equation:
RT\nP^
= v.2-vYA •
(22)
Incomplete dissociation of the molecules does not affect the value of an interfacial potential, and the distribution potential determined from Eq. (6) is established at the interface. The standard distribution potentials of the individual ions (see Eq. 5) remain unchanged in exactly the same way. However, owing to the dissociation equilibrium and the interfacial partition of undissociated molecules, the constants of these processes are related to the difference between the Gibbs standard energies of the ions (and hence between the standard potentials):
RT
^-£%n* =A*A +AlGl =FAX- ~FAX+ •
(23)
"•BA'BA
1.5. Complex formation in one of the phases Suppose that the binary electrolyte BA is distributed between the phases a and (3 and that phase a contains in addition a neutral compound T which can combine with the cation B+ to form the neutral charged complex BT+ (Fig. 5). The reaction A
+
B (a) + T(a)
/ff+
o
BT+(a)
(24)
takes place in phase a and the dissociation constant K\T+ is defined by the equation: RTlnKaBT+ = VTT+ - VT+ ~ VT •
(25)
Apart from the homogeneous dissociation constant KaBT+ of the complex BT+, it is also possible to introduce the constant of the heterogeneous or surface dissociation K^+ referring to a process in which the complexing agent T and the complex BT+ are in the phase a while the ion B+ is in the phase (3. This constant is defined by the equation
102
A.G. Volkov and V.S. Markin
P T l n Val^ — n O a , ,°.P Kl [nKBT+-[iBT+-[iB+-[iT
..".a .
(")£.\ (26)
A simple relation between the two dissociation constants is thereby introduced, and the partition coefficient of the ion B+ is - ^ BT+ ~ •**• BT+ ' rB+
\LI)
•
The equation for the heterogeneous dissociation equilibrium is cM=^^ r+ exp(^).
(28)
The case where the complex BT+ is very stable (i.e. its homogeneous dissociation constant KgT+ is small) while the ion B+ is only slightly soluble in the a-phase (the partition coefficient Pfia+/p is small) has special interest. One can then assume that, apart from the charged complex BT+, the a-phase contains only one charged species, the counterion A". It is possible to find the interfacial potential difference for this case:
I
1
"
,
-,1/2
f cp V 2K
\ B+ cacom denotes
)
F
ca
cp
K
2K
A- BT+\
BT+
where the overall concentration of the free and bound complexing agent in phase a and the activity coefficients for simplicity are assumed to be unity. We will consider two limiting cases arising from Eq. (29). Suppose that the concentration of the A"B+ electrolyte is low. Then, by expanding Eq. (29) as a series, we obtain
A
^«fl ta -~fefA r
r
(30)
A- ^BT*
Eq. (30) can be written more conveniently by introducing the interphase partition coefficient of the ion B+ taking into account complex formation: p«/p _ a /J^a/P
r
B+
~ '-con, ' IS-BT+ •
f3 1 \
K-1 y)
Electric Properties of Oil/Water Interfaces
103
Figure 5. The distribution of a binary electrolyte B+A" in two immiscible liquids in the presence of a complexing agent T that selectively binds B" ions.
The expression for the interfacial potential is then RT P a / p A°(h = — l n ^ - .
(32)
This is a typical expression for the distribution potential which does not include the complexing agent concentration. The form of the relation is determined by the valences of the electrolyte, and Eq. (32) is valid only for a 1:1 electrolyte. For arbitrary valences, a more general expression is: F 1 ca — A"d> = — - — I n — T ^ - T T K 1
Z
B~ZA
^A
(33)
A
BT+
The dependence of FA^fy/ RT on lnc"om is described by a straight line with a slope of 1/2 for a 1:1 electrolyte, 1/3 for 1:2 electrolyte, and 1/4 for 2:2 electrolyte. We will now proceed to another limiting case. Suppose that the electrolyte concentration is so high that the concentration of the complexing agent in the phase a is close to saturation. We then obtain from Eq. (31) ca
RT
A"<|) = — l n - ^ V r
r
A-
L
A-
(34)
104
A.G. Volkov and V.S. Markin
Fig. 6. The distribution potential versus electrolyte concentration in phase p when complex formation occurs in phase a.
Fig. 6 illustrates the dependence of FA^/RT on lnc"_, which consists of two linear sections. The slope of the second section is -1 for a univalent anion A" or (ZA)"1 in the arbitrary case. The point of intersection (Fig. 6) of two sections yields 0.5\n(c"omKg^+/P^), which makes it possible to determine the heterogeneous dissociation constant of the complex BT+. 1.6. The Donnan potential The Donnan potential [6]is the name given to the interfacial potential difference that arises when certain ionic solutes cannot cross the interface between two immiscible electrolyte solutions while the remaining ions are free to move reversibly from one phase to the other [1-4]. Suppose, for example, that an ion R with a charge number zR has such a high partition coefficient / ^ t h a t it is almost totally concentrated in the phase a. Apart from the ions R, the system contains a binary electrolyte B+A" with a concentration in the p-phase of
a, i?-\B+A" | B+A",
p
(35)
Electric Properties of Oil/Water Interfaces
105
This type of equilibrium was first considered by Donnan for two aqueous (a/(3) electrolyte solutions separated by a semipermeable membrane. By definition, the partition coefficient of a salt is given by Q
AB
which defines at the equation relating the activities of ions B+ and A" in both phases: U
B+UA~
\rAB
U
A-UB+
!
•
\-> I)
Using this equation and the two electroneutrality conditions for the phases zRcaR+caB+-caA_=0
(38)
cl-cA=0
(39)
one readily obtains the Donnan potential [3]: RT
f P a/p v p v a ^ B+ B+JA
Ag4> = — I n p
IF *"r
J
a
p
P Y v \rA-
(
zaca^\
+ — I n l +^ s - .
-
a/(!
RT
IB+lA-J
\ IF *-r
c V
(40)
a
L
B+ J
Equation (40) is similar to Eq. (15), but differs by the salt distribution potential given by the first term on the right-hand side of Eq. (40). The Donnan potential can also be written as a function of the concentration of the dissolved electrolyte AB: Aa6-—In p IF
B+ a/
+ — I n 1+ ZRC^BA v" ) 2F V 2Pa p a p
1B+1A
~
P K
a
+
Z C
R RYBA
2Pa pap
(4l)
As shown in Eqs. (40) and (41), the Donnan potential is the sum of the distribution potential and the term dependent on the concentration of the impermeable ion R. If this concentration is low, the second term vanishes, and Eqs. (40) and (41) are reduced to the expression for the distribution potential. If the concentrations of R ions is high, the potential tends to another limit. If we setzR> 0 then
106
A.G. Volkov and V.S. Markin
RT z cava Ag4> = — In ™ - .
(42)
In this case the concentration of counter-ions A" in phase a is no longer dependent on the partition coefficient and interfacial potential, and approaches the limiting value caA_*zRcaR.
(43)
In this case the co-ions B+ are almost completely expelled from phase a: a
M^W Z Ca
(44) Y"
1.7. The Nernst potential We now consider an especially important particular case of the Donnan potential where only one ion B~B can pass across the interface. The interfacial potential difference arising in such a system is called the Nernst potential [7]: RT
n^
zBF
aB
A^ = A^°fi+—-In-f-.
(45)
If the phases a and [5 are identical but are separated by a membrane permeable to B+, then the potential difference is given by RT
/J' 3
zzBrF
aaB
A"d) = — l n ^a . P
(46)
Sometimes the potential difference determined from Eq. (45) is also referred to as the Nernst-Donnan potential, because this equation was first obtained by Nernst, but corresponds to a special case of the Donnan equilibrium [1-4]. 1.8. Gibbs free energy of electron and ion transport coupling Suppose that each of the two phases, a and p, contains its own oxidationreduction couple Red/Ox (Fig. 7a). The exchange of n electrons across the interface results in the redox reaction
Electric Properties of Oil/Water Interfaces
v.Red, +v2Ox2 <^v3Red2 + v4Oxl.
107
(47)
The following relation holds under equilibrium conditions: —a
—p
— EJ5
—a
v,^ Red , +v 2 |i O ; t 2 =v 3 ^ Rerf2 +vA\iOxl
(48)
from which the interfacial redox potential can be calculated:
(49)
A^ = AX 0 +^ln^4-^T\uOx\)
\uRed2)
Here kffiRI0 denotes the standard interfacial redox potential of reaction (47), which is equal to
A»K/0 = ( v , ^ . +v 2 ^ 2 -v 3 nL°£ -v 4 ^,)/«f,
(50)
where n is the number of electrons transferred from Red] to Ox2. This is the interfacial potential for equal activities of the oxidized and reduced solutes in each solvent phase. It should be noted that the interfacial redox potential of the reaction (47) is defined by equations (49) and (50) even though one particle is neutral in each of the redox couples. If the pH of one of the phases changes during reaction (47), vlRed] +v2Ox2
(51)
then from the equilibrium condition -a lhterf,
V
+V
-p -EP -a -a 2^O,2 = V 3 h ^ 2 +V4^O,l +™VH+
(52)
one can find the interfacial redox potential [8]:
(
\ VI /
r.
\ V2
^ _ m — a ^ ) ya°f— \UOx\)
\UKtd2)
\UH+)
t
(53)
where ApK/0 = ( v , ^ , + V 2 LC 2 - v 3 ^ 2 - v 4 ^« - m^InF.
(54)
108
A.G. Volkov and V.S. Markin
Fig. 7. The distribution of redox components in a system of two immiscible liquids.
The influence of the liquid interface on the standard redox potential of an electron-exchanges reaction has been studied by Samec [9] and Volkov [10]. 1.9. The mixed potential If each phase of two immiscible liquids contains common ions and redox couples, then the interface acquires a mixed potential, which is contributed by both the interfacial redox potential and the ion distribution between the phases. A two-phase system allows various combinations of electron-exchange reactions. For example, if each of the phases a and (3 contain a redox couple and one of the couples is soluble in both phases (Fig. 7) electron-exchange reactions can take place both at the interface Reda + Oxp <-> Red\ + Oxa 1
2
l
1
(55) V
'
Electric Properties of Oil/Water Interfaces
109
and in the bulk phase P: Rerff + Ox[ <-> Rec/2p + Obcf.
(56)
The equilibrium condition implies that
hLl " hLl =0 -P
(5V)
- a
Ho*l - Maxl =
0
(58) and it follows that
Ag(fr = — h ^ +—In ^"T* 1 . " ^
/
Rerfl
"
r
(59)
"Orl"Re
The standard mixed potential for reactions (55) and (56) is given by A
^ ° - o ^ l ^ .
(60)
Equation (59) is equivalent to Eq. (11) for two-electron reactions ifzRedi = -1, zOxi = +1, and aledl'aMa = a^d^a9M . The last condition can be true only at a single point for specified activities of all the ingredients. This condition does not follow from phase electroneutrality because both phases contain not only the redox couples but also other ions. Equation (59) does not explicitly contain the second redox couple Red2/Ox2, but it appears in this equation in the second term on the right-hand side. The second redox couple must match the first couple because equilibrium in the system is described by the four equations
uLi^Li n L = M-Li -a -p -a -p ^ReJl + Vo*2 = M-0.1 + I^Rerfl
(61) (62) (63)
110
A.G. Volkov and V.S. Markin
-P -P ~P -P ^Rerfl + V-Oxl = Hart + h ^ 2 •
,,.. (64)
If a common electrolyte BpAt is introduced into both phases, this will not affect the general form of Eq. (59). However, the electroneutrality condition for each of the phases
X « = O;I^ = o '
(65)
J
will lead to a change in the distribution of the redox component concentrations in each phase, which in turn will contribute significantly to the second term on the right-hand side of Eq. (59). Most redox reactions occurring at water-oil and water-biomembrane interfaces proceed with a change in pH, for example NADPH o NADP+ + 2e" + H+ .
(66)
Consider a system of two immiscible liquids a and p (Fig. 7). Let each of the phases have its own redox couple, and let one of the couples be common to the two phases: Re
(67)
Rerfp + Ox[ <-» Red\ + Ox" + mH* .
(68)
If in the course of the reaction n electrons are transferred from Red| to OX2 and if m •*• n, then from the equilibrium condition one obtains as expression for the mixed potential A a (b-A a (f)°
—
In
{n-m)F
Rerfl
M
(69^1
a R e d l <,
where the standard potential is given by A«<|»« /o =—^
In^-.
(70)
Electric Properties of Oil/Water Interfaces
111
From the condition that the phases are in thermodynamic equilibrium and from the electroneutrality condition, one can write for reaction (63) 4 + 2 equations for two redox couples, which exceeds the number of independent variables in the system. If in the course of reaction (63) proton equilibrium is established between phases a and f5, the mixed potential can be written in the form ^-Ap^o-—In—
p — + —\pH,
(71)
.
(72)
where A"<|)°/o = — I n " " ^ " ^
}
r
Ked\
Another example of a redox reaction at the oil/water interface has the form Red <-> Ox + we" + mH+
(73)
The electrons which are the products of reaction (73) can be accepted at the interface by a second substance dissolved in one of the two phases. The standard Gibbs energies of the reaction (73) for each phase, a and p, are: AGa° = X e r f - a\x°Ox - n\i°e - myH+ A
G p = P Kerf " p V°ax - nK
-
m
(74) (75)
P V°H+ •
Subtraction of Eq. (74) from Eq. (75) gives the change of the standard Gibbs energy at the interface if the electron acceptor is located in one phase only, or localized at the phase boundary:
AG;-AG: = (^ld-
yKed)-(^Ox-
tt[x°Ox)-m(yH+-
yH+)
(76)
or AG°=RT\n—^—, P (P Y
(77)
112
A.G. Volkov and V.S. Markin
where P{ is the distribution coefficient of the z'-th ion: RT\nP, = ^ , - y ,
(78)
In the case of a /7-electron reaction, the standard redox potential A£° at the interface is determined by: AEo =
_^ln_?krf
(79)
It is possible to shift the redox potential scale in a desired direction by selecting appropriate solvents, thereby permitting reactions to occur that are highly unfavorable in a homogeneous phase. If the resolvation energies of substrates and products are very different, the interface between two immiscible liquids may act as a catalyst [11-14]. 1.10. The tetraphenylborate hypothesis and interfacial potentials It is known from thermodynamics that the absolute value of a potential difference can be measured only in conductors of identical composition. Therefore, the difference of Galvani potentials between a point in an aqueous phase and a point in the bulk of an organic solvent cannot be measured experimentally, so that non-thermodynamic assumptions must be made. It is a simple matter to measure the interfacial potential difference by varying one ion of the dissolved salt. It is also possible to measure the partition coefficient for each salt, P^Jf". By virtue of Eq. (5), the interfacial potential can be represented in the form Ap^A^L-^lnP^.
(81)
In this way a scale of interfacial potentials can be constructed, but this scale will include an unknown constant term, Ap(|)°K, the standard distribution potential of the invariant ion. The problem, therefore, is to choose the origin of the scale correctly. The origin cannot be chosen by thermodynamic methods, so auxiliary methods are required. The historical aspects of the problem are considered in a review by Kolthoff [15].
Electric Properties of Oil/Water Interfaces
113
The idea of the tetraphenylborate method is to choose a cation and an anion with equal standard energies of transfer between two solvents, so that the interfacial distribution potential will be zero. In this case the ion partition coefficients will equal the salt partition coefficient which is easily determined experimentally. Once the latter coefficient is found, we can find the standard free transfer energy and the standard potential distributions of individual ions. These potentials are equal in magnitude but have opposite signs: A^:=-A«(|)°.
(82)
In one application of this approach, Koczorowski [16] worked with tetraethylammonium picrate (Et)4NPi, while others used tetraphenylarsonium tetraphenylborate (Fig. 8). The ions comprising the latter salt, TPhB" and TPhAs+, are very much alike. They are symmetric, almost spherical particles of sufficiently large radius, that can be assumed to have the same hydrophobic properties. Their charges are equal in magnitude, are located at the centers of the spheres, and are screened from the solvent, so that the electrostatic contributions to the free resolvation energy are equal. Such ions have a relatively small polarizability, a low surface charge density, and do not participate in a specific interaction with solvent molecules. For these reasons the standard transfer energies of these ions will be nearly equal for any two solvents. The tetraphenylborate method allows one to determine the standard distribution potentials using cyclic voltametry, chronopotentiometry, polarography at a dropping electrolyte electrode, chronoamperometry, and solubility and extraction data. Since
we can choose a zero point on the potential scale and pass from the scale of applied potentials to the scale of Galvani potentials or distribution potentials. The ion standard distribution potential depends both on the nature of solvent (primarily on the optical and static permittivities) and on the ionic radius (Fig. 2). Figure 2 shows the distribution potentials versus the ion radius for water-nitrobenzene, water-1,2-dichloroethane and water-dichloromethane system. The radii of inorganic ions are taken in the Gourary-Adrian scale [17] and those of tetraalkylammonium salts are found according to Robinson and Stokes [18]. As can be seen from Fig. 2, the ion distribution potential drops sharply with increasing ion radius and the permittivity of the non-aqueous solvent. Solid curves in Fig. 2 are from theoretical calculations. For ions of radius less than 0.3 nm the main contribution to A°<> | " is due to the electrostatic
114
A.G. Volkov and V.S. Markin
portion of the free resolvation energy, and for ions of radius greater than 0.3 nm the main contribution is due to the solvophobic effect [1-4].
Fig. 8. Structural formula of tetraphenylarsonium tetraphenylborate.
1.11. Hung's method of Galvani-potentials calculation Interfacial potential values for systems with small volumes were analyzed by Hung [19-21], who considered the distribution of charged particles between phases and the corresponding interfacial potentials. In particular, the finite volumes of contacting phases which affect the distribution of charged particles were taken into account. Such relationships are complex and generally do not permit analytical solutions. However, solutions can be obtained with the help of simplifying nonthermodynamic assumptions or by numerical methods. Hung considered equilibrium distributions of ions /, between two liquid phases a and P: a
/;•
/;• p
(84)
where Va is the volume of phase a. From the mass conservation law it follows that
Vac« + Fpcf = m,
(85)
where mx is the amount of ion I\ in both phases. Hung also assumed electroneutrality of each phase
:>>.<=°
(86)
Electric Properties of Oil/Water Interfaces
2>,cf=0
115
(87)
i
and it follows from Eq. (85) - (87) that
ZZ'W'=0
(88)
Generally speaking, the total system consisting of phases a and p is electroneutral, but electroneutrality of a single phase (86) and (87) cannot be assumed if one of the phase's volume is small. According to Kakiuchi [22] if the diameter of a given phase is less than 10"6 m, the Poisson equation can be used to correlate local ionic concentrations with the interfacial potential, both of which vary with the distance from the phase boundary and depend on the shape of the interface. Combining Eqs. (85)-(88) and (7) gives Hung's equation: y
z
-£i
+
t r i + (Y«/vyf)exp[(z,F/i?rXA^-A^;)] j
y
7r P.o
^
'
( 8 9 )
^=o
^ l + v(Y;VY?)exp[(z,F//?rXA^-A^)] where v = y°IV® and c° denote the initial concentration. Equation (89) has no analytical solution when the number of ions in the system is more than 3 and must be solved numerically. From Hung's equation (89), it follows that the value of Ap<(> depends on temperature, the volume ratio of the two phases, initial concentrations of components, activity coefficients and standard electrical potentials of ion transfer between phases a and p\ If the system contains only one binary electrolyte A+B", Eq. (89) gives the distribution potential (14). Detailed analysis of Hung's method can be found in reviews [4, 8, 16, 21-24]. 1.12. Distribution potential in small systems As described above, the distribution potentials were very well studied in the systems where both phases had macroscopic dimensions. The condition of electroneutrality is obeyed in the bulk of both phases and the value of the potential is given by equation (14). However, this is not the case in the systems where one of the phases has microscopic dimensions like in microemulsions. If the size of the droplets is comparable with Debye length neither the value of distribution potential nor the solute concentration in the microphase can be
116
A.G. Volkov and V.S. Markin
found from Eqs. (6) or (7). Importance of this problem was underlined in a number of papers [1-4, 8, 16, 21-23], but the solution of the problem was not yet given. In this chapter, we shall analyze this issue and derive appropriate equations. Let us consider a microemulsion with droplets of oil (D) in water (W) (Fig. 9). Let the radius of the droplet is equal to R and the aqueous phase contains a uni-univalent electrolyte with concentration in the bulk (far away from the droplet) equal to cm. The electrolyte can partition into the oil droplet with partition coefficients for cations and anions equal to PCD/W and Pa /w , respectively. If these values are not equal to each other then the distribution potential builds up at the interface. To investigate the effect of geometry we shall use the Poisson-Boltzmann equation for electrical potential \\i, which is a function of radius r. Due to the spherical symmetry of the system the angles are not involved. In the aqueous phase (r > R) the equation is: 1 d_( 2 dy_\ = e^J fe_Mr)>| _ 2 r dr{ dr) zozD[ \ kT ) \
f # ) | kT )\
In the oil droplet (r < R) it has a similar form, but also includes the distribution coefficients:
These equations imply that concentrations of anions and cations in the aqueous phase are
c.(r) = « p [ S ^ ) ) and C , W = e x p ( - 2 ^ ) ,
(92)
correspondingly, while in the oil they are
«.
(93)
We would like to obtain an analytical solution of the problem. To do this, we assume that potentials are small and the exponentials in the PoissonBoltzmann equations can be expanded into series:
Electric Properties of Oil/Water Interfaces
r dry
117
dr J
J_ d( tdyjA^i r dry dr J
(r)
]t
ifr<^
(95)
Here we provided potential \\> in the water or oil with subscripts W or D respectively and introduced the Debye parameters in the water and oil:
£Q Syp Kl
S Q S Q Kl
Additional parameter v|/o that appeared here after expansion into series represents the distribution potential in the macroscopic phases. As was shown above it is equal to O
kT PDIW =—In-^TF-
(97)
However, because we assumed that potentials are small, it can also be expanded into series taking into consideration that the difference between partition coefficients is not very large. Then expression (97) reduces to ( pD/W
pDIW\
M>o-(p/w+r,)eo.
(98)
The set of equations (94) and (95) should be supplemented with two boundary conditions. The first one is the condition of continuous potential at the water-oil interface: Vw(R) = yD(R).
(99)
The second one is the condition of electroneutrality of the total system, which can be presented by the integral: \cc(r)-ca(r)] o
r2dr = 0.
(100)
118
A.G. Volkov and V.S. Markin
Solution of equations (94) and (95) can be found as a combination of the modified Bessel functions of the order 1/2: = r- | / 2 [4/ 1 / 2 (K w r) + ^ I / 2 ( K w r ) ]
(101)
v|/o(r) = r- 1/2 [5!/ 1/2 (K w r) + 52Jf1/2(Kwr)] + v|/0.
(102)
¥^(r)
Both potentials should be limited and this immediately gives Af= B2 = 0. The remaining coefficients A2 and B\ should be found from the boundary conditions (99) and (100). Substituting functions (101) and (102) in (99) and (100) and performing integration, one obtains the following system of equations for coefficients A2 and B\.
e-K»* = Rf2-
\h^-
sinh{KDR) +WoR
(103)
2A2 p i ^ ± I e - - « + V r
(r+r)s,J-
(104)
1
AC0Sh(Koi?)_^_Sinh(K^)
=0
Solving this system and substituting coefficients A2 and B\ in (101) and (102) one finds the potential profile in the oil droplet and in surrounding aqueous media:
(105)
(106)
Electric Properties of Oil/Water Interfaces
119
The most interesting parameter is the potential in the center of the droplet. It can be found from Eq.105:
(107)
The expression in the square brackets is always less than 1. Therefore, the potential in the oil droplet is always less than \\i0 and the value of the distribution potential would be established if the oil was present in macroscopic amount. Dependence on radius. The potential (107) depends on the droplet radius R. Let us consider two limiting cases for very small and very large radii. For
(108) Therefore, in the very small droplet the potential disappears and becomes zero. Then again for R —> 0
(109)
In a very large droplet the distribution potential approaches its macroscopic value \|/0. The total dependence of the distribution potential on the droplet radius is presented in Fig. 10. There are two parameters that define this function. In this example, they are selected as: (110) The potential is normalized by \\i0 and the droplet radius is normalized by Debye parameter, KD . One can see the quadratic dependence of the potential on KDR at
120
A.G. Volkov and V.S. Markin
small radii. The half of the maximum potential is achieved when KDR « 2 and then the potential asymptotically approaches the maximum value of v|/0. Dependence on electrolyte concentration. In macroscopic phases the distribution potential does not depend on l:l-electrolyte concentration if one neglects the difference in activity coefficients of ion. The same is true for small droplets. It can be easily seen from Eq. (107) if one introduces a dimensionless radius p = KDR .
(111)
Debye parameters KD and KW depend on electrolyte concentration as 4c. However, these parameters appear in the equation only as a ratio of one to another. This ratio does not depend on concentration, and hence the distribution potential also does not depend on electrolyte concentration.
Fig. 9. Oil droplet (D) in aqueous solution of electrolyte (W). Radius of the droplet is R.
Electric Properties of Oil/Water Interfaces
121
Fig. 10. Distribution potential in the center of the droplet.
Potential at the surface of the droplet. The value of the potential at the droplet surface \\iw(R) = yD(R) = v|/(7?) can be found from Eq. (106):
(112)
Of course, it is only a fraction of the total distribution potential \|/ D (0). We have analyzed the case of small potentials, which is certainly justified for small droplet. We have also obtained analytical solution of the problem. If the droplet is not small and the potential exceeds kT/e0, then the solution can be found numerically.
122
A.G. Volkov and V.S. Markin
2. SURFACE POTENTIALS Mechanical and electrical properties of all interfaces - from pure water to elaborate surfaces of cellular membranes - depend on ion adsorption at these surfaces [25]. This phenomenon plays a very important role in colloid and physical chemistry, biology and medicine. The structure and stability of large biomolecules and membranes depend on the distribution of counterions in the aqueous phase. The transport of ions through ion channels in cellular membranes is determined by the surface potential generated by adsorption of ions from aqueous phase. Conduction of nerve impulses and effect of anesthetics strongly depend on this adsorption. Therefore, it is very important to know the structure of aqueous interfaces and distribution of ions in them. Other fields in which this knowledge is indispensable are seen in environmental aquatic chemistry and chemistry of the atmosphere. Chemical reactions involving aerosolized particles in the atmosphere are derived from the interaction of gaseous species with the liquid water associated with aerosol particles and with dissolved electrolytes. For example, the generation of HONO from nitrogen oxides takes place at the air/water interface in seawater aerosols or in clouds. Clouds convert between 50 and 80% of SO2 to H2SO4 which contributes to the formation of acid rain [26]. Another example is the release of chlorine atoms from sea salt aerosols reacting with the gaseous components. Molecular chlorine, a photolabile precursor of Cl atoms, is a product arising from a heterogeneous reaction of ozone with Cl" ions. Measurements of inorganic chlorine gases indicate the presence of reactive chlorine in the remote marine boundary layer; reactions involving chlorine and bromine can affect the concentrations of ozone, hydrocarbons and cloud condensation nuclei. The heterogeneous reaction between HOBr and HCl converts significant amounts of inactive chlorine (HCl) into reactive chlorine (CIO). These are only a few examples of global processes in environmental chemistry occurring at the air/water interface. In all these cases, the adsorption of simple inorganic ions is especially interesting. In the beginning of the last century, Heydweller [27] found that surface tension of water increases with addition of inorganic electrolytes having relatively small ions. Some organic molecules (sodium formate, glycine, amino acids at the isoelectric point when both the amino and the carboxyl groups are ionized in the form of zwitterion) also increase surface tension. The surface tension of solutions increases almost linearly with increasing concentration of salts, and for inorganic salts there is an empirical equation: y = y0 + be, where b = 1.7 mNttV'M"1 for LiOH, 1.8 niNn^'M"1 for NaOH and KOH [28], 1.63 mNm' 'M"1 for NaCl and 1.33 m N m ' V for KC1 [29]. For organic tetraalkylammonium salts the constant b decreases with increasing of cation
Electric Properties of Oil/Water Interfaces
123
radius: for tetramethylammonium chloride b is positive, for tetraethyl ammonium chloride b is about zero, and for bigger cations b is positive. This phenomenon was explained as a result of negative adsorption of small ions. Wagner [30] was the first to describe this phenomenon as an electrostatic effect of image forces. Onsager and Samaras [31] derived equations, which predicted the increase of surface tension with increase of electrolyte concentration, but only for the very dilute solutions. Later Buff and Stillinger [32] presented the statistical mechanical formulation of the same problem by also employing the conventional model for very dilute solutions. They calculated the molecular distribution functions from a generalized form of Kirkwood's integral equation and found the surface tension directly from the molecular theory for this thermodynamic parameter. These approaches were quite successful, although they used a simple theory for point charges and did not take into account effects of ion radii. It was Frumkin [33,34] who indicated that the increase of surface tension and surface potential of aqueous solutions of inorganic salts depended on ionic radii. However, it was believed for a long time that equivalent concentrations of different salts gave very similar variations of surface tension^ at least at small concentrations [35]. Later it was found that the differences are rather pronounced even at small concentrations [36]. At higher concentrations, different ions exhibit specific differences: the higher the ion hydration, the higher the surface tension increment. The contributions of two ions of the electrolyte to the surface tension appear almost additive. There was a series of attempts to advance the theoretical explanation of the surface tension given by Onsager and Samaras [31], but without much success. A critical discussion of these attempts can be found in Randies [37]. Krylov and Levich [38] considered the discreetness of the charge of adsorbed ions, but found this effect to be rather small. Having in mind all these improvements, we still have to stress that the original works of Wagner [30] and Onsager and Samaras [31] correctly identified the major players determining the surface tension of the aqueous electrolyte - image forces at the interface and negative adsorption of ions. Using the well-known Gibbs adsorption equation, they calculated the value of surface tension without ascribing too much importance to the difference between anions and cations [30, 31]. The surface tension increment was explained at least for the small concentration of electrolytes. In contrast to this, the situation with the surface potential was completely hopeless. In Onsager-Samaras approach, if there is no difference between anions and cations (beyond the charge sign), there is no reason to expect that they would generate any surface potential because they would equally be negatively adsorbed at the interface and their charges completely compensate each other. Therefore, the whole effect of change of the surface potential of water in the presence of electrolytes originates from the
124
A.G. Volkov and V.S. Markin
difference between anions and cations. In this sense, this effect is more subtle and elusive than surface tension. The theoretical explanation of this phenomenon was published recently by Markin and Volkov [39]. The explanation of the surface potential can be found in the difference in ionic radii and their manifestations at water surface. Because of finite radii, different ions can interact differently with interfaces. This results in a different interfacial distribution of energy, concentration and charge density of these ions. We shall use the model of ions with finite radii to describe both surface tension and surface potential at the wide range of concentrations. 2.1. The model: Ions of finite radii at the air/water interface Let us consider an air/water interface as depicted in Fig. 11. Coordinate x is directed from the interface to the bulk of water. The energy of an ion with radius a at the interface was calculated by Kharkats and Ulstrup. They accounted for the Born solvation energy and the energy of interaction with the image. When x > a, the energy of an ion can be presented as WKU{x) = (zeof \ 327te0E(ra|
Uw-zA\la
Uw-zA~J\
\KEW+EA)X
\EW+EA)
2 [_\-{2xlaf
^a^lx 2x
+ a^ 2x-a_\
(H3)
The first term in curls in (113) represents the Born solvation energy, and the last two terms give the interaction with the image. In the region 0 < x < a:
327iEoSw,a [V
\\ + xla){\-2xla) \ + 2xla
|
a)
a
f
\zw+zA)\ |
2x \
2xVjj a)\\
a) |
{EW+EAJ
(zeof f 2s, Y ^ x~\ \6%zazAa\zw +e 4 J ^
a)
When the center of the ion is right at the interface (x = 0), the electrostatic Gibbs energy is equal to WKU(x = 0) =
(ze Y K -^
4mo(Ew+sA)a
.
(115)
Electric Properties of Oil/Water Interfaces
125
Fig. 11. Distribution of water and ion concentration at distance* from the air/water interface; XG is the position of the Gibbs dividing surface.
When the ion is located in the air, the electrostatic free energy is obtained from equations (113) and (114) by exchanging Zw <=> %AThe image energy can be derived from eqs. (113) - (115) by subtracting Born energy W^om = JFKu(+GO) • One can see an important difference: the energy of a point charge becomes infinite at the interface, while the Kharkats-Ulstrup function eliminates this feature and makes the curve continuous. Equations (113)-(114) give the energy of a lone ion interacting only with its image. However the presence of other ions in the solution will modify this interaction due to Debye screening of electrical field. Therefore the ion/image interaction should decay much faster than in equations (113) - (114). To take this effect into consideration, we shall introduce the screening function J[x) as follows: (116)
126
A.G. Volkov and V.S. Markin
with Debye constant K and ion radii aan and acat. For binary electrolyte the Debye constant is given byK= jegCg^zf /zoekT . This definition consistently takes into consideration the model of ions as hard spheres so that cations (cat) and anions (an) cannot approach each other closer than at aan + acal. Therefore the image energy can be presented as
Wbnv(x) = f(x)[WKU{x)-WKU(^)].
(117)
This function would work rather well for calculations of the surface tension. However, for calculation of the surface potential one would need to take into account more subtle effects that help to distinguish between cations and anions. One of them is the solvophobic effect that takes into consideration the work of creating the cavity to place the ion into the solvent. If one assumes that there is surface tension ycavily in this cavity, then in the bulk of the solvent the solvophobic energy is 4na2ycavity. If the ion crosses the interface from right to left (Fig. 11) the area of the cavity decreases and with it the solvophobic energy. If the ion center is positioned at x, then in the range -a < x < a the energy is 2na(a + x)ycmily; at x<-a, it is 0. If the reference point is selected in the bulk of water, then these quantities should be decreased by 4na2ycmjty, and one finally obtains the relative solvophobic energy as 2a, if x < -a Wsolvophoblc = -2%aycavily-a-x,if
-a<x
(118)
0, if x>a The surface tension in the cavity is not very well known; it is believed to be close to the surface tension of water, but is probably less than that due to the small radius of curvature as was argued by Tolman [42]. We shall estimate this value in the range of 40 - 50 mN/m. One can see that the solvophobic effect decreases the energy at the interface. The total energy of ion i with charge number z, includes image and solvophobic components and electrical potential of the electrical double layer cp(x): W, (x) = Wwmge (x) + WsohophohK (x) + z,e0 [q,(x) - 9(00)] .
(119)
Electric Properties of Oil/Water Interfaces
127
Fig. 12. Surface tension of water in the presence of NaCl: (dots) experimental data; (dashed line) Onsager limiting law; (solid line) according to Markin and Volkov [39].
Notice that the image energy and the solvophobic energy also depend on the type of ions. To calculate the spatial distribution of ions we used standard methods for the electrical double layer. The Poisson-Boltzmann equation can be presented in the following way:
£2=-S&2> l O J-S&l. dx
sos i
\_
(120)
kT
where dielectric constant 8 assumes the value eA or ew depending on coordinate x. We assume that the aqueous phase contains a binary 1:1 electrolyte with bulk concentration c0. Energy W\ is given by equation (96) and summation includes both cations and anions. For boundary conditions, we assume that the electrical potential
^
=0.
(121)
Equation (97) with boundary conditions (98) can be solved numerically and both surface tension and surface potential can be found.
128
A.G. Volkov and V.S. Markin
2.2. Concentration dependence of surface tension In the first theoretical description of the surface tension of electrolyte solutions Wagner and Onsager and Samaras [31] presented the energy of interaction of an ion with its image at the distance x from the interface as
W(x) =
^
exp(-2Kx)>
(122)
l6m0Ewx where e0 designates the charge of the ion, s0 - the permittivity of free space, sw - the dielectric constant of water, K - Debye constant. This function is presented in Fig. 12 by the dashed line. The surface excess of ions (/) with respect to water (w) was calculated as CO
,(W) = o J W h W(XV kT] - l)dx .
r
c
(123)
o Using surface excesses of ions, Onsager and Samaras calculated the surface tension as a sum of a series and tabulated this sum [31]. For very dilute solutions of 1:1 electrolytes, they found an analytical expression for the surface tension, which they called the limiting law: Ay = y-y 0 =1.012clog(1.467/c)-
024)
This equation is presented in Fig. 12 (dashed line), together with experimental data for NaCl. As one can see, the limiting law coincides with the experimental data only at concentrations smaller than 0.1 M, after which the discrepancy becomes very large. The tabulated values improve the situation only a little. And of course, the limiting law does not distinguish between different ions describing them by the same equation. The main drawback of previous approaches was the absence of an appropriate model of ions at the air-water interface. To obtain a radical improvement of the theory of surface tension in the presence of simple electrolytes, we shall use the new expression for energy (119) that accounts for both finite radius of ions and solvophobic effect. We shall begin with calculation of the surface tension, because surface potential at low electrolyte concentrations has some peculiarities and cannot be found directly from the standard Poisson-Boltzmann equation. Fortunately, this does not influence the calculation of surface tension because at low
Electric Properties of Oil/Water Interfaces
129
concentrations the potential is small and has no effect on the distribution of ions that are determined by image forces and solvophobic effect only. Let us calculate surface excesses of ions i, with respect to water Fj(W) . The Gibbs dividing surface corresponding to zero-excess of water (Fw = 0) does not necessarily coincide with the surface of water as demonstrated in Fig. 11, where its coordinate is designated xG . If the bulk concentration of ions in the air and water are correspondingly 0 and CQ, then the ion surface excesses [39] are
F, ( w ) = c ;|exp[-^lj dx + c 0 + |[exp[-^ll-l \dx.
(125)
Position of the Gibbs dividing surface, xG, was found in the following manner. Water concentration in solutions of simple salts depends virtually linearly on salt concentration as one can see in Fig. 13 for solutions of NaCl and KC1. The data for these examples were taken from CRC Handbook of Chemistry and Physics [29]. The trend lines of these dependences can be presented as cwaler = 5 5 . 6 - a csah.
(126)
The slope a was found for NaCl to be equal to 1.18 and for KC1 equal to 1.73. One has to have in mind that at the far right end of the concentrations scale (about 3 M) the data taken from CRC Handbook of Chemistry and Physics [29] probably are not very reliable because of the saturation effect. However, at smaller concentrations, the data ideally fit a straight line and this is the only result we take from these data. Besides, as indicated later in the text, the effect of the displacement of the Gibbs dividing surface is rather small. Therefore, any small inaccuracy resulting from saturation effect will become negligible in the final result. Because concentrations of anions and cations are slightly different at the interface, we shall define the salt concentration as the mean of the two. Then the spatial distribution of water concentration in the solution can be presented by
c,^(x) = 55.6-^S> jexp " ^ ^
+«xP - ^ ^
j.
(127)
Assuming that the actual surface of water is located at coordinate xD = 0, the water excess at this dividing surface is
130
A.G. Volkov and V.S. Markin
r».,,,« ) =|]{2-expf-2^W] + e x p [ - ( ^ ] } ,
(12S)
and the position of the dividing surface corresponding to zero water excess is
Fig. 13. Water concentration in aqueous solutions of NaCl and KCI. Experimental points are taken from ref. [29].
Fig. 14. Dependence of the surface tension of aqueous solution of KCI on concentration. Experimental points are taken from ref. [29]. The solid line is calculated using Eq. 131.
Electric Properties of Oil/Water Interfaces
131
However, position of this Gibbs dividing surface is different from 0 only at very high salt concentrations. In the isothermal case the surface tension is related to the surface excesses by the Gibbs adsorption equation:
rfy = - t r w 4 .
(130)
In the simple case of a uni-univalent electrolyte the increment of the surface tension can be presented as C
j
Ay = y - Yo = - j(r flmon(w) + rcalmniw))—. o
(131) c
To calculate this quantity we have to solve the Poisson-Boltzmann equation (123) and find the integral in Eq. (131). We carried out numerical integration for a number of different electrolytes and compared them with experimental data. Results for KC1 are presented in Fig. 14. One can see that the present model provides radical improvement comparative to the Onsager limiting law, although at very high concentrations there is a certain discrepancy between experimental and theoretical data. 2.3. Surface potential The elegant theory by Onsager and Samaras for surface tension predicted a zero surface potential because it did not envision any difference between the way cations and anions interacted with the interface. However, experimental observations clearly demonstrated rather pronounced surface potential depending on electrolyte concentration. The sign of this potential with respect to the bulk of water very often is negative, although some salts give positive potential. In the absence of a rigorous theory, a number of semi-empirical attempts were made [37, 43]. The increase of the surface tension with electrolyte concentration was interpreted as the presence of a solute-free layer 4-5 A thick at the surface. Randies [43] assumed that this surface layer is completely inaccessible to cations in solutions of all alkali metal salts, but anions can penetrate more closely to the surface. In other words, Randies postulated two different planes of closest approach for anions and cations. The anions built up negative charge at the surface that was balanced by positive charge of cations deeper in the bulk and hence a negative surface potential was established. Using this simple model and the standard kinetic theory of the electrical double layer,
132
A.G. Volkov and V.S. Markin
Randies calculated concentration dependence of the surface potential. However, comparing his theoretical results with experimental data he had to admit that his model was quite unable to account for the observed surface potentials. The problem was not simply in numbers: theoretical predictions were qualitatively wrong. The Randies theory predicted that the surface potential at low concentrations should increase as a square root of concentration, while experimentally this dependence was either linear or even superlinear. Therefore the theoretical and experimental lines had opposite curvature which was unacceptable. This finding seemed rather strange because the model qualitatively cannot be wrong: there is definitely spatial separation of positive and negative ions at the interface that creates surface charge and this charge should be balanced by the other part of the electrical double layer. At low concentrations the electrical potential is small and does not influence the ion distribution at the interface. Therefore one can conclude that the surface charge Q should grow linearly with the electrolyte concentration: Q~c0 . If the double layer capacitance is CDL , then the surface potential is (? = Q/CDL. According to Gouy-Chapman theory, at low concentrations the capacitance is equal to the capacitance Cdiffuse of the diffuse part of the layer and hence is proportional to the square root of the electrolyte concentration: CDL
~ Cd,ffuse ~ V C 0 •
(132)
That is why Randies came up with the result
(133)
However, there is an interesting anomaly in the capacitance of the double electric layer: at low electrolyte concentrations it does not obey the theory of Gouy-Chapman. This phenomenon named the Parsons-Zobel effect [44] was discovered by Samec et al. [45] and later studied by Wandlowski et al. [46]. These authors analyzed the relationship between inverse capacitances CD£X and Cdiffuse'X- According to Gouy-Chapman-Stern or Vervey-Nissen theory this
Electric Properties of Oil/Water Interfaces
133
Fig. 15. Parsons-Zobel (PZ) plots for the nitrobenzene/water interface. The solid line was calculated according to the Gouy-Chapman theory with the concentration independent capacitance of compact layer equal to 0.74 F/m~. Experimental points are taken from Samec etal. [45].
relationship is given by a straight line with cut-off at the ordinate determined by the compact layer capacitance [47]. However, experimentally observed, the total capacitance at low concentrations of electrolyte (large Q , ^ / 1 ) strongly departs from the straight line and displays the tendency to become constant (Fig. 15). We shall discuss this peculiarity later but now just accept this fact 'at face value'. If relationship (132) does not hold and the double layer capacitance is rather constant, then relationship (133) should rather transform to cp~c 0 .
(134)
This corresponds much better to the experimental observations. Following this line of arguments, we calculated the surface potential for different salts. First, using previous equations we calculated surface excesses of cations and anions and found surface charge as Q = F(rcatl0HK)-ranwniw)),
(135)
134
A.G. Volkov and V.S. Markin
Fig. 16. Theoretical estimation of the surface potential on concentration for NaCI aqueous solution based on Gouy-Chapman (GC= - 4.77-y/c) model, with Parsons-Zobel (PZ= - 10c) effect at liquid interfaces, and the actual curve (Approximation= x(0.0001 + 0.00193x2 J
).
Fig. 17. Dependence of the surface potential on concentration. Theoretical lines for KF (triangles), KC1 (filled circles), and NaCI (diamonds). Experimental points are taken from Frumkin and Randies [33, 34, 37]. The curves were drawn according to the following equations: KF = |"(2.092V^ +0.261c)' 4 +cT4 /1296~| NaCl= -c(0.0001 + 0.00193c 2 )"" 4 .
, KC1 = -c(o.OO666 + O.O393c 2 )~"\
135
Electric Properties of Oil/Water Interfaces
Fig. 18. Surface potential of an aqueous solution of KI. Experimental points are taken from Frumkin [33, 34] (V) and Jarvis and Scheiman [36] ( • ) . Radius of I" was taken 0.206 nm (dashed line, equation = radius
P
of
0.22 V4
-[~(l8.839Vc - 2 . 5 1 4 C ) " 4 +
(thin
continuous
line,
), or crystallographic equation
=
-1-1/4
30.42Vc-5.429cJ + c~4/83521 ), or estimated as 0.25 nm (solid line, equation = -x [2.744* 10~6 +0.000138x 2 (9.23-4.724cVc ) * 1
where F is a Faraday. Using the data for the interface nitrobenzene-water [45] we assumed that at low concentration the limiting value of the double layer capacitance is about 0.1 F/m2. The results for NaCl are presented in Fig. 16 by the straight line marked PZ. Calculations based on the Gouy-Chapman theory, as described in the section 'Surface tension', are presented by the line marked GC. One can see that the Parsons-Zobel effect works only at small concentrations where it predicts bigger capacitance and hence the smaller magnitude of the surface potential than Gouy-Chapman theory. Later on, this relationship changes and the Gouy-Chapman theory determines the limiting value of capacitance. Therefore the actual curve should smoothly change from PZ to GC as presented by the 'Approximation' line. Similar calculations were performed for a number of salts and the results are presented in Fig. 17. Notice that the surface potential for KF is positive while all the others are negative in total agreement with experimental data. At the separate Fig. 18 we presented data for KI. Iodide salts often display a number of peculiarities that put them aside from other simple salts of alkali metals. For example, iodide in aqueous solutions in the presence of oxygen has the tendency to form complexes like IO3~ and/;. Therefore, it came as no surprise that calculated curves predicted much smaller (in magnitude) surface
136
A.G. Volkov and V.S. Markin
potential than experimental data. To account for the presence of complexes I3 , we took the radius of anions in this case as a fitting parameter and found that the data can be nicely described with r_ = 0.250 nm. We have calculated both surface tension and surface potential for aqueous solutions of simple electrolytes. For this purpose we used a new model of ions at the interface. It considered the solvent in the classical way as a dielectric continuum. The model took into consideration finite radii of ions, which permitted to eliminate the divergent function, describing the energy at the interface for point ions and providing a way to distinguish between different types of ions. By applying standard methods of the electrical double layer for ions of finite radii and solving the ensuing equations, we were able to find the surface tension in a wide range of concentrations in agreement With experimental data. Previous investigations could explain the surface tension up to centimolar concentrations, while in the present work we extended this range to molar concentrations. The surface potential was explained for the first time both qualitatively and quantitatively. In this model [39], the sign of the surface potential was determined by the larger ion that was less repelled from the surface than the smaller one. However, if only the finite radii of ions were taken into account, then the calculated potentials were much smaller in amplitude than experimental data. This discrepancy happened because the surface potential is much more sensitive to the subtle details of the model than is the surface tension. While the latter is determined by the sum of anion and cation adsorptions, the former is determined by their difference. This is a classical case of a small difference of big numbers: even a very small inaccuracy of any of these big numbers brings a large error in their difference. Therefore we looked for additional effects that could help to distinguish between cations and anions. This happened to be the solvophobic effect, which does not change much the adsorption of each ion but considerably contributes to their difference. As a result, the amplitude of calculated potentials increased significantly and became closer to the observations. Calculation of the energy of the solvophobic ion-solvent interaction is based on phenomenological or statistical mechanical molecular models in which the parameters describing both the ion and solvent are not always known. To avoid uncertainty, the solvophobic contribution to the solvation energy is sometimes estimated using the semi-empirical solvophobic formula [48]. The main idea was formulated and discussed in a number of papers [49-53]. Surface energy at the interface of cavity in water created to accommodate the ion with radius a, when expressed in terms of surface tension at the air/water interface Ycavity, is equal to ApG°(svp/!) = 47ra2ya[isgn(Ya - y p ). This relation infers that the molecules of the two solvents water and air do not mix or react
Electric Properties of Oil/Water Interfaces
137
chemically with each other. The hydrophobic contribution to the Gibbs free energy of ion or molecule is obviously greater for particles with larger radii, a. One also has to take into account that ions are very small entities and the surface tension in cavities should be smaller than at the planar interface. The effect of curvature on the surface tension at a molecularly sized sphere was estimated by Tolman [42]. The surface tension at radius a can be written as
Jcav^ = YP/awar(l + 25/ a ) > where S according to Tolman, is the distance from the surface of tension to the dividing surface for which the surface excess of fluid vanishes. Parameter S can vary between 0 and a few angstroms. We found that this effect has optimum contribution to the adsorption of ions at ycavity = 50 mN/m. From a theoretical point of view, the limits of applicability of the solvophobic formula are not quite clear. Nonetheless, it works surprisingly well for calculation of the partition coefficients of a system of two immiscible liquids [54]. The only problem that remained with surface potentials was the very fast growth of calculated potentials at small concentrations: they grew as a square root of concentration, meaning that it is infinitely fast, while experimentally potentials changed rather linear with concentration. This last problem was resolved by taking into consideration the Parsons-Zobel effect [45, 46] according to which at small electrolyte concentrations the capacitance of the double layer did not vary like Jc^, but rather became constant in an obvious contradiction with the classical point of view. We modified our calculations at small concentrations taking into account the constant capacitance and providing the smooth transition from one limiting case to another. The results are presented in Figs. 17 and 18. So, our analysis was based on the model that accounts for finite radii of ions, the solvophobic effect and the Parsons-Zobel effect. The model worked rather nicely for simple ions like Na+, K+, Cl" and F ' as one can see in Fig. 17. However, with ions having the tendency to redox and photoredox reactions with oxygen (air), like iodide that is known to form IO3" and I3" ions, the situation is different. Following the same method we have calculated the surface potential for the solution of KI, assuming the radius of I" equal to 0.205 nm according to Gourary and Adrian or the crystallographic value of 0.220 nm. The theoretical curves are presented by the broken and thin continuous lines in Fig. 16. As one can see, both of them predict much smaller potential than was actually observed. We believe that it is due to the fact that a fraction of iodide ions exists in the form of I3" or IO3" complexes and has a bigger radius. We did not try to estimate its value independently but rather considered it as an effective radius and a curve-fitting parameter to achieve the agreement between the theory and
138
A.G. Volkov and V.S. Markin
experiment. We found that it can be done with effective radius of 0.250 nm. This number does not seem unrealistic and hence it provides an estimate of an important parameter for this salt. We found a reasonable agreement between the theory and experiment, although not an ideal one. There are a few reasons for certain discrepancies. First, the experimental data on the surface potential is not very numerous and often are contradictory. Measuring the surface potential is a rather delicate procedure and it is also method-dependent. Surface potentials at the air/water interface can be measured by the dynamic capacitor method, the radioactive probe method, and the jet electrode method [8]. Each method gives the opportunity to measure surface potential at the air/water interface with the accuracy about ±1 mV, but difference in measured values between different authors can reach more than ten percent. There can be a problem with elimination of diffusional potential in the salt bridge between the measuring cell and the reference electrode [33, 34] or effect of charge generation by the radioactive probe. For example, for 0.5 M KI aqueous solution surface potential was measured to be equal to 22 mV [33, 34], 10 mV [37], or 12 mV [36]. Second, the present theory is based on a rather simple (although improved with respect to previous ones) model of interface. The early publications [55, 56] merely assumed that there is a layer of a few water molecules at interface completely devoid of ions. This naive approach could explain qualitatively the increments of the surface tension, but was unable to explain the surface potential. For this purpose, Randies [43] arbitrarily postulated that this layer is accessible to only one type of ion, and hence the surface potential can build up. However, even in this case, results were qualitatively unsatisfactory. Our model overcomes these difficulties. It is formulated in a classical way where the solvent is modeled as a continuum dielectric characterized by its macroscopic dielectric properties. This approach invokes a simple, easily interpretable physical picture that does not involve many adjustable parameters. However, it applies the concept of the macroscopic dielectric response to microscopic distances where it may not have a clear physical meaning. There is a number of other aspects where our model may be vulnerable to criticism. It is obviously naive to visualize the air/water interface as a flat and sharp dielectric border between two phases. Fluid interfaces are usually treated in electrochemistry like that, but they are known to be never ideally flat because of the thermal excitation of capillary waves. Capillary waves induce roughness of fluid interfaces and may influence their electrochemical properties [57]. So, the surface is dynamic. Its dielectric constant changes with distance and so on. One could not expect from this simple model to provide an ideal agreement with the experiment. It is rather surprising that we were able to go this far predicting the surface tension with discrepancy of only about 10% and about the same for the surface potential. It is rather amazing because the model
Electric Properties of Oil/Water Interfaces
139
completely ignores many important properties of the system like the discrete nature of the solvent, dynamic variability of the interface and many other features. Any future development should include all these features and account for the detailed structure of the interface. Unfortunately, this structure is not very well known. There are a number of publications where authors try to elucidate the structure of water interface and the manner in which ions interact. One interesting question is, are the water molecules oriented at a certain degree at the interface? If they are, then they could create a permanent dipole potential at the interface that could influence the distribution of ions at the interface, and hence generate the increment of the surface potential in the presence of electrolytes. However, opinions about this issue are controversial. There are no reliable experiments to determine this orientation. Computational methods of molecular dynamics might be helpful, but different authors come to different conclusions. Wilson and Pohorille [58], as well as other authors cited in this paper, found that the molecular dipoles of water tend to lie parallel to the surface with a slight asymmetry of this distribution. This leads to a net dipole moment of the interfacial layer of water pointing to the liquid. Wilson and Pohorille [58] studied Na+, Cl" and F" ions at the water-vapor interface. They came to the rather unexpected conclusion that the free energy curves of two anions at the interface are qualitatively similar to one another, and qualitatively different from the free energy curve of the cation, which is more strongly repelled from the interface. Since Na+ and F" are about the same size, while Cl" is larger than F", Wilson and Pohorille concluded that the free energy required to move these simple, monovalent ions to within one molecular layer of the interface depends primarily upon the sign of their charges, and not upon their sizes. They ascribed this effect to the asymmetric orientational distribution of water molecules at the pure water liquid-vapor interface. Another interesting conclusion is related to the solvation of ions at the interface. The ions at the surface were found to retain their solvation shells. As a consequence of this, a bulge is formed in the surface above the ion. The size of this bulge was found to depend more on the sign of the charge of the ions rather than on their relative sizes. 3. ELECTRIC DOUBLE LAYERS 3.1. Polarizable and nonpolarizable liquid interfaces The electrical double layer at the oil/water interface is a heterogeneous interfacial region that separates two bulk phases of polarized media and maintains a spatial separation of charges. Electrical double layers at such interfaces determine the kinetics of charge transfer across phase boundaries, stability and electrokinetic properties of lyophobic colloids, mechanisms of
140
A.G. Volkov and V.S. Markin
phase transfer or interfacial catalysis, charge separation in natural and artificial photosynthesis, and heterogeneous enzymatic catalysis [8, 59-70]. The interface between two immiscible electrolyte solutions (ITIES) can be either polarizable or non-polarizable, depending on permeability to charged particles. If the interface is relatively impermeable it is called polarizable. Otherwise it is called non-polarizable or reversible. Planck [71, 72] introduced a rigorous thermodynamic concept called a completely polarizable interface, defined as "an interface whose state is completely determined by the charge that has passed through it beginning from a given instant." Planck's definition assumes that direct current cannot flow through the interface [71]. There can only be the transitive current in the boundary layer. However, the definition says nothing about how charged particles from different phases are distributed in this transitional layer. The distribution of particles in the boundary layer can have any configuration and the tales of this distribution for different particles can even overlap in the boundary layer (Fig. 19 a,c). At a completely polarizable interface charge transfer between bulk phases becomes impossible. A different concept of an ideally polarizable interface was formulated by Koenig [73] and represents a particular case of a completely polarizable interface (Figure 19 b,d). Two homogeneous phases a and p are separated by a transitional layer whose properties gradually change from the properties of phase a to those of phase p. A Koenig's surface, which is impermeable to all charged particles (both ions and electrons), is drawn in the transitional layer. This surface represents an infinitely thin and infinitely high-energy barrier for all the charged particles involved. It is this assumption that distinguishes an ideally polarizable interface from a completely polarizable surface, which makes it irrelevant that charged particles are contained in each phase. Koenig's definition cannot be described thermodynamically, because it suggests a definite structure of the interfacial region and employs non-thermodynamic assumptions about the existence of an infinitely high-energy barrier for all charged particles. The term "ideally polarizable interface" was introduced by Grahame and Whitney [74], whereas Koenig used Planck's term. Frumkin demonstrated that a completely polarized electrode, unlike an ideally polarizable one, can include systems with local adsorption charge transfer. Although local charge transfer exists at the interface and Koenig's surface cannot be drawn, there is no actual charge transfer between the phases, so the electrode is completely polarizable according to Planck, but non-ideally polarizable according to Koenig. If phases a and P contain at least one common ion that can freely pass across the interface, the interface may be called reversible or nonpolarizable (Fig. 20). Although the latter term is commonly accepted, it does not correctly describe the state of the system. A reversible (non-polarizable) interface can be partially polarizable or completely non-polarizable.
Electric Properties of Oil/Water Interfaces
141
Figure 19. The structure of a perfectly polarizable interface between two immiscible liquids a and p (a, b, c, d) and an ideally polarizable interface.
142
A.G. Volkov and V.S. Markin
Fig. 20. Scheme of the structure of a nonpolarizable ITIES (a) and the distribution of the concentrations of ions (b) capable of passing from one phase to another.
Fig. 21. Distribution of the electric potential at ITIES.
A completely non-polarizable interface containing at least one common ion can pass a high current in either direction without causing a deviation of the
Electric Properties of Oil/Water Interfaces
143
interfacial potential difference from the equilibrium value. Although in practice we encounter neither ideally polarizable or completely nonpolarizable interfaces, under certain conditions the properties of some interfaces are very close to being ideal. 3.2. Verwey-Niessen model Verwey and Niessen first described the electrical double layer at ITIES as two non-interacting diffuse layers, one at each side of the interface [75]. Both solvents were assumed to be structureless media with macroscopic dielectric permittivities, and potential distribution in the electrical double layer was defined by Gouy-Chapman theory [76-78]. Gavach et al. [79] extended the Verwey-Niessen model by introducing an ion free transition layer at the interface between two immiscible electrolyte solutions. This is directly analogous to the compact layer (or the inner Helmholtz layer) in classical electrochemistry. Stern theory was extended to ITIES, and the final model is referred to as the modified Verwey-Niessen model (MVN). In the MVN model, the electrical double layer consists of two diffuse ion layers back to back, which produce a compact inner layer between the two phases (Fig. 21). Dielectric permittivity of the medium at any point in the diffuse layer is assumed to be constant and equal to the bulk phase value s. The compact layer or inner Helmholtz layer is located between -8" and +8P. In a more detailed analysis the dielectric permittivities in both parts of the compact layer are s^ and z\. In the MVN model, the boundary of the compact layer (called the outer Helmholtz plane) is at the distance of closest ion approach to the interface. For ions with different radii, a few outer Helmholtz planes can be introduced as necessary. In the absence of specific adsorption the compact layer located between two nearest outer Helmholtz planes disposed on different sides of an ideal interface is usually called the ion-free layer. The maximum electrical potential in the compact layer (A"<|>.) includes a dipolar potential (Apg), which is shown schematically as a narrow region at the sharp interface. A dipolar layer can be located not only in the compact layer but can also occupy part of the diffuse layer. The amplitude and sign of Apg can differ from the total interfacial potential. Figure 21 illustrates four possibilities for potential distribution at ITIES. Generally speaking, the dipolar potential depends on the total interfacial potential, A°<j). The interfacial potential difference consists of the sum of the potential drops:
144
A.G. Volkov and V.S. Markin
£$ -
•
(136)
Charge density at each side of the electrical double layer is usually called the interfacial free charge density, and depends on the model chosen for the interphase. In an ideal polarized electrode, the free charge equals the total thermodynamic charge in the Lippmann equation of electrocapillarity if all components of an interface have the same charges as they have in the bulk phases. We will denote surface charge density in the diffuse layers as qd, qd, and in the compact layer as q^. From electroneutrality of the interphase we have: <£+<7*+?5 = O.
(137)
At the potential of the free zero charge (fzcp) the compact layer has not adsorbed ions and the charges of the diffuse layers are equal to zero: < j ) X = 0 and A ^ = A « ^ .
(138)
The dipolar potential A^h also needs the index fzcp due to its dependence on total interfacial potential. Usually the dependence is weak and equation (137) is used to determine the dipolar potential. From the Gouy-Chapman theory for a 1:1 electrolyte it follows that 2z0RT q.= Hd
F
. F$d SKSinh—— ,
(139)
2RT
where K, the Debye length, is: J2F2c° K= J ^ -
(140)
\ se o i?r and c° is the bulk electrolyte concentration in the corresponding phase. If a Helmholtz layer does not have charges, potential drops in the two diffuse layers have a simple relation: saKasinh^ + 8pKpsinh^ =0 . 2RT 2RT
(141)
Electric Properties of Oil/Water Interfaces
145
The relation between potentials of two diffuse layers depends on a parameter a/p _
S K
_
£
C
which is the reverse ratio of the diffuse layer capacitance. There is a direct proportional dependence between ^ and <j>^ with coefficient r|a/p when the potential drops are small:
4>5 / 4>5 = - n a / p .
(143)
From equations (142) and (143) it follows that the potential drop (<j)d) in the diffuse layer will be less if the dielectric permittivity (s) or electrolyte concentrations (c°) are increased or, if the diffuse double layer capacitance (Cd) is increased. At the contact between aqueous and organic phases virtually the entire drop of the potential occurs in the organic phase. However, with increasing interfacial potential the situation changes dramatically. When potentials are large enough equation (141) can be presented in a different approximation: RT C = " ^ - — sgn^.lnTf /p .
(144)
Here the coefficient of proportionality is equal to one. From equations (136) and (141) it is possible to find the dependence of the potential drop in both diffuse layers Ap<j>- A^<> | A on potential drops ^ and ^ in each diffuse layer: , ^ and
=
RT F
(s V n
/ s a K«) + exp[F(A"p4> - A ^ , ) / 2RT]
( s P K 3 / s a K a ) + exp[-F(A a p (t)-A a p^)/2i?r]
(145)
146
A.G. Volkov and V.S. Markin
=
"
RT (£ K K a /E p K p ) + e x p [ F ( A ^ - A > J / 2 / ? r ] _ F n (E a K a /s [ 5 K p ) + exp[-F(Aap<|)-Aap(t)J/2JRr]~
_
RT CM
V K )+
ln F
(B p K p /(s°K°) + e x p [ F ( A ^ - A ^ J / 2 J ? r ] • ( 1 4 6 ) ( E P K P/( 8 « K ") + exp[-F(Aap<|)-Aapfi)/2J??r]
The spatial distribution of potential near the interface is accompanied by changes in electrolyte concentrations that produce surface excesses of ions. In each phase the surface excess T\ of an ion i can be divided into two components in the inner layer F* and in the diffuse layer Ff :
r; = r*+r^
(HV)
The charge of each contacting phases is equal to:
.
(148)
For a binary electrolyte the ion excess in the diffuse layer can be written as
r
'-V [o ^-Mr ) - 11 = ^ [ J 1 -fei? + iidfe y " 11 - (149)
where c° is the electrolyte concentration in the bulk phase. The interfacial capacitance C consists of the capacitance of the compact layer, C/,, and capacitance of two diffuse layers Q and Cf,. Differentiating equation (148) with respect to charge qa of phase a, and assuming that the drop of the potential in the compact layer Aa^h does not depend on an electrolyte concentration, we have: ^ =^ ' . +^ + C(q,c) Ch{q) Cad(q,c) Cpd{q,c)
(150)
Diffuse layer capacitance depends on potential or surface charge q and electrolyte concentration: |2s n sc Ffo, F I rCOSh = y^U 2RT 2RT^°ERTc
Cd = F
+q
7 •
(151)
Electric Properties of Oil/Water Interfaces
147
3.3. The electrocapillary equation A comprehensive thermodynamic theory of the electrocapillary phenomena at polarizable and nonpolarizable liquid interfaces was developed by Markin and Volkov [64-69] using Hansen's method [80]. Electrochemical processes occurring at the interface between two immiscible liquids are traditionally described based on Gibbs thermodynamics of surfaces. However, Hansen's method, by extending and generalizing Gibbs method gives us a better understanding of the nature of interfacial phenomena and provides us with an improved method for describing them. Consider two phases in contact, a and P, with rh neutral components, whose chemical potentials are designated by \ih, rc types of cations and ra types of anions with chemical potentials Uj. Using Hansen's reference system and choosing the solvents a and p as reference substances one can write down Hansen's rendition of Gibbs adsorption equation: d
J = S(«fi)dT
+r
Ufi)dP- Z r *(a, P )4^ - Z r / ( « . P ) J ^ . fcta.p
(152)
/
Here ^ a P ) , t*ap) and rh(<Xip) are the surface excesses of the entropy, volume and substance h in the reference system, when the absolute excesses a and p are zero. The terms corresponding to different components of the system are divided into two summations, one for neutral and the other for charged particles, the reference substances a and P being excluded from the sum of neutral particles. Individual components of the system, both charged and neutral, can be present either in both phases or in only one. The electrocapillary equation in its final form, which includes the electromotive force of the measuring cell can be written as:
dy = -(ss(afi)-(VzrF)Qas'Ra
-(Vzm,F)Q\)dT
+
+ (i; r (i/z j ,F)e"v;-(i/z,,F)^;)*-^r %pl rf f i r hzafi
-X(i/v;,jr t(a , P) 4i /t - Za/vy )r,
}*}'
1/v
" E ( rjr m(ap) <% m -X(1/vL')r/(a,p)^/,,,' -
-(l/^,v:,r)(^ril(aP) + 2z/,
zJ^»M*-QadE$&
148
A.G. Volkov and V.S. Markin
Note the symmetry of the last term with respect to the transposition of the a and (3 indexes: QadE;H] = &dl%$
.
(154)
Equation (153) contains rh + rc + ra - 1 independent differentials of intensive variables, whose number is equal to the number of the degrees of freedom of the system. The impedance or electrocapillary properties of the interface between two immiscible electrolyte solutions can be measured with the following cell:
RE1
1 H2O(a) M+C1" a, AAga4>REi
2 A + B"(p) oil a2 A >
3 A+C1" (a) H2O a3 A$D
RE2
(I)
-A A g a KE2
Silver/silver chloride electrodes are usually used as the reference electrodes (RE), water as phase a , and nitrobenzene or 1,2-dichloroethane as phase p . T h e interface between phases 1 and 2 is the polarized oil/water interface serving as a working electrode (interface). T h e common cation A + is usually the terabutylammonium ion, and B" is tetraphenylborate or dicarbollylcobaltate. The potential difference measured in the cell (I) consists of the s u m of two interfacial potential drops:
-E = A»4> + AM> W1 - A > f f £ 2 + A(»D .
(155)
A<|>D is the Nernst-Donnan
potential difference between the nonaqueous fraction and the aqueous solution of RE2 with the common ion A+: RT
A^=-A
a
o
pC+7Tln^,
(156)
and Aap<|>°, is the standard potential difference for A+ known from literature data or calculated theoretically. By suitable selection of the Cl" concentrations and ionic strengths of the electrolytes into which Ag/AgCl - electrodes are immersed, we obtain:
Electric Properties of Oil/Water Interfaces
RT
149
n
A?4>™-Aj
(157)
Substituting (156) and (157) into (155) we obtain
d£?$ = -dE-
a
dA ^D
HT n +-—dln^ r ax
(158)
With equation (158) the electrocapillarity equation at the flat oil/water interface can be written as follows:
dy=-Yjrid»,-QadE + Qa^dln^^ i
r
(159) ay
where Q is the charge which must be supplied to each side (a or p, correspondingly) of the polarizable interface when expanding it by a unit area in order to maintain the potential difference between the phases [81]. At the ideally polarizable interface, the ions of each group are separated by the Koenig barrier, which rules out a possible overlap between the groups (Fig. 19). In this case the free charges qa and qp of the phases are equal to the thermodynamic charges: q a = Q a = -q p = - Qp.
3.4. Zero-charge potentials Potentials of zero charge of the interface can be found reliably by the same independent methods that are used at the metal/water interface. These include finding the differential capacitance minimum of the electric double layer from electrocapillary curves, with a flowing-electrolyte electrode, vibrating boundary method, radiotracers, or by measuring the second harmonic generation. Potentials of the zero free charge at nitrobenzene/water and 1,2dichloroethane/water interfaces, obtained from the differential capacitance minimum of the electric double layer in solutions of a surface-inactive electrolyte do not necessarily correspond to the potentials of thermodynamic zero charge. They can depend on electrolyte concentration when the capacitance of the compact layer is affected by surface charge as a result of nonlinear double-layer properties.
150
A.G. Volkov and V.S. Markin
The potential difference between two immiscible electrolyte solutions can be written as the sum:
Aap4> = A-
(160)
where Aap(j) and <|)d are the potential drops across the compact and diffuse layers, respectively. In the linear approximation and in the absence of specific adsorption, the electric double layer is equivalent to three capacitors in series:
d^ dq
_ dAa^h dq
d$l +
dq
d^ +
(161)
dq •
One can use the following approximation when modelling the electric double layer:
dq1
dq2
dq2
dq2 '
where j-p
Z
i ( ^ ) = — [y{x + d)+\y{x1
rp
d)\-——
-. X + d
\\[>(x')dx' l U
(163)
x-d
We find the position of the minimum in the C vs. Aap<)) curve as:
A>"=Aap(t);'+(t)«"+tf"=O.
(164)
It follows from the Gouy-Chapman diffuse-layer theory that if qa= q$= 0, then((|>°)" = (
Electric Properties of Oil/Water Interfaces
151
the capacitance curve on electrolyte concentration. Therefore, a correct determination of the zero free charge potential for the nitrobenzene-water system in the presence of a binary surface-inactive electrolyte is possible only at base-electrolyte concentrations less than 0.01 M. For the water-nitrobenzene system, this quantity Awnb<j> pzfc = -0.29 V and for the water-1,2-dichloroethane system AWde<)>pZfc= -0.27 V. For many systems the maximum of the electrocapillary curve is located in the region of ideal polarizability of the electrode, and the maximum potential corresponds to the potential of zero total (thermodynamic) as well as zero free charge. At the mercury/water interface the region of ideal polarization is a few volts when soluble mercury salts are absent (2.2 V), but at the interface between two immiscible liquids the region is about a tenth of that value. Under conditions where current flow does not significantly alter the compositions of the liquid phases and where the equilibrium at the interface is not disturbed by the current, the electrocapillary curves yield the potential of zero total charge. However, under realistic conditions where the region of ideal polarization is narrow, the maximum of the electrocapillary curve corresponds to the potential of zero total charge, rather than zero free charge. This closely resembles systems consisting of amalgams and liquid electrolyte solutions that are described by Frumkin [82]. The structure of the electric double layer at ITIES has been investigated using such common electrochemical methods of capacitance measurements as impedance, galvanostatic and potentiodynamic techniques. More recently, it has become possible to measure electric double-layer capacitance at the interface between two immiscible electrolyte solutions using a four-electrode or twoelectrode potentiostat. Commonly studied systems are nitrobenzene/water and 1,2-dichloroethane/water, in which the organic phase has a relatively high dielectric permittivity and a high dissociation constant. In typical investigations, cyclic current-potential curves are usually determined before impedance measurements in order to find the potential range where the contribution of faradaic impedance is low. The electric double-layer capacitance is measured in the potential window between extremes in the cyclic voltamogram. Over this range of potentials, the water/dichloroethane interface has properties close to those of an ideally polarizable electrode. A "diffuse" picture of the compact layer has been considered [83]. According to this hypothesis adsorbed ions can penetrate into the compact layer, which consists of solvent dipoles, analogous to nonlocalized electron gas penetration from a metal electrode to an aqueous electrolyte solution. Here a penetrating ion can act as a hydrophobic anion in the organic phase and potential distribution §(x) near to the point of zero charge can be calculated using the Poisson-Boltzmann equation:
152
A.G. Volkov and V.S. Markin
If X<X2W,E = EW: f = (K,) 2 (*-Ay)
(165)
If x " Z < x < x ° 2 r g , z = z h :
(166)
If x>x^,z
(|)" = ( K J 2 ( ) )
= zorg: V = (Karg)\
(167)
In each of these three areas, the solution can be written as: <j)(x) = A+e" + A_e-",
(168)
where six coefficients A+ and A. can be determined from the six boundary conditions:
K-oo) = AWOJ (j)(oo) = 0
(169)
= S/i(t)'(x2w+0)
8^'(xr-0) = 8o^(x^+0). As a result the equation for the capacitance of the electric double layer is:
, (170)
fl + _E*!S*_>|exp(KA6) + fi__!*!S*_l eX p(_ KA5 ) V
E
orgKorg
)
^
E
orgKorg
)
fl + ^ ^ Y l + -^ 1 ^]exp(K / ,5) + |l-^ K ^lfl-^ K ^lexp(-K / ,5) where 8 = xfx - x j . It should be noted that the size of a hydrophobic penetrating anion is large in relation to the compact part of the electrical double layer and it is not clear if the Poisson-Boltzmann equation can be used in this case.
Electric Properties of Oil/Water Interfaces
153
3.5. Specific adsorption at liquid-liquid interfaces Ions can be adsorbed specifically if the main contribution to their interaction with the interface (ions, dipoles) is caused by non-coulombic shortrange forces. Specific adsorption cannot be explained using only the theory of diffuse double layer. Specifically adsorbed ions penetrate to the compact layer and form a compact or loose monolayer. The surface passing through the centers of specifically adsorbed ions is usually called the inner Helmholtz plane. If several kinds of specifically adsorbed ions are present, each ion can have its own inner Helmholtz plane. It was found while studying the mechanism of electron transfer across the interface between two immiscible liquids, that specific anion adsorption occurs at the octane/water interface in the presence of metalloporphyrins. This adsorption increases in the order of Cl", Br", I", and is caused by coordinative bonding of the anions as ligands of the porphyrin metal atoms. Specific ion adsorption at the polarizable nitrobenzene/water interface containing monolayers of phosphatidylcholine, phosphatidylserine, octaethylene glycol monodecyl ethers, tetraethylene glycol monodecyl ether, and hexadecyltrimethylammonium (HTMA+) were studied in detail [84]. HTMA+ exhibited no specific adsorption in the potential range where the aqueous phase was positive. However, a strong adsorption occurred in the potential range, where the electric potential in the aqueous phase was negative with respect to the nitrobenzene. Another example of specific ion adsorption was discussed in terms of the formation of interfacial ion pairs between ions in the aqueous and the organic phase. The contribution of specific ionic adsorption to the interfacial capacitance can be calculated using the Bjerrum theory of ion-pair formation. The results show that a phase boundary between two immiscible electrolyte solutions can be described as a mixed solvent region with varying penetration of ion pairs into it, depending on their ionic size. The capacitance increases with increasing ionic size in the order Li+ < Na+ < K+ < Rb+ < Cs+. Yufei et al. [85] found that significant specific ion adsorption occurs at the interface between two immiscible electrolytes and the potential dependence of the capacitance is strongly influenced by the adsorption isotherms due to the interfacial ionic association. When there is specific adsorption of ions dissolved in phase a the condition of electroneutrality can be written as: q*=-qa=-{q*+qad)
,
(171)
where q, is the charge of the inner Helmholtz plane. The separation of qa from g° and qad cannot be done without introducing a model of the interface.
154
A.G. Volkov and V.S. Markin
Although capacitance of the interface can be calculated as before using the equation:
dq
dq
dq
dq
^
'
The second term in the right side of the equation is not the diffuse layer capacitance since the charge of a diffuse layer in phase a is equal to
€ =-%-Yu(lTJ .
(173)
j
The diffuse layer capacitance is equal to:
c
<174>
«'-%-
and one can write
C-j=(C')-l+{C%)-1
1+ V
d
,,
+(C*)~'.
(175)
/
The drop of a potential in the compact layer depends on the surface charge qa and the charge of the inner Helmholtz plane q"-J. It shows that
or
dq> = crWl,
Hw-'iw
d
^ •
(177)
Here the index q"J means that all q"J are constant except one. In the absence of specific adsorption the capacitances of the interface, compact and diffuse layers are always positive. The capacitance of the compact layer in the presence of specific adsorption can be either positive or negative.
Electric Properties of Oil/Water Interfaces
155
3.6. Adsorption isotherm Traditional models for calculation of adsorption isotherms are based on the assumption that surface-active compounds at the interface can substitute for adsorbed molecules of one solvent, but cannot penetrate the second phase [34, 61, 86, 87]. Although these models are useful for metal/water interfaces, recent interest has focused on the surface chemistry of amphiphilic compounds which can penetrate both phases and replace adsorbed molecules of both solvents, for example water and oil. We present here a theoretical analysis of the generalized Frumkin adsorption isotherm for amphiphilic compounds [61]. The interface between two immiscible liquids may be considered to be a surface solution of surfactant in a special kind of solvent. In order to calculate the entropy of such a solution, we will adopt a simplified lattice model and use lattice statistics, a widely used method for describing surface solutions. The transition from three-dimensional (3-D) to two dimensional (2-D) geometry may cause errors in statistical formulas, if some peculiarities of 2-D solutions are overlooked.The main difficulty when dealing with a monolayer of a surfactant, one can consider this monolayer as a 2-D system. The solvent molecules do not form a monolayer, but rather a multilayer. Therefore the transition from 3D- to 2D-geometry should be specified. Consider molecules of both solvents which are substituted by a surfactant (Fig. 22). Suppose that these molecules can be assembled into columns consisting of m 0 molecules of oil and m w molecules of water. Suppose that one column of oil molecules matches the n w molecules of water. This match of 1 oil column and n w water columns will be considered in what follows as a quasi-molecule of solvent Q. These quasi-molecules constitute a "monolayer" of solvent. They consist of m 0 oil molecules and n w niw water molecules. Designate the molecules of surfactant in the bulk as A, and in the monolayer as B. At the interface aggregation of surfactant molecules can take place, rA <=> B, such as dimerization of porphyrin or pheophytin molecules at the octane/water interface. Let the surfactant B replace p quasi-molecules at the interface. Therefore one can write pQ + rA = B+p (oil) +/?« w (water). The chemical potentials for (178) are
(178)
156
A.G. Volkov and V.S. Markin
Fig. 22. The structure of the oil/water interface with adsorbed monolayer of amphiphilic surfactant. Taking the 2-D solution as ideal, we have:
y£> =u°Q + R T l r i $
(180)
and
ufs = u RS + R T lriXS.
(181)
In the bulk phase we have
A = u ? + R T l n XhA,
(182)
uh0 = u ? + RTln X^ ,
(183)
u^ = u° w b +RTln X^ .
(184)
Electric Properties of Oil/Water Interfaces
157
In all these equations X designates the mole ratio of corresponding substances. Substituting these equations into (175), one obtains: p& + ry?* - n 0 / - p\x°f - pnywb + RT In
/
X
^
= RT In -%-
(185)
Using the standard Gibbs free energy of adsorption:
AIG° = ft - rtf + ptf + pny* - n%
(186)
one obtains the adsorption isotherm: J^=(^l exp(-^) (X^)" (X*)"(Xt)'"1" RT
(187)
We considered the 2-D solution of surfactant B in the solvent of quasi-particles Q, in which the mole ratios were defined as N
Y .«
N
.y»
B
X =
Q
X
' V7iF-' °-N-B+Nl
0 88) (188)
Some authors prefer another set of definitions when real particles in the interface are considered. The equation for this state with real particles A, O, W becomes:
X
>^fe'
Xs
K
Q
N'A+N'O
+
K'
(190) x,
_
" (191)
K N'A+N^+N'W'
and we can obtain:
O89)
158
A.G. Volkov and V.S. Markin
X;=—^—;X!;=—^—.
(192)
The adsorption isotherm can then be presented in the form
-^MX*
+XT1
/X'^
=
s
b
{^ Qy
exp(-^^)
b
(x onx H,r»
pv
RT
(193)
In the past the adsorption isotherm was presented in terms of the fraction 6 of the surface actually covered by the adsorbed surfactant. If we introduce r| as the ratio of areas occupied in the interface by the molecules of surfactant and oil, the mole fractions in surface solution can be presented as follows: B
© 0 + n(i-0)
s =
°
nO-Q) ©+Ti(i-©)
(194) V ;
The adsorption isotherm takes the form: 0 p
n C\ —Pi\p V \l U 7
[Q + , ( l - e ) r
=
/ ^ ^ (Y \P(Y V"' \^o) y-^w)
exp(-^ RT KI
(195)
In this isotherm the mole fractions X*,X*,X* of the components in the bulk solution are presented. In the general case, they must be substituted with activities: 0
npC\—CV\p
[0+ ,(l- Q ) r- = y*\
exP(-^ RT
K1 J \ao> \aw) (196) If the molecules B can interact as pairs in the adsorbed layer and the energy of each new particle is proportional to its concentration, then their chemical potential, \x*B, instead of equation (181), should be presented as:
V U
U
(nb\p(nb\p"'-
VSB = \xf + RT In X-2aRTX^
( 197 )
where a is so called attraction constant. Then after some algebra we obtain, instead of Eq. (196), the isotherm [61, 88]:
Electric Properties of Oil/Water Interfaces
F
^(l-©)"
'
(x*)"(x*f)p"«
159
i?r
v
'
Recall that ri was introduced as the ratio of areas occupied in the interface by the molecule of surfactant to the same of oil and p was introduced as the number of columns of oil, which could be supplanted with one molecule of surfactant. Therefore, p is a relative size of the surfactant molecule in the interfacial layer. It is reasonable to suppose that: T\=P
.
(199)
If the concentration of surfactant in the solution is not high and the mutual solubility of oil and water is low, then we can use the approximation x ! = x * = I s o that the general equation (196) simplifies to: 0 [ / 7
-^"l)0]
ex P (-2a0) = (X*)' e x p ( - ^ - )
(200) This is the final expression for the isotherm that we will call the amphiphilic isotherm. It is straightforward to derive classical adsorption isotherms from the amphiphilic isotherm (200): 1. The Henry isotherm, when a — §,r=\,p=\,
0 « I:
© = X*flexp(-^). 2. The Freundlich isotherm, when a = 0, p= I,® 0 =(X*)rexp(-^).
(201) «l: (202)
Kl
3. The Langmuir isotherm, when a = 0, r = \,p= I: - ^ - =X*exp(-^). 1-0 RT 4. The Frumkin isotherm, when r— l,p= 1:
(203)
160
A.G. Volkov and V.S. Markin
-^— exp(-2a©) = X"a exp(- ^ - ) .
(204)
Therefore, the amphiphilic isotherm (200) could be considered as a generalization of the Frumkin isotherm, taking into account the replacement of some solvent molecules with larger molecules of surfactant. Of course, the amphiphilic isotherm includes all the features of the Frumkin isotherm and displays some additional ones. To elucidate them, it will be convenient to change the variable x* to the relative concentration y= X* / X ^ ( 0 = 0.5), where x*(0.5) is the concentration corresponding to the surface coverage 9 = 0.5: y=
:
exp(a-2a&)
(205)
This equation gives the coverage fraction 8 as a function of relative concentration y, while a and p are the parameters of this isotherm - the first being the attraction constant and the second, the size of surfactant. These parameters play an important role because their effect on the shape of amphiphilic isotherm is very strong. Amphiphilic isotherm (200) analysis can be used for the determination of the interfacial structure. An amphiphilic molecule, which consists of two moieties with opposing properties such as a hydrophilic polar head and a hydrophobic hydrocarbon tail, should be used as an analytical tool located at the interface. Pheophytin a is a well-known surfactant molecule that contain a hydrophobic chain (phytol) and a hydrophilic head group. The value of p less than 1.0 indicates that adsorbed molecules of w-octane are parallel to the interface between octane and water. Substitution of one adsorbed octane molecule requires about 4-5 adsorbed pheophytin or chlorophyll molecules. These are supported by molecular dynamic studies in the systems decane/water, nonane/water and hexane/water. The structure of both water and octane at the interface is different from the bulk. Adsorbed at the interface octane molecules have a lateral orientation at the interface. 3.7. Image forces Any charge near phase boundary interacts with the interface due to the different polarization of two phases. The force of this interaction can be found by solution of the Laplace equation, but it is more convenient to use the method of images. This method employs a fictitious charge {image) which, together with
Electric Properties of Oil/Water Interfaces
161
the given charge creates the right distribution of electric potential in the phase [8, 90]. All ions interact with all images giving rise to the so called image forces. The image forces are largely responsible for both positive and negative adsorption. Important contribution to this effect is given by the change of the size of solvation shells of ions in the boundary layer. This has an impact on the planes of closest approach of ions and dipolar molecules to the phase boundary. The electrostatic Gibbs free energy for an ion in the vicinity of a boundary between two liquids with dielectric constants sj and E2 is determined by the Born ion solvation energy and by the interaction with its image charge [8]. Dealing with the energy of image forces and with the interactions of charges at the oil/water interface, approximate models of the interface are often employed, which are based on the traditional description of the interface between two local dielectrics. In the oil/water system, the force of attraction (or repulsion) of charge in the organic (P) phase with its image in the aqueous phase (a) sitting on the same axis perpendicular to the dividing plane is given by: F(h) = -^^S.-^-j,
(206)
where h is the distance from the interface. If ea > sp, the charge in the nonpolar phase is attracted to its image, but if za < 8p there is repulsion between the charge in an organic phase and its image. From equation (206), it follows that charges in the organic phase are attracted to the oil/water interface. Image forces attract the diffuse layer on the organic side, making it thinner, and repel, thus thickening the aqueous diffuse layer. The Kharkats-Ulstrup model [40, 41] incorporates the finite radius of an ion a, which is assumed to have a fixed spherically uniform charge distribution and can continuously pass between the two phases. For a point charge at long distances from the interface, electrostatic Gibbs free energy can be written as
AG =
j £ ^ L f 4 + 5iZft2£l 327ieoeaa^
(207)
ea+sp h )
where a is an ionic radius, h is the distance from the interface and ze is the charge of an ion. The potential <> j for the spherically-symmetrical charge distribution is:
162
A.G. Volkov and V.S. Markin
Z6
°
+—^
4jiE0ear1
£g l
~£p
(in region a)
4jre 0 E o #' 1 e a + Ep
• —^ 4TIS 0 S^
^L_ 8a+sp
(208)
(in region P)
Using standard procedures of calculations, Kharkats and Ulstrup obtained equations for the electrostatic part of the free Gibbs energy of a finite size ion [40, 41]. When h> a AG
&of \ 327i808aa|
UaSi\2a Ua-z^\ 2 [ s a + e j / ? [e a +E p J \\-(2hlaf
^ ^ 2 / z + alj- (209) 2h 2/z-aJj
The first term in (209) is the Born solvation energy, and the second is the interaction with the image. For a charge located in region p, the electrostatic free energy is obtained from (209) by exchanging s a <^> ep Kharkats and Ulstrup [40, 41] obtained the expression for the electrostatic free energy in region 0 < h< a:
AG=325T£ ^ £ a {[V f 2a)3 l Vs + ^+EYJV4 - ^a)+ 0 a
a
(za-zA2U\+h/a)(l-2h/a) E
Ua+ JL
\ + 2h/a
p
^ a Ju2hVJ 2h V
a) \\
(ze0)2 f 2sp Yf
hV
167i£0Epa l v s a +E p J V a)
If h - 0 and center of an ion is at the interface, electrostatic Gibbs energy will be equal to (ze \2 AG(h = 0) = ^ ^ . 47i8 0 (s a +s p )a
(211)
Image forces play a significant role in electric double layer effects. The excess surface charge density is: K
q = ec0 ]{exp[-*g(h)/kBT]-l}dh, K
(212)
Electric Properties of Oil/Water Interfaces
163
where Ag(h) = AG{h)--^—.
(213)
AG{h) is the electrostatic contribution to the Gibbs energy of solvation, and hd and hc are the distances of closest approach of the anion and the cation. If image forces are taken into account, the diffuse layer capacitance can be calculated by the equation Cd = Cf exp[Ag(h,)/2kBT),
(214)
where C'f is the Gouy-Chapman diffuse layer capacitance without image terms. As noted by Kharkats and Ulstrup [40], equation (214) always gives diffuse layer capacitance corrections toward higher values. The first step to the statistically correct Gouy-Chapman theory for the diffuse double layer uses a restricted primitive electrolyte model. This model considers ions to be charged hard balls of identical radii in a structureless dielectric continuum with constant dielectric permittivity. There are three main approaches to creating a statistical theory with this model. The first approach uses the Gouy - Chapman theory, there the average electrical field in the diffuse layer is calculated from the Poisson equation. The structure and physical properties of the double layer are then calculated from the restricted primitive electrolyte model. As a result, the modified Poisson-Boltzmann theory (MPB) was developed. The second statistical approach incorporates correlation functions and integral equations from the theory of liquids. In this case, the well-known uniform Ornstein-Zernike equations were modified to calculate the ion distribution function near a charged interface. Modified in this way, the Ornstein-Zernike equations became nonuniform and were solved using hyperneted chain approximations (HCA) incorporating a mean spherical approximation for correlation functions (HCA/MSA). A third statistical approach used to describe the electrical double layer was achieved by using the integral Born-Green-Ivon equations. These correlation function approaches resulted in the theory of ionic plasma in semi-infinite space. The next advance in describing the double layer was achieved by upgrading the restricted primitive electrolytes model to a "civilized" or "non-primitive" model. This was accomplished by addition of hard spheres with embedded point dipoles to the restricted "primitive" model.
164
A.G. Volkov and V.S. Markin
This theory takes into account the presence of long-range Coulomb interactions, image effects and external electrical fields. The new "non-primitive" model became the basis for the theory of ion-dipole plasma. The serious mathematical difficulties encountered in describing the "non-primitive" models necessitated the development of less statistically rigorous intermediate models. In these intermediate models, the first monolayer of solvent molecules at the charged interface was considered as a complex of discrete particles and an electrolyte outside this layer was described by the Gouy-Chapman theory. This intermediate approach was able to provide good theoretical agreement with experimental results. The concept of constant dielectric permittivity in the double layer is also of concern. Different factors such as high ionic concentration or strong electric fields can affect the dielectric permittivity. In addition, description of the dielectric properties of solutions by a single parameter, the dielectric permittivity, also came under criticism and led to the development of nonlocal electrostatics. Non-local electrostatics takes into consideration discrete properties of solvent molecules. It assumes that fluctuations of solvent polarization are correlated in space. This means that the average polarization at each point is correlated with the electric displacement at all other points and, therefore, uses the solvent dielectric function which couples polarization vectors throughout the solvent. 3.8. Modified Poisson-Boltzmann (MPB) model The popular Poisson-Boltzmann equation considers the mean electrostatic potential in a continuous dielectric with point charges and is, therefore, an approximation of the actual potential. An improved model and mathematical solution resulted in the modified Poisson-Boltzmann (MPB) equation [91]. This equation is based on a restricted primitive electrolyte model that considers ions as charged hard spheres with diameter d in a continuous uniform structureless dielectric medium of constant dielectric permittivity s. The sphere representing an ion has the same permittivity s. The model initially was developed for an electrolyte at a hard wall with dielectric permittivity ew and surface charge density a. The charge is distributed over the surface evenly and continuously. This theory takes into account the finite size of ions, the fluctuation potential, and image forces in the electrolyte solution next to a rigid electrode, but it still an approximation. The MPB theory begins with the Poisson equation for the mean electrostatic potential \\> in solution:
Electric Properties of Oil/Water Interfaces
d2y
-*
r =
1
v
165
0
-wZ**»,g«W.
(215)
where s0 is the electrical constant, z\ is the charge number of ion i, e0 is the elementary charge, n° is the bulk number density, and go\(x) is the wall-ion distribution function representing the ratio of the local ion density to the ion density of the bulk electrolyte solution. The distance x is measured from the electrode so that the distance of closest approach of an ion / is d/2. According to Kirkwood the distribution function is given by the following equation:
go,X*) = S,X*)exp[- —z,.e o v|/(*) + T|,-].
(216)
The factor tfa) accounts for the excluded volume of the ion and r\, is the fluctuation potential that takes into consideration the ionic atmosphere around an ion created by other ions. The fluctuation potential can then be presented as n =
' ~~kf -T W * ' -^.b)d(z,eo),
(217)
where
*:=i-7
£ l 5 L
-,
(218)
<> | j is the fluctuation potential created by the ion / and its atmosphere, r is the distance from the center of the ion. The difference <>|* is the potential created by the ionic atmosphere only and
}
ifr>^,x>0,
(219)
where/is the image factor determined by the difference between two dielectric constants of the electrolyte solution and the electrode:
166
f
A.G. Volkov and V.S. Markin
£
~ Sw
/ =
.
(220)
The local Debye-Huckel parameter at a given point is:
£Q£/C-/
,
In equation (219), 5 0 is a constant and r* is the distance from the mirror image in the interface. Hence, equation (219) takes into account both the ionic atmosphere and image forces. If r < d and x > 0, the equation for the fluctuation potential should be written as: d2\\>
vy2,
Ib^'
r
(222)
The term Cfe) can be presented as a power series expansion in the bulk density:
;,(x)=i+5(x)+^-5>;,
(223)
where X+
\[(x'-X)2-d2]goj(x<)dX',
D/
\
V
^<x<^,
0 rf/2
5(x) = ^ ^ L ,
(224)
Now we can combine these equations and obtain the modified PoissonBoltzmann equation for the diffuse layer:
^dx = —EL Ex ;^ o « , ° C , W e x[p j kT - ^ U[_ ( V ) + -87t£ ^ - eaf ( F - F o ) jJJ l x>± (225) 2 O
o
Electric Properties of Oil/Water Interfaces
167
Since the distance of the maximal approach of ions to the interface cannot be less then d/2, the usual linear dependence of potential in the compact layer can be introduced: ¥(*)
= V(0)
+
* « .
(226)
Additional functions in equation (225) are:
C,(*) = l+ ^ & 0 + * 2 > , 0 '•
m^)
"{ \(x<-x)2-a>}goi(x')dx\
(227)
(d A max I — ,x-d I
= —[w(x + d) + \i(x-d)]—-— 2
1
j\\i(x')dx'. "
(228)
x-d
a
2(l + KX)exp(-Kti) + 2(l - Kt/)exp(-icc) + 2K266C jr" 1 exp(-Kr)t/r + <
F=
X
+ — [exp(-K) + 2K(CI - x) exp(-icc) - exp(-K(2x + d))] «*) r = 2[exp(-KC?) (l + K(X + d)) + exp(-Kx)| -
>-,
-<x
— [2o: exp(-KJc) + (1 + Krf)(exp(-2icc) - 1 ) exp(-Kd)] .
[KX
F=
(iJrtW-fsy)>
52 — 1
K^
ifx< d
~-
exp[K.(d - 2x)]sinhKc/
(K J cosh Kd - sinh K J ) 5 2 ~ (l + K^)sinhKJ • = lim K2 = - ^ _ V z,2«°.
(230)
(232) (233)
(234)
168
A.G. Volkov and V.S. Markin
The boundary conditions for these equations are: 1) \\i(x), \\i'(x) -> 0, if x —» oo, 2) \|/'(0+) = -5/S 0 E,
(235)
3) \\i(x), \\i'(x) are continuous at x > 0. This set of equations completes the mathematical formulation of the problem, but the equations are very cumbersome and can only be solved numerically. The major difficulty is related to the finite ion size. If we consider the limiting case of point charges, the equation for potential in the double layer would be:
^Y = - s— I»!Vo exp{-^Lx) + - ^ { , c 0 -K +f exp(-2Kx)}lj . (236) dx e , [ kT [_ 87ie e [ 2x JjJ o
0
In comparison with the Gouy-Chapman theory, the MPB equation for point charges contains two improvements, in that the Debye-Huckel parameter K depends on distance and the ion image is screened. In solving the MPB equations, different estimates of C, \{x) were considered. Depending on the type of C,\(x), the equations were called MPB1, MPB2, MPB3, and so on. The MPB theory overcomes certain limitations of the Gouy-Chapman model by taking into account ionic volume, ionic atmosphere and image forces. Results of both theories coincide at low electrolyte concentrations and small surface charges, but beyond this limiting case there is a very significant qualitative and quantitative discrepancy. For instance, the MPB theory predicts a thinner diffuse ion layer and higher capacitance than the Gouy-Chapman model. Near a charged interface MPB predicts considerable adsorption of co-ions. The major difference is that the decaying potential distribution of the GouyChapman model transforms into the decaying-oscillating solution of the MPB theory, with oscillations beginning at high electrolyte concentrations when Kad> 1.24. This oscillation was also predicted by Stilinger and Kirkwood [92] who estimated a different value KOd> 1.03. Other modern theories predict oscillatory behavior of the electrostatic potential, but under different conditions. In the hyperneted chain approximation (HCA) theory the condition is K.od > 1.23, which is very close to MPB theory. The Bogolyubov-Born-Green-Ivon theory puts the condition at K0 1.72. Therefore, oscillatory behavior is a consistent finding of all theories. It is worthwhile to mention that damped oscillation in the
Electric Properties of Oil/Water Interfaces
169
MPB theory is due to a fluctuating potential and its existence does not depend on the excluded volume. One of the major drawbacks of MPB theory is that it does not take into account the discreteness of a solvent. In the diffuse layer the discreteness of solvent molecules strongly influences ion-ion interactions at small distances. Furthermore, an understanding of the structure and properties of the compact layer is impossible without an adequate model of solvent molecules and their interactions with the electrode and between themselves. Application of MPB model to ITIES was made first by Torrie and Valleau [93]. Cui et al. [94] applied the modified Poisson-Boltzmann theory (MPB4) to the interface between two immiscible electrolyte solutions and found that MPB4 describes the experimental results at the nitrobenzene/water and dichloroethane/water phase boundaries. It reproduces the Monte Carlo calculation more accurately than the MVN theory for 1:1 electrolytes over a wide range of conditions including variations in the image forces, ion size, the inner layer potential distribution, electrolyte concentration, surface charge density and solvent effects. They used the equation for electrostatic potential from the MPB4 model for the inner layer at ITIES when x < all: V|/(x) = V|/(0) + xv|/' (0) - xkc\ qaq |"2 qaq I eh ,
(237)
where \\i(x) is the electrostatic potential at coordinate x, i|/(0) is the electrostatic potential at the surface, vj/'(O) is the derivative of y(0), c is the electrolyte concentration, qaq is the surface charge density of the aqueous phase, A, is the parameter determined by fitting the experimental values of the Galvani potential differences and Eh is the dielectric permittivity of the compact layer. The third term in equation (237) represents the dependence on electrolyte concentration, surface charge density, and the solvent effect on the inner-layer potential distribution. These effects can be ascribed to ionic penetration into the opposite phase and ion-ion correlations across the interface. Cui et al. [94] obtained good agreement between theoretical calculations and experimental data and came to the conclusion that the structure of the 1,2-dichloroethane/water interface is similar to that of the water/nitrobenzene interface, except that the effects of the image force and ion size are more pronounced and the inner-layer potential drop plays a more important role. 4. ELECTROCHEMISTRY OF EXTRACTION In recent years, there has been a great surge of interest in the experimental determination and theoretical calculation of the standard Gibbs free energy of ion transfer for individual ions moving between two solvents (Gibbs energy of
170
A.G. Volkov and V.S. Markin
resolvation). The Gibbs free energy of ion resolvation is a key concept intimately related to the electrochemistry of the interface between two immiscible electrolyte solutions, ion transport across biological and artificial membranes, the mechanisms of interfacial and surface catalysis, the kinetics of ion transfer across an oil-water interface, the coupling of heterogeneous reactions in bioenergetics, extraction, and the design and manufacture of ionselective electrodes [95-97]. By solvation, we mean the sum of all structural and energy changes occurring in a system when ions pass from the gaseous phase into solution. It has become customary to divide the interaction between ion and solvent and the corresponding energy, into several components. To understand the energies involved in transferring an ion or dipole between two solvent phases, one must take the following effects into account: (i) electrostatic polarization of the medium; (ii) production in a medium of a cavity to accommodate the ion also known as the solvophobic, hydrophobic or neutral effect; (iii) changes in the structure of the solvent that involve the breakdown of the initial structure and the production of a new structure in the immediate vicinity of the ion; (iv) specific interactions of ions with solvent molecules, such as hydrogen bond formation, donor-acceptor and ion-dipole interactions; (v) annihilation of defects: a small ion may be captured in a "statistical microcavity" within the local solvent structure, releasing energy of this defect; (vi) the correction term for different standard states. This sub-division is purely conditional, since many of these different effects may overlap: for example, electric polarization of the medium may significantly influence its structure. Assigning such a division does permit analysis of the individual components. However, sometimes, components can be grouped into "blocks". For example, one may speak about the solvophobic effect which combines the formation of a cavity and structural changes of the solvent in the vicinity of the new particle. This is quite justifiable because the solvophobic effect, together with the electrostatic effect, provides the major contribution to the solvation energy [48, 98]. On the other hand, sometimes it becomes necessary to consider each component of a given effect. For example, the total solvation (or resolvation) energy can be divided into electrostatic and non-electrostatic parts:
AapG,° = AapG°(e/) + A ° G , W 0 .
(238)
Electric Properties of Oil/Water Interfaces
171
To find the Gibbs standard free solvation energy, one needs to compare the Gibbs free energies for a given ion in each media. If one of the phases ((3) is a vacuum {vac), the difference: vacG, = Gi
+G ,
(239)
is called the solvation energy of the ion / in the a phase. The resolvation energy can be represented as the difference of two solvation energies: 101O A
G
P ,
=
— = AvacG,
~ KacGi
,
(240)
z:r where A"p<> | ° is the standard potential for transfer of ion / from a to (3 phase. Table 2 lists the standard free energies of ion transfer from water into different organic solvents obtained by cyclic voltammetry, chronopotentiometry, polarography and spin-lattice relaxation of quadrupole nuclei of ions, using solubility and extraction data. As can be seen from Table 2, the standard free energy of ion transfer from one solvent to another strongly depends on the nature of both ion and solvent. If two immiscible liquids a and p are mutually saturated, a certain amount of one solvent will be dissolved in the other solvent and vice versa. The corresponding energy of resolvation between two mutually balanced solvents is called the free partition energy. For most ions and solvents the standard free energies of transfer and partition coincide within experimental error. However, for some solvents and ions these quantities may differ significantly. When considering the standard free resolvation energy attention should be paid to whether one deals with ion transfer between pure or mutually saturated solvents. 4.1. Electrostatic contribution to the solvation energy Two sets of models exist to calculate the electrostatic portion of the solvation energy. In the first set the medium is considered as a structureless continuum, while in the second set it is represented by a set of individual particles having either realistic or simplified properties. In early works the solvent structure was taken into account by directly calculating the energies of particular configurations of solvent molecules in the vicinity of the ion. The configuration and the number of molecules in it were chosen with a certain degree of arbitrariness, proceeding from some physical considerations, which ensured an excellent fit between experimental and theoretical data. Such models completely ignored the statistical properties of the solution which are very important for obtaining a correct description of solvation. The modern approach
172
A.G. Volkov and V.S. Markin
lies in developing a statistical theory of ion-dipole plasma which describes both the energy and the statistical aspects of the phenomenon. There is also an intermediate approach, in which, while remaining within the framework of continuum theories, attempts are made to take account of the influence of the discrete nature of the solvent on the effective parameters of the model. For this purpose, account is taken of non-linear dielectric effects, and the mutual correlation of the polarization vectors of the solvent molecules, situated at short distances from one another, is analyzed using the theory of non-local electrostatics. Each of these approaches, reflecting the role of different effects, has its own advantages. 4.2. The Born model The first continuum model, developed by Born [99], considered the ion as a hard charged sphere of a given radius a immersed in a continuous medium a of constant dielectric permittivity s". This was a macroscopic theory, for it used macroscopic laws to describe the properties of a solvent at a small distance from the ion. In the Born theory, the solvation energy (its electrical part, to be exact) is given by the well known relation:
where e0 is the electric constant and s a , the permittivity of the medium. Equation (241) gives the energy per particle. Proceeding from this relation, the free energy of resolvation during the transfer of an ion from medium a to medium p or the difference between the energies in media p and a can be represented in the form:
^•^-iS^?-?'-
< 242 >
The solvation energies calculated by the Born formula differ noticeably from experimental values [48, 98]. Since, in most of the cases, the resolvation energy is the difference between two comparatively large solvation energies, even a relatively small error in each energy may give rise to considerable error in the resolvation energy, even an incorrect sign. For example, equation (219) implies that if ea > sp there is a higher probability for the ion to reside in solvent a, irrespective of ion size. In practice this is not always the case. It is known that small ions of radius a < 0.2 nm reside mainly in a polar solvent of high permittivity, while large organic ions reside preferably in the hydrophobic
Electric Properties of Oil/Water Interfaces
173
phase. Data on the partition coefficients, extraction, solubility, and currentvoltage characteristics show that the standard free energy of ion transfer from water into a less polar solvent is positive for small radius ions and negative for large radius ions, while equation (242) implies that the sign of this energy does not depend on the ion radius. 4.3. The Non-local electrostatics method Another rapidly developing semi-macroscopic approach is to calculate the electrostatic part of the solvation energy using the non-local electrostatics method. Non-local electrostatics takes into account that fluctuations of solvent polarization are correlated in space, since liquid has a structure caused by quantum interaction between its molecules. This means that the average polarization at each point depends on the electric displacement at all other points of the space correlated with a given point [98]. In non-local electrostatics the electric displacement D and electric field E are related by the tensor emn(r):
£ m (r) = ZJ*'£ o e m n (r-r')£"(r'), (m,n = x,y,z).
(243)
n
It should be noted that, although, this relation is spatially complicated, it is linear. Further calculations are carried out in terms of the Fourier transform of the tensor emn(r), which is called the static dielectric function e(k):
eflO = YMr\ m,n
d(r - V )e~lk^\
K
mn(r -
r').
(244)
The potential produced in a medium by a charged sphere of radius a at a distance r from its center is given by:
ze0
"t dk sin kr sin ka
Hence, we can easily find the electrostatic contribution to the solvation energy:
174
A.G. Volkov and V.S. Markin
Polarization of the medium can be divided into three main groups: optical, vibrational and orientational. To evaluate equation (246), one has to specify the function s(k). This can be done, for example, by subdividing fluctuations of the medium polarization into three modes relating to different degrees of freedom: namely, (1) electronic or optical; (2) vibrational or infra-red; (3) orientational or Debye. If the radius of the correlation of the zth mode of a fluctuation is Xx, then: j . ^ )
j =
1
" S
_ O
/
i (
B
O
i ^ B
i )
2
1
+
^
i _ i + ( 2
S 2
" B
i )
3
1 + ^
2
3 -
( 2 4 7 )
In this expression the correlation length of the electronic mode is set equal to zero. The exact values of the correlation lengths X2 and X,3 cannot be determined a priori, but they can be approximated from physical considerations. In the infra red region, the length ^3 depends on liquid type. In the case of non-associated liquids the correlation length for orientational vibration is approximately equal to the intermolecular distance, while for associated liquids (water for example) A,3 is equal to the characteristic length of the hydrogen bond chain, i.e. 0.5 - 0.7 nm [48, 98]. Integration of equation (246) with (1 - l/s(k)) expressed by equation (247) yields:
A—cm=T^ ( • - - • ( - - - ) #)•{ Snsoa[ where: \\)(x) = l-(l-e~x)-
sopt
.
\sopl
sj
\A2J
L \s2
- -Mr)L <*»> s3j
V/Vj'v
(249)
The agreement between theory and experiment obtained by Kornyshev and Volkov is more than satisfactory [48]. Yet, the question arises whether the results obtained are sensitive to the fitting parameters Xt and whether such a choice of X\ is justified. In addressing this question, one can conclude that the results are most sensitive to parameter A,2 and depend very little on A,3. A detailed analysis of this situation is presented in [48]. The non-local electrostatics method refers to the continuum models, but the effective parameters needed for calculation are chosen by analyzing the solvent structure. This method gives rather accurate values for the solvation energy for small ions, and also permits calculation, by virtue of equation (248), of the resolvation energy. For large ions
Electric Properties of Oil/Water Interfaces
175
there remains considerable discrepancy between theory and experiment, which makes it necessary to take into account other effects. In our case, these are the work done in the formation of a cavity in the solvent and the solvophobic effect. The largest value for the solvation energy is obtained in the Born limit A.2 = A,3 = 0, i.e. in a structureless solvent. The major disadvantage of the continuum and semi-continuum approaches to the solvation problem lies in the solvent model itself. An allowance for dielectric saturation or dipole correlation is an attempt to partially describe the discrete properties of the solvent within the framework of the continuum model. Other analogous approaches attempt to take into account the unknown effect of the solvent molecular structure on the thermodynamic properties of a system. However, the problem can only be solved correctly using a statistical model of the solvent and taking into account its discrete structure. 4.4. Statistical solvent models A consistent approach to the solvation problem is to develop a statistical theory that will consider both ions and solvent molecules as discrete species, though such an approach demands the introduction of some model potentials for particle interaction. Unlike "primitive" electrolyte models which use continuum descriptions of the solvent, the modern statistical analysis employs a "nonprimitive" or a so-called ion-dipole plasma model in which the electrolyte solution is considered as a system of hard spheres of radius a and aQ each carrying at their center a point electric charge or a constant point dipole, respectively. Ordinary electrostatic interactions are assumed to exist between the charges and dipoles. The system is described in terms of many-particle correlation functions satisfying the corresponding equations [100, 101]. To solve these equations, one has to introduce various simplifications, the mean-spherical approximation being the most widely used [102]. The ion-dipole plasma model permits calculation of the ion solvation energy, and also the properties of the pure dipole liquid. Wertheim [101] calculated the permittivity of the dipole liquid in the mean-spherical approximation: e = q(2Q/q{-Q,
(250)
where: <7(x) = (l+2x) 2 /(l-x) 2 ,
(251)
and the parameter C, is a solution of the equation: q(2Q-q(Q=nd]x2/3E0kT.
(252)
176
A.G. Volkov and V.S. Markin
Here nA is the number of solvent particles per unit volume and \x is the solvent molecule dipole moment. This result is a considerable improvement over the well known Onsager relation. For highly associated liquids there exists a noticeable discrepancy between theory and experiment. This is not surprising because the model ignores a number of intricate forces which act between real solvent molecules. For example, for a dipole liquid with density and dipole moment (1.8 D) equal to the corresponding parameters of water, equation 250 yields, at room temperature, the permittivity value 48. To obtain a value of 80, the dipole moment should be set equal to 2.62 D. Naturally, a system of hard, constant-dipole spheres can hardly seem as a proper model for water. Yet, such a system may be a reasonable model of an organic solvent. In any case, it is the first step for constructing a relevant model of water. In particular, the application of the model to studying solvation and the interaction of ions has revealed a number of essential effects caused by solvent discreteness. The energy of ion solvation (the "Born" energy) was calculated by Chan et al. using the mean spherical approximation:
KacG° = -^—r^—TTV^1"")' 87ieo(a + ad IX)
(253) e
where A, is found using the equation: A. 2 (l-A,) 4 =16e. (254) The parameter A, ranges from 1 to 3; for E = 80 it is equal to 2.6. Together with the radius of the solvent dipole moment, this parameter determines a new characteristic length for the problem, aj'k which naturally arises in the solution of the equations. This characteristic length is solely determined by the properties of the dipole liquid and depends on two parameters: the molecule radius and permittivity. Formally, equation (253) is analogous to the Born energy with the renormalized radius increased by aJX. It is interesting to note that an empirical relation exists, which is identical in form to equation (253) and which can be used to describe the solvation energy and the free resolvation energy [103]. If the solvent molecule radius is set equal to 1.5 A, the aforementioned relation gives the value 0.58 A for aJX. An apparent increase of the ion radius in equation (253) is caused by solvent discreteness. It is obvious that the permittivity cannot have the bulk
Electric Properties of Oil/Water Interfaces
177
value immediately at the ion "surface". Equation (253) implies that this value is effectively attained at a distance of a + aJX from the ion center. One should, however, bear in mind that this idea of a stepwise change in the permittivity is only an approximate description of sufficiently rigorous physical results obtained in the discrete model. Furthermore, an analysis of the model shows that at small distances the concept of local permittivity is simply invalid and a more correct concept of medium polarization must be used. If a dielectric is treated macroscopically, its polarization at a distance r from the ion is given by the well known relation: PnJir) = (e - 1 W47re 0 r 2 .
(255)
The results obtained in the mean spherical approximation model are mainly of a qualitative character, since the model is approximate. Nevertheless, they provide a more accurate physical representation of solvation and provide another way with which one can evaluate other theories and understand how these theories can agree with experimental results. 4.5. Contribution of the solvophobic effect to the resolvation energy Modern approaches to the calculation of the energy of the solvophobic ion-solvent interaction are based on phenomenological or statistical mechanical molecular models in which the parameters describing both the ion and solvent are not always known. To avoid uncertainty, the solvophobic contribution to the resolvation energy is sometimes estimated using the semi-empirical solvophobic formula [48, 99]. The main idea is as follows. Surface energy, when expressed in terms of surface tension at the ionsolvent interface y0 a , is equal to Ana y0a and the difference in the surface energies in media a and P is Ana (y0 a - y0 p ). According to Antonov's rule [104] Yo.a " Yo,p = YaPSgn(Ya " Yp),
(255)
where sgn(ya - yp) = +1 for ya > yp and sgn(ya - yp) = -1 for ya < yp, y ap is the interfacial tension at a flat boundary between solvents, and y aj yp are the surface tensions at the boundaries of solvents a and p, respectively. With air under the same pressure and temperature, we obtain the formula for the solvophobic contribution to the resolvation energy: AapG°(svph)
= -47w 2 y a p sgn(y a - y p ) .
(256)
178
A.G. Volkov and V.S. Markin
Relation (256) infers that the molecules of the two solvents w and m do not mix or react chemically with each other. The hydrophobic contribution to the Gibbs free energy of ion or molecule (dipole, multipole) transfer from water to hydrocarbon is negative and the absolute value is obviously greater for particles with larger radii, a. The effect of curvature on the surface tension at a molecularly sized sphere was calculated by Tolman [42]. The surface tension at radius a can be written as
1+ — a where 8 is a parameter, which according to Tolman is the distance from the surface of tension to the dividing surface for which the surface excess of fluid vanishes. Parameter 5 can vary between 0 and a few angstroms. From the theoretical point of view, the limits of applicability of the solvophobic formula are not quite clear. Nonetheless, it works surprisingly well for calculation of the partition coefficients of a system of two immiscible liquids. Some deviation from the solvophobic formula is observed if the interfacial tension of two pure immiscible solvents is less then 10 mNm"1 and if the size of the dissolved particle is less than 0.2 nm. An important advantage of the solvophobic formula is that it does not involve fitting parameters because yaP is determined experimentally. 4.6. The total resolvation enegy In the preceding sections we considered different effects which contribute to the energy of ion resolvation. However, for most solvents which are of practical interest the greatest contribution is due to the electrostatic and solvophobic effects. Therefore, the resolvation energy can be represented as a sum of all contributions: AG(tr) = AG(e/) + AG(solv) + AG(si),
(258)
where AG(el) is electrostatic contribution, AG(solv) is the hydrophobic effect and AG(si) is caused by specific interactions of the transferred particle (ion, dipole) with solvent molecules, such as hydrogen bond formation, donoracceptor and ion-dipole interactions. The calculation results are in excellent agreement with experimental data. The discrepancy observed for small radius anions may be attributed to the formation of hydrogen bonds between anion and solvent or to defect
Electric Properties of Oil/Water Interfaces
179
annihilation. The energy corresponding to this discrepancy is on the same order of magnitude as the energy gain obtained in the formation of a weak hydrogen bond between anion and solvent. It should, however, be noted that for small anions resolvation energies reported by different authors differ considerably. 4.7. Dipole resolvation Bell [105] calculated the electrostatic part of Gibbs energy of dipole molecule transfer from a vacuum to a medium of dielectric constant SJ:
^G{el) = --^-[^-\
(259)
12ne0/ l ^ . + l j where (j. is the dipolar moment and / is the effective dipole size. Now one can calculate the electrostatic ("Born") Gibbs energy of dipole transfer from phase w to phase m using a thermodynamic cycle [8]:
A'Giel) = -^-J ^^- - -^±) = - J ^ J
^ >
]-
(260)
Generally speaking, the dipole moment, |^, can vary from solvent to solvent. The concentration of dipole molecules in an oil phase can be calculated from the estimated free Gibbs energy of a dipole molecule transfer from water to the hydrophobic phase: cot, = cwcxp(-AGlr/RT).
(261)
The partition coefficient of an ion or dipole Kj can be also calculated from the free Gibbs energy of transfer: Kt = exp(-AG? (tr) IRT)
(262)
Equations (260) - (262) are very useful in the theory of extraction, partitioning and passive transport of small neutral molecules across liquid membranes. The partition coefficient Kj is also important in the formulation and delivery of pharmaceutical agents because it describes relative lipophilicity and solubility in lipid/water systems. For many drugs the inequality can be written: 0 < log Kj < 4.
(263)
180
A.G. Volkov and V.S. Markin
The choice of a given drug delivery method can be guided by the partition coefficient of the agent. For instance, a drug with a low partition coefficient will likely be given by injection, while a skin patch or ointment could be used for drugs with high partition coefficients. Oral consumption would be appropriate for drugs with coefficients in the intermediate range. REFERENCES [I] V.S. Markin and A.G. Volkov, J. Colloid Interface Sci, 131 (1989) 382. [2] V.S. Markin and A.G. Volkov, Russ. Chem. Rev., 57 (1988) 1963. [3] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 23 (1987) 1405. [4] V.S. Markin and A.G. Volkov, Adv. Colloid Interface Sci., 31 (1990) 111. [5] E. Lange and K.P. Miscenko, Z. Phys. Chem., 149 (1930) 1. [6] F.G. Donnan, Z. Electrochem., 17 (1911), 572. [7] W. Nernst, Z. Physik. Chem., 4 (1889) 129. [8] A.G. Volkov, D.W. Deamer, D.J. Tanelian and V.S. Markin, Liquid Interfaces in Chemistry and Biology, J. Wiley, New York, 1998. [9] Z. Samec, in: Liquid-Liquid Interfaces. Theory and Methods (A.G. Volkov and D.W. Deamer, eds.) pp. 155-178. CRC-Press, Boca Raton, New York, London, Tokyo, 1996. [10] A.G. Volkov, J. Electroanal. Chem., 205 (1986) 245. [II] A.G. Volkov and Yu.I. Kharkats, Chem. Phys., 5 (1986) 964. [12] A.G. Volkov and Yu.I. Kharkats, Kinetics and Catalysis, 26 (1985) 1322. [13] Yu.I. Kharkats and A.G. Volkov, J. Electroanal. Chem., 184 (1985) 435. [14] Yu.I. Kharkats and A.G. Volkov, Biochim. Biophys. Acta, 891 (1987) 56. [15] I.M. Kolthoff, Pure Appl. Chem., 25 (1971) 305. [16] Z. Koczorowski, in: Liquid-Liquid Interfaces. Theory and Methods (A.G. Volkov and D.W. Deamer, eds.) pp. 375-400. CRC-Press, Boca Raton, 1996. [17] B.S. Gourary and F.S. Adrian, Solid State Physics, 10 (1960) 127. [18] R.A. Robinson and R.H. Stokes, Electrolyte Solutions, Butterworths, London, 1959. [19] L.Q. Hung, J. Electroanal. Chem., 115 (1980) 159. [20] L.Q. Hung, J. Electroanal. Chem., 149 (1983) 1. [21] L.Q. Hung, in: Interfacial Catalysis, A.G. Volkov (Ed.), pp.83-112, M. Dekker, New York, 2003. [22] T. Kakiuchi, in: Liquid-Liquid Interfaces. Theory and Methods (A.G. Volkov and D.W. Deamer, eds.) pp. 1-18, CRC-Press, Boca Raton, 1996. [23] T. Kakiuchi, in: Liquid Interfaces in Chemical, Biological and Pharmaceutical Applications, A.G. Volkov (Ed.), pp. 105-121, M. Dekker, New York, 2001. [24] A.G. Volkov and D.W. Deamer (Eds.) Liquid-Liquid Interfaces: Theory and Methods, CRC-Press, Boca Raton, 1996. [25] A.G.Volkov, Sov. Electrochemistry, 23 (1987) 90. [26] A.G. Volkov, J. Mwesigwa, A. Labady, S. Kelly, D.J. Thomas, K. Lewis and T. Shvetsova, Plant Sci., 162 (2002) 723. [27] A. Heydweller, Ann. Phys., 33 (1910) 145. [28] J.J. Bikerman, Physical Surfaces, Academic Press, New York, 1970. [29 D.R Lide, Ed., CRC Handbook of Chemistry and Physio?, 77th Edition, CRC-Press, New York, 1996. [30] C. Wagner, Physik. Zeitschr., 25 (1924) 474.
Electric Properties of Oil/Water Interfaces
181
[31] L. Onsager and N.N.T. Samaras, J. Chem. Phys., 2 (1934) 528. [32] F.P. Buff and F.H. Stillinger, J. Phys. Chem., 11 (1955) 312. [33] A.N. Frumkin, Sbornik rabot po chistoy i prikladnoy khimii, Scientific Chemical Technical Publishing House, Petrograd, pp. 106-126, 1924 (in Russian). [34] A.N. Frumkin, Z. Phys. Chem., 109 (1924) 34. [35] R.M. Suggitt, P.M. Aziz and F.E.W. Wetmore, J. Amer. Chem. Soc, 71 (1949) 676. [36] N.L. Jarvis and M.A. Scheiman, J. Phys. Chem., 72 (1968) 74. [37] J.E.B. Randies, in: Advances in Electrochemistry and Electrochemical Engineering, P. Delahay, Ed., Vol. 3, pp. 1-3 0, Interscience, New York, 1963. [38] V. S. Krylov and V.G. Levich, Doklady Acad. Nauk SSSR, 159 (1964) 409. [39] V.S. Markin and A.G. Volkov, J. Phys. Chem. B, 106 (2002) 11810. [40] Yu.I. Kharkats and J.J. Ulstrup, J. Electroanal. Chem., 308 (1991) 17. [41] J.J. Ulstrup and Yu.I. Kharkats, Russ. J. Electrochem., 29 (1993) 299. [42] R.C. Tolman, J. Chem. Phys., 17 (1949) 333. [43] J.E.B. Randies, Discuss. Faraday Soc, 24 (1957) 194. [44] R. Parsons and F.G.R. Zobel, J. Electroanal. Chem., 9 (1965) 323. [45] Z. Samec, V. Marecek and D. Homolka, J. Electroanal. Chem., 187 (1985) 31. [46] T. Wandlowski, K. Holub, V. Marecek and Z. Samec, Electrochim. Acta, 40 (1995) 2887. [47] A.G. Volkov, D.W. Deamer, D.I. Tanelian and V.S. Markin, Progress Surf. Sci., 53 (1996) 1. [48] A.A. Kornyshev and A.G. Volkov, J. Electroanal.Chem., 180 (1984) 363. [49] B. Sisskind and J. Kasarnowsky, Zh. Fiz. Khim., 4 (1933) 683. [50] H.H. Uhlig, J. Phys. Chem., 41 (1937) 1215. [51] V.S. Markin and A.G. Volkov, J. Electroanal. Chem., 235 (1987) 23. [52] L.S. Bartell, J. Phys. Chem. B, 105 (2001) 11615. [53] T.V. Bykov and X.C. Zeng, J. Phys. Chem. B, 105 (2001) 11586. [54] V.S. Markin and A.G. Volkov, Electrochim. Acta, 34 (1989) 93. [55] W.D. Harkins and E.C. Gilbert, J. Am. Chem. Soc, 48 (1926) 604. [56] W.D. Harkins and H.M. McLaughlin, J. Am. Chem. Soc, 47 (1925) 2083. [57] L.I. Daikhin, A.A. Kornyshev and M.I. Urbakh, J. Electroanal. Chem., 483 (2000) 68. [58] M. Wilson and A. Pohorille, J. Chem.Phys., 95 (1991) 6005. [59] V.S. Markin and A.G. Volkov, Progress Surf. Sci., 30 (1989) 233. [60] V.S. Markin, A.G. Volkov, Electrochim. Acta, 34 (1989) 93. [61] V.S. Markin and A.G. Volkov, in: Liquid-Liquid Interfaces. Theory and Methods (Eds.: A.G. Volkov and D. W. Deamer), CRC-Press, Boca Raton, pp. 63-75, 1996. [62] V.S. Markin and A.G. Volkov, in: The Interface Structure and Electrochemical Processes at the Boundary between Two Immiscible Liquids (V. E. Kazarinov, Ed.) pp. 94-130, Vol. 28. VINITI, Moscow, 1988. [63] A.G. Volkov, M.I. Gugeshashvili and D.W. Deamer, Electrochim. Acta, 40 (1995) 2849. [64] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 24 (1988) 318. [65] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 24 (1988) 325. [66] V.S. Markin and A.G. Volkov, Progress Surf. Sci., 30 (1989) 233. [67] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 24 (1988) 478. [68] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 24 (1988) 579. [69] V.S. Markin and A.G. Volkov, Russ. Chem. Rev., 57 (1988) 1963. [70] V.S. Markin and A.G. Volkov, Electrochim. Acta, 35 (1990) 715. [71] M. Planck, Ann. Phys., 44 (1891) 385. [72] M. Planck, Ann. Phys. Chem.N. F., 40 (1890) 561.
182
A.G. Volkov and V.S. Markin
[73] F.O. Koenig, J. Phys. Chem., 38 (1934) 111. [74] D.C. Grahame and R.W. Whitney, J. Amer. Chem. Soc, 64 (1942) 1548. [75] E.J.W. Verwey and K.F. Nielsen, Phil. Mag., 28 (1939) 435. [76] G. Gouy, Compt. Rend. Acad. Sci., 149 (1910) 654. [77] D.L. Chapman, Phil. Mag, 25 (1913) 475. [78]. A.G. Volkov (Ed.) Liquid Interfaces in Chemical, Biological, and Pharmaceutical Applications, M. Dekker, New York, 2000. [79] C. Gavach, P. Seta and B. d'Epenoux, J. Electroanal. Chem., 83 (1977) 225. [80] R.S. Hansen, J. Phys. Chem., 66 (1962) 410. [81] A.G. Volkov, Langmuir, 12 (1996) 3315. [82] A.N. Frumkin, Zero Charge Potentials, Nauka Publ., Moscow, 1979. [83] Z. Samec, V. Marecek and D. Homolka, J. Electroanal. Chem., 187 (1985) 31. [84] T. Kakiuchi, M. Kobayashi and M. Senda, Bull Chem. Soc. Jpn, 60 (1987) 3109. [85] C. Yufei, VJ. Cunnane, D.J. Schiffrin, L. Murtomaki and K. Kontturi, J. Chem. Soc. Faraday Trans., 87 (1991) 107. [86] Freundlich, H. (1926). Colloid and Capillary Chemistry, Methuen, London. [87] A.N. Frumkin, Electrocapillary Studies and Electrode Potentials, Saposhnikov Publ. House, Odessa, 1919. [88] V.S. Markin and A.G. Volkov in: Encyclopedia of Electrochemistry, Eds. E. Gileadi and M. Urbakh, Vol. 1, pp. 162-187, Wiley-VCH, Weinheim, 2002. [89] A.R. Vanbuuren, S J. Marrink and H.J.C. Berendsen, J. Phys. Chem, 97 (1993) 9206. [90] L.D. Landau and E.M. Lifshitz, Electrodynamics of Continuous Media, 2nd Ed, Pergamon, New York, 1984. [91] C.W. Outhwaite, L.B. Bhuiyan and S. Levine, J. Chem. Soc, Faraday Trans. II, 76 (1980) 1388. [92] F.H. Stilinger and J.G. Kirkwood, J. Chem. Phys, 33 (1960) 1282. [93] G.M. Torrie and J.P. Valleau, J. Electroanal. Chem., 206 (1986) 69. [94] Q. Cui, G.Y. Zhu and E.K. Wang, J. Electroanal. Chem, 383 (1995) 7. [95] V.S. Markin and A.G. Volkov, Sov. Electrochemistry, 23 (1987) 1105. [96] V.S. Markin and A.G. Volkov, Russian Chem. Rev, 56 (1987) 1953. [97] V.S. Markin and A.G. Volkov, J. Electroanal. Chem, 235 (1987) 23. [98] A.G. Volkov and A.A. Kornyshev, Sov. Electrochemistry, 21 (1985) 814. [99] M. Born, Z. Phys, 1 (1920) 45. [100] D. Henderson, Progress Surf. Sci, 13 (1983) 197. [101] M.S. Wertheim, Ann. Rev. Phys. Chem, 30 (1979) 471. [102] D.J.C. Chan, D. Mitchell and B.W. Ninham, J. Chem. Phys, 70 (1979) 2946. [103] A.K. Govington and K.E. Newman, in: Modern Aspects of Electrochemistry, Eds. J.O'M. Bockris and B.E. Conway, Vol. 12, pp. 41-129, Plenum Press, New York, 1977. [104] G. Antonow, J. Chim. Phys, 5 (1907) 372. [105] R.P. Bell, J. Chem . Soc, 32 (1931) 1371.
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 5
Deformation of fluid particles in the contact zone and line tension V. Starov Department of Chemical Engineering, Loughborough University, Loughborough, Leicestershire, LE11 3TU, UK 1. TWO DROPS/BUBBLES AT EQUILIBRIUM IN A SURROUNDING LIQUID The hydrostatic pressure in thin liquid films intervening between two drops/bubbles differs from the pressure inside the drops/bubbles. This difference is caused by the action of both capillary and surface forces. The manifestation of the surface forces action is disjoining pressure, which has a special S-shaped form in the case of partial wetting (aqueous thin films and thin films of aqueous electrolyte and surfactant solutions). Disjoining pressure solely acts in thin fiat liquid films and determines their thickness. If the film surface is curved then both disjoining and the capillary pressure act simultaneously. Disjoining pressure (DP), Tl(h), is a manifestation of colloidal forces action. In Fig. 1 S-shaped DP isotherm is presented, which correspond to a sum of three component: dispersion, electrostatic and structural [1]. DP pressure isotherms can be directly measured, however, not in the whole range of film thickness but only those where Tl'(h)<0, that is, at h
184
V. Starov
Fig. 1. S-shaped isotherm of disjoining pressure.
of two drops/bubbles immersed in a liquid. The theory is based on a consideration of the shape of a liquid interlayer between emulsion droplets or between gas bubbles of different radii under equilibrium conditions taking into account both the local DP of the interlayer and the local curvature of its surfaces. The model of solid non-deforming particles is frequently used, when carrying out an analysis of the forces acting between colloidal particles. However, real droplets/bubbles and even soft solid particles within the contact zone can deform. The latter changes the conditions of their equilibrium. In this case, interaction is not limited only to the zone of a flat contact, but is extended onto the surrounding parts within the range of action of surface forces. For elastic solid particles, such a problem was treated in [4-6]. Below, the case of droplets/bubbles is considered (e.g., emulsions, gas bubbles in a liquid), whose shape changes especially easily under the influence of surface forces (Fig. 2). For the first time the equilibrium position of droplets in a surrounding liquid was studied earlier in [7-10]. However, the finite thickness of a liquid interlayer between droplets/bubbles and variation in its thickness in the transition zone between the interlayer and the equilibrium bulk liquid phase have not been taken into account in [7-10]. The effect of the surface forces in a thin interlayer was taken into account only as a change in the magnitude of the interfacial tension [8, 10]. The problem in using this approach arises if we try to utilize Navier-Stokes equation for description of flow/equilibrium in thin liquid films: a thickness dependent surface tension results in an unbalanced tangential stress on the surface of thin films.
Deformation of Fluid Particles in the Contact Zone and Line Tension
185
Fig. 2. Two identical drops/bubbles, 1, at equilibrium in a surrounding liquid, 2. h(x) is the thickness of the liquid film between two drops/bubbles.
Below a different approach is used, which takes into account the interlayer thickness and the effect of the transition zone between the thin inter- layer and the bulk liquid. A low slope approximation and constant surface tension approximations are used. Then, as was shown earlier in [2, 11] it is possible to use the equation taking into account both the disjoining pressure and the capillary pressure in the interlayer.
Fig. 3. The equilibrium profile of interlayer, h(x), between two cylindrical droplets of the same radius, R.
186
V. Starov
1.1. Two identical cylindrical drops/bubbles At first, let us consider a simplified case of two identical cylindrical drops/bubbles (Fig. 3). It is assumed for a simplicity that the radius of action of surface forces is limited to a certain distance, tsf. Beyond this distance, the surface of droplets retains a constant curvature radius, R, and is not disturbed by surface forces. The interacting droplets are considered as being surrounded by a liquid with constant pressure, Pt. The thickness of the liquid interlayer, 2h(x), varies from 2h0 on the axis of symmetry at x=0 to 2h=tsj at x=xn (Fig. 3). For the drop profiles not disturbed by the surface forces we use the designation hs at yB (Fig. 3). The liquid in droplets is assumed incompressible, which leads to the condition of the constancy of the volume of droplets per unit length: V = 4 } (Hs~hsjdx = nRQ = const , where Ro is the 0 radius of an undisturbed isolated droplet prior to contact. Note, that in general the system of two droplets is thermodynamically unstable with regard to coalescence. Therefore, the following calculations give the conditions of the meta-stable equilibrium of droplets separated by a thin interlayer of the surrounding liquid. The excess free energy of the two cylindrical drops (per unit length of cylindrical droplets), F, is equal to F = yS + F + PV , where S is the total interfacial area per unit length, y is the interfacial tension, FD is the excess free energy per unit length determined by the surface forces action, P is the excess pressure and V is the volume. In the case of cylindrical drops/bubbles S, FD and V are as follows:
S = 4^\+H'/dx 0 x
+ 4 ^l+tifdx x0
+4 \ 0
i\+h'ldx
o[~°°
F = 2 | \H{h)dh dx 0 I2h V = 4 J \Hs-hs}dx x0
+4 \{Hs-h)h= n RQ = const 0
(1)
where h(x), hs (x), Hs (x), x0 and radius, R, are determined in Fig. 3; tsfis the radius of surface forces action. Substitution of the latter expression into the excess free energy results in
Deformation of Fluid Particles in the Contact Zone and Line Tension
187
R I rRI XQ / ^ 1 X 0l~°° l F = 4y j^l+H'fdx+4 l^l+h'f dx+4 J i\+h' dx + 2 \ \U(h)dh dx + [O x0 0 J 0 I2h + P 4 \{Hs-hs)dx+4 xQ
\{Hs-h)k 0
A variation of the excess free energy with respect to h(x), hs (x), Hs (x) and two values xn and R results in the following equations: r oi-3/2 yh"\\+h'l\ +U{2h)=-P
?r3/2
r
^[1+Vj r
=- P
?r
(2)
3/2
where the last two equations give the equation of a circle of radius R. The latter y
immediately determines the unknown excess pressure as P=-—. After that the first Eq. (2) takes the following form r
T~I~3/2
yh"\ l+h'z
y
+U(2h)=L
(3)
where the first term in the left hand site is due to the capillary pressure, the second term is determined by the local value of DP, n(2/z). The boundary conditions for Eq. (3) are as follows: h{xOhhs{xo)
h{Xo)=hs{xo\
h{xo) = tsf/2
(4)
hs{R)=Hs{R), hs{R)=-H^{R)=cc,
(5)
A1(0)=//jl(0)=0.-
(6)
Reminder, hs describes the profile of the droplet at 2hs > tsf .
188
V. Starov
Fig. 4. Partial wetting. S-shaped disjoining pressure isotherm (left side, a) and the liquid profile in the transition zone (right side, b).
If the droplets are located at distance, 2h§>tsf, then n(2/z)=0; and in this case, the solution of Eq. (3) gives the profile hs (x), which corresponds to the circular form of the cross-section of non-deformed droplets that do not interact with one another. At2fiQ
(7)
Substituting the expression for H(2h) from Eqs. (2) and (3) into Eq. (7) and then carrying out integration, we obtain:
Deformation of Fluid Particles in the Contact Zone and Line Tension
« = 2 r L- r ^I-|
189
(8)
In view of the boundary condition (4), at the point x=x0 both values of h(x)=hs(x) and their derivatives, h'=h's, are equal. This enables one to express h (XQ) through the central angle cp (Fig. 3) and values xn and R: h'(xo)=tm
.
(9)
Substituting the latter expression for h (XQ) into Eq. (8), we obtain 0 = 0 , as should be at the equilibrium. Thus, as one should expect that the conditions of equilibrium preset by Eq. (3) and by the boundary conditions (4)-(6), correspond to zero interaction forces between the droplets. It should be noted that angle cp has a value, which is very close to that of the contact angle 0, to be determined at the point of intersection of the continuation of the undisturbed profile of a droplet with axis x (Fig. 3). The values of 8 and cp practically coincide when the interlayer thickness is small as compared with R, and when XQ is not too small. This enables one to use Eq. (9) for calculation of the contact angles. Thus derived values of x0, B, 0, and function h(x) give the full solution of the problem, where the distance between the centers of droplets, B, (Fig. 3) is: B=tsf+2^R2-x^
=t^+{2xo/taa
(10)
1.2. Interaction of cylindrical droplets of different radii Let us now consider a more complicated case of interaction of cylindrical droplets of different radii R2 >R/ (Fig. 5a). Applying he same method of minimization of the excess free energy as above, we obtain the following equations:
7\
r1+
2
T3/2
(^)
r
+n(t) = y/Rl;
(11)
2r3/2
yh2 U + {h2)
-U(t)=-r/R2
(12)
190
V. Starov
Fig. 5. The equilibrium profile of interlayer t(x) = h/.(x) - I12 (x) between two droplets of different radii Ri, and R2 in the general case (a); and in the simplified case (b).
where h]{x) and h2ix) are measured from an arbitrary plane perpendicular to the axis of symmetry, t(x)= hi(x)-h2(x), is the thickness of the interlayer. The latter two equations enable one to determine two profiles, hj (x) and h2(x). These equations with the boundary conditions
/i|(0)=^(0)=0;40)=y; = tan
^= x o ^i-*oJ
h
2(xo)=tan(P2=xolR2-xo)
(13)
; (14)
and with the conditions of constancy of volumes give a solution to the problem. After addition of Eqs. (11) and (12), we obtain
{r?RiHr/R2}=M>k-<
(15)
The latter means that the curvature of the whole interlayer equalizes the capillary pressure drop between the droplets. In a similar way may be solved the same problem for two droplets of different composition with different interfacial tensions, /jand y2, of the first and the second droplet, respectively. In this case, even for the droplets of the same radius, the interlayer on the whole proves to be curved owing to the appearance of a capillary pressure drop, AP/c={/-^-y2)/R •
Deformation of Fluid Particles in the Contact Zone and Line Tension
191
In the case of not strongly curved interlayer between droplets the term ( h '>
2
may be neglected as compared with 1. Then, we obtain from Eqs. (11),
(12), and (15), taking into account that t{x) = h,(x) -h2(x) and t"(x) = h, \x) h2"(x): r-t"+2-u(t)=r
-j-+^-
(16)
This equation should be subjected the following boundary conditions
/(0)=0; 4 o R / ; A^O^^O [R\-4]
+ R
{ 2~4)
(17)
and coupled with the conditions of constancy of volumes, which determine the unknown radii R: and R2. The latter conditions and (16)-(17) enable one to calculate t(x), determining the variable thickness of the interlay er. Thereafter, on substituting the known dependence n[?(x)]=n(x) into Eq. (12), it is possible to obtain from y-h2-n(x)=-y
IR2
(18)
the profile h2(x) of the lower surface of the interlayer. The boundary conditions for Eq. (18) are as follows:
h2(0)=0; h2(x0)=Rcos
h'2(xQ)=tan
(19)
It has been taken into account in the latter expression that angle cp2 is determined from Eq. (14). The sum h2(x) + t(x) gives the profile of the upper surface of the interlayer. Calculation of the interaction of cylindrical droplets of different radii (R2 > Ri) can be simplified, if we assume that the interlayer is of a constant thickness. This means that the effect of a transition zone is neglected, which is justified only at xQ»t. In this case, the curvature of each surface of the interlayer is constant (Fig. 5b), and Eqs. (11) and (12) may be rewritten in the following way:
192
V. Starov
{y I r)+U(t)=y I Rx
(20)
I//(r+t)]-Yl(t)=-y / R2
(21)
This system of equations enables one to determine two unknown values: t, the interlayer thickness; and r, the radius of curvature of its surface on the side of the smaller droplet. Summing and subtracting by terms Eqs. (20) and (21), we obtain (at r»t):
r-lRfo/fa-Rx) Il(t)~y IRy-yl r~y{R2 +RX )l 2RXR2 •
(22) (23)
If the shape of disjoining pressure isotherm, Il(t), is known, then Eq. (23) determines the equilibrium thickness / = const of the curved interlayer. At R2 » R, Eqs. (22) - (23) results in: r~2Rl
(22')
Tl(t)~y/2Rl
(23')
Eq. (22') coincides with that derived earlier by Princen and Mason [8]. Let us note that Eqs. (22) and (23) may be derived by another method using the concept of disjoining pressure [1]: K = Pd-P,
(24)
where Pd is the pressure just under the interlayer surface, and Pt is the pressure in the bulk phase, which the interlayer is at the equilibrium with. On the side of a smaller droplet, we have Pd = P, +(y IRt)-(y Ir); on the side of a larger droplet,Pd = P, +(y/R2)+ y(r + h). As the disjoining pressure does not depend on which side of the interlayer it is determined, then from Eq. (24) it follows: n = ( / / Rx )-(y I r)={y IR2)+[/ l(r+t)\
(25)
which coincides with Eqs. (20)-(23). However, determination of r and t does not solve completely the problem, because the position of the center of the interlayer curvature remains unknown. Its
Deformation of Fluid Particles in the Contact Zone and Line Tension
193
position can be determined by minimizing the value of the free energy of the system
F=2yR2{7t-(p2y2yRx{7r-(piy
2r
L
(26)
t
and by taking into account the condition of the constancy of the volume of droplets:
i( 1 \ l( \ \ 2 R\ n-(py\—sin2^j \+r (p—sin2^ \=n-R^Q-const R2\7t-(p2^—sin2
\-r \cp—sm2q> \=n-R2Q=const
(27)
where Rt0 and R2Q are the radii of the undisturbed droplets (at/?Q>^). It is possible to express all the values via XQ: XQ =R\ sin
i+(/,;)
, i+[*;)
\Ah2j
r l-^frjJ
194
V. Starov
The boundary conditions for Eqs. (28) and (29) are given by (13)-(14), where x0 should be replaced by r0. In this case, conditions of the constancy of the volume of droplets can be written as:
V =2n \ {Hls-hls)dx +2n \{Hls-h)ix= -nR^const 3 rQ 0 V = In f(H2s-h2s)dx rQ
+27T ](H2s-h)ix= 3 0
\^=const
At (h1)2 « 1, Eqs. (28) and (29) can be simplified. On summing these equations we obtain:
(3O)
where t(r) = h\(r) -h2{r). Solution of Eq. (30) describes the variations in the thickness, t{r), of the interlayer between the droplets. The distance between the centers of the droplets can be obtained using the same methods as are used in the case of cylindrical droplets. Thus, the use of the isotherms of disjoining pressure of thin interlayers enables one to solve the problem of the equilibrium of droplets when they are in contact with one another, and to calculate the shape of the deformed droplets and of the interlayer. 1.3. Shape of a liquid interlayer between interacting droplets: critical radius As has been shown above, the shape of a liquid interlayer between interacting droplets depends on their size, interfacial tensions, and the shape of DP. The calculations presented below describe the shape of an interlayer between droplets under equilibrium conditions, until its possible rupture. In carrying out further calculations, we assume that the interlayer thickness, h(r), varies within the contact region but not very sharply, that is, approximation (h )2« 1 can be used. This enables one to use the isotherm of disjoining pressure of flat interlayers, TJh), in equations of equilibrium, as well as to simplify the expression for the local curvature of the interlayer surfaces. Let us consider the interactions of two identical spherical droplets (Fig. 3). In this case, the equation of the interlayer profile, h(r), can be derived by solving Eq. (30)
195
Deformation of Fluid Particles in the Contact Zone and Line Tension
Fig. 6. The disjoining pressure isotherm. n(h) as used in calculations.
Y
- /?"+— \+n(h)=2y / R=P
2{
(31)
r)
where h' = dh/dr ; h = d 2h/dr2; y is the interfacial tension, and R the radius of droplets. For carrying out quantitative calculations, we use below the simplified isotherm of disjoining pressure, IJ(h), in the following form:
K
' \a(to-h),
0
(32)
Such an isotherm (Fig. 6a) provides a possibility for obtaining an analytical solution of the problem, and, at the same time, it possesses the main properties of real isotherms (Fig. 6b), corresponding to the attraction of droplets at large separations, i.e., t0 < h < tsy, and to their repulsion at small distances, at h < toHere, the thickness tsf corresponds to the radius of action of surface forces. Beyond its limits, at h > ts/, the interaction forces vanish: 77= 0. Parameter a=(rii-ri2)/^/- determines the slope of the isotherm. Solution of Eq. (31) together with isotherm (32) has the following form:
196
V. Starov
h(r)=to-(2y/aR)+A-Io(z),
(33)
where z = r(2a/y)'/2, and Io is the Bessel's function of an imaginary variable. The constant A is determined from the boundary condition, h(rn) = tsf, where rn is the radius of the zone of deformation (Fig. 3), and z^-r^al' A=[tsf-t0+(2r/aR)\/I0{z0).
yf (34)
Substituting Eq. (34) into Eq. (33), we obtain an equation determining the profile of the interlayer between droplets:
Accordingly, the minimum distance between the surfaces of droplets is equal to:
In Eqs. (35) and (36), the value of r0 remains still to be determined. For this purpose let us use the condition (7):
0= ]n{h}2xr-dr=0 0
(37)
Substituting into Eq. (37) the equation of isotherm (32), and replacing in the latter h(r) by its expression from Eq. (5), we obtain the following expression: ? AaR(
°
r\sf0
2rY"0
Ift(z)
aR)]Q 1 ^ )
In the case whenzg>5 (sufficiently large drops) the integral in Eq. (38) is equal to TQ (// 2a) . In this case Eq. (38) results in the following expression for the radius of the contact zone:
r O =<2«r)" 2 [^ + i]
(39)
Deformation of Fluid Particles in the Contact Zone andLine Tension
197
Using Eq. (39) the relative extension of the contact zone, where the effect of the surface forces is pronounced, can be expressed as:
|Wto + i] R
\
2y
At 1/aR«\t s r-tQJ/2/,
,40)
aR\ or R»2y Ia(tss-tQ) the ratio rJR tends to the value
(tsf-to){al2y) , which is independent of the radius of the drop, R. This means the geometric similarity of all the large fluid droplets that are deformed by the contact interaction. For the droplets of small radius, R, one may use another approximation for Io(z), which is valid at z < 1: I0(z) = 1 + (z2/4). In this case, integration in Eq. (38) yields: 2_
2#^/-'oJ
r =
(4i)
° I-M/2;>(V-,0)
Let us compare the latter expression with the solution for solid spheres. Their equilibrium state (Fig. 7b) can also be characterized by the interaction region of radius rn, within which the surface forces exert their effect. The profile of the solid sphere can be represented as h=hQ+2R- \—^l-(r/R.)
. At r0 «R,
its
approximate form may be used: h = ho+{r1IR)
(41')
The minimum distance, h0, between the solid particles can be determined from the same condition (7) and the known Derjaguin's approximation [1]:
<&(ho)=x-R]n{hydh
(42)
Replacing the lower integration limit by 4/and using the model DP isotherm (32), after integration in Eq. (42) we obtain the following expression: h0 - 2ta -tx. On substituting this expression into Eq. (41) and taking into account that h(r0 )=?s/we get:
198
V. Starov
Fig. 7. The schematic representation of the contact interaction of the fluid (a, c) and the solid (b) particles.
r:=2R{tsf-t0)
(43)
The same value of r0 is also obtained by solving Eq. (41) for "small" fluid droplets. The latter means that very small emulsion droplets and gas bubbles practically behave themselves as solid particles: they practically do not deform in the contact zone. Let us call the critical size of fluid droplets, R*, such that at R« R* their interaction does not differ from that of solid spheres. Comparing Eqs. (41) and (43), note that these coincide under the condition that the second term in the denominator of Eq. (41) is much smaller than unity. This condition allows one to determine R * as R'=2r/{t^-to)-a.
(44)
The distance between the centers of droplets can be determined using Eq. (10) and simple geometrical considerations: B = tsl +2{R2-r02)"2
(45)
Accordingly, the contact angle, 6, can also be determined at the point of intersection between the non-deformed surface of a droplet and the r axis (Fig. 7a): cos<9 = BI2R = (tsfl2R)+ [l - (r0 /R)1]'1
(46)
Deformation of Fluid Particles in the Contact Zone andLine Tension
199
This expression holds only for relatively large droplets. As has been already pointed out above, small droplets (R < R*) are practically non-deformable, and their undisturbed, circular profile does not intersect the r axis. In the case of small droplets, like that of complete wetting [2], the contact angle is absent. Let us now numerically calculate the profiles of fluid droplets in the contact zone using Eq. (35), and compare them with two other known models: the model of solid non-deformable particles (Fig. 7b) and the model of a flat interlayer between similar droplets (Fig. 7c) (in the latter model the effect of a transition zone between the surrounding bulk medium and the flat portion of the interlayer is not taken into account). In the case of flat interlayers let us use the equations of equilibrium of thin flat films (see below): 2 • /(cos0 -1) = ]n(/j)-dh + 2P- hc
(47)
2/1,
where Of is the equilibrium contact angle determined at the point of intersection between the continuation of the non-deformed part of a sphere and the symmetry plane. In that case, the thickness of the flat interlayer, 2he, is determined using DP isotherm, UJi), as 77 = P, where P = 2ylR is the capillary pressure drop at the spherical interface. The contact area of droplets at he « R is equal lorrr* = nR2 • sin2 6 where r0 is the radius of the contact zone. Substituting Eq. (32) of the disjoining pressure isotherm into Eq. (47), we obtain the following expression: cos9 = \-(a/4y)\(tl
-tj
-{2y/aR)2\+(l/R)-[t0 -(2y/aR)]
(48)
Fig. 8 represents the calculated profiles of droplets in the contact zone for the model of solid particles (curves 3), the model of a flat interlayer (curves 2), and real fluid profiles (curves 1), that is, the profiles, while taking into account the deformation of the profile under the simultaneous effect of both capillary and surface forces within the contact zone. The calculations were carried out for the model DP isotherm, FI(h), (Eq. (32)) preset by the following parameters: y= 30 dyne/cm; ts/= 3 x Iff6 cm; t0 = 2 x Iff6 cm; 77; = 77(0) = 3 x 106 dyne/cm2, and I7min = 77(tsf) = -1.5 x 106 dyne/cm2, which gives a = 1.5 x 10u dyne/cm3. According to Eq. (47) the adopted values correspond to, the contact angle 6 = 9°. As has been shown by Aronson and Princen [10], the emulsion droplets, separated by a thin inter- layer of a dispersion medium, form finite contact angles with the bulk phase within the range of 0 to 90° The left-hand part of the plots in Fig. 8 relates to the spherical droplets having radius R = 10"3 cm; the right-hand part to R = 10"4 cm. As appears from the left-hand part me large-sized droplets deform considerably thus forming the
200
V. Starov
Fig. 8. The profiles. h(r/R), of the contact zone of identical spherical particles, R = lO'Jcm (to the left), .and the R= 10"4 cm (to the right), calculated while taking into account the transition zone (I), and in the approximation of a flat interlayer (2), and solid particles (3).
practically flat contact zone. Profiles I and 2 are close to each other, and the transition zone occupies rather a small region immediately near the contact perimeter. Now, deviations from the profile of solid particles are very large (curve 3). Thus, in the case of R » R* the conditions of the droplets equilibrium may be described with the framework of the theory of flat interlayers, i.e., on the basis of Eq. (47). A decrease in the size of droplets (the right-hand side of the graph in Fig. 8) makes the profiles of solid particles (curve 3) and of droplets (curve I) approach to one another. In the central part of the contact region, the flat interlayer regionreduces, while the differences from the profile of the flat interlayers (curve 2) increase. A further, decrease in the droplet radius causes a still greater approaching of the profiles of the fluid and the solid particles to one another, and disappearance of the flat portion of the interlayer. As has been shown above at R « R* one may use the known solutions for the interaction of solid spheres. In this connection, it is important to evaluate values of R *. This allows one to determine the regions of applicability of different solutions; namely, flat
Deformation of Fluid Particles in the Contact Zone and Line Tension
201
interlayers for R »R*; solid particles for R «R*; and, finally, the approach developed above for the intermediate values of R. As appears from Eq. (44), the critical radius R* depends on the parameters of the disjoining pressure isotherm (a, tsft and tn) and the interface tension. Under otherwise equal conditions, a decrease in the interface tension reduces the values of R*, thus limiting the region of the applicability of the theory of solid spheres. Let us evaluate R*, assuming y = 50 dyne/ cm, and (%-/«) = Iff6 cm. The values of parameter a will be varied, which, in accordance with Eq. (48), is equivalent to a change in the contact angle Of cos^=l-(a/4r)-(/t/-/2)2
(49)
The calculations show that the values of R* decrease as 0 increases. Thus, at small values of 0close to zero, R* = 10"3 cm; at 0 = 5-6°, R* = 10"4 cm; at 0 = 2030°, R* = 5 x Iff6 cm; and at 0 = 90°, R* = 5 x Iff7 cm. In this way, an increase in the contact angle, corresponding to the enhancing of the droplets interaction, causes a decrease in the critical radius R * Now, this means that in the case of a strong interparticle interaction, the approach of solid spheres becomes ever less applicable; yet in the case of weakly interacting droplets, that approach may be used even for relatively large-size droplets. The solutions obtained allow one to evaluate R* and choose corresponding equations for calculation of the equilibrium shape of the droplets within the contact zone. 2. LINE TENSION The presence of the transition zone between the drop/bubble and thin liquid interlayers can be described in terms of line tension r, which has been introduced by Gibbs [16]. In the case of surface tension the transition zone between the liquid and vapour is replaced by a plane of tension with excess surface energy y. In the same way the transition zone between the drop/bubble and the thin liquid interlayer may be replaced by a three-phase contact line with excess linear energy r. In contrast to surface tension y, the value of the line tension, r, may be both positive and negative. In the case of positive value it makes the wetting perimeter to contract but to expand in the case of the negative value of the line tension. When the line tension is introduced, an additional terms should be incorporated in the Neuman-Young equation [13]: my- (cos # - cos # , ) = + — I
U
dr
(50)
o)
where r0 is the radius of the wetting perimeter, m and plus or minus depends on the system geometry: m=l and "-" correspond to the drop on the solid substrate,
202
V. Starov
m=2 and "+" correspond to the two identical drops/bubbles in contact (Fig. 9); contact angle Of in the case of big cylindrical drops (see below). In the case of liquid drops on the flat solid substrate, the positive values of x cause an increase in the values of contact angles 6 , whereas the negative value results in their decrease. In the case of flat films in a contact with a concave meniscus, the influence of the line tension ris inversed, because "-" sign should be used now in Eq. (50). For water and aqueous electrolyte solutions the line tension values are of about 10"6-10"5 dyne [15]. Thus, the terms in the right-hand side of Eq. (50) becomes noticeable at r < 10~4cm. The values of the line tension, r, for drops on solid substrates has been calculated as a difference between the values of y-COSo, calculated in two different ways (i) neglecting the transition zone and (ii) taking it into account [15]. Since the line tension arises due to the existence of the transition zone, it is clear that this difference is just associated with the additional terms in the right hand site of Eq. (50). An expression for x was obtained in the case of a model isotherm of disjoining pressure [15] and the line has been estimated as 10~5 •*-10~6 dyn and negative. De Feijter and Free [17] have considered the transition zone between Newton black film and bulk liquid. According to their estimations the line tension value is also negative and has the same order of magnitude.
Fig. 9. Magnification of the upper part of the transition zone between the drop/bubble and thin liquid interlayer. I- real deformed profile of the drop/bubble, 2 - ideal spherical profile, when the influence of DP pressure has been ignored, 3 - thin liquid interlayer of thickness 2he.
Deformation of Fluid Particles in the Contact Zone and Line Tension
203
T. Kolarov and Z.M. Zorin [18] have measured the line tension value. They have used Sheludko's cell for measurements of properties of free liquid films. Aqueous solution 0.1% NaCl with SDS 0.05% concentration (CMC=0.2%) hasbeen used. They have calculated line tension for this system using the Neuman-Young Eq. (50). The value of line tension has been found -1.7- l(T6dyn. That is, negative and in the good agreement with the theoretical predictions [15]. D.Platikanov et al. [19] have carried out an experiment measurements of line tension dependency on salt concentration and present experimental evidence of line tension sign change. Let us consider two identical drops/bubbles in contact (Fig. 9) under equilibrium conditions. It is convenient to specify the reference state to write down an excess free energy of the system (Fig. 9). This reference state is introduced as "a flat equilibrium liquid film of the thickness 2h". From mathematical point of view it means addition/subtraction of a constant to/from the excess free energy. The latter constant does not influence the final equation, which describes the profile in the transition zone. However, as we see below, it influences substantially the definition of the line tension. The latter means that the choice of the reference state is very important and we use below the same choice of the reference state as in [15], which is the uniform flat equilibrium film of thickness 2he, where he is the half-thickness of the equilibrium film. Using that choice the excess free energy, F, of the system (curve 1 in Fig. 9) has the following form: R
.
GO
OO
2h
2he
F=lx\r 2y(Jl+h'2 -1)+ \Tl{h)dh- \n(h]dh+2P-(h-he) dr 0
(51)
where h(x) is the half-thickness of the liquid layer, he is a half-thickness of the flat equilibrium thin liquid film; P=2ylR is the excess pressure, u(h) is the disjoining pressure, h(R)=H is the position of the drop's end. The lower limit of integration correspond to the end of the transition zone there h=he. We can use infinity as the upper limit of integration instead of R, because at this stage we are not interested in the upper part of the drop/bubble. Under the equilibrium condition the system is at the minimum free energy state, that is two the following conditions should be satisfied: SF = 0
(52)
The latter condition results as above in the equation for determination of the liquid profile in the transition zone:
204
V. Starov
However, the condition (52) is not the only one required for the equilibrium. Let us remind the second condition. If the end of the transition zone can move along a curve
-\® + (
(54)
where cj> = A 2y{^j\ + h'2 -1) + )u(h)dhl_
)n(h)dh + 2P-(h-hc) '
. Substitution -
of the latter expression into condition (54) and taking into account that in the case under consideration (p(r)=he=const, hence,
2
\
h'2
2yr U\ + h' -\)—==
=0
(55)
Condition (55) can be satisfied only if h' -> 0,
(56)
at the end of the transition zone, which means a smooth transition from the transition zone to the flat thin film. Let us introduce an ideal profile of the liquid interlayer in the transition zone, 2hid, which is a spherical part up to the intersection with the equilibrium liquid interlayer of thickness 2he (2 in Fig. 9). The excess free energy of such ideal profile differs from the exact excess free energy given by Eq. (51) because the presence of the transition zone is ignored in the case of the ideal profile. The latter means that the line tension, r, should be introduced to compensate the difference:
F = 2n\r 2y{j\ + (h'J - 1 ) - )u{h)dh + IP• (hid -he)dr
+ Inr.r (57)
Under equilibrium conditions the following two conditions should be satisfied: 5^=0
(58)
Deformation of Fluid Particles in the Contact Zone and Line Tension
205
which gives an equation for the ideal liquid profile in the transition zone
f
L± , rdr
K
=p
r
2^
(59)
\ + \h.
v L
v
J
\ )
and the transversality condition which in the case of ideal liquid profile with excess free energy given by Eq. (57) is as follows:
-**-O^l dhid \r_r
+ ^ =0
(60)
dr
°
where cDrf = r 2y(^l + hut -1) - "\U{h)dh + IP • (hd - hj
. Substitution of the
2/1,
latter expression into condition (60) results in the following equation at r=r0 and h=he:
2r{j^hJ-1)- )n(h)dh--pL
-{*L+ L]=o.
After a rearrangement the latter condition becomes 2 r (cos<9-cose / )= - + - ^ 1 Vro vr0 J
(61)
where 2/ cos 0 =2y+ ]u (h)dh.
(62)
2/1,
Eq. (62) gives the expression for the contact angle of a big cylindrical drop (see below). Note that integration in the right hand site is over h from he to infinity, which does not depend on the profile (real or ideal profile) but only on the
206
V. Starov
integration limits. Eq. (61) coincides with Eq. (50) for the system under consideration (Fig. 9). Excess free energy given by Eq. (51) and (57) should be equal. The latter gives the following definition of the line tension, x: )r 2yUl + (h',d)2-l]-)n(h)dh
+ 2P-(h,d-he) dr + rox =
r
(63)
i 1
= JV 2y (VlT/7 -1) + ^U{h)dh- \ll(h)dh + 2P-(h-he) 0
2*
dr
2b,
The latter equation is an exact definition of the line tension, r in contrast to Eq. (61), where the value of line tension is unknown. In Eq. (63) the real liquid profile, h{r), is the solution of Eq. (53) and the ideal liquid profile, hic/(r), is the solution of Eq. (59). Dependency of the line tension on the radius r0 has been investigated in [15] in the case of a model DP isotherm. Below we focus on the absolute value of the line tension and a possible comparison with experimental data. For this purpose let us consider the line tension in the simplest possible case: contact of two identical cylindrical drops/bubbles. In this case the corresponding excess free energies (51) and (57) take the following form OC
~Z
I
F= | 2/(i\+h 0
00
00
-1)+ \U{h)dh- \U(h)dh+2P-(h-he) dx Ih 2he
(64)
and F = ]\2y{jl
+ (h'J-\)-
)n{h)dh + IP • (hld - ht)\dx + r ,
(65)
where now F is an excess free energy per unit length. From Eq. (64) we conclude rh
"
[i+ (h')'}
+n(2h)=p.
(66)
Deformation of Fluid Particles in the Contact Zone and Line Tension
207
Eq. (66) describes the whole range of the liquid profile including the lower bulk part of the drop/bubble, the thin flat liquid interlayer in front of it and the transition zone in between. The boundary conditions for Eq. (66) are h{R)=H,
h-+he, ti
h'{R)=v
(67)
x^-0
= 0,
(68)
x^O
h=\,
We can integrate Eq. (66) once using boundary condition (67), which yields:
r
,
v = w,
(69)
[i + w} where L(h) = P(H-h)-
^U(h)dh. Eq. (69) can be rewritten as: 2*
h=
' iWf~
y
<70)
The left hand side of Eq. (69) is always positive and less than y. That means the same should be true for the right hand site of Eq. (69):
0
(71)
where L(h) = y, if h = he and L(h) = 0, if h = H . Condition (68) results in:
p-(H-he)=r+)u(h)dh.
(72)
The capillary pressure can be expressed as before: P =
L R' (73)
208
V. Starov
where R is the radius of the curvature of the cylindrical drop. Simple geometrical considerations show: R = -^-r
(74)
cos 9 With the help of the latter condition we can conclude that:
/•cos 0 P
=
(75)
H Using Eqs. (72) and (75) we conclude:
1 + — • )n(h)dh cos
# =
" Y
J.
(76)
H <-——
= n{2ht)
{ii)
ti
If he/H«l the contact angle in Eq. (76) is refereed to as d^and coincide with that given by Eq. (62). In the case of partial wetting contact angle is in the following range 0 < 0 < — , o r O < COS 6 < 1. Using condition (71) the latter unequally can be rewritten as
oo
y-h \U{h)dh<-—^.
-y<
H
2he
Hence, the integral, \Tl(h)dh, should be negative in the case of partial wetting. 2/1,
In the case of the ideal profile we should use Eq. (65), which results in
r
i
, 2 f= P
2
1 + fc)
(78)
Deformation of Fluid Particles in the Contact Zone and Line Tension
209
The boundary conditions for Eq. (78) are hid{R) = H,
h'id{R) = «>
(79)
Eq.(78) can be integrated once using boundary conditions (79), which yields:
h id=
' i~P~~l
(80)
V '-'id
where Lld(hld) = P(H-h,d). Line tension r can be expressed using Eqs. (64) and (65) as:
r = J ly^X + h'1 -1) + ]n(h)dh - )ll{h)dh+ 2P-(h-hJ\ dxo|_
2/i
Ih,
J
- f 2 ^ 1 + fc)2 -\]-"\n{h)dh+2P\hd -h^dx Using Eqs. (70) and (80) we can rewrite the latter equation as
T =2)[yJl + h'2 -L(h)]dx-2)[r^
+ {h'J -4,(AjJfc
(81)
x,,
0
Using Eqs. (70) and (80) we can switch from integration over x to integration over thickness. The latter transformation of Eq. (81) gives:
x = 2{(Vy2-Z2(/z) - V Y 2 ~ 4 W ) dh
(82)
K
The similar equation was obtained earlier [20]. Eq. (82) can be rewritten as \2
/„ m 2P(H
t =- 1
,
- h) \u{h)dh -
\l\(h)dh 2
. ^"
^ dh
(83)
In the case of h«H inside the whole transition zone expressions for L(h) and Lid(h) can be rewritten as
210
V. Starov
/ \ L(h) = r\cos0f-e(h)\
L(h)
1 °° e(h) = - \U(h)dh
= ycosd Using the latter expression Eq. (83) can be rewritten as 2 2 cos 6,e{h)-E y{h) ' y ' ' dh 2 2cosdfs(h)-s (h)
2y " r =- = 2 — { sin#, K •
i
i+
+
(84)
1— 2
"y
sin 6f
In the case of small contact angles s ( h ) « l and the latter equation takes the following form: r =- ^ - f tan 0f I
,
£{k)
dh
(85)
2cos$fe(h) l+ + 2
f
sin ^
It is possible to show that 0.5 < 1+
, <1 in the case under I , 2cosd,E{h) "\( sin 2 ^
consideration. Let the mean value of the latter expression be co, where 0.5
J
J()u(h)dh\ih 0f
ih\ih
(86)
)
Eq. (62) can be rewritten as c o s ^ = 1 + s(hj, or sintf. * 0 = 2 ^ - e(h ) = 2 I- - ]n(h)dh V Ylh-
(87)
Combination of the latter expression and Eq. (86) results in
r=
2C
°
] (]u(h)dh\lh =
2C0 Y
^
J f ]n{h)dh\h
(88)
Deformation of Fluid Particles in the Contact Zone and Line Tension
211
We can compare the experimental data by D. Platikanov, M. Nedyalkov [19] to the theory predictions according to Eq. (88) (at a>=l). We used below only dispersion and electrostatic components of the disjoining pressure. The following expressions for different component of disjoining pressure are used: for dispersion component: n,=--^,
(89)
where A ~ 10~udyn • cm is the Hamaker constant and for electrostatic component n , ; = 6 4 -c-R -T -tan/2 2 ((^/4)-exp(-K- • h),
(90)
where c, R, T.F, \u = ——, K-J are concentration of the electrolyte, the RT V ERT universal gas constant, temperature in °K, Faraday number, dimensionless zeta potential of the film surfaces, and the inverse Debye length, correspondingly; 8 is the dielectrics constant of water. Hence, the total DP is:
Fig. 10. DP according to Eq. (90) with cut-off thickness, ' * .
212
V. Starov
U(h) = ~
+ 64-c-R-T-
Xwh2{y7lA) • exp(-/c • h)
(91)
h Unfortunately the DP in the form given by Eq. (91) does not allow any equilibrium liquid films at low thickness (a - films). To overcome this problem we introduce a cut-off thickness, K (Fig. 10). According to this choice the equilibrium thickness 2he does not depend on the pressure inside the drops/bubbles and is always equal to K . D. Platikanov and et al. [19] have determined both contact angle and line tension on the NaCl concentration in the range of concentrations 0.2-0.45 mol/1. These two dependencies are used below. 2.1. The comparison with the experimental data [19] and discussion Details of experimental measurement and system under consideration are given in [19]. Line tension of free liquid film between two bubbles was investigated in the range of NCI concentration 0.2-0.45 mol/1. The film and bubble surfaces were stabilised by surfactants [19]. Zeta-potential of the film surface was \\i= 17 mV or (//=0.68 according to [19].
Fig. 11. Calculated dependency of line tension on electrolyte concentration.
Deformation of Fluid Particles in the Contact Zone and Line Tension
213
The cut-off thickness, tt, was used as a fitting parameter. We used experimental values of contact angle from [19] on salt concentration to determine the cut-off thickness tt according to Eq. (87). A reasonable agreement between experimental dependency of the contact angle on the salt concentration and the calculated according to Eq. (87) has been attained. The fitted dependency of the contact angle on the salt concentration was much weaker than the original experimental data [19]. However, we tried to compare our calculation of line tension according to Eq. (88) and the corresponding experimental data of line tension from [19] using already calculated cut-off thickness tt. Calculated dependency of line tension on the electrolyte concentration is shown in Fig. 11. The conclusions are as follows: • line tension dependency on the salt concentration is in a qualitative agreement with experimental dependency in [19], that is line tension goes from positive to negative values at the increase of the electrolyte concentration; • the absolute values of the calculated line tension were found considerably different from the corresponding experimental values, the calculated electrolyte concentration (0.022 mol/1) at which line tension switches from positive to negative values does not match the experimental value (0.36 mol/1); • the calculated line tension remain almost constant in the range of electrolyte concentrations used in [19]; • line tension decreases much faster with electrolyte concentration (Fig. 11) than experimental values [19]. The discrepancy between the measured and calculated line tension dependencies can be caused by one or both of the following reasons: • only dispersion and electrostatic components of DP pressure have been used to compare with the experimental data. It looks like these two components are not enough to describe adequately behaviour of thin liquid films and the transition region in the system under consideration. Influence of both structural component (caused by the water dipoles orientation in a vicinity of free film surfaces) and steric (caused by the direct interaction of head of the surfactant molecules on the film surfaces) can not be ignored. The theory of these components of DP is to be developed; • if the experimental system used in [19] was really an equilibrium one. According to the definition of line tension (83) it is determined by the equilibrium liquid profile in the transition region from the thin flat liquid interlayer to the bulk surface of bubbles. It is difficult to prove that under experimental conditions used in [19] the liquid profile in the transition region was equilibrium one in spite of the efforts undertaken to reach the equilibrium [19]. Note that in the case of partial wetting, which is under consideration, the contact angle hysteresis [2] is unavoidable. The latter phenomenon can completely mask non-equilibrium state of the system. Unfortunately except for [19], where considerable efforts have been undertaken to reach the equilibrium
214
V. Starov
state, and other few publications, the possible deviations from the equilibrium state of the system under consideration usually even does not discussed. We would like to emphasise again: if the liquid profile in the transition region from thin liquid interlayer to the bulk drop/bubble interface is not at the equilibrium then an arbitrary values of line tensions can be measured, which have nothing to do with the equilibrium value of line tension. 3. ACKNOWLEDGEMENTS This research has been supported by the Royal Society, UK.
REFERENCES [I] Derjaguin, B.V., Churaev, N.V., and Muller, V.M. "Surface forces", Plenum Press, New York, 1987. [2] Starov, V.M. Adv. Colloid Interface Sci., 39, 147 (1992). [3] Churaev, N.V., Sobolev, V.D. Adv. Colloid Interface Sci., 61, 1 (1995). [4] Derjaguin, B. Y., Colloid J.(USSR Academy of Sciences), 69, 155 (1934). [5] Derjaguin, B. Y., Muller, V. M., and Toporov, Yu. P., Colloid J.(USSR Academy of Sciences), 37, 455, (1975); 37, 1066 (1975). [6] Muller, V. M., and Yushchenko, Y. S., Colloid J.(USSR Academy of Sciences), 42, 500 (1980). [7] Princen, H. M., J. Colloid Sci, 18, 1978 (1963). [8] Princen, H. M., and Mason, S. G., J: Colloid Interface Sci. 20, 156 (1965); 20, 246 (1965). [9] Princen, H. M., J: Colloid Interface Sci. 71, 55 (1979). [10] Aronson, M. P., and Princen, H. M., Nature (London) 286, 370 (1980). II1] Churaev, N.V., and Starov, V.M. J. Colloid Interface Sci., 103 (2), 301 (1985). [12] De Feijter, J. A., and Vrij, A., J. Eleclroanal. Chern. Interface Eleclrochem, 37, 9 (1972). [13] Starov, V. M., and Churaev, N. V., Colloid J. (USSR Academy of Sciences), 42, 703 (1980). [14] Churaev, N. Y., Starov, Y. M., and Derjaguin, B. Y., J. Colloid Interface Sci, 89, 16 (1982). [15] Derjaguin, B. Y., and Churaev, N. Y., J. Colloid Interface Sci, 38, 438 (1976); 66, 389 (1978). [16] Rowlinson, J.F., and Widom, B. "Molecular Theory of Capillarity", Clarendon, Oxford, 1984. [! 7] de Feijter, J.A., and Vrij, A. J. Electroanal. Chem., 37, 9 (1972). [18] Kolarov, T., and Zorin, Z.M. ColloidPolym. Sci., 257, 1292 (1979). [19] Platikanov, D, Nedyalkov, M., and Nasteva, V. J. Colloid Interface Sci., 75, 620 (1980). [20] Dobbs, H.T., and Indekeu, J.O. PhysicaA, 201, 457 (1993).
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 6
Hydrodynamic interactions and stability of emulsion films E. Mileva" and B. Radoevb "Department of Colloid and Interface Science, Bulgarian Academy of Sciences "Acad. G. Bonchev" Str., bl. 11, 1113 Sofia, Bulgaria E-mail: [email protected] b
Department of Physical Chemistry, Sofia University 1 "J. Bourchier" Str., 1126 Sofia, Bulgaria E-mail: [email protected]
1. INTRODUCTION The hydrodynamic interactions are among the most important factors that determine the properties and the stability of emulsion systems. Diverse aspects of this topic might be found in numerous publications (see e.g. Refs. 1-6). One specific feature that distinguishes emulsions from other complex disperse systems, like foams and suspensions, is the coupling of fluid motion in the contiguous phases. This interdependence of the flows is closely related to the mobility of the fluid/fluid interfaces and to the presence of surfactant species. The interfacial mobility is the most prominent characteristics: due to viscosity of the adjoining fluids, the flow in one liquid phase is transferred through the boundary into the neighboring phase and the fluid motions mutually influence one another. Consequently, the mathematical model of an emulsion system inherently comprises the coupling of the hydrodynamic equations that describe the flow fields in the contiguous phases, and they all have to be solved simultaneously. The kinetic stability of emulsions is particularly related to the hydrodynamic interactions at close approach of two fluid particles. If the gap between them becomes of the order of their dimensions or smaller, the fluid motion is mostly localized in the narrowest parts of the system. An anizodimetric region might be distinguished where the major dissipation of energy takes place [5,7-10]. This zone generates the leading term of the drag
216
E. Mileva and B. Radoev
force between the two droplets and is termed a thin liquid layer or an emulsion film. Therefore, the close distance hydrodynamics of two axially approaching droplets presents the basic flow patterns that could outline the governing effects correlating flow and stability in emulsion systems. Within this model, two aspects are of particular interest: the drainage peculiarities of the emulsion film, and the flow pattern inside the droplets as caused by the squeezing outflow in the film. In most of the studies it turns convenient to assume two major regimes of interacting droplets: constant approach velocity (e.g. Refs.[11,12]) and constant driving force (e.g. Refs. [13-15]). The specific emphasis in this chapter is related to the second case. The emulsion film investigations came out as a further development of the close-distance hydrodynamics of foam films. The latter was vigorously investigated in the seventies and the eighties of the last century [16-22] and is very much allied to the film microinterferometric technique, which is one of the most popular instrumentations in thin liquid layer studies [18,23,24]. The leading idea is that at small gap widths between the particles, the fluid motion in the film is retarded and the governing flow equations are, as a rule, of creepingflow type. The anizodiametry of the gap often allows additional simplifications, like a lubrication flow mode, named sometimes Reynolds law, or similar [1621]. Insofar as the flow inside the droplets is concerned, various simplifications of the general Navier-Stokes equations have been proposed ranging from lubrication-type relations [12,25], through full creeping-flow model [4,7,11,15, 26-30] and, to hydrodynamic boundary layer presentation [31]. Another very important factor for the behavior of the emulsion systems is the presence of surfactants. The most popular observation is the so-called Bancroft rule: "the phase in which the stabilizing agent is more soluble will be the continuous phase" [see e.g. Ref.32]. There are different interpretations of this rule, related either to energy considerations [32], or to some hydrodynamic reasoning [3,31,33-37]. In the latter case, the presence of surfactant is always interpreted in view of its impact on the interfacial mobility of the fluid interfaces. This point of view stems from the physicochemical hydrodynamics of Levich [38] and was developed for the specific interactions in emulsion films by Lee and Hodgson [3]. The key idea is that the surfactant mass transfer additionally couples with the hydrodynamics in the emulsion systems. The geometrical anizodiametry of the emulsion films modifies further on both the flow and the mass transfer regime [31,33-37]. One substantial property of the emulsion systems is their coalescence stability. The stability of the thin liquid layer, most generally presented by the so-called 'thickness of rupture of the film', plays a decisive role. The theory of film rupture was first formulated for the case of foam systems [17,39]. With a slight modification, it is also applicable for emulsion films [31]. The studies
Hydrodynamic Interactions and Stability of Emulsion Films
217
show that the true stability of the emulsion films is related to their thermodynamic properties (surface tension, disjoining pressure). As for the intrinsic coupling of hydrodynamics and mass transfer of surfactants, it concerns the so-called kinetic stability, related to the time evolution of the draining emulsion film. It may be expressed by the interfacial mobility of thin liquid layer alone (see Section 4). The basic result is that while for some characteristics (such as lifetime of films) the interfacial mobility is of utmost importance, the thickness of rupture is actually independent on mobility factors. The aim of the present chapter is to give a survey on some particular aspects of the influence of interfacial mobility on the hydrodynamic interactions in emulsion films, as well as to follow the effect of the mass transfer of surfactant species on these interactions. A special attention is paid on the possibility to develop a systematic scaling procedure for the analysis of the general dynamic equations which model the fluid motion in the contiguous liquid phases. The procedure offers universal approach to estimate the relative importance of the possible coupling cases of hydrodynamics and mass transfer in emulsion systems. It permits the proper outline of the key asymptotic cases that are most important for experimental practice and could be solved explicitly. In Section 2 is presented an overview on the surfactant-free case of closedistance hydrodynamics of two fluid particles. The basic steps in the scaling procedure are traced. Two important asymptotic cases: (i) a solid sphere and tangentially mobile flat interface, and (ii) a solid sphere and a droplet, are analyzed. In Section 3, the scaling procedure is modified to account for the effect of the surfactant fluxes on the flow coupling in the adjacent phases. In Section 4 are summarized the basic elements of the theory of thickness of rupture. The influence of the interface mobility and the presence of surfactant are discussed, as well as the impact of film radius on film stability. A special attention is paid on the random character of the perturbations and on its consequences for the behavior of the emulsion systems. 2. EMULSION HYDRODYNAMICS The major element in modeling of hydrodynamic interactions in emulsion systems is the study of the axial approach of two emulsion droplets in an incompressible viscous medium (see Fig. 1). In this case a mutual deformation of their hydrodynamic fields occurs, as compared to the motion of a single droplet in otherwise undisturbed infinite fluid. This deformation results in additional forces acting on the particles. The two-droplet model presents the 'primitive model of emulsion hydrodynamics' and gives the basic relations that govern the peculiarities of the fluid motion and the coupling of flows in the adjacent phases.
218
E. Mileva and B. Radoev
Let us start with the surfactant-free case. The flows exterior and interior the fluid particles obey the steady Navier-Stokes equations and the continuity equation [40]: p'(v' • v)v' = -Vp' + n ' V V V-v'=0
(2.1a) (2.1b)
The following notations are used: p', v' - for the pressure and the velocity fields; u ' - f o r the viscosity and p ' - for the density of the fluid; V,V-,V~ are the gradient, divergence and the Laplacian operators. From now on all quantities referring to the droplet will be supplied with superscript (rf), and those concerning the film (continuous phase) - with (/). The basic boundary conditions that represent the coupling of the contiguous phases are [4]: • the velocity vectors outside and inside the droplets are continuous on the interface yf = \d
(2.2a)
• the tangential component of the normal stress vectors of the fluids interior and exterior the droplets are continuous through the interface:
K=K
(2-2b)
with n,\ being normal and tangential coordinates in a local coordinate system, bound to the droplet's surface. The normal components of the normal stress vectors have discontinuities but for simplicity, we shall regard the fluid particles as spherical and this boundary condition won't be necessary. 2.1. Scaling concepts As already stated, if the separation between the liquid particles becomes of the order, or thinner than their dimensions, the flow is mostly localized in this gap anizodimetric region, which generates the leading term of the drag force for two droplets and is called a thin liquid layer (an emulsion film). Due to the inherent interfacial mobility of the fluid particles, the flow outside causes a fluid motion inside them. At small gap width, the 'degeneration' of the flow pattern in the thin layer leads to the formation of a respective 'effective zone of interaction' inside the droplets, occupying regions adjacent to the thin layer outside [7-9]. With the help of this basic scaling concept, and by means of dimensional and similarity theories [41], the general
Hydrodynamic Interactions and Stability of Emulsion Films
219
hydrodynamic equations are rendered dimensionless and the range of several physically reasonable simplifications in the statement of the interaction problem are outlined. Based on the flow equations and the boundary conditions (2.1-2), the idea about the existence of this 'effective zone of interaction' is specified by introducing the following scaling parameters (Fig. 1) [8-10]: • Hf,Rf - characteristic dimensions of the liquid film, s = HJ /RJ « 1; • Hd,Rd ~ Rf - characteristic dimensions of the effective zone inside the droplets, adjacent to the thin liquid layer; • Uf - characteristic radial velocity of the film outflow; • Ud - characteristic radial velocity of the interface; Besides, due to (2.1b) the relative approach velocity of the droplets is f V =-dl/dt, and Vf ~zUf
(2.3)
Thus V1 is the characteristic drainage velocity of the film fluid motion.
Fig. 1. Two emulsion droplets of radii R at small gap widths: / - distance of closest approach of the droplet surfaces; hd(r,t) - generant curve of the droplets; (r,vf,z) - cylindrical coordinate system.
220
E. Mileva and B. Radoev
Insofar as the pressure should balance the inertia and the viscous terms in the Navier-Stokes equations, the scaling of this quantity in the film and droplet phase is [9,10]:
,/.^[,+J1,/f*l]\^r*Vl f
f
R[
f
{H )
(2.4)
f
U {H )
pd = VdK l + ReJ + {^j) f
P
(2.5)
d
R
[H ) J
AU Uf-U" , Here — - = — , and f f U U
HfRf
\if
\xd
are the Reynolds numbers for the film and the droplet phase. These scaling parameters are used for rendering dimensionless the respective quantities in the differential equations and the boundary conditions (Eqs. (2.1-2)). In a cylindrical coordinate system, the governing flow equations acquire the form of [9]: ReJ vrJ—— + —fT v J — — = dr U * dz J 1 dpf\ - ~~ J
2
n Jf
8 +Re
L
d?
AC/1 d 1 d ( , \J AUd2vf + — - + s2 [rv + — r—
&
J
2 dpf\ 2 D , AC/1 41 d _dv? 25 v/ J - ^ — 8 +Re + —j- + s r—— + s fdz L Uf \ r dr dr dz2 for the emulsion film, and
(2.7a)
(2.7b)
221
Hydrodynamic Interactions and Stability of Emulsion Films
Red\vd^
dr
+
v
d
^ dz
z
=
dp'l, nc (Rd)2] -— l + Re+\—dr [Hd)
d 1 d id\ (Rd)2d2v* + [rvf + — T \ dffdrK ' {Hd j dz2
J
{Hd) r
z
' dr
d dp" D d (R ) -J— 1 + Re +\—T dz [Hd)
,1
r
-
dz (2-8b)
,
(Hd) 1 d _dvd d2vd + \—r\ r-^ + =\Rd ) r dr dr dz2
for the droplet phase. All quantities captioned by a short line are dimensionless. The particular goal here is to track the flow coupling in the adjacent phases. The basic result is that this coupling is incorporated in an interface mobility scaling parameter. As is seen from Eqs. (2.7-8), this mobility may be expressed via the already introduced dynamic and geometrical scaling parameters, namely [8-10] Ud=-^ 1 + fn
with
^ = ^ Uf l + fh
(2-9)
where fh is a hydrodynamic factor related to the viscosity in the film and the droplet, the geometric scaling parameters of the layer and of the 'effective zone of interaction' interior the droplets, and it reflects the coupling of the flows in the contiguous phases. In the specific case of tangentially mobile but undeformable spherical droplets [hd(r)=l + R-R^jl acquires the form of [9,10]: f,=^ Jh
H
+ E2\ l + ^+u + uH\ \ H * * )
with
E = ^T V
and
-(r/R)2
], this factor
H=^— HJ
(2.10a)
For the simple case of creeping flow inside the droplets (see further (2.16)) fh specifies into a characteristic complex, involving only the geometrical dimensions of emulsion film and the ratio of the viscosities in both phases. fh=ZV
(2.10b)
222
E. Mileva and B. Radoev
Instead of fh, it is often very convenient to work with the so-called mobility of the fluid interface (see also Section 4): Mo = —
(2.10a')
Jh
Then the scaling of the interface velocity becomes:
u" = v^-
with
^L = _L_ UJ
l + Mo So, Ud =0, AU/Uf =l,ifMo
(2 . 90
1 + Mo
= 0;andUd =Uf, AU/Uf = 0, if Mo ->oo.
2.2. Thin liquid layer The mutual approach of the droplets in a regime of a constant driving force results in an overall slowdown of the flow motion in the emulsion layer and therefore, to Ref < 1 in the set (2.7). Hence, the fluid motion in the thin layer is adequately modeled by the full Stokes equations [40], which in dimensionless form are expressed as [9]: dpf\ 2 At/1 AU d2vf 2 d 1 d (__r\ J —~ £ + — r = £ \rv;)+—r T df L UJ \ dffdr' Uf dz2 f 2 dp A t / 1 e 4 1 d _dvf 2 2d v! J \ r - — 8 + —Jr = ~^- + s fdz I U \ f df dr dz2
..... (2.11a) /O11U,
(2.11b)
The structure of this set of equations allows the introduction of a general criterion for the appraisal of the lubrication-theory approximation as a model for the emulsion film outflow, namely [9,35-37]: ^ >
E
'
(2.12)
Provided the latter relation is valid, (2.11) might be reduced to the well-known lubrication-theory equations:
SLJ-!iL . ^-o or
dz
dz
(2.H)
Hydrodynamic Interactions and Stability of Emulsion Films
223
The set (2.13) has been traditionally used to model the flow in thin liquid layers within the general presumption of their geometric anizodiametry and of decreased tangential mobility of the film surfaces [3,11-12,16-22,27-28]. Note that if (2.12) is valid, the lubrication model is always adequate, no matter what actually causes the tendency for effective immobilization of the fluid interfaces. The reason might be either the higher viscosity of the droplets compared to that of the continuous phase, or the presence of surfactants (e.g. termed 'controlled slip' in Ref.[3]), or the combined effect of both (see also Section 3). If Eqs. (2.9) and (2.10b) are inserted in Eq. (2.12) the criterion for the validity of the lubrication model becomes fh >e2 [8,9]. Two main subcases are of particular interest [8]: (i)Hd[Red <1J~ RJ (see also Eq.(2.16)) - a creeping flow in the droplet, and the expression (2.12) acquires the form of pT > s , or (ii) Hd(Red >l)~RJ
/^fRe7
(see Eq.(2.16)) - a hydrodynamic boundary layer
interior the fluid particles, and (2.12) is reduced to [x>z/^Red . These relations show that the larger the viscosity ratios and the thinner the layers, the more adequate are the lubrication theory equations (2.13) as a model of the flow pattern in the emulsion film. It is also obvious from case (ii) that these equations might be used even if hydrodynamic boundary layers are formed in the contiguous phases, as is the case e.g. in Ref.[31]. What is more, higher Reynolds numbers in the droplets lead to a relative widening of the viscosity-ratio values for which Eqs. (2.13) are asymptotically correct. The lubrication theory cannot represent correctly the fluid motion in the layer only when \AU/s2Uf )< 1. The appropriate asymptotics then is presented by the following set of equations [36]:
df
dz2
d f r d r '
^L = ^lL dz
dz
dpf
did
df
,
f
Uu/e'U')*!
(2.14a)
( A £// E '£/')< 7
(2.14b)
\
df f df
^Lj^lL dz
for
dz
for
224
E. Mileva and B. Radoev
2.3. Flow inside the droplets Generally the scaling of the fluid flow inside the droplet in the direction normal to its surface might be presented as [9]: H d ~ . R" ~ ,Rf ; d 4l + Re 4l + Red Hd(Red
and
~Rf
(2.15)
or
Hd (Red > l)~-^=
(2.16)
Upon substitution of Eq. (2.15) in (2.8), the dimensionless form of the Navier-Stokes equations interior the droplets acquire the form of [9]:
ffi L d? J -®L(l + Re<)+°-L°-(rv')+(l+ Re'pl-
(2.17a)
Re^^^^L5F
L dz
V
5z
^ ;
"J
(2.17b)
r 5r
5r
v
;
dz2
This system reduces to the full Stokes equations for the case of Red < 1 [7-9]:
^1 = A Z A ( ^ ) + ^ df f dfv
df
®L = L°-F*L dz
r dr
dr
r
'
(2.18a)
dz2
+ ?*L
(2.18b)
dz
and to the hydrodynamic boundary equations, for Red > 1 [9]: ^
+
dr ^ dz
=0
v - ^ =- ^ dz dr
+
^ dz
(2.19a) (2.19b)
Hydrodynamic Interactions and Stability of Emulsion Films
225
It is evident from the dimensionless Eqs. (2.17), that the anizodiametry of the flow in the thin layer does not directly affect the order of magnitude of the various terms of the hydrodynamic equations inside the droplets. It is the Reynolds number alone, which determines the flow pattern inside the droplet phase. This is true for any interface mobility insofar, as the source of droplet's motion is exterior the fluid particles, in the thin gap between them. 2.4. Important asymptotic models The scaling analysis shows that for a broad range of cases the fluid flow in the thin layer might be modeled by the lubrication theory equations (2.13). Therefore, the essence of the effect of film drainage on the overall coupling of flows in adjacent phases might be adequately traced by analyzing simpler models, with one of the particles being just a solid sphere. Here are presented the major steps in constructing an asymptotic model for the flow interior the droplet, when it is at close proximity of another particle (solid sphere) and a thin liquid layer is formed between them [7,29-30]. The onset of this thin layer might be regarded as a restricted 'stirring' of the droplet's surface [7,8]. Insofar as at small gap widths the perturbation of the interface covers only a small portion of its surface (~ Rf), the droplet phase does not set any other geometric constraint to the 'natural propagation' of the flow pattern inside it. The realm of this perturbation is insignificant as compared to the dimensions of the droplet itself [Rf «R). Hence, the motion inside the fluid particle might be asymptotically modeled as a flow in a liquid semi-infinite space [7,8,29]. So, the basic characteristics of the flow coupling might be adequately described by the results of the first simple model: a solid sphere moving towards a flat tangentially mobile interface bordering a semi-infinite space (see Fig. 2a) [7,29]. In order to account for the effects of 'opening of the gap', another limiting case is also examined: two spheres of equal radii, one solid and the other liquid, approach along their common axis at close separations [30] (see Fig. 2b). The method of solution of these two models is based on the potential theory approach. The classical potential theory is a general methodology for finding solution of boundary-layer problems involving differential equations of elliptic type [42,43]. Odquist [44] and Ladyzhenskay [45] have specified this approach for the case of steady creeping flows, by constructing the so-called hydrodynamic potentials. The rapid development of the computer techniques has revived the use of this method [26-28,46]. Following this procedure, the flow model defined by Eqs. (2.1-2), is reformulated into velocity and pressure fields, given by integrals, whose densities are velocities and shear stresses on the mobile thin layer interfaces [45]. Hence, the problem of flow coupling is reduced to the solution of integral equations involving these densities.
226
E. Mileva and B. Radoev
Fig. 2. Asymptotic models: (a) a solid sphere and a flat tangentially mobile interface, (b) a solid sphere and a droplet.
The particular case of the motion of a particle towards a fluid interface, and of two particles along their common axis, additionally simplifies the mathematical formulation of the hydrodynamic interaction problem. Because of the axial symmetry of the model, the Stokes equations may be expressed as a biharmonic equation for the stream function [7,29-30]. In particular, for undeformable spheres and plane surfaces, the general solution might simply be presented via the usual harmonic potentials [7,28-30]. In this case, the issue of hydrodynamic interactions at small gap widths is reduced to solving integral equations for the tangential mobility on the fluid interfaces. 2.4.1. Sphere and flat interface (Fig. 2 a) The outflow of the thin liquid layer is modeled by the lubrication theory equations (2.13), while the creeping motion inside the semi-infinite liquid phase is presented as a biharmonic equation for the stream function *¥ d(r,y,z) [7,29,36,37,40]:
V 2 V 2 f—cosy} = 0 { r
with
V 2 = - ^ - + - — + ^- + J-^dr2
)
rdr
dz2
(2.20)
r2 5cp2
The velocity components are related to the stream function via the relations [40]:
*=l?f. , t = -L*f. r oz
r or
(2.21)
Hydrodynamic Interactions and Stability of Emulsion Films
227
The general solution of differential equation (2.20) might be written as [7,47]: d
XL/
cosip = zud
Aud = 0
where
(2.22)
r The specification of the respective boundary conditions is [29,36]:
(a)
vf[r,h'(r)]=0
(e)
P,{ (r ,0) = P,'(r ,0)
(b)
v{[r,h'{r)]=V/ = - —
(f)
Vd(r,6)=0
(c)
v{(r,0)=U(r)
(g)
p'(co,z) = 0
(d)
v{(r,0)=0
(2.23)
where hs(r) = -l - Rfl - ~Jl - (r/R)2 ] is the generant curve of the solid sphere. Most often in thin film studies, this curve is reduced to a parabolic profile [29,30]: hs(r) = -l-r2/2R . Equation (2.20) and the conditions (2.23) result in the following presentation of the harmonic functions in (2.22) [7,29,36]:
«'(r,q,,z) = -M"f T 2n
^ 2
^
^
V/2
2
(2.24)
2
o o [r + r - 2rr, cosfa -
as a double layer potential or, equivalently u"(r Q
A -
l
Tf\U°(r>)-U(n)] \™ Sol h V')
J [r
cosyf.dr.dy, > ~ > COS(V " ( P/)+ z J
+r
2rr
as a single layer potential. Here the notation is used:
U°{r) = -^f-
(2.26)
4h (r) The last expression is the radial velocity on a liquid/gas interface (jl = 0,z = 6).
228
E. Mileva and B. Radoev
The coupling problem as formulated by Eqs.(2.13) and (2.20), together with boundary conditions (2.23), result in the basic expression for the tangential velocity on the flat interface U{r) [29]:
£/(,)_ ir^feir^ii
=uo{r)
(2.27)
2 cos ip{_ dz ) z = n
According to the potential theory [42], the latter equation might be written in two equivalent forms [29]: —i \
hHrYc'cK 2K
\u{r,)\cos(n .r.dr.dy,
"2V
U(r) + eii^n\l
V n
V"
o o [r + ff - 2rr,
_.. .
'Z'2=U0{f)
(2.28)
cosy,)
and
U(r)=~—W^J^X
C0S9 F d d(?
' ' "' '
v/2
(2.29)
Here the effective parameters, used in rendering the respective quantities dimensionless, are as follows: Rf =^2Rl;z - •yJ2l/R;UJ =\VJ/s). Once either of the two integral equations (2.28-29) has been obtained, the problem of hydrodynamic interaction is solved, at least in principle. Any flow property might then be found by substituting the solution for U(r) in the respective expressions. Below the steps in calculating the drag force are traced. The integral equations (2.28-29) are of Fredholm type [43]. Most often similar equations are solved numerically [26-28,46]. The scaling concepts that have been introduced in the subsections 2.2-2.3 however, allow the application of an iterative procedure for finding asymptotic solutions in several important cases. Thus, at low and high viscosities in the contiguous phase, the results are presented as: (1) sfT < 1 (low viscosity in the semi-infinite space z>0). Here more convenient is the form (2.28) and the solution is developed as a power series in the hydrodynamic factor fh (as defined by (2.10b)):
t7(r)=XW^)
(2.30)
Hydrodynamic Interactions and Stability of Emulsion Films
229
and more explicitly:
C7(r)=t7»(r)-Bp:^y/(r)+^r)
(2-3D
with J/(r) shown in the Appendix (Eq.(A.l)). In the above expression, the notation O{°) is used, showing the order of magnitude of the accuracy of the asymptotic solution. (2) ejl > 1 (high viscosity in the semi-infinite space z>0). The integral equation (2.29) turns to be the more convenient initial form and the solution is given in power series of the mobility factor Mo (as defined by (2.10a')):
U(r)=YWuXr)
U(r)=-^J,(r)+ci
-4
or
U(F)=YJ(MO)UI(F)
(2.32)
or
U(r) = MoJ3(r)+a([Mo]2) (2.33)
[1^] J
en
where the explicit form of J3(r) can be found in the Appendix (Eq.(A.2)) For creeping flows with axial symmetry, the resultant action of the shear stress on a moving particle is represented by one integral property - the drag force, acting parallel to the axis of symmetry [40]. In the present model (Fig. 2a) the expression becomes [29,36]: °°J~
yfr
1
2
<234)
*--«*'fcn+wm$* The insertion of Eqs.(2.31) and (2.33) in (2.34) results in [36,37]: Fplane=-2it\ifVfIj-{l
+ 3.57zvi)
for
efi < /
(2.35)
The factor outside the brackets is the drag force for a solid sphere, approaching a flat gas/liquid interface. This is the minimal drag force value of the model system on Fig. 2a. The second term inside the brackets concerns the influence of tangential mobility of the interface. Eq.(2.35) shows that in this particular case (low viscosity in the adjacent phase), the effect of nonzero \xd leads to a relative increase in the minimal value of the drag force.
230
E. Mileva and B. Radoev
As for the case of s|d > 1, one has [36,37]:
Fplam = 6n»fVf ~{l-0.39-L)
or Fplane = 6n^Vf ^-(l-0.39Mo)
(2.36)
The factor outside the brackets is the drag force for the solid sphere, approaching a flat solid interface. This is the maximal drag-force value of the model system on Fig. 2a. The second term inside the brackets again concerns the influence of the tangential mobility of the fluid interface. It accounts for the decrease in the maximal drag force when the solid phase is replaced by a liquid semi-infinite space of finite viscosity \id. These results are in accordance with calculations of other authors as well (e.g. Refs.[26-28,48]). These references concern the surfactant-free case for similar systems and the numerical coefficients in the respective cases are comparable with our results. One should also note that the method of calculations proposed here is universal for axially symmetric hydrodynamic interactions in emulsion systems and with slight modifications might readily include additional effects, like deformation of the fluid interface, etc. 2.4.2. Solid sphere and a droplet (Fig. 2b) The goal here is to propose the correct estimation of the impact of the finite dimensions of the droplet on the flow pattern inside it. Two spheres, a solid and a fluid one, approach each other at small separations. It is important to have them with identical radii, so that the effect of the curvature to be maximal. The general procedure is completely analogous to that in 2.4.1 and details might be found in [30]. Here only an outline of the major results is given. The general expression for the biharmonic function is presented in a form proposed for the first time by Schroder [47]:
Vd
p2-R2 r
cosmr
d
d
u ; vn = 2R
p
1
5T rf
p2 sinQ 59
„
/
dVd
, vfl =
(2.37) p sinS dp
where (p,cp,9) is a spherical coordinate system, centered in the droplet, *F d is the stream function and, u is a harmonic function (Aud = 0). The boundary conditions are completely analogous to (2.23) [30]. They are only slightly modified by replacing the flat interface (z - 0) with the droplet's surface hd{r)= R[ 1 - •N/7 - ( r / R ) ]. The final integral equation for the radial velocity vf [f,hd(r)j is presented in two alternative forms [30]:
Hydrodynamic Interactions and Stability of Emulsion Films
231
^ [^, /7- (F)] - spr^^ ^') - ^ ^ ^^f? "ff ^ ^ - ^^ ^ ) 1 - ^ - ^ + ^(s )j 2n
{ a o \r2 + ff -2r,r, cosy,)
=wwm
'
\
<238a)
and
J
z
0[r
+r,
-2rlr,cos(pl)
where y*[r,hd(r)]=
V
'f'h
^
. The term on the right-hand side (RHS) of the
last equation is the radial velocity on a tangentially free interface, i.e. when one has a bubble, instead of a droplet. At hd(r) = 0, Eqs.(2.38a-b) turn into the integral equations for the radial velocity U(r) on a flat liquid/gas interface (2.28) and (2.29), respectively. One peculiarity, besides the familiar coupling factor sjl (hydrodynamic factor fh as given by Eq. (2.10b)), is the additional term that appears only in Eq. (2.38). It is of the order of the geometrical factor s , related to the dimensions of the thin layer (~£>(e)). Thus, there is a slight difference in the obtained solutions, and in the respective drag-force expressions as compared to the flat interface case in subsection 2.4.1. Two asymptotic cases are of particular interest: (1) sjl < 1, with two possibilities which reflect the relative importance of the factors fh and s : (1.1) s j l « e «1 , i.e. j l « 1 and, (1.2) s « s p T « /, i.e. jl > 1. For both cases the solution of the integral equation is expressed as a power series in (ejl):
(2.39) where J',(f) might be found in the Appendix (Eq.(A.3)).
232
E. Mileva andB. Radoev
(2) sjl » 1. Insofar as one has always e « 1, this is possible only for high viscosity in the droplet. Thus, the tangential velocity is suitably presented as a power series in (sjl)"' or (Mo) :
vrrf0[r,^(r)] = -4[^(r)+^s)]+ 4f4:Tl \o [r. h " (F)] = MO[J'3 (F) + O{z )] + o\Mo)2 J
(2.40)
with J'3(r) given in the Appendix (Eqs.(A.4)). For the two major cases of low and high droplet viscosities the drag-force expression acquires the form of [30]: Fvhere=-^fVf^-{l
+ ^[Q2+^)]}
for
efi«7
(2.41)
ej!»7
(2.42)
and F^=^'V'^\I--L[R2+C?V^
Fsphere=-2izWVf^{l-Mo[R2+a{z)}}
or
for
The numerical quotients, g 2 and R2, are shown in the Appendix (Eqs.(A.5-6)). The force expressions (2.41-42) need some additional comments. The essential functional dependence on the parameters of the system is totally determined by the presence of (at least) one tangentially immobile boundary in the thin liquid layer. If all the rest conditions (namely Vf ,1, \if ,\xd,R) are the same, the drag force has a maximum value when the mobile interface is flat. The 'bending' of this interface into a droplet surface, results in a relative decrease of the shear stress on it. This effect (might be termed 'an opening of the gap' [30]) is characteristic of the lubrication-theory cases. It manifests itself regardless of whether the adjacent phase is gaseous, liquid or solid. The ratio of the factors outside the brackets in the drag force expressions is constant, irrespective of F • Fplam/Fsphere ~ 4 (see (2.35) and (2.41); (2.36) and (2.42)) [29,30,36,48].
Hydrodynamic Interactions and Stability of Emulsion Films
233
If one neglects this purely geometric effect and examine more closely the structure of the terms related to the viscosity in the contiguous phase, two effects are clearly outlined. The first effect is 'the pure influence of the viscosity in the droplet' and it is found in both asymptotic models (3.57 in (2.35) against Q2 in (2.41); 0.66 in (2.36) against R2 in (2.42)). The second effect (~ 6>(z)) appears only in the case of a spherical droplet (2.41-42). It is related to 'bending of the streamlines' inside the droplet, as compared to those in the semi-infinite liquid space. Insofar as this effect results in additional dissipation of energy, it leads to a relative (slight, however) increase in the drag force, as compared to the flat interface case. Thus, the main conclusion, to be drawn from the presented model investigations, is that at small gap widths between the particles (s «1), the effective flow zone inside the droplet shrinks uniformly around the symmetry axis and to the surface bordering the emulsion film. Therefore, the fluid flow interior the droplet could be regarded as a motion in a semi-infinite space (Fig.2a) [7,29-30]. This model situation constitutes the true specification of the 'primitive model of emulsion hydrodynamics', always when the flow in the emulsion film is of lubrication type. 3. ROLE OF SURFACTANTS The surfactants have a profound influence on the tangential mobility of the liquid interface [38]. The geometric anizodiametry of the thin layer between the droplets ( s « l ) results in a specific coupling of their mass transport and the hydrodynamics of the system [3]. One manifestation of this coupling is the relationship between the kinetic stability of the emulsion films and the preferential solubility of the surfactant species either in the film, or in the contiguous phase. As already mentioned, this effect is a firm experimental fact and is named a Bankroft rule [32]. An attempt to link it to hydrodynamics of emulsion films was first made by Lee and Hogdson [3], and was extensively studied for different regimes of the flow interior the droplet [11,12,25,31,33-35]. The essence of the mutual influence of fluid motion and surfactant mass transfer in the adjacent phases is incorporated in the boundary conditions (2.2). They are specified to include the Marangoni stress component that arise due to the interfacial tension gradient, and the surface viscosity effect [31,38,49]: (3.1) where y is the interfacial tension. In what follows, the issue of surface viscosity [Is [50,51] is neglected. The latter might routinely be included in our analysis. It
234
E. Mileva and B. Radoev
was shown in [49], that this effect results in a slight modification of the interface mobility expression. Upon the application of the potential theory approach an integrodifferential equation is obtained for the interface mobility, instead of the basic integral equation (3.27) (see Ref.[49] for details). As has been mentioned also in [31], for emulsion films, this surface viscosity effect is usually small compared to the other terms in (3.1). So, it won't be dealt here further on. Another important boundary condition is the surface mass balance of the surfactant species:
v t (rv/)-DX 2 r=y n
AT-) with jn= M^f)
(3.2)
-'' =-/„
(3.2')
Only the case of diffusion controlled fluxes towards the droplet's surface is presented. It was established that this possibility is of utmost importance for the interface mobility when thin liquid layer is formed in the gap between the fluid particles [21,22]. Depending on the preferential solubility of the surfactant in the respective cases, jn might acquire any of the forms in (3.2'). F is the surface concentration of the surfactant; Ds,Df,Dd are the surface and the bulk diffusion coefficients; cf ,cd are the bulk concentrations of the surfactant in the phases and they obey the steady mass transfer equations [38]: f
_ / / d dcf d2cf) f vf + v{ =D\ r + r dr dz \r dr dr dz~ J fdc
dr
f
fdc
dz
yr dr
dr
dz J
(3.3)
(3.4)
3.1. Scaling analysis The analysis, already presented in subsection 2.2, has to be modified to include additional scaling parameters related to the surfactant properties, as well
235
Hydrodynamic Interactions and Stability of Emulsion Films
Fig. 3. Mass transfer fluxes: (a) and (b) for a surfactant, soluble in the film; (c) for a surfactant, soluble in the droplet phase; c" - surfactant concentration in the meniscus;
options (see Fig. 3). The following new notations are introduced [34]: • Ac{., Acf - maximum value of the characteristic changes in the surfactant concentrations along the r, z -axes for the thin liquid layer; • Acd, Act - maximum value of the characteristic changes in the surfactant concentrations along the r, z -axes inside the droplets; • 5 f ,8 d - characteristic lengths normal to the interface in the film and in the droplet, within which the major concentration change occurs. The goal further on is to correlate these new characteristic changes with the length and dynamic scaling, already introduced in Section 2.1. The major result is the estimate of the tangential mobility of fluid interfaces, in the presence of a stabilizing surfactant in the film (the continuous phase) or/and in the droplets. 3.1.1. Mass transfer of a surfactant, soluble in the thin layer For a surfactant, soluble in the emulsion thin layer (continuous phase), the estimation of the diffusion flux in (3.2') is [34]:
ii~H~&^-
(3-5)
Additional assumptions are made: (i) The local value of the interfacial tension is related to the local concentration change as at equilibrium [3,34,49]: Vj~^=% ^ dr
dcf „ dr
-
^
^L
dcf „ Rf
(3 .6)
236
E. Mileva and B. Radoev
(ii) The scaling of the various terms in the surfactant balance equation (3.2) is as follows [34]:
V.-CrvJ-^-
and V , r — ^ c ' - ^ ^ -
(3.7)
All the quantities marked with a subscript | concern the respective equilibrium properties [3,34]. Following Levich [38], the surface concentration might be written as
r=r°+r'
(3.8)
where F " is the value in the absence of fluid motion, and F ' is the perturbation of F °, due to the flow in the system. Thus the maximal value of the changes in F isF". The equation of bulk mass transfer (3.3) is used to find the relation between the scaling parameters Ac/ and Ac/, namely [34]: Pef +(Rf/hf)\ f Ac = -,—• —Ac, ; Pe'+l A / y
UfRf Pe = — ; Df f J
s /
h' -
Rf , = yll + Pef
,.m (3.9)
Here Pef is the bulk Peclet number inside the thin liquid layer. Two asymptotic cases are particularly important: (1) Pef « 1. Due to the geometrical constraints in the film \Hf «Rf), one always has 8 f ~ Hf « Rf. Besides, Ac/ ~ c"' - cf and Ac/ ~ cf - c{ , where cm is the value of the bulk concentration in the meniscus far from the film; cf is the surfactant concentration in the film bulk; c/ is the concentration in the subsurface layer near the interface (see Fig. 3a). The mass balance equation is reduced to the simple Laplace-form (RHS of the Eq. (3.3)) and,
^ - f f r ) Ac/=s2Ac/
(3.10)
Thus, due to the anisodiametry of the thin liquid layer (e « / ) , the scale of the surfactant flux in a direction, normal to the film interface is considerably smaller than the scaling of the concentration change in the radial direction,
Hydrodynamic Interactions and Stability of Emulsion Films
237
Ac^ Ac^ namely: i{ ~ Df—z—~Dfz—r7-~ej{. Thus, conditions for the depletion of surfactant are created in the film bulk. In order to maintain the normal flux towards the interface new quantities of surfactant are required to enter the layer. They come from the meniscus region, thus moving along longer distance [3,34]. This effect is also familiar from foam films [20,21]. Rf (2) Pef » 1. In the case of Hf >bf—j^=, a diffusion boundary
(3.11)
The geometrical anisodiametry of the thin layer would not cause a difference in the scaling of the surfactant flux in the normal and radial direction inside the film bulk. So, the order of magnitude of the interface flux is the same as if there is no another interface in the vicinity [34]. This is a 'classical' diffusion boundary layer [38]. Based on the above-defined additional scaling parameters the following general estimate for the tangential mobility of the layer/droplet interface is obtained [34]: Uf
U<=
with
f
' + /*+//
™
V
f +f
» <
f
(3.12)
l + fh+ff
where fh is the same hydrodynamic factor as in (2.9) and, for the case of undeformable droplet interface, is given by (2.10); / / might be termed a concentration or Marangoni factor:
ff
T" c 5 / dj \ifDfZRfdc\ +
DfRf{Rf
It simplifies to:
)dcf
(Rf/5f)2+Pef l + Pe' n
1 + Pef
238
E. Mileva and B. Radoev
r" ay // =
fl
^
Pe7
for
(3.14)
and
r°
ay
E_ v
p*
+
ar s 7 7 7 D / / ac 0 V P ^ 7
;
Note that in the scaling relation (3.12), the effects of hydrodynamics and surfactant influence are delineated. Upon the insertion of (3.12) in the criterion for the validity of the lubrication approximation (2.12), one has
fo+//V('+/*+//)>*'• Alternatively, the interface velocity scale might be expressed via mobility factors: Moi=\
and
M/=
//
7
/„+//
=
M M
° °'\
(3.13')
Mo + Mofc
with Mo defined by (2.10a'). Therefore, UfMof
At/ 1 U = 7- with —- = l + Mof Uf 1 + Mof d f f and U =0, AU/U =1, if Mo =0; or , Ud =Uf, So, the validity criterion (2.12) acquires the form of Tjd
l
-^>z2 1 + Mof
(3.12) AU/Uf =0, if Mof ->oo. (2.12')
3.1.2. Mass transfer of a surfactant, soluble in the droplet Following similar procedure as in subsection 3.1.1, scaling relations are obtained for the case of a surfactant, soluble in the droplet phase [34]:
239
Hydrodynamic Interactions and Stability of Emulsion Films
Ud =
Uf
,
with
^ =
f
>+f"d
(3.16)
where the concentration factor is:
r" 5 ^ ^ (Rflbd)2+PeJ ff = V1***'*'* 7+ P ^ +
and
Ds (bd\dT Da'Rf[Fj'd?0
(Rflhd)2 + Ped 1 + Ped
P*J™-,
5 - = - ^ - - ^
(3.18)
Note that unlike the case of Pef < 1 when the maximum value of the concentration scaling normally to the film interface is 8f ~Hf, here 5 d ~ Rf{~ Rd ~ Hd) for Ped < 1 (compare Fig. 3a and Fig. 3c). Then:
r " ay d=_v/D?_dc\
Ds
F"
//=
for
dy
s
*'»'*'o'Jrt ;+
ped<J
dT
J
^
5 r
for
^
(3.20)
> 7
£
Alternatively, the interface velocity scaling (3.16) might be expressed via an interface mobility factor Mod: UfMod
TTd
U
.t, -T
with
d
.,
U MoMod
1
d
=
-r =
fh+ff
d
c
—r
Mo + Mo
d
c
and
1
— -= f
1 + Mo Mo
Af/
T
(3.16')
d
1 + Mo 1
Moc - —j.
f
(3.1V)
240
E. Mileva and B. Radoev
with Mo given by (2.10a'). Therefore, the criterion (2.12) is specified as
T^>z'
<2 12
- ">
Putting together the scaling relations ((3.19) and (3.14); (3.20) and (3.15); (3.19) and (3.20)), it is clearly outlined that a surfactant, soluble in the droplet would always has a less pronounced immobilizing effect on the interfacial mobility as compared to a surfactant, soluble in the emulsion layer. This influence is further diminished in thinner films, due to the factor s , irrespective of the particular mass-transfer-feeding pattern of the interface ((3.19) for Ped <1; (3.20) for Ped > 1). These scaling considerations show that the coupling of film flow conditions with the surfactant flux peculiarities might add the appropriate hydrodynamic background in understanding of the experimentally observed Bancroft rule. Only in the specific case of a diffusioncontrolled mass transfer of the surfactant, soluble in the emulsion film, it is possible to ensure the necessary conditions for the so-called 'controlled slip' [3] of the tangentially mobile interfaces that could efficiently immobilize these interfaces. In all the rest of the cases, the surfactant feed-up flux toward the film surfaces is very powerful and could not sustain low tangential mobility of the interfaces. The formation of the classical diffusion boundary layer alone, on either side of the mobile interface, is of no particular importance. It has none special effect on interfacial mobility and consequently, on the flow regime in the system (compare (3.15) and (3.20), which are of the same order of magnitude). The important issue for the surfactant, soluble in the fluid particles, is that the onset of a powerful flux from inside the droplet phase towards the film surface masks all the rest details. Thus, the film drains almost like in the surfactant-free case and even for a surfactant soluble in both phases, the effect would be exactly as by the preferential solubility inside the droplets (see Fig. 3c and find the details in Ref. [34]). 3.1.3. Interface concentration balance of the surfactant Let us focus on the surface concentration balance of the surfactant species, i.e. on the boundary condition (3.2). In itself, this balance is quite similar to the bulk mass-transfer equations (3.3-4). Therefore, it is possible to introduce additional dimensionless complexes like the bulk Peclet number, which account for the relative importance of the different factors included in (3.2). For details see Refs.[35-37,52], where the surface analogue of the bulk Peclet number, named surface Peclet number is defined as:
Hydrodynamic Interactions and Stability of Emulsion Films
241
This parameter is a measure of the relative importance of surface convection and surface diffusion of the surfactant species. Other authors [12,25] have also introduced similar complexes related to other characteristic interfacial changes due to the mobility of the film interfaces. Upon insertion of Eq. (3.21) in the boundary condition (3.2), the scaling of F ' is obtained in the form of [35]:
r ' =r ° f ^ - ^ l [Pe' + lJ
with
j=
4^7 DT°/(RfY
(3-22)
Here instead as in (3.7), the scaling of the surface diffusion flux is linked directly to the maximal possible change in F , i.e. DsV2zF ~DT" /(R1)2. Thus, the surface concentration change ( F ' ) is generally related to both the fluid flow and to the mass-transfer fluxes. The characteristic parameter (j) is the ratio of the scales of bulk and surface diffusion changes, in the case of soluble surfactant [35-37]. So, the boundary condition (3.2) might therefore, be written in a dimensionless form:
[( Pes+1
J
tT
(Pes+l)'
T
J [Pe'+lJ
J
'
where F = F / F °. For full analysis of (3.23) and the consequences from this particular formulation for the validity of the lubrication approximation, including the case of insoluble surfactants, see Refs.[35-37,52]. 3.2. Important example The explicit solution of the basic surfactant-free case, which imparts the essence of the close-distance emulsion hydrodynamics, has already been presented in section 2.2.1 (Fig. 2a). The aim here is to specify this asymptotic model to account for the presence of surfactants, soluble either in the thin liquid layer, or inside the semi-infinite liquid space [36,37]. The additional scaling parameters from subsection 3.1 are used for the mathematical formulation of the interaction problem. The basic equations and boundary conditions are: • in the thin layer - the lubrication theory equations (2.13); • inside the liquid semi-infinite space - the creeping flow equations (2.18);
242
E. Mileva and B. Radoev
• mass transfer of the surfactant in both phases is described by a reduced version of (3.3-4), namely V 2 c ; = 0, or V2cd =0 (Laplace equations); • on the flat fluid interface the conditions (3.1-2) are specified as follows: pf(r,0)=Prdz(r,0) + ~tdr
(3.24)
(
\-D'°f(r,z = 0) j,=\ f
(3.25)
dz
• additional relations are also designated, namely
^-[r,h'(r)]=0,
(')=(/)(');
r(r^)=r"
(3-26)
The Stokes stream function is presented as a biharmonic function (2.22). Therefore Eqs.(2.13) and (2.22), together with the boundary conditions (2.23,3.24-26) result in an integral equation for the tangential mobility on the plane interface, which is a modification of (2.27) [36,37]:
U(r)-,p±m
^
2cosq y dz ) _n
= U'(r)
4\iJ or
(3.27,
This constitutes the basic result for the model on Fig. 2a in the presence of a surfactant in the system. 3.2.1. Surfactant, soluble in the liquid semi-infinite space Several additional simplifications are introduced here. They do not change substantially the essence of the effect of the surfactant on the interfacial mobility of the fluid interface, but make the interaction problem solvable in terms of the potential theory. First, the perturbation in the surface concentration is small, i.e. F = F " + F ' , r ' « r °. Second, the surface diffusion flux is neglected. This effect is taken into account in [49] and it does not change drastically the results. The tangential mass balance (3.25) becomes: Flrd
rnVrU{r)=DJ—(r,z dz
= 0)
(3.28)
243
Hydrodynamic Interactions and Stability of Emulsion Films
Eq.(3.28) may be viewed as a normal derivative of cd(r,(p,z) on the plane fluid interface. Insofar as the surfactant concentration balance obeys the Laplace equation, and in combination with (3.28), its solution might be presented as a single layer potential [36,37]: c (r,ra,z)=
f
=r—
(3.29)
2xD+rj-2rr,co!fal-v)+z3]/2
Besides,
(dr)
dc'Xar)^
The second multiple on the RHS of (3.30) is the tangential derivative of the single layer potential (3.29) at the plane interface. The integral equation (3.27) might be written in two forms which are modifications of Eqs.(2.28) and (2.29): _h'(r) YjA,, [U {r, )]cosy ,r,dr,d
and W
sa-JJlL
2
7c[n-f2cos(
4ndc"aii'Dd
h'fa) J
-
II
\_2
oo
[r, +r2 - 2r,r2cos{q2 -
-,
/
\V
/2
cosyfldrldql I I— 1
—?
J [r' + rf -
~,
. \
l/2
.. '
2rr,cosy,)
The respective solutions for the asymptotic cases of low and high viscosities of the liquid semi-infinite space are:
244
E. Mileva and B. Radoev
(i)6jr U^U^-ei^J^-s^)^-,^ J2(f) 2K 4n nJ D dc „
(3.32)
with J,{r) and J2{r) presented in the Appendix (Eqs.(A.l),(A.7)), and F = -%\ifVf — \l + 3.57&ix+0.66s-^——-%
(3.33)
The mass transfer of the surfactant acts towards an effective decrease of the tangential interface mobility, and consequently, increases the respective drag force (compare (3.33) with (2.35)). (2) Ejl > 1
Ui^-Llj^
+ e-j-^-rpj 4n \iJD
8(j. |_
J4{r)\
(3-34)
dc 0
J3(r) and J4{r) could be found in the Appendix (Eqs.(A.2),(A.8)). F = 6n^fVf—
7-0.39— + 0.05-^—,T -%-
(3.35)
As judged against to the surfactant-free case (Eq.(2.36)), here only to a slight decrease of the interfacial mobility term, related to the finite \\.d-value is observed. 3.2.2 Surfactant, soluble in the emulsion layer The calculation procedure is completely analogous to that in the previous subsection. The respective results are [36,37]: (l)efT<7
£/(,)= U-if)- *!pJ,lf)2%
Zp-^r? fy JAr)
4% \iJDJ
dc1 „
F = -n]ifVf —[ 1 + 3.57z\i+ 0.66-^-— •%•
2
H
l{
*
vfDf dcf J
(336) (3.37)
Hydrodynamic Interactions and Stability of Emulsion Films
245
In the present case the impact of the surfactant species is most prominent, as compared to all the other options (Eqs. (3.32-35), and see further (3.39)). The difference between (3.37) on one hand, and (3.33) and (3.35) on the other, goes through the film geometric parameter e . While it is present in the Marangoni factor in the case of the surfactant, soluble in the adjacent phase, here for a surfactant soluble only in the emulsion layer, it is absent. This is just the same situation, as has already been shown in the scaling considerations (see again Eqs. (3.14) and (3.19)). (2)sp:>7
4% |V DJ dcJ „
sjj. |_ F = 6iz\xfVf—
1-0.39^ / [
+ 0.05 Ejl
T / d
. -% E\i Df dcf
(3.39) 0\
4. RUPTURE OF THIN EMULSION FILMS It is well known that when the gap widths between the dispersed particles get too small, the onset of additional specific interactions is observed. These are usually termed surface force interactions [53]. In turn, such surface forces per unit area represent the so-called disjoining pressure 77 [53,54]. A simple relation exists between this disjoining pressure and the Gibbs energy (per unit area) G(h) of a thin layer: n(h)
= -^on
(4.1)
where h is the thin layer thickness. If the layer is a film of constant thickness, then h is simply its thickness; if the layer thickness is variable then h must be considered as the local value dependent on the position in the gap between the two particles. The relation (4.1) permits to assign the respective component of the disjoining pressure FI } , to a respective component of the energy G7. The particular type of interaction that governs the film behavior depends on its thickness. At relatively large separations (10'8m
246
E. Mileva and B. Radoev
Other types of interactions (steric, hydrophobic, undulation, etc.), known as non-DLVO interactions [53], become important at thickness h ca. 10'8 m. Insofar as the problems discussed in the present chapter concern mainly the hydrodynamics of relatively thick films (h>J0'7m), when considering the problems of the stability of emulsion films we shall further deal predominantly with the DLVO-interactions. 4.1. Linear instability condition Two emulsion droplets approach at a small separation, so that a thin layer is formed between them (see also Section 2, Fig.l). A particular profile of the thin layer is established under the action of the hydrodynamic and capillary pressures. Two limiting cases are possible: (/) quasi-nondeformed, with spherical surfaces (at relatively high surface tensions) and (//) profiles comprising plane-parallel films in their thinnest sections, i.e. thin layers of quasi-constant thickness. The transition between these two states passes through the so-called dimple formation [58-60]. Further on we shall focus on the stability of plane-parallel films that are the simplest and the most comprehensively studied system. At the end of this Section (see below Eq.(4.2"')) we shall try to generalize the results for the cases of more complex profiles. As it is known, the instability of the thin layers is most often manifested by breaking or rupturing of thin liquid lamellae - a phenomenon which is encountered in both real and model systems. Let us consider a plane-parallel film formed between two emulsion droplets that approach against one another. Since the liquid surfaces are not even but corrugated by thermal fluctuations [61], the problem of the overall film stability is focused upon the stability of the individual fluctuation modes. The universal approach to describe the fluctuation modes is the Fourier analysis. Further on we shall employ the term Fourier component as equivalent to a fluctuation mode. Corrugation results in larger surface area and consequently, to an increase in the energy, proportional to the surface tension. The resistance to this increase is expressed in the appearance of capillary pressure. For this reason the thermal fluctuations on the surface of deep liquid are stable, since only capillary forces are operating in such a case. With thin films, the fluctuations of the surface shape are related not only to the increase in area but to changes in local film thickness, as well (see also Eq. (4.2)). As was mentioned above, the energy of a thin film is a function of its thickness. Hence, upon surface corrugation the interactions between the surfaces are also perturbed. Two cases can be observed with the decrease in thickness: (i) Gibbs interaction energy decreasing, which leads to progressing surface corrugation and (ii) increase in the Gibbs interaction energy, which is equivalent to damping of the surface corrugation. In the first case, unstable Fourier components are present, while in the second case all Fourier components are stable. Since the
Hydrodynamic Interactions and Stability of Emulsion Films
247
film has two interfaces, the correlation effect of their fluctuations has to be taken into account. If there are no specific force interactions in the layer, the fluctuations on the two surfaces are independent. When, however, disjoining pressure is present, correlation arises and is expressed in the frame of two collective modes, bending mode - also called undulation mode - and squeezing, or peristaltic mode [62,63]. The bending mode preserves the thickness of the film and does not contribute to its rupture. Important for rupture is the squeezing mode, since it leads to change in local thickness. It is the squeezing mode, which we shall have in mind as a model here. Therefore, it is very important to study the dynamics of the squeezing waves (see Fig. 4) through the pressure difference Ap, of the perturbed state and the initial state in the film [64]: -A/? = p T ( O + n ( / 2 + 2 C ) - n ( / 0
(4.2)
Here py is capillary pressure, C is deformation of a single film surface from its initial state; n (h) and n (h + 2^) are the disjoining pressures in the initial state and in the perturbed one, respectively. Note that the factor of 2 before
(4.2')
where yV% is the familiar expression for the capillary pressure at small perturbations, with y being surface tension. The linear relation between driving pressure Ap and perturbation C gives the name to the linear stability analysis.
Fig. 4. Squeezing mode in a parallel film (a sketch): h - film thickness; Q - deformation amplidude.
248
E. Mileva and B. Radoev
The comparison of the role of the two perturbation pressures in the interaction energy, i.e. the estimation of the relative weight of the two terms in (4.2') is commonly performed via Fourier transformation: ^p(q) = {-iq2 + 2dn/dh)i,
(4.2")
where q is the wave number of the respective Fourier component, and Ap,C, are Fourier images of Ap.t^ . As we shall show further on (see Eq. (4.7)), film stability follows directly from the sign of the factor in brackets in front of (, (-yq2 + dYl /dh). Before that, however, we shall pay attention to the important mode with wave number qc, for which Ap, or what is the same, Ap becomes zero: q2_2dYl/dh
(43)
Y
In the literature [18] the wave qc is called critical mode, or critical wave but, in fact, this is an equilibrium steady wave; its amplitude neither grows nor decreases. It follows from (4.3) that the spectrum of the short waves (q>qc; q=2n/X, X is wave length) is stable, since the capillary forces prevail over the van der Waals interactions. In contrast, the long waves (q
Hydrodynamic Interactions and Stability of Emulsion Films
249
observed case is a thin film with a dimple. Essential in both cases is that the thickness (the profile) of the layer becomes a function of the radial coordinate r, hence, the interactions in the layer vary, i. e. n=TI[h(r)]. Consequently, the perturbations in such a layer do not spread in a medium with constant disjoining pressure, a fact that may become decisive for the thin layer stability. The problem with the variable thickness cannot be figured out in a simple manner because of the complexity of the balance (4.2"): CO
2
.
Ap(q)=-yq ( + 2 \—[n (q - q,)]((q^q,
(4.2'")
-QO
where the second term on the RHS is no longer an ordinary product but a convolution. We shall not go deeper into analysis of the integral equation, but shall do with an illustration of a typical distribution of corrugations around a local minimum of the film profile. This minimum can be an element of the thin layer at the center between two quasi-nondeformed droplets, as well as a part of the barrier ring around the dimple at the film periphery. The idea of Fig. 5 is to show that the condition Ap = 0 does not correspond to a single wave (qc) but is realized through a corrugation of variable (increasing) length and (decreasing) amplitude towards the film periphery. Moreover, at a sufficiently steep profile the thin layer may turn being stable, even when the condition of local instability (4.3) is fulfilled in the thinnest portions (see, for instance [66]).
Fig. 5. Squeezing corrugation in a parabolic gap (a sketch). Due to the variable interactions (IT[h(x)]) the deformations are with decreasing amplitude toward the gap periphery.
250
E. Mileva and B. Radoev
4.2. Role of the film drainage on the thickness of rupture The spectral analysis of (4.2") is very instructive for understanding the reasons and conditions of thin film stability. However, it does not answer the very important, from experimental and practical viewpoint question of the film lifetime or, what is the same: of the thickness of rupture. Vrij [39,67] has proposed a solution to the problem. The essence of his model is that the film ruptures when its two surfaces touch locally. The identification of the moment of touching with rupture is not a general condition and requires additional specification with respect to the action of the non-DLVO forces (for example, the appearance of films of higher stability (see e.g. instance [64]). In the more general interpretation, Vrij's model is applicable to estimating the time of appearance of black spots [23,68]. Therefore, all the important characteristics (lifetime of the film, including the velocity of coalescence in emulsion, thickness of rupture, etc.) directly depend on the estimate of the time of reaching the state of local contact of the surfaces. In other words, this is the question of evolution, i.e. of the kinetics of growth of the unstable waves. These are surface waves, a particular case of the relatively well-studied waves on liquids [38,69]. Bearing in mind (see 4.3) that the equilibrium waves Xc are as a rule larger than the thickness h {Xc>h) [17,64], the hydrodynamics of the unstable waves is adequately approximated with the already considered lubrication approximation (see Subsection 2.1, Eq.(2.10'); see also [67]): ^ = -^T(l dt 12\iJ
+ Mo)V2Ap
(4.4)
where Ap is the driving pressure (4.2'), dC, Idt is the velocity of deformation and Mo is a coefficient of the type already introduced in the previous subsections, accounting for the mobility of the film interfaces. Analogous mobility coefficient is introduced also in [70,71], where miscellaneous effects influencing the surface mobility of thin films are commented. Here we shall focus on the effects that are typical for emulsion systems. As is already shown in previous sections of this chapter, the surface mobility is controlled by the viscosity in the droplets and by the Marangoni effect (see e.g. Eqs.(2.10) and (2.10a'); Eqs.(3.12') and (3.17'); Eq.(3.37)). Note that both, hydrodynamic and Marangoni effects, might lead to Mo=0, i.e. practically to immobilization of the surfaces. The evolution of the perturbation C,(t) follows directly from the solution of (4.4), which, as we have found (see (4.2")) is suitably presented through Fourier transformation. Combining (4.4) with (4.2') yields the equation (4.5):
Hydrodynamic Interactions and Stability of Emulsion Films
^--^{l
+ MoWU-l^-V
12\iJ
dt
y
251
(4.5)
dh J
The solution of (4.5) poses two problems for thinning films: the initial condition ^{q.t = 0) and the time dependence h(t). The specificity of the initial condition is related to the fluctuation (random) character of the perturbation waves. In [65] this problem is reduced to the solution of the respective FokkerPlanck equation of the kind (see also Ref. [72]):
P^fy-4—1^W^4 \
dt
q- dh ) 8C,
(4.6)
DC,
where XCO is a distribution function (probability density) of the process, normalized on f a s J/(£ ,t]dC, = 1. In (4.6) p = 12\if jh\l + Mo)q4 is the drag coefficient, kB is Boltzmann constant, T is temperature. Since the fluctuation mean value is zero, i.e. (£) = 0, the analysis should be carried out on its dispersion (C,2(k,t)). Applying the standard stochastic methods, from (4.6) one obtains for the dispersion of the surface wave amplitudes the following evolution equation [65]:
dt
y
q dh p
'
with zero initial condition (C 2(k,t = 0)) = 0. Note that in contrast to (4.4) where zero initial condition is equivalent to trivial (zero) solution, Eq. (4.6) has non-zero solution due to the inhomogeneous term kgT. Thus, for example, for a steady state film (/?=const), where all factors in (4.6) are constant, its solution takes the form:
Eq. (4.7) demonstrates explicitly the conclusion from (4.2") that the long wave spectrum {q
252
E. Mileva and B. Radoev
while the short wave spectrum (q>qc) is stable (£2{k,tJ) (decreasing with time). It is worth pointing out that initially, i.e. .at
\f - 2(dn /dh)/q2
t«l,
the
process for both spectral branches acquires the character of Brownian motion: ((,2{q,t^0)) =
k
-^t.
(4.7')
For the kinetics of the process, the resulting equation (4.7') means that the longest time is spent around the initial state, followed by an exponential growth (or recession). For emulsion systems, the important question is whether coalescence would occur at the collision of two droplets. Translated into the language of the thin-layer model this is equivalent to the problem of the wave evolution in a thinning film. As we already pointed out, the solution of (4.6) for these cases is complicated because of the dependence of /? and IJ on the film thickness h, which in its turn is a function of time, h{t). This problem finds an elegant solution within the frame of the linear perturbation approach, i.e. in the absence of coupling between the film dynamics and the perturbation waves. Indeed, in this case the film thickness h does not depend on the perturbation £"(0 and, as a unique function of time h(t), provides the correctness of the substitution: d/dt = Vfd/dh, where Vf --dh/dt is the velocity of film thinning (Eq. (2.3), Section 2.1) (see e.g. Ref.[64]). By means of this substitution (4.6) becomes:
dh
\
q an y
'
In Eq. (4.8) time is eliminated and the amplitude of perturbation is presented as a function of the film thickness (C2\(h), which is convenient for obtaining an explicit solution for the thickness of rupture h=hrupt. Following Vrij [39], the formal condition of rupture, reads: ^2)(h
= Km) = hnipt/2
(4.9)
Here, however, comes another problem related to the random nature of the surface waves. It can be seen that the direct inverse Fourier transform of
Hydrodynamic Interactions and Stability of Emulsion Films
253
(4.8.) is divergent and the reason for this is the formal infinite correlation of the surface perturbation £(r). The correct approach requires taking into account the finite spatial correlation when estimating the mean square amplitude (f,2) (t, (r)C, (r,)). Already Vrij refers to the presence of spatial correlations of the fluctuation waves [39]. More detailed analysis is presented in [73,74]. There the term 'correlated subdomain', i.e. the region in which ((^(r)^(r ; )):£0, is defined, as well as the important for the film quantity number of uncorrelated subdomains. The main result of the stochastic analysis in the above cited studies is that the larger (by area) films contain larger number of uncorrelated subdomains, which leads to shorter life-time and, respectively, to larger thickness of rupture hrupt. Besides, these results confirm the well-known fact that the larger (macro-) systems are less stable. The authors in [73] give a numerical solution for the particular case, which corresponds to the approximate relation: K,,P,~R'fiL
(4-10)
For the traditionally studied microscopic films (film diameter from 0.1 to 1.0 mm, see Refs. [65,75]) the already commented stochastic effects, i.e. the effects of correlation, are negligible with respect to the effect of the so-called most rapidly growing wave [39]. The reason is the sharp maximum of the factor (~ I2 v^2 ~ 2{d£l /dh)\) in Eq. (4.6), which determines the most rapidly growing wave in the entire Fourier spectrum causing the film rupture. The literature abounds of formulae for estimating the thickness of rupture, based on the above considerations, whose results do not differ significantly [64,67,76]. An example of an excellent agreement with the experiment is the result obtained in [65]: *.(*!/**«, yu7
(4
.n)
It is worth pointing out that the relation (4.11) does not require the use of a thinning-film model and experimental data can be directly applied instead. The latter is a great advantage of (4.11), as with larger films, the kinetics of thinning is rather complex and its simplification within the scheme of a given model increases the danger of erroneous hrup{ estimate [77]. The fact that fluctuation wave growth competes with film thinning, i.e. that the wave velocity DC, /8t is scaled with the film velocity Vf (see (4.8)), is of a decisive importance for the role of surfactants in the film stability. As we have pointed out above, the influence of the surfactant on the wave kinetics is
254
E. Mileva and B. Radoev
accounted for by the mobility factor Mo (see (4.4)), but the mobility of the film surfaces affects to the same extent the film velocity Vf, as well. Moreover, since Vf ~ 7/p ((3 is proportional to the factor inside the brackets on RHS of Eq. (3.39), Section 3.2.2), the product F7(3 in Eq.(4.8) is independent of Mo. An important consequence of this theoretical result is that hrupt must be independent on the surface mobility. From the experimental viewpoint, the rupture thickness independence on Mo is equivalent to its independence of the surfactant concentration. This, however, is not consistently supported by the experimental results [64,78]. Such a discrepancy between theory and experiment is still an open question. Most frequently, its solution is sought in non-linear effects [76,79], a hypothesis that has not found convincing evidence so far. Another possibility for such an influence of the surfactant on hrupl is the effect of macroscopic heterogeneities in the film thickness caused by hydrodynamic factors. At higher drainage rates Vf, these non-homogeneities are more pronounced [65,77]. Since lower surfactant concentrations correlate with higher Mo-coefficient or lower coefficient fc (see Eqs. (3.19),(3.33),(3.35), (3.37),(3.39)), i.e. with higher Vf, it is to be expected that generally, lower surfactant quantities will incite stronger (macro-) heterogeneity. On the other hand, more and larger non-homogeneities are registered as higher mean film thickness [65,77]. This may be the true reason for obtaining higher than the actual values oihrupl, and not the effect of decreasing surfactant content [80]. 5. CONCLUDUNG REMARKS One very important dynamic characteristic feature of the emulsion systems is the mobility of the interfaces of the droplets. The treatment in the present chapter shows that it is determined by two major factors: (1) the relative viscosity of the droplets and the continuous phase; (2) the presence of surfactants. The viscosity factor fT was first introduced by Rybczinski and Hadamard [38]. At close separations between the fluid particles, this effect is modified due to the hindered outflow in the thin liquid layer. In flow characteristics, it appears always in a combination with the geometry of the emulsion film in the form of the hydrodynamic factor fh. The impact of the surfactant mass-transfer on the hydrodynamics of emulsion systems is related to the onset of flow-driven surface-tension gradients that tend to immobilize the interfaces, and to the responding surfactant fluxes. At small separations, the coupling of the surfactant balance and the fluid motion is modified by the formation of thin emulsion layers, thus leading, in some particular cases, to an enhanced suppression of the tangential mobility of the
Hydrodynamic Interactions and Stability of Emulsion Films
255
surfaces. This effect might be accounted for as a specification of the interfacial mobility scaling by introducing the so-called Marangoni factor: fc. For the majority of the cases of correlation between the flow motion and the surfactant fluxes, both effects, fh and fc act in an autonomous additive manner. Their overall impact on the emulsion film hydrodynamics might be accounted for via a unified parameter termed mobility of the fluid surfaces (Eqs. (3.12'),(3.17'))As for the overall stability of the emulsion systems, it is determined both by the capillary forces and the surface force interactions, exactly as by the foam systems. The interfacial mobility of the droplets is important, insofar as it ensures sufficient slowdown of the emulsion film drainage. For higher interface mobility, the drainage velocity is also increased. This increase is related to the onset of well-defined non-homogeneities of the film thickness, and therefore, in a more rapid coalescence. As a rule, higher surface mobility is observed when surfactants are solvable in the droplets, or almost equally solvable in the contiguous phases.
REFERENCES [I] J. Sjoblom, Emulsion and Emulsion Stability, Marcel Dekker Inc., NY, 1996. [2] T. Gurkov and E. Basheva, in Encyclopedia of Surface and Colloid Chemistry, A. Hubbard (ed.), Marcel Dekker Inc., NY, 2002. [3] J. Lee and T. Hodgson, Chem. Eng. Sci., 23 (1968) 1375. [4] S. Haber, G. Hetsroni and A. Solan, Int.J.Multiphase Flow, 1 (1973) 57. [5] E. Rushton and G. Davis, Int.J.Multiphase Flow, 4 (1978) 357. [6] A. Sharma and E. Ruckenstein, Coll&Polym Sci., 266 (1988) 60. [7] E. Mileva and B. Radoev, Theor.&Appl. Mechanics, Bulg. Acad. Sci., 12 (1981) 49. [8] E. Mileva and B. Radoev, Coll&Polym.Sci., 263 (1985) 587. [9] E. Mileva and B, Radoev, Coll&Polym.Sci., 264 (1986) 823. [10] E. Mileva and B. Radoev, in Particulate Phenomena and Multiphase Transport, T. N. Veziroglu (ed.) Vol.4, Hemisphere Publ.Corp., (1988) 261. II1] E. Klaseboer, J. Ph. Chevaillier, C. Gourdon and O. Masbernat, J.Coll.Interface Sci., 229 (2000) 274. [12] L. Y. Yeo, O. K. Matar, E. S. P. de Ortiz and G. F. Hewitt, J.Coll.lnterface Sci., 257 (2003)93. [13] S. Jeelani and S. Hartland, J.Coll.lnterface Sci., 156 (1993) 467. [14] S. Jeelani and S. Hartland, J.Coll.lnterface Sci., 164 (1994) 296. [15] A. Chester and I. Bazhlekov, J.Coll.lnterface Sci., 230 (2000) 229. [16] A. Scheludko, G. Dessimirov and K. Nikolov, Ann.Sof.Univ., 49 (1954/55) 127. [17] A. Scheludko, Proc. Konikl. Ned. Akad. Wetenschap., B65 (1962) 87. [18] A. Scheludko, Adv.Coll&Interface Sci., 1 (1967) 391. [19] B. Radoev, E. Manev and I. Ivanov, Kolloid Z., 234 (1969) 1037. [20] I. Ivanov, D. Dimitrov and B. Radoev, Colloid Journal, 41 (1979) 36 (in Russian). [21] B. Radoev, D. Dimitrov and I. Ivanov, Coll&Polym.Sci., 252 (1974) 50.
256
E. Mileva and B. Radoev
[22] I. Ivanov and D. Dimitrov, in Thin Liquid Films, I. Ivanov (ed.), Marcel Dekker Inc., NY (1988) 379. [23] D. Exerowa and P. Kruglyakov, Foams and Foam Films, Elesevier, Amsterdam, 1998. [24] R. Sedev and D. Exerowa, Adv.Coll&Interface Sci., 83 (1999) 111. [25] L. Y. Yeo, O. K. Matar, E. S. P. de Ortiz and G. F. Hewitt, J.ColI.Interface Sci., 241 (2001)233. [26] K. Jansons and J.Lester, Phys.Fluids, 31 (1988) 1321. [27] S. Yiantsios and R. Davis, J.ColI.Interface Sci., 144 (1991) 412. [28] R. Davis, J. Schonberg and J. Rallison, Phys.Fluids, Al (1989) 77. [29] E. Mileva and B. Radoev, Coll&Polym.Sci., 266 (1988) 368. [30] E. Mileva and B. Radoev, Coll&Polym.Sci., 266 (1988) 359. [31] T. Traykov and I. Ivanov, Int. J. Multiphase Flow, 2 (1976) 397. [32] J. Davies and E. Rideal, Interfacial Phenomena, Academic Press, NY, 1960. [33] T. Traykov, E. Manev and I. Ivanov, Int. J. Multiphase Flow, 3 (1977) 485. [34] E. Mileva and B. Radoev, Coll&Polym.Sci., 264 (1986) 965. [35] E. Mileva and B. Radoev, Coll&Surf., A74 (1993) 259. [36] E. Mileva and L. Nikolov, Coll&Surf., A74 (1993) 267. [37] E. Mileva and L. Nikolov, in Proc. First World Congress on Emulsions, Paris, France, Vol 2 (1993) 310.1-6. [38] V. Levich, Physicochemical Hydrodynamics, Prentice Hall, Engelwood Cliffs, 1962. [39] A. Vrij, Disc. Faraday Soc, 42 (1966) 23. [40] J. Happel and H. Brenner, Low Reynolds Number Hydrodynamics Prentice Hall, Engelwood Cliffs, 1965. [41] L. Sedov, Methods of Similarity and Dimensional Theory in Mechanics, Nauka, Moscow, 1977 (in Russian). [42] N. Gunter, Die Potentialtheorie und ihre Anwendungen auf Grundlagen der Mathematischen Physik, Leipzig, 1957. [43] V. Volterra, Theory of Functionals and of Integral Equations, Dover, NY, 1959. [44] F. Odquist, Math. Z., 32 (1930) 329. [45] O. Ladyzhenskaya, The Mathematical Theory of Viscous Incompressible Flow, Gordon and Breach, NY, 1969. [46] S. H. Lee and L. G. Leal, J.ColI.Interface Sci., 87 (1982) 81. [47] K. Schroder, Math. Z., 49 (1943) 110. [48] V. Beshkov, B. Radoev and I. Ivanov, Int.J.Multiphase Flow, 4 (1978) 563. [49] E. Mileva and D. Exerowa, in Proc. Second World Congress on Emulsion, Bordeaux, France, Vol.2 (1997) 2-2-235/1-10. [50] L. Scriven, Chem.Eng.Sci., 12 (1960) 98. [51] T. S0rensen, in Dynamics and Instability of Fluid Interfaces, T. S0rensen (ed.), SpringerVerlag, Berlin (1979) 1. [52] E. Mileva and B. Radoev, Commun.Dept.Chem., Bulg.Acad.Sci., 24, (1991) 513. [53] J. Israelashvili, Intermolecular and Surface Forces, Academic Press, London, 1992. [54] B. V. Derjaguin, Kolloid Z., 69 (1934) 155. [55] B. V. Derjaguin and L. Landau, Acta Physicochim. URSS, 14 (1041) 633. [56] E. J. W. Verwey and J. Th. Overbeek, Theory of Stability of Liophobic Colloids, Elsevier, Amsterdam, 1948. [57] B. V. Derjaguin, Theory of the Stability of Colloids and Thin Films, Consultants Bureau, New York, 1989. [58] R. S. Allan, G. E. Charles and S. G. Mason, J.ColI.Interface Sci., 16 (1961) 150. [59] D. Platikanov, J.Phys.Chem., 68 (1964) 3619.
Hydrodynamic Interactions and Stability of Emulsion Films
257
[60] [61] [62] [63]
K. A. Burrill and D.R. Woods, J.Coll.Interface Sci., 42 (1973) 15. L. Mandelstamm, Ann.Phys., 41 (1913) 609. A. Vrij, J.Coll.Interface Sci., 19 (1964) 1 J. G. H. Joosten in Thin Liquid Films, Surfactant Science Series, vol. 29, Marcel Dekker Inc., NY, 1988. [64] I. Ivanov, B. Radoev, A. Scheludko and E. Manev, Trans.Faraday Soc, 66 (1970) 1262. [65] B. Radoev, A. Scheludko and E. Manev, J.Coll.Interface Sci., 95 (1983) 254. [66] K. W. Stockelhuber, B. Radoev, A. Wenger and H. J. Schulze, Langmuir, 20 (2004) 164. [67] A. Vrij and J. Th. Overbeek, J.Am.Chem.Soc, 90 (1968) 3074. [68] D. Exerowa, A. Nikolov and M. Zacharieva, J.Coll.Interface Sci., 81 (1981) 419. [69] J. Lucassen, van den Terapel, A. Vrij and F. T. Hesselink, Proc. Konikl. Ned. Akad. Wetenschap., B73 (1970) 108. [70] R. Tsekov, H. J. Schulze, B. Radoev and Ph. Letocart, Coll&Surf, A142 (1998) 287. [71] R. Tsekov and B. Radoev, Int.J. Min. Processing, 56 (1999) 61. [72] E. W. Montroll and J. L. Lebowitz (eds.), Fluctuation Phenomena, North-Holland, Amsterdam, 1989. [73] R. Tsekov and B. Radoev, J.Chem.Soc.Faraday Trans., 88 (1992) 251. [74] R. Tsekov and B. Radoev, Adv.Coll&Interface Sci., 38 (1992) 353. [75] A. Scheludko and E. Manev, Trans Faraday Soc, 64 (1968) 1123. [76] D. Valkovska, K. Danov and I.B. Ivanov, Adv.Coll&Interface Sci., 96 (2002) 101. [77] E. Manev, R. Tsekov and B. Radoev, Disp.Sci.Technol., 18 (1997) 769. [78] E. Manev, A. Scheludko and D. Exerowa, Coll&Polym.Sci., 252 (1974) 586. [79] A. de Wit, D. Gallez and C. Christov, Phys.Fluids, 6 (1994) 3256. [80] J. Angarska and E.Manev, Coll&Surf., A190 (2001) 117.
APPENDIX
jfrh)f^U'%)V°SV^'t 2
2
oo \r +r,
-2rrlcos(pl)
cos
's&h-Yi^Nl, J',(r)= "fc L A
II
* **>**>v/2
^
\FdF)
(A.2) <»*<*'*>>
\2fr{r,)-h%)]\(r'+r?-2rrlCosvy2
(A3)
25 8
E. Mileva and B. Radoev
Q
(A 5)
^-'l\u*(-l%(-Y>^
/?,= f7
—
-
^J'(r)
(A.6)
. ,.>. _ Yfi - r, cosy, )Vr;U° (r, frdr.dy, J
2\r)-}}
rr2 oo
z~2 i n
^A-^
W3
[r +r, - JrrjCOSip,)
i (-\-TflTrfc ^''"JJlJJ no [no
~T'cosW-
/2
f,dr,cosyld<s?l \(-2 -2 -,— J \r + r, - 2rr,
V2 cosy,)
(A.8) —
Id
Vr=-f-r
(A.9)
r or
\ = Vr-^-4r or
r
(A- 10 )
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 7
Structure and layering of fluids in thin films D. Henderson a , A. D. Trokhymchukab and D. T. Wasan0 a
Department of Chemistry and Biochemistry, Brigham Young University, Provo UT 84602, USA b
Institute for Condensed Matter Physics, National Academy of Sciences of the Ukraine, Lviv UA 79011, Ukraine c
Department of Chemical and Environmental Engineering, Illinois Institute of Technology, Chicago IL 60616, USA 1. INTRODUCTION During the past century there have been many investigations of the properties of electrolytes near electrodes and of aqueous and organic solvents near surfaces and confined by dispersed colloidal particles. Further, the last decade has seen an intensified interest in understanding the nature of those processes that are carried out in many real systems and nontraditional systems of technological importance that comprise colloidal particles of various sizes and are under various external fields and spatial confinements [1-3]. This is the case of the large number of chemicals and consumer products, such as polymeric latexes, paints, inks, coatings, emulsions, gels, foams, cosmetics and Pharmaceuticals, food and beverages. The properties of confined colloidal suspensions are essential for processes such as sedimentation, lubrication, paper making, soil remediation, detergency, wetting of solids by micellar solutions, etc. In general, we can talk about a complex system characterized from the point of statistical mechanics by four distinct length scales: a molecular scale (a few angstroms) of the primary suspending medium due to water, some organic solvents and electrolyte ions; a submicroscopic scale (1 nm to 100 nm) that encompasses nanoparticles or surfactant aggregates called micelles; a microscopic scale (1 ^m to 100 fim) that characterizes the size of liquid droplets or bubbles in emulsions or foam systems; and a macroscopic scale. The most popular and widely used statistical mechanical model for colloidal systems, which we can call the colloid primitive model, results when the
260
D. Henderson, A.D. TrokhymchukandD.T. Wasan
molecular nature of the suspending fluid is neglected. This model has a long history and is grounded in the fact that a small size ratio of suspending fluid species to colloidal particles (typically 10H and smaller) allows one to neglect the discrete nature of the molecular solvent under some circumstances. Due to this small size ratio, it is assumed that the solvent molecules and electrolyte ions and/or the smaller (e.g., submicroscopic) species are invisible on the length scale of the giant micro- and macrocolloids and may be considered to be a continuum. The theory of colloidal dispersions that treats the suspending medium as a continuum is that due to Derjaguin and Landau [4] and Verwey and Overbeek [5] (DLVO). The DLVO theory balances the repulsive electrostatic double layer and the attractive van der Waals forces between the particles and has long been the basis of an explanation of the structure and stability of colloidal dispersion and is a classic. However, the advent of new instrumentation during recent years, such as the surface force apparatus and the thin film capillary force balance as well as atomic force microscopy and total internal reflection microscopy, has resulted in a new area of research directed to an understanding the nature of the properties of thin fluid films formed from aqueous latex suspensions, surfactant micellar solutions, and microemulsions. A direct forcedistance measurement in fluids confined between two mica sheets in the surface force apparatus made by Horn and Israelachvili [6,7] clearly showed that the interaction between two surfaces becomes oscillatory as the macrosurfaces are brought together because of the finite size of the solvent molecules. Furthermore, these authors showed that the measured interaction energy is more repulsive than that given by the DLVO theory. In other sets of the surface force apparatus experiments by Kekicheff et. al [8-10] with confined supramolecular fluids, such as micellar solutions and microemulsions, the oscillatory force caused by the supramolecule species has been measured. Further experiments by Wasan and Nikolov [11] using differential and more common interference methods, low-angle transmitted light and Kossel diffraction techniques, digital optical imaging, and video microscopy have clearly established the presence of structural forces and depletion forces as have theoretical simulation studies of particle layering and in-layer particle ordering in colloidal suspensions confined between the film surfaces. These important experimental and theoretical findings revealed the role of excluded volume forces and it is becoming apparent that it is time to move beyond the colloidal primitive model and consequently move further beyond the classical DLVO theory. A great advantage of a statistical mechanical approach is that once an adequate model is formulated and an appropriate level of an approximate description is established, a broad set of properties (including thermodynamics, structure and dynamics) and their origin and interdependence may be examined
Structure and Layering of Fluids in Thin Films
261
simultaneously within the framework of the same (theoretical) "experiment". The recent development of theoretical approaches and their results have already exposed the inadequacy of the DLVO model. This has been discussed previously [12] for the dynamics of colloidal suspensions and illustrates that the continuum has a strong effect via Stokesian friction and hydrodynamic interactions, but it is still widely believed that the discrete nature of the suspending fluid has no influence on the static properties. However, the statistical mechanical approaches [13-15] for aqueous and organic solutions, with the solvent and solute particles taken into account explicitly, agree with the experimental observations. We hope that even such a brief discussion of the state of the art of confined colloidal suspensions and molecular solutions highlights sufficiently the applied and basic impact of the layering phenomenon and show the existing problems and the perspective for future development. The particular goal of this chapter is to highlight the important role that the non-traditional structural forces, seen originally in experiment, play in the confined colloidal suspensions and properties of electrochemical interfaces and show the link between statistical mechanical theory and experimental observations. We argue that an adequate understanding of confined colloidal suspensions cannot be based on a colloid primitive model that neglects the discrete nature of the suspending medium. To proceed with this we have applied the hard-core models for the suspending medium to electrochemical interfaces and confined colloidal suspensions and developed a formalism based on the Ornstein-Zernike relation that treats homogeneous bulk fluids, electrochemistry, and colloidal science in a unified manner. In particular, we report some statistical mechanical results for the local density, electrical potential, and disjoining pressure and interaction energy of a film formed from a colloidal suspension. The scope of this chapter is as follows. In the next section we introduce the general statistical mechanics model of the complex colloidal system and present the way in which the Ornstein-Zernike equation can be applied to the dispersion problem. Section 3 is dedicated to the non-charged systems while in Sec. 4 we discuss the application to charged systems. The layering phenomenon in the films formed from non-charged and charged colloidal suspensions is discussed in Sec. 5. A summary and outlook are collected in Sec. 6. 2. STATISTICAL MECHANICS OF LAYERING PHENOMENA The system we are interesting can be modeled in a general manner as a homogeneous (or bulk) M-component fluid mixture whose components have diameters dk and densities Pk - Nk IV ; where Nk is the number of particles of species k, V is the volume, k-\,...,M. There is also a giant (G) particle, whose diameter is D, that represents a confining surface or colloidal particle.
262
D. Henderson, A.D. TrokhymchukandD.T. Wasan
We assume that there are only one or two giant particles in the mixture. Thus, the giant particle is present in zero concentration, PQ = 0 . Further, we assume pGD = 0 ; otherwise, the giant particle would affect the entire fluid. In other words, the giant particles are part of the bulk fluid. The system as whole can be called a suspension with the giant particles referred to as the species belong to the suspended phase, while the M-component fluid mixture forms the suspending fluid. 2.1. Ornstein-Zernike equation The properties of a suspending fluid near one or two giant particles can be obtained by computer simulation, using either the Monte Carlo (MC) or molecular dynamics (MD) techniques, or by means of the integral equation theory (IET) based on the Ornstein-Zernike (OZ) equation [16]
hxfli.R\2) = c^i.R\2) + Y,Pv\hXvi.Rn)c^i.R%i)^,
(1)
v
where the Greek subscripts range over all the suspending fluid components 1, ..., M, as well as giant particle G, and r a is the position of particle « . Finally, The functions hAu(R)= h^x(R) = gA/l(R)-\ are called the total correlation functions (TCFs) for a pair of particles A, and M that are separated by the distance R. The functions 8xn (R)= 8M (R) are the pair or radial distribution functions (RDFs). These functions give the local probability of a pair of particles being separated by the distance R , and are normalized so that the RDF is unity when R is large while the TCF vanishes when R is large. The functions C A/Y (-^) = C/U (^) are the direct correlation functions (DCFs) that specify the direct correlations between two particles. The integral on the RHS of Eq. (1) (a convolution integral) is the indirect part of the total correlation function. The idea behind the OZ equation can be seen most easily by considering a fluid consisting of only one component. At low densities, there are only direct correlations. Thus, h(Rn) = c(Rn).
(2)
At higher densities, a third particle comes into play. As well as being directly correlated, particles 1 and 2 are indirectly correlated by particle 3 that is directly correlated with either of particles 1 or 2. It is convenient to write this indirect term by a convolution integral. Thus,
Structure and Layering of Fluids in Thin Films
h(Ri2) = c(Rl2) + pjc(Rl3)c(R32)dr3
.
263
(3)
Continuing when more extended clusters contribute, we obtain h{Rn) = c(Rx2) +
p\c{RX3)c(R32)dr3
7 r + p2 \c(R]3)c(R34)c(R42)dr3drA
(4) +...
A convolution integral in real space is an algebraic product in Fourier space. It is common to call a series such as that in Eq. (4) a chain sum. We define the Fourier transform (FT) of a function f(R) by
J(k) = - ^ \Rf(R) sin(kR)dR k o
(5)
and take the FT of Eq. (4), yielding
h(k) = ?(k) + p£2(k) + p2c\k)
+ ....
(6)
Summing
m=
Hk)
•
(7)
l-pc(k) Equation (7) can be written as )i(k) = c?(k) + ph(kMk),
(8)
which in real space is h(Rn) = c{R{2) + p\h{R,3)c{R32)dr3.
(9)
This is the OZ equation for a single-component fluid. Equation (1) is a straightforward generalization of Eq. (9) for a multi-component fluid. The OZ system of integral equations is only an identity that defines the DCFs. By itself this system of equations is not the basis of a theory. However, by coupling the OZ equations with an approximate closure, a potentially useful theory results.
264
D. Henderson, A.D. TrokhymchukandD.T. Wasan
In this chapter, three closures will be mentioned. First, there is the PercusYevick (PY) equation [17], h^{.R)-c^{R) = y^(R)-\,
(10)
with yAM(R) = exV[J3uZu(R)]gAM(R)
(11)
called the cavity function or background function. The function u^ (R) is the interaction potential between a pair of particles and ji = \l kBT, with kB and T being the Boltzmann constant and temperature, respectively. A second closure is the hypernetted chain (HNC) approximation K(R)-cx^R) = \nyx^R).
(12)
The H N C approximation is an offspring with many parents. For a full set of references, see Barker a n d Henderson [18]. T h e name comes from t h e fact that, in contrast to the simple chains of Eq. (4), more complex chain integrals with cross links have been included. Many model systems consist of particles with a hard core Too,
R < dAu
"*<*>-Ucn «>
The infinity, oo, for R
impenetrability of the pair of interacting particles due to the presence of a hard core while wx (R) is responsible for the interaction beyond the hard core. Of course, real particles do not have an infinitely hard core. However, the repulsive part of the pair potential is steep so that approximating this repulsive part by a hard core is not unreasonable. A third closure that is useful for hardcore systems is the mean spherical approximation (MSA) [19], h^(R) = -\,
R
(14)
R>dXM.
(15)
together with
%W = - M , ( 4
Structure and Layering of Fluids in Thin Films
265
Equation (14) is an exact statement of the fact that particles with a hard core cannot overlap. Equation (15) is an approximation and can be regarded as linearized form of the HNC approximation. The advantage of the MSA is that useful analytic solutions can be obtained for a variety of model systems, including pure hard spheres, charged hard spheres and hard spheres with a dipole. The PY approximation is identical to the MSA for hard-sphere fluid, where w^M (R) = 0. Perhaps this is why the PY approximation yields a useful reliable analytic result for hard spheres but is not particularly useful for most other fluids. The HNC approximation does not yield analytic solutions. It tends to be useful for many model systems with long range interactions but not for pure hard-core systems. For many model systems, the MSA approximation combines the advantages of the PY approximation and HNC. Because the MSA yields analytic results, our attention here will be directed to this approximation. For historical reasons, when applied to hard-sphere systems, the MSA will be called the PY approximation. 2.2. Interfacial applications of the Ornstein-Zernike equation Returning to the OZ equation, Eq. (1), and letting for the density pG of giant particles be vanishingly small, we obtain a system of three equations
htJ{Rn) =Cij{Rn)+Y,Pk
,
(i6)
+ Y.Pk lhGk(Ru)cki(R32)dr3,
(17)
K(*,3:M*32)^3
k
h-Mi)
= cM2)
k
and hGG(RX2)
= cGG{Rn)
+ Y^Pk lhGk(Rl3)ckG(R32)dr3,
(18)
k
where subscripts are range over the suspending fluid components only, i.e., i,j,k = 1,2,...,M. The first equation, (16), is just the OZ equation for the homogeneous M-component fluid mixture. Equations (17) and (18) involve the giant particles and, due to this, play an important role in the statistical mechanics of inhomogeneous fluid systems, i.e., systems involving mesoscopic objects. Namely, Eq. (17) describes the fluid inhomogeneity by means of the local density profiles,
266
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
of the suspending fluid species / near the giant particle or surface. In turn, Eq. (18) describes the correlations between two giant particles mediated by the suspending fluid, and yields information about the effective interaction W(R) between a pair of giant particles via the exact relation W(R) = -kBT\ngGG(R).
(20)
The function W{R) also is known as potential ofmean force. Equations (17) and (18) are applicable to any size D of the giant particle. If D » dj, for practical purposes it is more convenient to rewrite Eqs. (17) and (18) using as variables different definitions of particle separations for each equation. First we use the particle center-to-center separations, Rt , as variables. The volume element is dr3
=2*Y23dRudR23,
(21)
where Ry are the three sides of a triangle. Since diameter of giant particles D is large, it is better to change variables so that the distance from the surface of the giant particle is used as the variable. Thus in Eq. (17), we can introduce variable z = Rn-D/2,
(22)
that denotes the distance of the center of a fluid particle from the surface of a giant sphere, and s = Rl3-D/2.
(23)
Dropping the subscript G for an easier notation, Eq. (17) yields the equation due to Henderson, Abraham and Barker [20], 00
h-Xz) = c,(z) + In^pj j
00
\ hj{s)ds \tClJ{t)dt. -oo
(24)
\:-s\
Equation (24) is often called the HAB equation. To transform Eq. (18), we introduce variable x = Rn-D,
(25)
Structure and Layering of Fluids in Thin Films
267
that is the normal distance or gap width between the surfaces of the two giant spheres, and s = Ru-D/2
(26)
and t = R23-D/2.
(27)
Thus, Eq. (18) has the form 00
GO
h(x) = c(x) + xDYJPj \hj(s)ds \cj(t)dt, j
—00
(28)
X—S
a result obtained by Attard et al [21] and Henderson [22]. Again the subscript G has been dropped for a more convenient notation. The HAB equation, (24), in accordance with the definition (19), usually serves to yield the local density profiles /?, (z) of the suspending fluid species in the vicinity of a giant particle (or near a single wall). In addition, it yields the local density profiles pt (z, x) between the two surfaces separated by a normal distance or gap width x [sometimes instead of x we will use notation H that is more common to denote the thickness of a slit-like film]. Statistical mechanics provides a relation between the local density pt(z,H) and the force/unit area f(H) that the species of suspending fluid exert on the inner side of the slit-like film surfaces in the direction normal to the surfaces, f(H) = -kT^ i
)du'^H>)Pl{z,H')dZ.
(29)
o
where u^z,!!) is the "bare" interaction between fluid species / on the distance z from confining surfaces fixed at separation H. Very often the force / ( / / ) is referred to as the solvation force; f(H = <x>) corresponds to the bulk fluid pressure p , i.e., pressure of a homogeneous phase of suspending fluid that is in an equilibria with film fluid. The solvation force measured relative to the bulk pressure defines the so-called disjoining pressure, Tl(H) = f(H)-p.
(30)
268
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
In practice, the disjoining pressure can be measured by displacing one of the surfaces a distance dH and calculating the work per unit area to bring the surfaces from the separation H to infinite separation. Thus, the film energy per unit area, E(H), can be calculated as CO
E{H)=
\U{H')dH'. //
(31)
Alternatively to the HAB equation, Eq. (28) usually serves to obtain the interaction energy between a pair of giant spheres. In particular, applying the HNC closure (12) to describe the correlations between a pair of suspended giant spheres, Eq. (28) becomes OO
In g(x) + pu{x) = TTD^PJ j
CO
\hj{s)ds \Cj{t)dt, 0
(32)
x-s
where u(x) is the "bare" interaction energy of the two giant spheres at the gap width x. Taking into account relation (20), the effective interaction between the two giant spheres mediated by a suspending media, has form QO
00
fiW{x) = pu(x) + «£>£PJ \ hj (s)ds jCj (t)dt. j
(33)
X-S
-OO
It is useful to note that Eq. (28) and consequently Eqs. (32) and (33) can be easily differentiated to give d\h(x) - c(x)]
n *^ = ~~zDTPj
'
^
^
j
"r. , , ,
, ,
\hj (s)cj(*"
s ds
)
.. .. (34)
•
-00
Therefore, the force F(x) between the two giant spheres separated by gap distance x is given by „ „
Fix) =
dW(x)
du(x) n -^
%, . . ,
.,
- ^=
^ - -£>£pj \hj(s)cj(x - s)ds.
dx
dx
2
, J
._,.
(35)
J -00
Thus, the HAB equation, (24), in conjunction with Eqs. (29)-(31), provides a description of a slit-like geometry while Eq. (28) combined with Eqs. (32)-
Structure and Layering of Fluids in Thin Films
269
(35) provides a route to the sphere-sphere geometry. The connection between these two geometries can be obtained from the Derjaguin construction [23],
nD
nD
dx
which gives the energy per unit area of two flat surfaces in terms of the force per radius between two giant spheres. We recall that here x = R-D is the gap width between two spherical surfaces and H is the separation between two plane parallel surfaces. Another form of the Derjaguin approximation concerns the disjoining pressure between two flat walls can be expressed through the second derivative of the potential of mean force between two giant spheres
dH
nD
dx2
Equations (36) and (37) are obtained by regarding the giant spheres as being composed of a collection of planar integration elements and integrating. If we suppose that the same approximation has been used in the set of OZ equations (16)-(18), then Eqs. (36) and (37) can be used to validate the Derjaguin construction. Summarizing this section, we wish to make some general and useful comments. First of all, we note that the initial set of the three OZ equations (16)-(18) form a hierarchy. The homogeneous equation, (16), can be solved without reference to Eqs. (17) and (18) that involve the presence of the giant particles. Equation (17) requires as input the DCFs of the homogeneous fluid either from Eq. (16) or some other source, such as a computer simulation. Equation (18) requires as input the correlation functions (or local densities) for the suspending fluid near a giant particle, either from Eq. (17) or some other source. Both Eqs. (16) and (17) can be solved numerically using an iterative algorithms. In contrast, a numerical solution of Eq. (18) does not require an iterative algorithm. Secondly, in applications to a particular problem, it is not necessary to use the same closure for each equation of the set (16)-(18). For example, in electrochemical applications it is often convenient to use the MSA closure for Eq. (16) and the HNC closure for Eq. (17). This procedure is called the HNC/MSA theory. Finally, in Eq. (18) the correlation functions for giant particles appear only on the LHS of this equation. Thus, applying a particular closure for the giant particle correlations need not relate to the closures used for the correlation functions of the suspending fluid on the RHS of Eq. (18). For example, using
270
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
the PY closure (10) to describe the correlations between giant spheres, Eq. (18) becomes CO
y(x)-l
= nD^PJ j
GO
\hj(s)ds -oo
Jc,(t)dt.
(38)
x~s
By comparing Eq. (38) with Eqs. (32) and (33), we obtain /3W(x) = -\ngHNC(x)
= -hPY{x),
for x>0,
(39)
where use has been made of the fact that for hard spheres y(R) = g(R) outside the hard core. The result (39) is consistent with the energy of interaction W{x) being proportional to the diameter D of the giant spheres. Thus, the HNC approximation predicts correctly that W{x) is proportional to D whereas both the PY approximation and MSA lead to the incorrect prediction that W(x) is proportional to the logarithm of D. This is because both the PY approximation and MSA are linearized versions of the HNC approximation. If the PY approximation or the MSA are used for the giant particle correlations then the ansatz, /3W(x) = -h(x),
(40)
should be applied. This ansatz is consistent with Eq. (39) and is equivalent to using the HNC for the correlation between giant spheres. In this case, using the MSA or PY closures for Eqs. (16) and (17), yields the HNC/MSA/MSA or HNC/PY/PY approximations, respectively. 3. UNCHARGED FLUIDS NEAR SURFACES 3.1. Simple hard-sphere fluid The simplest system to which the equations of the previous section can be applied is giant hard spheres dispersed in a one-component (or simple) hardsphere fluid. Wertheim [24] and Thiele [25] have solved analytically the OZ equation (16) with the PY closure (10) for a homogeneous hard-sphere fluid and obtained the DCF and thermodynamic properties. Subsequently, Wertheim [26] obtained the Laplace transform (LT) of the RDF for this fluid. Baxter [27] has given an alternative, and in some ways simpler, solution in terms of FTs and contents himself with the DCF and thermodynamic functions. Barker and Henderson [18] used Baxter's method to obtain the RDF in real space for d < R < 2d (d is fluid particle diameter) analytically. This method can be
Structure and Layering of Fluids in Thin Films
271
extended to greater values of R as has been done numerically by Perram [28]. Alternatively, Smith and Henderson [29,30] have inverted analytically the Wertheim's LT of RDF for d < R < 6d using a zonal expansion technique. However, our main interest here is not a homogeneous hard-sphere fluid but the suspending hard-sphere fluid with the presence of giant particles. The starting point for such study is a binary mixture of hard spheres. Lebowitz [31] has extended Wertheim's analysis to obtain, within the PY approximation, the LTs of RDFs and thermodynamics of a mixture of hard spheres. Using Lebowitz's result, Henderson [14] has obtained the LTs of RDFs for the system composed of giant spheres diluted in a single-component (call it species 1, since M = 1) hard-sphere fluid. Together with Wertheim's result [26], these LTs are
and
£|#)1,'"|!'"l-"A'!';-M'ti|lO, -Lx(s)e~sd' +S(s) where
(43)
variable s is the LT variable and 77, = npx 16. The fluid particle
diameter is dx=d, and t]xd\ = > is the volume fraction of the suspending fluid. The functions L\(s) and S(s) will be defined shortly. The inverse LTs (41)-(43) can be evaluated by means of a line integral in the complex plane, evaluated analytically usually by residues. Since the giant spheres are present in extreme dilution (pG = 0), the basic behavior of all correlations in the system must be that for a suspending fluid. This yields the result that the denominator in LTs (41)-(43) is the same. Henderson [14] exploited this fact and has related all three LTs to a single LT that he inverted analytically using a zonal expansion. These results are plotted in Fig. 1. Indeed, the profiles of all three curves are quite similar having an exponentially
272
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
decaying oscillatory shape with the same "periodicity" and decay length but with a different magnitude and phase. As is seen in Fig. lb, the fluid density near the giant sphere (or, for that matter, a surface) is stratified or layered just as it is in the vicinity of a fluid sphere in Fig. la. These results, first given in Ref. [20], were perhaps the earliest indication of layering near a surface. Later, this has been observed experimentally [7,11,32]. An alternative, easier but numerical, method to calculate the correlation functions in real space can be developed by converting the LTs into FTs that can then be inverted numerically by simple quadrature. With this method one can obtain numerical results for any value of R. An advantage of the inversion by means of residue theorem is that it provides one with simple analytical results for the correlation functions at short distances as well as in the asymptotic regime of large R. In particular, for the correlation function between a pair of giant spheres at their contact, x = 0, we obtain h(0)=^^.D,
(44)
2(1 -
h(x) ~ 2 | yHS | cos{coHSx + arg{yHS})exp(-/cHSx),
(45)
where coHS and KHS are the real and imaginary parts of the pole of c[g{R)\ with the smallest imaginary part, and yHS is the residue of £.[g(R)] at the same pole. The latter result (45) is plotted in Fig. lc as well and we can see that the agreement between a leading asymptotic term (45) and a complete FTNC/PY/PY result (43) is rather good except for short distances of about one layer of suspending fluid, where an asymptotic formulae should not be expected to be applicable. The results of Eqs. (44) and (45) are valuable, since, in accordance with ansatz (40), both provide information on the effective interaction between the giant spheres mediated by the suspending fluid. Recently, Roth et al [33] exploited the fact that the asymptotic term (45) describes accurately the oscillatory structure of the total correlation function and have developed a simple parametrization for the interaction energy W(x) between two macrosurfaces suspended in an ordinary hard-sphere fluid. Their parametrization provides an accurate fit to the density functional theory (DFT) results as well as to the existing computer simulation data.
Structure and Layering of Fluids in Thin Films
273
Fig. 1. Correlation functions for a pair of giant spheres of diameter D suspended in a simple hard-sphere fluid. The suspending fluid consists of spheres of diameter d whose volume fraction is <j> = 0.35 . Part a gives the RDF of a bulk phase hard-sphere fluid. Part b gives the fluid-sphere RDF, which is the normalized density profile of the suspending fluid near a giant sphere or a flat surface. Part c gives the TCF of the pair of giant spheres which is related to effective interaction or mean force potential; the dashed line represents the calculations using a leading asymptotic term (45).
In Fig. 2 we plot the interaction energy, W{x), obtained from Eqs. (40) and (43) together with DFT-based results of Roth et al [33]. Note that the energy is oscillatory and exhibits stratification. As the two giant spheres are brought closer together, layers of suspending fluid species are ' squeezed' out. At very close separations, all of the suspending fluid species have been squeezed out. When the gap between the pair of giant spheres is depleted of the species of suspending fluid one may speak of a depletion force that is attractive. This effect was first noted by Asakura and Oosawa (AO) [35], who examined the depletion force at low densities of the suspending fluid and found W(0) = -3D(#/2J. We can see that in the low density limit, expression (44) reduces to the AO result [35]. This is not surprising since the PY theory is known to be correct at low densities.
274
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
Fig. 2. Interaction energy W between two giant spheres whose diameter is D, and disjoining pressure n , between two plane parallel surfaces suspended in a simple hard-sphere fluid. The suspending fluid consists of spheres of diameter d whose volume fraction is <j> = 0.314 . The dashed lines are the results of the HNC/PY/PY approximation while the solid lines represent the parametrization of Roth et al [33] for the interaction energy, and that of Trokhymchuk et al [34] for the disjoining pressure. The symbols for the disjoining pressure correspond to the MC data of Wertheim et al [38],
The agreement of the HNC/PY/PY result [ Eqs. (40) and (43) ] with the DFT-based result is remarkably good for all range of separations except for those separations near contact at x = 0 , where the agreement is only qualitative. The result of Eqs. (44) is larger in magnitude by roughly a factor (1 - £,y2. This is the penalty for using the PY approximation to treat the fluidfluid and fluid-giant sphere correlations. Beyond contact, as the gap width between macrosurfaces increases, the interaction energy calculated within HNC/PY/PY approximation approaches the DFT-based data. Particularly, as we can see from Fig. 2, beyond the depletion region, defined by the position of the main repulsive maxima in the energy profile, both results are practically indistinguishable. This is an expected result, since, as we already noted [36,37], both approaches have common roots. Trokhymchuk et al [34], using the asymptotic term (45) for the interaction energy W(x) and Derjaguin construction (36) and (37), have given simple analytical expressions for the film energy per unit area E(H) and disjoining pressure Tl(H) of a hard-sphere fluid between two flat surfaces, as a function of suspending fluid density. These expressions are parametrized to satisfy with some known exact relations for a confined hard-core fluid and are designed to be easily implemented in the calculations of film properties in the various
Structure and Layering of Fluids in Thin Films
275
applications of the type reported recently by Wasan and Nikolov [3]. Figure 2 shows the disjoining pressure Tl(H) exerted by a hard-sphere fluid film that is calculated from parametrization of Trokhymchuk et al [34] and compared with MC data for the same system studied by Wertheim et al [38]. 3.2. Binary and many-component hard-sphere fluid and size polydispersity It is interesting to generalize Eq. (39) to an arbitrary number of components in the suspending fluid. Here we present such a generalization that is intuitive but not unreasonable. The result [37] is 3^?]le-sd!-(3S2s2/2-3S]s
£fe(*)] = ^
sld+d)
^
i<j
+ 3S0) d
D,
(46)
i
where
hiJ=36TJiTiJ(di-dJ)2,
(48)
Li:(s) = 12f7,.[(1 + ^3 /2> 2 J,. + (1 + 2S3)s] ^ , » + £ [ 1 8 7 ^ / 4 ( 4 " dj)s2 + ^ ( 1 - sdtJ)]
(49)
j*i
and S(s) = \25Q(l + 2S3)s-\SS22s2
-6S2(\-52)s3
-(l-S3)2sA (50)
+ ^dhiJ[l-S(di+dj)] The definitions of L^s) and S(s) are based on the definitions of Lebowitz, changed slightly for an easier generalization to an arbitrary number of components, and differ somewhat from the definitions used by Wertheim. Because a large number of components leads to a large number of zones, Eq. (46) is not suitable for a zonal expansion. However, numerical inversion is
276
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
not difficult. The energy and force in a case of binary suspending fluid (M = 2) with a size ratio, small-to-large 1:10, are plotted in Fig. 3. Because of the large size asymmetry, the depletion interaction scenario observed for two giant spheres in a one-component suspending fluid composed of only large particles, is significantly affected when the fine species of suspending fluid are taken into account. The gap between two giant spheres, that is depleted of the large particles in a one-component fluid, becomes filled by the fine particles in the case of a two-component fluid. Now the small species of diameter dx = ds exhibit stratification since they are confined by giant spheres of diameter D and the large suspending fluid species of diameter d2= d. This changes qualitatively the shape of both the interaction energy and force profiles that now show fine oscillatory structure at separations near the contact of the giant spheres; the oscillations are governed by the diameter of small species. To a reasonable approximation, the energy and force between two giant spheres suspended in a bidisperse fluid can be viewed as a superposition of single-component results obtained using the diameters and, perhaps, effective densities of the both fluid components. Similar trends were seen previously for a bidisperse colloidal suspension between two flat parallel walls [39]. This, of course, is to be expected since energy per unit area, E(H), and disjoining pressure, Yl(H), for a pair of flat parallel surfaces are related simply, by means of Eqs. (36) and (37), to W(H) and its derivatives.
Fig. 3. Interaction energy, W, and corresponding force, F, between two giant spheres whose diameter is D, and dissolved at infinite dilution in a two-component hard-sphere fluid thatconsists of the small spheres of diameter d\ = ds and volume fraction ^ =0.15 and the large spheres of diameter d2 = d = \0ds and volume fraction ^ 2 = 0-20.
Structure and Layering of Fluids in Thin Films
277
To a reasonable approximation, the energy and force between two giant spheres suspended in a bidisperse fluid can be viewed as a superposition of single-component results obtained using the diameters and, perhaps, effective densities of the both fluid components. Similar trends were seen previously for a bidisperse colloidal suspension between two flat parallel walls [39]. This, of course, is to be expected since energy per unit area, E(H), and disjoining pressure, Yl(H), for a pair of flat parallel surfaces are related simply, by means of Eqs. (36) and (37), to W(H) and its derivatives. Similar results are plotted in Fig. 4 for a four- and ten-component suspending fluids. Each of these multi-component fluids include the fine particles as in the case of a binary fluid. Again, there is layering but the "period" of the layering reflects the sizes of all the constituent species of suspending fluid. In general, an increase of the number of components of the suspending medium shows a tendency to destroy the fluid layering. As the giant spheres are brought together, the interfacial region becomes depleted of the larger fluid particles but can still be filled with smaller particles. Thus, oscillations are present at a wide range of separations determined by the size of the smallest and the largest suspending fluid particles.
Fig. 4. Interaction energy, W, and corresponding force, F, between two giant spheres whose diameter is D, and dissolved at infinite dilution in a four and ten-component hard-sphere fluid. In each case suspending fluid consists of fine species of diameter d\ = ds and volume fraction
whereas in the ten-component case (dashed line) the , (]±0A)d
, (1 ± 0.6)rf , and (l±0.8)J . In both cases,
the total volume fraction of the suspending species, other than fine component 1, is 0.2.
278
D. Henderson, A.D. TrokhymchukandD.T. Wasan
At this point we wish to show how our results for a many-component hardsphere suspending fluid can be extended to describe the polydisperse hardsphere suspending fluid with a continuous size distribution /(<x) that is normalized to be a probability density, i.e., \f{cr)d(j = 1. In a rigorous way the concept of polydispersity has been discussed by Salacuse and Stell [40] as well as others and applied to the polydisperse fluid of hard spheres by Blum and Stell [41]. Following their recipe, the introduction of probability density /(cr) projects a countable infinitude of the fluid components into a continuous distribution of sizes, replacing the species number densities pt by the average density //(cr,) where p is the fluid total number density. Then Eq. (46) has the form £[g(R)] =
5
, , , 3?7\f{x)e~sxdx - (\S2s2 - 3S]S + 3S0) 2*12 \\f{x)f{y)h{x,y)e-s{x+y)dxdyr,\f{x)L{x;s)e-sxdx
,
(51)
+ S{s)
where Sn=tjjf(x)xadx,
(52)
h(x,y) = 36(x-y)2,
(53)
L(x; s) = 12[(1 + —> 2 x + (1 + 2 S3 )s] 2
(54) 2
+ n J/O0[l W (x ~y)s + h{x, y){\ - s
*^)]dy,
S(s) = US0(\ + 2S3)s - 18£2 V - 652(\ - S3)s2 - (1 - 53)2s4 + \l2 \\f(x)f(y)h(x,y)(\ In these equations rj = npl6.
- s[x + y])dxdy
Structure and Layering of Fluids in Thin Films
279
Figure 5 shows the effect of polydispersity on the interaction energy between two giant spheres suspended in a fluid with the Shultz size distribution
( V+z
where (a) is the average diameter and the degree of polydispersity is controlled by parameter z. For the sake of comparison, the results of AO polydisperse model at a low density are shown as well. The AO results are calculated from oo
pWA0{x)=7^pd\f{G){(T-xfda,
(57)
x
where x = R-D. We can see that results of both, Eqs. (51) and (57), are quite similar showing that an increase of the polydispersity at low volume fraction of the suspending medium leads to an increase of the range of depletion attraction.
Fig. 5. Effective interaction energy, W, between two giant spheres whose diameter is D, and dissolved at infinite dilution in a hard-sphere suspending fluid with different degrees of size polydispersity. Part a shows results at a low volume fraction 0.01; the dashed lines represent the results of the AO polydisperse model. Part b shows results at a high volume fraction 0.3925.
280
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
As we already learned from the results of Fig. 2, a high volume fraction of monodisperse suspending medium promotes an oscillatory structural interaction between suspended species. The results presented on part b of Fig. 5 shows that introducing size polydispersity at a fixed volume fraction results in a decrease of both the attractive well at x - 0 and structural repulsive peak as the polydispersity goes up. Eventually, this leads to a smoothing the oscillatory features of the interaction energy profile. The effects of size polydispersity, similar to those of Fig. 5, have been observed by Walz and Sharma [42]. 3.3. Fluid of dumbbells, flexible linear chains and more complex aggregates The equations in the previous subsections are appropriate for a suspending fluid composed of spherical hard-core species with different diameters. An extension of the analysis to the suspending fluids comprising non-spherical particles can be provided by generalizing the set of OZ equations (16) - (18). One of the ways to do this is by using the ideas and models from the theory of site-site associating fluids. We will illustrate this by considering three models of a suspending fluid: a fluid formed from dimerizing hard spheres or dumbells, a fluid of flexible linear chains and a network-forming fluid. To formulate the fluid models we assume that the hard-sphere particles that compose a suspending fluid are of the same diameter d and each may have one, two or four attractive sites placed on the particle surface. The potential of interaction between two fluid particles for such model fluids becomes angulardependent and consists of a hard-core repulsion as in an ordinary hard-sphere fluid, plus a contribution describing a highly directional associative (AS) attraction between the sites of two different fluid particles,
fljn
^
\~£ab>
I 0,
x
ab
<x
c
r,-Q\
xab>xc
where 1 and 2 denote the position and orientations of the two fluid particles while a and b are the notations for sites on the particle 1 and 2, respectively. The variable xab is the distance between the sites a and b of the particles 1 and 2; an associative bond between two particles is formed if their two sites are within the distance xc. The parameter sah reflects the strength of an associative interaction; when sab - 0, the models converge to the ordinary hard-sphere fluid. Models with one and two sites describe the dimerizing and polymerizing fluids, respectively [43-45]. The four-site model mimics symmetrical and nonsymmetrical network-forming fluids that may consist of very complicated
Structure and Layering of Fluids in Thin Films
281
aggregates, e.g. linear branches, loops, crossing rings, etc [46]. Therefore, the four-site model is able to reproduce various kinds of macromolecules and a network of bonds by appropriate choices of the energy parameter sah . The properties of these model fluids in a bulk phase have been investigated intensively during the last decade [45-49] using a multidensity version of the OZ equation (16) that is usually called the Wertheim OZ (WOZ) equation [43,44]. For a simplest case of a dimerizing fluid, i.e., with only one site per fluid particle, WOZ reads
hf{\,2) = cf (1,2) + X \Kr {\,l)PrEcf (3,2)3,
(59)
ye
where the Greek superscripts a,/3 denote the bonding status (equal to 1 when site is bonded and 0 when site is non-bonded, i.e., free) of the sites located on the pair of particles 1 and 2, respectively. In a Wertheim multidensity formalism [43,44] the fluid particle total and direct correlation functions //,-,-(1,2) and c,7(l,2) become matrices with the elements, h"p{\,2) and c,^(l,2), called the partial total and partial direct correlation functions, respectively; the fluid density p also is replaced by the density matrix whose elements represent the number densities of the fluid particles with a bonded status of their sites
k"]=[^ *"].
(60)
where p° refers to the density of non-bonded particles. In contrast, none of the partial correlation functions has the physical meaning; as in the case of a simple hard-sphere fluid the physical meaning is attributed to the total correlation function that in this case is a linear combination of partial correlation functions
^(l,2) = ^ 0 (l,2) + 2[ P ° k°1(l,2) + f /7 ° j ^'(1,2).
(61)
Moreover, the partial correlation function /z°°(l,2) is a part of the selfconsistency relation [43,44] that splits the total fluid density on the density bonded and non-bonded particles
282
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
p = p° + (p°J {[/z°°(l,2) + ll/;.f (\,2)d2,
(62)
where ftfs (1,2) - exp[-u°f (1,2)/kT]-1, is the associative analogue of the Mayer function. Of course, if there are more than one site per particle, i.e., in the case of polymer chain (two sites per particle) and network (four sites per particle) fluids, Eqs. (59) - (62) are more complex [45,46]. However, using the associative generalization of the PY closure and within the two-particle density formalism for a dimerizing fluid [43,44], and the ideal chain [45] and ideal network [46] approximations to the description of bulk polymerizing and network fluids, respectively, leads to a significant simplification in the theory. In particular, the WOZ equation (59) may be solved analytically in the zerolimit for the range of associative interaction (58), i.e., xc —>0, by a method analogous to that used by Baxter [27] in the case of hard-sphere model. This sticky limit for the association is considered with the angle-averaged associative Mayer function being the singular form:
(ff (1,2)) =-*"*(*> [(e-M-^y
1] = K8(R - d),
(63)
where K is the parameter uniquely determined by the strength of associative interaction between sites a and b. The calculations are simplified further by the assumption of energetic equivalence of attractive sites, i.e., sab = s. To show the effect of the non-spherical nature of suspending fluid on stratification phenomena, in contrast to preceding subsections 4.1 and 4.2, here we explore the route utilizing Eq. (17), i.e., route based on the HAB equation. The route that is based on Eq. (18) has not been elaborated for the associating fluids so far. The associative version of hard-sphere HAB equation has been proposed by Holovko and Vakarin [47] and for dimerizing fluid reads
Ko(Rn) = c%(Ru) +1 K f (Rn)pPrcJ?(h2)dr,,
(64)
Pr where hjG°(R) and cfg(R) are the partial total and partial direct fluid speciesgiant sphere correlation functions, respectively. The superscript 0, that corresponds to giant sphere subscript G, indicates the absence of any bonded states at the surface of giant spheres if an associative interaction of the type of (58) between fluid particles and giant sphere is not considered.
Structure and Layering of Fluids in Thin Films
283
Recently, Duda et al [50] have applied the associative HAB equations of the type of Eq. (64) to evaluate the local density p(z,H) = phlG(z,H) for the suspending fluid composed of dumbells, flexible linear chains and network aggregates confined to a slit pore. The fluid-wall total correlation function hiG{z,H) was calculated as a linear composition of the partial fluid-wall functions hiG(z,H) and h]G{z,H) in the way similar to the bulk fluid case, Eq. (61). Some representative density profiles are plotted in Fig. 6 as a function of fluid particle distance from the confining surfaces. To mimic the reality where confining surface can be lyophilic or lyophobic, the fluid species-giant sphere interaction, ujG(z,H), was used as a (9,3) Lennard-Jones potential with and without attractive term, respectively. For the comparison, the local densities of an ordinary hard-sphere fluid are shown as well.
Fig. 6. Density profiles p(z,H)of a simple hard-sphere fluid (thin solid line), fluid consisting of dumbbell-shaped particles (dashed line), linear flexible chains (short dashed line) and network-forming fluid (thick solid line) in a slit-like film. The film thickness is H = 8d.
284
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
The quite tall and narrow peaks that are seen in Fig. 6 next to the confining surfaces reflect the well defined surface layers for a hard-sphere fluid. These peaks decrease for a fluid of dimers, decrease and widen for polymer chains, and bifurcate and almost disappear for fluid composed of network aggregates, especially in the case of the repulsive fluid-surface interaction. As the distance from the substrate increases, the heights of the peaks decrease and their widths increase. Nevertheless, we still can observe at least three well-defined layers even for flexible linear chains. At the same time, it is worth noting that the left shoulder of the double first peak of the network fluid density profile, that we observe near the attractive surfaces, totally disappears near the repulsive surface. It means that the fluid comprised of species of irregular shape tends to segregate from the confinement diminishing the effect of stratification. To understand the qualitative trends in the film properties initiated by the shape of suspending fluid species and different level of fluid stratification, in Figs. 8 and 9 we show the results for the force between film surfaces that is mediated by suspending fluid. Experimentally the force/radius between the two crossed cylinders of radius a-D/2 can be determined directly in the method of Israelachvili [7]. To obtain results that are comparable with measurements we have added to the fluid mediated force the long range van der Waals contribution [the so-called Hamaker force] that arises from the direct interaction between the surfaces. This interaction is obtained by integrating the - R~6 term in (say) the Lennard-Jones interaction (a four-fold integration) over the volumes of the two interacting bodies. The result is complex. However, if the bodies are giant, the result is quite simple and in the case of the two crossed cylinders equals - ADlYlx. Thus, the calculated force, which can be compared with the observed data, has the form [see, e.g., Eqs. (31) and (36)]
F
^
a
=
^
\2x2
+ n [U(H )dH ,
(65)
I
where A is the Hamaker constant and H is the gap width between confining surfaces. Figure 7 shows, by a dashed line, the force/radius calculated according to Eq. (65) between a pair of macrosurfaces that are suspended in a hard-sphere fluid of density pd3 =0.85. Taking a hard-sphere diameter d =0.9nm such fluid roughly simulates octamethylcyclotetrasiloxane (OMCTS), which is the frequently investigated experimental system [7]. The inset shows the measured force law between two cylindrically curved mica surfaces in OMCTS [6]. OMCTS is an inert silicone liquid whose non-polar molecules are quasispherical in a shape. To take into account the effect of non-sphericality, we
Structure and Layering of Fluids in Thin Films
285
assumed that the OMCTS molecule is slightly elongated in one direction and this shape can be represented by a dumbbell composed of two fused hard spheres, each of diameter dd. Choosing the diameter dd from the condition that the volume of the dumbbell-shaped molecule is the same as the volume of the spherical molecule of diameter d = 0.9nm, we performed calculation of the force in a dumbbell fluid with diameter dd =0.73nm, maintaining fluid volume fraction the same as in a monomer hard-sphere fluid. We can see that two calculated force profiles differ with the dumbbell model improving an agreement with experimental data.
Fig. 7. Force between two macrosurfaces suspended in a hard-sphere fluid of diameter d = 0.9 nm [dashed line] and in a dumbbell fluid of diameter dd =0.73nm [solid line]. The dotted line shows the van der Waals force. The fluid volume fraction in both cases is the same, pd
= 0.7. The inset shows the measured force law between two cylindrically curve
mica surfaces in OMCTS with average molecular diameter - 0 . 9 nm. The inset is adapted from Fig. 5 of Ref. [6].
286
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
Figure 8 shows the predicted results for the force between two macroscopic surfaces in a chain and network-forming fluids. These two results can be compared with measured force laws between mica surfaces observed by Christenson et al [51] across straight-chained liquid alkanes such as ntetradecane and n-hexadecane, and by Gee and Israelachvili [52] measured in a branched alkane [iso-parafin] 2-methyloctadecane. We see, that the force in linear chain fluid still exhibits oscillatory behavior, similar to the film of spherical molecules, but is shifted to the attraction region. The force mediated by a network fluid is completely different: it is repulsive until a thickness of about 1 nm [of order of two-to-three single bead diameters] and then becomes attractive and almost totally monotonic. The overall qualitative agreement between the calculated and experimental results is quite good.
Fig. 8. Force between two lyophilic macrosurfaces in a fluid of linear chain (thin solid line) with d - 0.4 nm, and in a network-forming fluid (thick solid line) with d = 0.45 nm. The density of both fluids is pd =0.85 .The dotted line shows the van der Waals force. The inset shows the measured force laws between mica surfaces in straight-chained liquid alkanes such as n-tetradecane and n-hexadecane (molecular width —0.4 nm), and in the branched alkane (iso-paraffin) 2-methyloctadecane. The inset is adapted from Fig. 13.5 of Ref. [7], page 272.
Structure and Layering of Fluids in Thin Films
287
The repulsive force between lyophilic surfaces at small separations originates from the first adsorbed layer of suspending fluid and extends on the separations of order of the thickness of this layer. The fluid layering near substrate can be identified from the oscillations of local density [see Fig. 6]. Two density maxima that we observe in a network-forming fluid next to the surface can be treated as the spliting of a rather thick first layer, which remains disordered or amorphous. The local density in such layer is always higher than the corresponding bulk density, and the force exerted by the fluid on the surfaces is repulsive. The fact that the chain structure of the suspending fluid can result in an attractive or repulsive contribution to the solvation force has been pointed out by Israelachvili [7] who calls this effect a positive or a negative solvation force. 4. CHARGED FLUIDS NEAR SURFACES In electrochemistry and colloidal science, the suspending medium is usually an electrolyte containing charged species - ions and polar solvent molecules. Further, the confining surfaces - electrode and/or the colloidal particles are charged. The MSA can be applied to such complex systems. 4.1. Bulk electrolytes The simplest model of an electrolyte is a system of charged hard spheres (the ions) in a continuum dielectric medium whose dielectric constant is s (the solvent). To keep the model simple we assume that the dielectric constant inside the charged hard spheres is the same dielectric constant as that of the electrolyte so that we need not worry about induced charges on their surface. This model is called the primitive model (PM) of an electrolyte. If the charged hard spheres all have the same diameter, this model is usually called the restricted primitive model (RPM). Waisman and Lebowitz [53] have solved the OZ/MSA equations analytically for the RPM. Their results have been simplified by Blum [54], who has also solved the OZ/MSA equations for the more general case of differing diameters, i.e., for the PM. In the case of the bulk RPM, the ion-ion RDFs are given by
J
s(\ + Yd)2
R
where d is the ion diameter. For the moment, all ions are considered to have the same diameter. The function gHS(R) is the hard-sphere RDF, f(R) is a
288
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
function that will be defined shortly, and F is a parameter that is related to the Debye screening length, K , defined by
-2 =4 fl>,V,
(67)
through the relation K = 2Y(l + Td).
(68)
Note that at low concentrations, when K is small, F = K/2 .
The LT of f(R) is given analytically, L[f(R)] = -7
,
•
(69)
Henderson and Smith [30] have inverted this LT and obtained f(R) as an infinite sum that involves spherical Bessel functions. From Eq. (69), we see that for small K (low concentrations) L[f(R)] = — -
(70)
S+K
So that f(R) = e~KR.
(71)
Thus, at low concentrations, the MSA result for the ion-ion RDFs becomes fin n
a-K(R-d)
If we make the approximation, valid at low concentrations,
{\ + Ydf
( d\2 K =[\ + =l + xd,
(73)
Structure and Layering of Fluids in Thin Films
289
neglect d in the argument of the exponential, and use gHS(R) = l,
(74)
the MSA result becomes the Debye-Huckel (DH) result [55]. We have obtained the MSA result from an integral equation whereas the DH approximation is usually obtained from a differential equation [often called the Poisson-Boltzmann (PB) equation]. The DH or PB approximation is the zero-diameter version of the MSA for charged hard spheres. In the DH/PB approach, Eq. (15) is assumed to be valid for c(R) for all distances. The MSA is a systematic generalization of the DH theory (strictly speaking in its linearized version). The factor 1 of the DH/PB approach becomes gHS(R), the exponential function becomes f(R), and (1 + KT/) becomes 2 (1 + Yd) . We will see that the factor (1 + Yd)2 overcomes the asymmetry that is a problem in the DH/PB theory. In the DH/PB approximation, only the central ion is given a size. In the MSA theory, all ions are treated consistently. The appearance of g (R) leads to oscillations that reflect the stratification phenomenon. If the PM is used, the density of the hard spheres in an electrolyte is low and these oscillations are correspondingly small. However, if an explicit molecular model of the solvent is used, the hard-sphere oscillations become important. The function f(R) also oscillates but this is not very important as the concentration of ions is small, whether or not an explicit model of the solvent is used. The oscillations in f(R) become more important with divalent ions. In a molten salt, the oscillations of f(R) would be of even greater importance. The renormalized screening parameter, 2F, is smaller than K SO that the system does not become overscreened. In the DH/PB theory \lK can become smaller than d. This is an unphysical situation because it would mean that the screening atmosphere is inside the central ion. We have commented that the MSA is a generalization of the linearized DH or LDH theory. The HNC is an appropriate generalization for the nonlinear DH theory [which does not yield an analytic solution]. Our attention here is restricted to the MSA and LDH approximation. The generalization of Eq. (66) to the case where the electrolyte ions have different diameters is difficult. However, considerable progress has been made by Blum [54]. As long as most of the ions have the diameter d and there are only a very small number of ions with different diameters, the ion-ion RDF is given by
290
D. Henderson, A.D. TrokhymchukandD.T. Wasan
where gjj (R) is the RDF of a hard-sphere mixture. The symmetric treatment of the ions [lacking in the DH/PB theory] is apparent in Eq. (75) and is of particular importance in applications to charged colloidal suspensions. To progress beyond the PM, we must include a molecular model of the solvent. A particularly simple model of the solvent is obtained using a system of hard spheres together with the dielectric background. This model has been called the solvent primitive model (SPM). At first glance, this would seem to be an unpromising model of a polar solvent. However, this model does recognize that the solvent molecules occupy space. Within the MSA, results obtained for a bulk electrolyte using the SPM are similar to the results obtained above with the PM. The solvent molecules are just another species of charged hard spheres that happen to have zero charge. Thus, Eq. (75) still applies but the density of hard spheres now includes the density of solvent and is large and g{fs(R) has pronounced oscillations. The second term in Eq. (75), that controls the electrical response of the system, is unaffected by SPM. Of course, this separation of the charged and non-charged terms is a feature of the linearized nature of the MSA. It would not be occur in a nonlinear theory, such as the HNC approximation where the effect of the SPM would propagate to the charged terms. A somewhat more sophisticated model of a solvent is a system of dipolar hard spheres. Now there is no dielectric background. The dielectric constant is the result of the non-zero dipole moment, /z, embedded into the solvent spheres. Wertheim [56] has used the MSA to obtain the properties of the dipolar hardsphere model solvent. He found \6s = A2(l + A)4
and
(76)
Structure and Layering of Fluids in Thin Films
291
where ps is the number density of the solvent. Note that Eq. (76) is a cubic equation in X1 that can be solved analytically. For s=\, X=\ and for £-78, X-2.65. Thus, X is a weak function of s. These equations give the relation between dipole moment /j and s. Although approximate, this relation is an improvement over the earlier Onsager formula. The value of /J , corresponding to the dielectric constant of water, is greater than the dipole moment of water vapor molecules. The usual view is that at high densities Eqs (76) and (77) give an effective dipole moment that accounts for such effects as polarization. Blum [54] has solved the MSA for a mixture of dipolar and charged hard spheres [loosely called the ion-dipole model] and has obtained results for the thermodynamic properties for this model electrolyte. The MSA results for this ion-dipole model will be discussed further in the next section. 4.2. Electrolyte near an electrode Equation (75) can be applied to an electrolyte near an electrode. Many electrodes are planar but some, for example a hanging drop mercury electrode, are non-planar. The planar electrode can be regarded as a giant sphere. The RDF for a pair consisting of a giant sphere of charge Q and radius a - D/2 and an ion of charge qt and diameter d, or in other words the normalized density profile gt(R) = pt(R)f p of the ions near an electrode is
M w-a w- n ?*L,/(*-f-f) a s{\ + Td)2Ta
v
™
2 2J
or gi(z)
=g f ( z ) - ^ . O o / ^ - 0
(79)
where z = R-D/2 and % =- ^
(80)
is the electrostatic potential at the giant sphere. Equation (68) has been used to obtain Eq. (79) which can be rewritten as g,(z) =g -(z)-^/z-*a £K \ 2
(81)
292
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
where E is electric field at the electrode. The electrostatic potential of the electrode, V, is <J>0 plus the potential difference, Ed I'2s, across the region of thickness d 12 between the electrode and the distance of closest approach of the ions. If the PM is appropriate, the PB theory is called the Gouy-Chapman (GC) theory [57], which actually predates the DH theory. If the GC/PB theory is applicable, Ed
ir
.
Ed
E
^ = ^ + ^o = — + —• 2s 2s SK
(82)
Using the MSA, gives
v
( 83 )
= 4^'
s2F and with Eq. (68),
1 1+IW 1 d — = = - + - + ... 2T
K
K
(84)
2
the GC/PB theory is recovered. The region near the macrosurface 0 < z
(85)
where CD is the diffuse layer capacitance, as predicted by the GC/PB theory, and CH is the inner, or Helmholtz or Stern, layer capacitance that is independent of the concentration and other properties of the ions. As will be seen shortly, this picture is overly simplified. As seen in Eq. (84), two terms appear in the MSA expression for V only if the expansion is truncated at two terms.
Structure and Layering of Fluids in Thin Films
293
To bring Eqs. (82) and (85) into agreement with experiment, the properties of the inner layer must be adjusted. One popular scheme is to say that the dielectric constant, s *, in the inner layer is different from the bulk value. Thus, Eq. (82) becomes
2s*
SK
A value s * =3~6 seems appropriate. A reduced value for s * seems reasonable since the solvent is expected to be less polarizable near the electrode. However, this empirical approach is less than satisfying. Is it reasonable to confine the interfacial region of the solvent molecules to a region right at the electrode? Is a sharp dielectric boundary reasonable? Maxwell's equations require a polarization charge at a dielectric boundary. This effect has been satisfied only in part because no attempt has been made to ensure that the induced charge does not disturb the GC/PB ionic profiles. For that matter, induced charges have been ignored at the electrode surface also. However, this latter point may be reasonable there since 78 is well on the way to infinity [the value of the electrode dielectric constant, assuming the electrode to be metallic]. Since the value of s* is empirical, one might argue that the effect of the induced charges has been included in the value of s *. For this to be the case one would need to establish that the induced charge at the interface is independent of concentration [or at least only weakly dependent].
Fig. 9. Ion density profiles pt(z) calculated from the GC/PB theory (solid lines) compared with MC simulation data (symbols). Part a gives typical results for a monovalent salt. Part b gives typical results for a divalent salt.
294
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
As was the case for a bulk electrolyte, the PB theory is the point charge [or low concentration] limit of the MSA theory. For the double layer, the PB theory is called the GC theory. In contrast to the homogeneous version of the PB [the DH theory], the GC/PB equations can be solved analytically even when the electrode charge is large and the equations are nonlinear. The nonlinear version of the MSA theory, the HNC approximation, requires a numerical solution. Linear or nonlinear, the GC/PB results for concentration, charge, and potential profiles are monotonic and the term double layer is sensible. In contrast, in the MSA or HNC approximations these profiles can oscillate, again reflecting the presence of stratification phenomenon. As is seen in Fig. 9 where a comparison of the GC/PB ion density profiles with MC simulation data is shown, this is particularly apparent for divalent salts where the oscillations in the density profile are large enough that there is charge inversion where the density of co-ions can exceed the density of counter-ions. In this case the charge in the layer of counter-ions near the electrode exceeds the magnitude of the electrode charge and there is a further layer of co-ions. The total charge in the profiles is still equal and opposite to that of the electrode as overall charge neutrality requires. However, rather than a double layer, there is a triple layer of charges. The oscillations seen in Fig. 9 are relatively mild. Much more pronounced oscillations would be seen if a molecular model for solvent were used. One result of the GC theory is kTYpAd/2)
= kTp +
F2
,
(87)
where p = ^ pt is the total density of the ions. This is a special case of the exact, within the PM or SPM, result due to Henderson et al [59] kT^Pi(d/2) = p + ^ ,
(88)
where p is the bulk pressure (including both hard sphere and electrostatic contributions) of the suspending electrolyte. The term on the LHS in Eq. (88) is the momentum transfer to the electrode. This must equal the sum of the pressure and the second term on the RHS, which is the electrostatic (or Maxwell) stress. The GC/PB result is a special case of the exact result for the case where p = pkT (i.e., uncharged point particles with zero diameter). Neglecting terms that are higher order in E than linear, Eq. (88) becomes
Structure and Layering of Fluids in Thin Films
295
kTy£dPi(d/2) = p.
(89)
In contrast, the MSA and HNC theories give, instead of p, a geometric or arithmetic mean of the hard-sphere values of p and p/3(dp/dp), in Eqs. (89) and (88), respectively. At low concentrations and not too large values of the dimensionless charge parameter, flq^jlsd, the difference with p is small. However, if the concentration is significant, the valence of the salt is large or if the temperature, dielectric constant, or diameter are small, p can be large or, more usually, small, resulting in large or small contact values for the ion density profile. In contrast, the GC/PB, MSA and HNC results for the contact values of the ion density can only equal or exceed the bulk density. Thus far, only the PM and SPM have been considered. Carnie and Chan [60], Chan et al [61] and Blum and Henderson [62] have obtained an analytic solution of the MSA equation for the ion-dipole model of suspending fluid. For ions of diameter d and solvent dipolar hard spheres of diameter ds that all of the same size ds - d, the normalized ion density profiles have the form gl(z) = gHS(z) + Ahi(z)
(90)
and the normalized solvent density profile has the form gs(z,0) = gHS(z)+ l3Ahs(z)cos0, where gHS(x)
(91)
is the normalized density profile for neutral hard spheres (ions
plus solvent molecules) near a hard wall. Although analytic, the MSA expressions are not explicit. However, at low ionic concentrations an expansion may be made and explicit results can be obtained. For example, the contribution to the potential resulting from an integration of gt(z) is VI =E +Ed +..., (92) K 2 and the contribution to the potential resulting from an integration of gs (z, 0) is
K,=[^(l-I]l^ + ....
(93)
296
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
The total potential is
K . £ + flf, + £rl) + .... SK
2s V
(94)
1 J
For the more general case of differing diameters where ds •*• d, Carnie and Chan [60] have shown that Eq. (94) becomes
V = - + ^-[d + ^ d , ) SK
2s\
A
+
....
(95)
)
It is to be noted that there has been a cancellation between terms of order 1 and (1 - s) I s to produce a term of order 1 / s. This is no problem with analytic expressions but might be a problem in numerical calculations when s -80 as a small error in either term might result in an appreciable error in V.
Fig. 10. Values of Ahs(z) electrode.
from the MSA for a 0.1M monovalent electrolyte near an
The parameter Cs is a normalization constant that is not relevant to our
discussion. The solid curve is the MSA result, calculated using the ion-dipole model, whereas the dashed curve is a plot of exp[-xr(z - d 12)]. Adapted with permission from Schmickler and Henderson [63].
Structure and Layering of Fluids in Thin Films
297
When the higher order terms are neglected, Eqs. (94) and (95) are formally identical to Eq. (86). However, the interpretation is quite different. The potential consist of two terms only if the expansion is truncated at two terms. Further the "inner layer" term in Eqs. (94) and (95) results from an integration over all space. The interfacial region for the solvent molecules is as diffuse as that of the ions. As long as the screening of the ions is incomplete, there is an electric field and, as long as there is an electric field, the solvent molecules will be oriented. We do not have two capacitors in series. If anything, we have capacitors in parallel. Note in Eq. (95) that ds is weighted by the factor (s-\)/A. that for an aqueous solution is about 30. This means, that in agreement with experiment, the "inner" layer term is not affected appreciably by the nature of the ions in the electrolyte. The MSA function Ahs(z) is plotted in Fig. 10. Note that Ahs(z) oscillates and ultimately decays exponentially, which is consistent with our remark that the ion and solvent interfacial regions are of similar thickness.
Fig. 11. Potential profile for a 0.01M monovalent electrolyte near an electrode. The solid curve gives the MSA result, calculated using the ion-dipole model whereas the dashed curve gives the linearized GC result. For the region, 0 < z < d 12, the lower and upper curves give the GC result calculated using s*= 80.5 and 2.5, respectively. Adapted with permission from Carnie and Chan [60].
298
D. Henderson, A.D. TrokhymchukandD.T. Wasan
The MSA potential profile is plotted in Fig. 11 and is also seen to be oscillatory. In contrast, the linearized GC/PB profile is monotonic and near the electrode contact, z = d 12 is quite different from the MSA result. We have commented that the GC/PB capacitance can be brought into agreement with experiment by empirical adjustment. Without empirical fitting the MSA iondipole result is in fairly good agreement with experiment. By taking into account [63] the polarization of the metallic electrons in the electrode, the MSA ion-dipole prediction is in very good agreement with experiment not only for a mercury electrode but for a wide variety of metal electrodes. It is not just that the lack of any need for an empirical fit that makes the MSA theory preferable to the GC/PB theory. We see from Fig. 12 that even if s* is adjusted so that the capacitance, a macroscopic quantity, agrees with experiment, the GC/PB values of the microscopic potential near the electrode are quite wrong. As a result, any description of electrochemical reactions that is based on the GC/PB theory, with or without empirical adjustment, will be misleading at best.
Fig. 12. Inverse capacitance of a monovalent electrolyte at the potential of zero charge as a function of the inverse diffuse layer capacitance obtained from the GC/PB theory. The points give the experimental results of Parsons and Zobel [58]. The light line gives the low concentration MSA results and the heavy line gives the full MSA results. Adapted with permission from Schmickler and Henderson [65].
Structure and Layering of Fluids in Thin Films
299
We have already pointed out Eq. (85) is thought to be in good agreement with experiment. This, after all, is the main justification for the division of the electrochemical interface into inner and diffuse layers. Figure 13 shows the full MSA result [65] for the ion-dipole model which is compared with experiment. We see that there is a small departure from linearity in the MSA results that is not predicted by the conventional theories. Note that the experimental results also show this departure from linearity, not just qualitatively but quantitatively. In summary, the conventional description of the electrochemical interface is misleading. There is no artificial "inner" layer. The interfacial region for the solvent molecules is as diffuse as the interfacial regions for the ions. The electrostatic potential close to the electrode is quite different from that given by the conventional picture. Finally, the density and potential profiles are not monotonic but oscillatory or stratified. 4.3. Charged colloidal suspensions Returning to Eq. (75), the TCF between a pair of charged giant spheres, i.e., colloids in this case, both of radius a=D/2, is
h(R-D) = hHS(R-D)
^ f(R-D) s{YD)2D
= h»s{x)--^f(x)
,
(96)
where x = R- D is the gap width between a pair of colloids. It is useful to note that we cannot obtain Eq. (96) from the DH/PB theory even with f(x) ~ e'** because of the asymmetric treatment of the spherical cores in the DH/PB theory, where only the central ion is given a nonzero size. Because the MSA is a linearized theory, we must use the ansatz (40) to obtain the interaction energy W{x) between colloid particles. Thus, the electrostatic and excluded volume contributions to W(x) are
J3W(X)
= -hHS (x) + ^
^ af(x).
(97)
300
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
In addition to these two contributions, in practice there is a short range dispersion interaction between the colloidal spheres, - Aa/\2x, where A is the Hamaker constant that can be measured experimentally. Thus,
™ = V ^ - - f + ^l /w . a
a
\2x
2
(98)
We have seen already [e.g., Eqs. (43) and (44)] that the hard-sphere correlation ETC"
function h between two giant spheres is proportional to the radius a. Hence, Wla is independent of a. Some years ago Derjaguin and Landau [4] and Verwey and Overbeek [5] proposed a theory (DLVO) for colloidal interactions. Their result is
^ a
=_
A+ ^ exp( _ ra) . \2x
m
2
Thus, the DLVO result is Eq. (98) in the limit of zero diameters for the ion species. The DLVO theory is another version of the PB theory. Of course, the derivation [4,5] differs from that given here. They solved the PB equation for an electrolyte between two plane parallel surfaces and then used the Derjaguin construction [15] to obtain the interaction between two giant spheres whereas we have obtained Eq. (98) from the OZ equation for two giant spheres using the HNC/MSA/MSA approximation. In the DLVO/PB theory, the interaction between colloidal particles is the superposition of an electrostatic repulsion that dominates at large separation and a dispersion attraction that dominates at small separations with a maximum at intermediate separations. This gives a pleasing account of colloidal stability. If the maximum is less than the thermal energy, the colloids flocculate whereas if the maximum is greater than the thermal energy, the colloids are stable. The assumption of zero ion diameters and the replacement of f(x) by an exponential that yields the electrostatic part of the DLVO theory are not too bad. However, neglecting the excluded volume contribution to the colloid-colloid interaction, -hHS(x), is a very poor assumption. The number concentration of the ions may be small but the density of the solvent molecule is not small. The presence of h (x) reflects the oscillatory structural force that results from the suspending fluid stratification in the gap between two colloids. In applications to colloids in an aqueous electrolytes an additional term, called as a dipole alignment contribution, -kThDW(x)la, should be added to Eq. (98) because of the reduced polarization of the water molecules near the
301
Structure and Layering of Fluids in Thin Films
colloidal particles. Henderson and Lozada-Cassou [4] have made a crude estimate of this term. Recently, Trokhymchuk et al [15] have made a more sophisticated derivation based on the MSA results for the ion-dipole fluid. Particularly, it has been obtained that similar to the excluded volume hardsphere contribution [see Eq. (45)], the decay of dipole contribution, h (x), has the same functional form
hDlp(x)~21
yDlp | c o s ^ x + arg{/D//>})exp(-/cZ)//>x),
where the coefficients coDIP, KDIP
and yDIP
(100)
are defined in the same way as
their hard-sphere counterparts a>HS, KHS and yHS but depend on both solvent density and dipole moment. The results of such calculation are presented in Fig. 13.
Fig. 13. The force between two giant charged spheres suspended in a 10
M aqueous
electrolyte. Part a shows the electrostatic contribution only: PB results [dotted line], HLC semiempirical approximation [long dashed line], MSA dipole alignment complete result [solid line] and asymptotic Eq. (100) [short dashed line]. Part b shows the total force (solid line) calculated according to Eq. (98) plus dipole alignment term, - kTh
I a.
302
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
Similarly to the excluded volume contribution, the dipole contribution has an oscillatory behavior and increases the total electrostatic repulsion at short separations, improving the quantitative agreement with experiment. The dipole alignment contribution extends until about 3-4 solvent layers near the surface and this region can be associated with an assumed region of lower dielectric constant [the inner layer] that has been discussed above to account for the observed capacitance of electrodes in electrochemical measurements. 5. FILMS FORMED FROM COLLOIDAL SUSPENSIONS Recently, considerable attention has been concentrated on the phenomenon of the layering of the nano-sized or submicroscopic colloids themselves when they are added to macrodispersions such as foams, emulsions, etc. The thin films formed from colloidal particles in such systems can be used as a tool to probe the stratification phenomenon within a colloidal substance [11]. It has been observed that thin colloidal films stratify, become thinner, in a regular step-wise manner. In this section we present some results obtained in the way when the OZ based formalism is combined with computer simulations to make a progress in studying a complex colloidal system. Colloidal films represent a complex system with the superposition of few (not just two as in the case of regular colloidal suspensions) distinct length scales. Such complex systems can be studied by an approach outlined in preceding sections using a many-component version of the OZ equations assuming that the submicroscopic colloids now are a part of a suspending medium. Proceeding in this way, we have found [66] that the hard-sphere colloidal suspension tends to be ordered in a monolayer structure next to a macrosurface. It has been argued that such enhanced layering or stratification is driven by the excluded volume forces that are entropic in origin and can be revealed only if the molecular nature of the primary suspending fluid, i.e., molecular solvent is taken into account. As we can see from Fig. 14, only 1% of the colloid particles dispersed in a hard-sphere solvent comprised of 15% of the fine (solvent-to-colloid size ratio is 1:10) species clearly indicates the formation of a surface-localized monolayer of colloid particles. This is a strong evidence of the prominent role that the fine particle medium plays in colloidal dispersions to enhance a structural forces both within the colloid particles and between the colloid particles and a macrosurface. We turn our attention to this fact, since no oscillations in correlation functions, i.e., no stratification, are expected when a colloid primitive model of a hard-core colloidal suspension at a low bulk volume fraction, such as 1% is used.
Structure and Layering of Fluids in Thin Films
303
Fig. 14. Normalized local density distribution of a suspension of hard-core colloids next to a film surface and its pictorial two-dimensional interpretation.
As the colloid volume fraction in a bulk phase increases, a well-defined second monolayer of the colloid particles next to a film surface also is observed; it is composed by the colloids that are adsorbed on a surface layer of the solvent species. In such an environment, the fine species prefer to be adsorbed on the surface of the colloids and fill the cavities made by the colloidal particles and the film surface, forming an effective film surface coverage. In practice, most colloidal suspensions are composed of charged colloidal particles or macroions. The explicit many-component modeling of such a system becomes progressively more complicated since besides the different length scales, similarly to non-charged colloidal suspensions, the total number of components present in suspension is increased: the molecular solvent is replaced by an electrolyte solution that additionally consists of cations and anions; plus there are counterions to maintain the electroneutrality of the system. Although the explicit many-component OZ approach still can be applied to such a system [15], here we wish to elaborate an approach that is based on the concept of mean or effective interaction [67-70]. This concept exploits the fact that essential forces experienced by colloidal objects in a suspending medium are those mediated by medium. Thus, the colloidal suspension can be viewed as a fluid of macroions interacting via the effective potential obtained by an averaging out the macroscopic degrees of freedom due
304
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
to suspending medium. Usually the DLVO potential (99) is used as an effective potential between colloids. The effects of the solvent and electrolyte ions are incorporated in the DLVO potential via a continuum approximation, i.e., they are present in determining only the screening length. The main goal of the study discussed here is to take into account the excluded volume effects that result from both the finite size of the simple electrolyte ions and from the molecular nature of the solvent. The basic theory of interactions between spherical macroions suspended in an electrolyte solution has been discussed in the preceding section. Due to this, we use the one component fluid model with an effective interaction between two macroions that includes both electrostatic and excluded volume interactions in accordance with Eq. (97). Trokhymchuk et al [71] have made such a study by applying canonical Monte Carlo method that was combined with a simulation cell where the colloid film region and a colloid suspension bulk phase are connected as has been first suggested by Gao et al. [72].
Fig. 15. Monte Carlo data for the normalized local density distribution of the macroions in a film formed at 0.049 volume fraction. The thick vertical lines mark the film left and right boundaries. The film contains four layers and has the thickness that corresponds to seven and half hard-core diameters of macroions. The dashed line shows the Monte Carlo data obtained from the simulation with DLVO forces only. The macroion charge number is 30 in both cases.
Structure and Layering of Fluids in Thin Films
305
Figure 15 displays MC data for the normalized local density distribution of charged macroions (presumably micelles) across the film region of thickness H. To reveal the role that the excluded volume forces due to suspending fluid are playing, the dashed line shows MC data for the macroion density distribution in a film formed under the same bulk conditions but assuming that the colloids interact only via the DLVO potential. Both models show that the colloid particles are forming the film that has a thickness between seven-to-eight hardcore colloid diameters and are organized into four well-defined layers. In both cases there is an evident difference between the layers next to the film surfaces and those in the middle of the film. Each surface layer is rather narrow with the higher density of the particles condensed directly on the film surface that steeply decreases going to the layer boundary. In contrast to the surface layers, the middle-film layers are thick and more diffusive, i.e., less organized. The thickness of the each of two middle-film layer in Fig. 16 is around two colloid particle hard-core diameters while the surface layer thickness only slightly exceeds the one particle diameter
Fig. 16. Snapshots of MC generated configurations of macroions at a bulk volume fraction 0.049 that corresponds to a high surfactant concentration, 0.10 mole/1. Number of charges carried by macroions is 30. The film with three layers has a thickness of five hard-core diameters of the macroions while film with two surface layers has a thickness of three and half hard-core diameters.
306
D. Henderson, A.D. TrokhymchukandD.T. Wasan
The main differences between the model with the suspending fluid contribution and the DLVO-like model are found in the surface layers. In the case of a model with a suspending fluid taken into account, the surface layers themselves show structuring with respect to the film surfaces. This results in each surface layer consisting of a well-defined mono-layer in the immediate vicinity of the film surface and one or two similar sub-layers that are less pronounced and are separated by an "effective" layer of a suspending solution. The shape of the density profiles of the surface sub-layers in this case has a 5 - like form indicating that surface sub-layers are the quasi-two-dimensional monolayers. The surface layers formed within the DLVO model, although thinner than the middle-film layers, are still far from being monolayers. As a result, the segregation of the middle-film particle layers from the surface layers is not observed in this case. As for the middle-film layers for both models only some quantitative differences in the particle local density distribution are observed. When the separation between surfaces decreases, we find that the next two films that provide the local minima of the film energy (per film particle) have a thickness around five and three colloid hard-core diameters, respectively. In particular, one of the films shown in Fig. 16 has thickness H - 5D and contains three layers of colloid particles, i.e. has one layer less than in the film discussed in Fig. 15. Again we observe that the middle-film layer is almost completely separated from the surface layers. The thickness of the middle-film layer decreases slightly when the separation between film surfaces decreases. In contrast, the surface layers do not change notably. This quasi-stability of the surface layers becomes even more evident by analyzing the local density distribution in the film that has thickness around three times the micelle hardcore diameter, H = 3.5D, and contains two surface layers only (Fig. 16). We conclude that the surface layers for three considered films remain largely unaffected and are almost identical. It follows that during the film thinning process the film changes its thickness by squeezing out one middle-film layer of particles (the so-called "squeezing layer" mechanism ). Then the height of the step-wise layer-by-layer thinning will depend on the effective thickness of the squeezed layer. To verify this assumption, the effective thickness of the squeezed layers have been calculated as the difference between the metastable thickness of the films containing four and three, three and two particle layers for both the low and high surfactant concentrations and compared with observation for the heights of the thinning steps. The film containing two surface layers is mostly stable and its thickness corresponds to the final film thickness in the film thinning process at the surfactant concentrations considered in the present study. These also agrees qualitatively well with the force measurements of Richetti and Kekicheff [8].
Structure and Layering of Fluids in Thin Films
307
6. SUMMARY A statistical mechanical approach to study the layering or stratification of the species of suspending fluid that occurs in a vicinity of suspended macrosurfaces has been presented and discussed. This approach is based on the solution of the system of Ornstein-Zernike (OZ) equations that relate the direct and total correlation functions of the giant solutes and the species comprising the suspending medium. In a self-consistent way this approach provides with the bulk phase properties of a discrete suspending medium, its inhomogeneous properties, i.e., local density distributions in an external field of the suspended species, and with the interaction energy between a pair of giant species that is induced by a discrete suspending medium. There are two main driving forces for suspending fluid stratification excluded volume, or entropy, and electrostatics. Excluded volume forces rely on the volume fraction occupied by the species comprising suspending medium, i.e., on their sizes while electrostatic forces contribute by the number of charges carried by suspending species. Within the mean spherical approximation (MSA) excluded volume and entropic contributions are additive. Using the MSA solution we have shown the way in which phenomena of the suspending fluid stratification are related and affect the properties of electrochemical interfaces such as double layer and electrostatic potential, as well as the stability of colloidal suspensions. In a transparent manner we have obtained that the conventional theories of suspending electrolytes such as due to Debye and Hiickel, Gouy and Chapman, and Poisson-Boltzman and DLVO theories all are zero-diameter limits of the suspending fluid species within the MSA theory. Although the assumption of zero ion diameters seems to be reasonable for electrolyte ions, the density of the solvent molecule is not small. Consequently, all these traditional theories neglect the excluded volume contribution to the stratification that results in a series of drawbacks related with their application. In particular, the conventional description of the electrochemical interface is misleading. There is no artificial "inner" layer. The interfacial region for the solvent molecules is as diffuse as the interfacial regions for the ions; the electrostatic potential close to the electrode is quite different from that given by the conventional picture. The excluded volume contribution to the effective interaction between two giant spheres in the case of a simple suspending fluid is characterized by (i) a monotonic depletion attraction for separations between confining surfaces that roughly are smaller than half of the fluid particle diameter, (ii) a repulsive maximum, located at about three quarters of the fluid particle diameter, (iii) a secondary minimum just after a separation of one diameter of the suspending fluid species, and (iv) has a shape of an oscillatory structural repulsion for separations larger than one suspending fluid particle diameter. In the limit of
308
D. Henderson, A.D. TrokhymchukandD.T. Wasan
low density of the suspending fluid this results are reduced to the AsakuraOosawa depletion interaction. We have shown that using a one-component, two-component and manycomponent modeling of a suspending medium there is a possibility to mimic the whole spectrum of colloidal dispersions from the submicroscopic to macroscopic scales. The real objects to which such approach can be applied include complex colloidal dispersions composed of solid surfaces, emulsion droplets, etc., all dissolved in aqueous or non-aqueous suspensions of colloidal particles, surfactant micelles, i.e. fluid systems where the constituents with a few competing length scales are involved. Numerical calculations have been performed for one-, two-, four- and ten-component model systems. A onecomponent model of suspending medium is essential and represents a crucial necessary step that allows one to move beyond the primitive modeling of colloidal suspensions, fn particular, it provides with an effective interaction energy between two mesoscopic surfaces in a molecular solvent. Proceeding in this way we have shown the importance of taking into account the molecular nature of suspending fluid that reveals the primary molecular solvent stratification in the vicinity of a single surface and in the film confinement formed by a pair of macrosurfaces. This theoretical picture is supported by the surface force measurements conducted by Israelachvili and his colleagues [7]. An extension of the simple suspending fluid to a bidisperse solution comprising species with highly asymmetric sizes, i.e., both nanosized colloids and species of molecular solvent, has revealed some new qualitative features for which the stratification phenomenon is responsible. The most notable is that the effective interaction between a pair of giant spheres in a two-component suspending fluid of colloid and fine species is no longer monotonic in the gap region depleted by colloid particles but shows an oscillatory repulsive behavior, reflecting the filling of this region by the fine species. This modeling prediction agree well with observed layering of water molecules known as a repulsive hydration forces [7]. Taking into account the presence of the molecular solvent component also affects the layering of colloid particles near the surface of giant sphere. It is interesting to note, that due to the large asymmetry in the particle size in a bidisperse suspending fluid, the contribution of each of the component to the interaction energy is split on the length scale. Due to this, it seems that to some extent the effect of each component with a competing length scales associated with particle diameters can be treated as a superposition of the results for two monodisperse fluids with corresponding particle size and, probably, with an appropriate effective density. Another interesting finding is that the contribution to the effective interaction between a pair of colloids due to the fine particles of bidisperse fluid medium seems to acquire characteristics that are similar to an adhesive sticky potential when compared to the
Structure and Layering of Fluids in Thin Films
309
contribution of the larger particles. This suggests that it may be possible to mimic the presence of the fine species in a simple manner by means of a sticky potential. Colloidal suspensions invariably contain particles with size polydispersity ranging from several percent in the very carefully controlled polymerization process to hundreds of percent in a typical emulsification process. An increase in the number of species that represent the colloid particles with different but similar diameters (four- and ten-component fluid models), i.e., a primitive attempt to explore the role of colloid particle size polydispersity, shows a tendency to diminish the effects observed when the colloid particle component is monodisperse; however, this does not affect the features introduced by the taking into account a monodisperse fine ("solvent") component. Additionally, we have presented the way in which continuous size polydispersity can be treated. Finally, we have combined the OZ/MSA approach with computer simulations techniques to study the complex colloidal systems involving more than two distinct length scales that is a case of emulsions and foam systems. In particular, we have employed an effective interaction energy between two giant spheres in a bidisperse suspending fluid, i.e., in a colloidal suspension in a MC study of the ionic micelle stratification. We have shown that the forces operating at the submicroscopic scale between colloidal particles are governing the stability of both micro- and macrodispersions. ACKNOWLEDGEMENTS This work was supported in part by the National Science Foundation under Grant No CTS 01-00854. We are grateful to Alex Nikolov and Yurko Duda in collaboration with whom some of the results reported here have been obtained.
REFERENCES [I] [2] [3] [4] [5]
J. Texter and M. Tirrell, AIChE J., 47 (2001) 1706. H. A. Stone and S. Kim, AIChE J., 47 (2001) 1250. D. T. Wasan and A. Nikolov, Nature, 423 (2003) 156. B. V. Derjaguin and L. Landau, Acta Physicochim., 14 (1941) 633. E. J. W. Verwey and J. Th. G. Overbeek, The Theory of Stability of Lyophobic Colloids, Elsevier, Amsterdam, 1948. [6] R. G. Horn and J. N. Israelachvili, J. Chem. Phys., 75 (1981) 1400. [7] J. N. Israelachvili, Intermolecular and Surfaces Forces, 2nd ed., Academic Press, London, 1992. [8] P. Richetti and P. Kekicheff, Phys. Rev. Letter, 68 (1992) 1951. [9] J. L. Parker, P. Richetti, P. Kekicheff, and S. Sarman, Phys. Rev. Letter, 68 (1992) 1955. [10] P. Kekicheff, and P. Richetti, Prog. In Colloid and Polym. Sci., 88 (1992) 8. II1] A. D. Nikolov and D. T. Wasan, J. Colloid Interface Sci., 133 (1989) 1.
310
D. Henderson, A.D. Trokhymchuk and D. T. Wasan
[12] T. Biben, J.-P. Hansen, H. Loven, in Structure and Dynamics of Strongly Interacting Colloids and Supramolecular Aggregates in Solution, (S.-H. Chen, J. S. Huang, P. Tartaglia, Eds.) Kluwer Academic Publishers, Dordrecht, Netherlands, 1992, 23. [13] D. Henderson and M. Lozada-Cassou, J. Colloid Interface Sci., 114 (1986) 180. [14] D. Henderson, J. Colloid Interface Sci., 121 (1988) 486. [15] A. Trokhymchuk, D. Henderson, and D. T. Wasan, J. Colloid Interface Sci., 210 (1999) 320. [16] L. S. Ornstein and F. Zernike, Proc. Acad. Sci. Amsterdam, 17 (1914) 793. [17] J. K. Percus and G. L. Yevick, Phys. Rev., 110 (1958) 1. [18] J. A. Barker and D. Henderson, Rev. Mod. Phys., 48 (1976) 587. [19] J. L. Lebowitz and J. K. Percus, Phys. Rev., 144 (1966) 251. [20] D. Henderson, F. F. Abraham, and J. A, Barker, Mol. Phys., 31 (1976) 1291. Reprinted as Mol. Phys., 100 (2002) 129. [21] P. Attard, D. Bernard, C. Ursenbach, and G. N. Patey, Phys. Rev. A., 44 (1991) 8224. [22] D. Henderson, 11th Symposium on Thermophysical Properties, Boulder CO, June 23-27, 1991; Fluid Phase EquiL, 76 (1992) 1; J. Chem. Phys., 97 (1992) 1266. [23] B. V. Derjaguin, Kolloid Z., 69 (1934) 155. [24] M. S. Wertheim, Phys. Rev. Letters, 10 (1963) 321. [25] E. Thiele, J. Chem. Phys., 39 (1963) 474. [26] M. S. Wertheim, J. Math. Phys., 5 (1964) 643. [27] R. J. Baxter, Aust. J. Phys., 21 (1968) 563. [28] J. Perram, Mol. Phys., 30 (1975) 1505. [29] W. R. Smith and D. Henderson, Mol. Phys., 19 (1970) 411. [30] D. Henderson and W. R. Smith, J. Stat. Phys., 19 (1978) 191. [31] J. L. Lebowitz, Phys. Rev., 133 (1964) A895. [32] M. Toney, J. N. Howard, J. Richer, G. L. Borges, J. G. Gordon, O. Melroy, D. G. Wiesler, D. Yee, and L. B. Sorensen, Surface Sci., 335 (1995) 326. [33] R. Roth, R. Evans and S. Dietrich, Phys. Rev. E, 62 (2000) 5362. [34] A. Trokhymchuk, D. Henderson, A. Nikolov and D. T. Wasan, Langmuir, 17 (2001) 4940. [35] S. Asakura and F. Oosawa, J. Chem. Phys., 22 (1954) 1255. [36] D. Henderson, D. T. Wasan, and A. Trokhymchuk, J. Chem. Phys., 119 (2003) 11989. [37] D. Henderson, D. T.Wasan and A. Trokhymchuk, Condens. Matter Phys., 4 (2001) 779. [38] M. S. Wertheim, L. Blum and D. Bratko, in Micellar Solutions and Microemulsions (S.H. Chen, R. Rajagopalan, Eds.) Springer: New York, 1990; Chapter 6, p.99. [39] A. Trokhymchuk, D. Henderson, A. Nikolov and D. T. Wasan, J. Colloid Interface Sci., 243(2001) 116. [40] J. J. Salacuse and G. Stell, J. Chem. Phys., 77 (1982) 3714. [41] L. Blum and G. Stell, J. Chem. Phys., 71 (1979) 1300. [42] J. Y. Walz and A. Sharma, J. Colloid Interface Sci., 168 (1994) 4851. [43] M. S. Wertheim, J. Stat. Phys., 42 (1986) 459; 42 (986) 477. [44] M. S. Wertheim, J. Chem. Phys., 85 (1986) 2929; 87 (1987) 7323. [45] J. Chang and S. Sandier, J. Chem. Phys., 102 (1995) 437; 103 (1995) 3196. [46] E. Vakarin, Yu. Duda and M. F. Holovko, Mol. Phys., 90 (1997) 611. [47] M. F. Holovko and E. V. Vakarin, Mol. Phys., 84 (1995) 1057. [48] M. F. Holovko and E. V. Vakarin, Mol. Phys., 87 (1996) 1375. [49] E. Vakarin, M. F. Holovko and Yu. Duda, Mol. Phys., 91 (1997) 203. [50] D. Duda, D. Henderson, A. Trokhymchuk and D. T. Wasan, J. Phys. Chem. B, 103 (1999)7495.
Structure and Layering of Fluids in Thin Films
311
[51] H. K. Christenson, D. W. R. Gruen, R. G. Horn and J. N. Israelachvili, J. Chem. Phys., 87 (1987) 1834. [52] M. L. Gee and J. N. Israelachvili, J. Chem. Soc, Faraday Trans., 86 (1990) 4049. [53] E. Waisman and J. L. Lebowitz, J. Chem. Phys., 52 (1970) 430; 56 (1972) 3086, 3093. [54] L. Blum in Theoretical Chemistry: Advances and Perspectives, 5 (H. Eyring and D. Henderson, Eds) Academic Press, New York, 1980, 1. [55] P. Debye and E. Httckel, PhysikZ., 24 (1923) 185. [56] M. Wertheim, J. Chem. Phys., 55 (1971) 4281. [57] G. Gouy, J. de Physique, 9 (1910) 457; D. L. Chapman, Phil. Mag., 25 (1913) 475. [58] R. Parsons and F. G. R. Zobel, J. Electroanal. Chem., 9 (1965) 333. [59] D. Henderson, L. Blum, and J. L. Lebowitz, J. Electroanal. Chem., 102 (1979) 315. [60] S. L. Carnie and D. Y. C. Chan, J. Chem. Phys., 73 (1980) 2949. [61] D. Y. C. Chan, D. J. Mitchell, B. W. Ninham and B. A. Pailthrope, J. Chem. Phys., 69 (1978)691. [62] L. Blum and D. Henderson, J. Chem. Phys., 74 (1981) 1902. [63] W. Schmickler and D. Henderson, J. Chem. Phys., 80 (1984) 3381. [64] L. Blum, D. Henderson, and R. Parsons, J. Electroanal. Chem., 101 (1984) 389. [65] W. Schmickler and D. Henderson, Progr. Surface Sci, 22 (1986) 323. [66] A. Trokhymchuk, D. Henderson, A. Nikolov, and D. T. Wasan, Phys. Rev. E, 64 (2001) 012401. [67] W. G. McMillan and J. E. Mayer, J. Chem. Phys., 13 (1945) 276. [68] L. Beloni, J. Phys.: Condens. Matter, 12 (2000) R549. [69] A. A. Louis, Philos. Trans. R. Soc. London, Ser. A359 (2001) 939. [70] C. N. Likos, Phys. Rep., 348 (2001) 267. [71] A. Trokhymchuk, D. Henderson, A. Nikolov, and D. T. Wasan, J. Phys. Chem. B, 107 (2003) 3927. [72] J. Gao, W. D. Luedtke, and U. Landman, J. Phys. Chem. B, 101 (1997) 4013.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 8
Theory of emulsion flocculation D. N. Petsev Department of Chemical and Nuclear Engineering, University of New Mexico, Albuquerque, New Mexico 87131 1. INTRODUCTION Emulsions are colloidal dispersions of liquid in liquid, for example, a mixture of oil and water. As a result of the mixing, one may obtain oil droplets in water (O/W) or water droplets in oil (W/O). The sizes of the droplets could be in the micrometer and even submicrometer range. A problem of both fundamental and practical importance is that of the emulsion stability against flocculation. Colloid stability is often analyzed in the framework of the Derjaguin-LandauVerwey-Overbeek (DLVO) theory [1-4]. DLVO theory suggests that the stability against aggregation in a colloidal dispersion (e.g., emulsion) depends on the balance between van der Waals attraction and electrostatic repulsion. However, these two forces do not represent a full account of the whole variety of interactions that may occur in colloidal systems. These include steric repulsion [5,6], depletion attraction [7-12], hydration and hydrophobic interactions, oscillatory surface forces, etc. [13]. In the case of emulsions, the colloidal particles are fluid. The droplet fluidity and interfacial mobility may have a strong impact on the emulsion behavior and particularly on their stability against flocculation [14-26]. Emulsion stability also always requires the presence of surface-active molecules or fine solid particles that adsorb at the droplet interface and prevent the droplets from rapid coalescence [4]. Even in the presence of a stabilizing additive, emulsions are thermodynamically unstable. The contribution of the interfacial free energy is proportional to the total area of contact between the two phases and is usually positive. Destroying the droplets and separating the phases macroscopically allows for considerable reduction of this unfavorable term, and therefore of the overall free energy. The time scales on which such event occurs may vary from seconds to years. Hence, emulsions are usually kinetically stable. The analysis of emulsion stability provides not only fundamental challenges of great scientific interest, but is also very important for various
314
D.N. Petsev
practical aspects. In some cases (as in food industry, paint production), the main concern is to obtain stable emulsions while in others (e.g. oil recovery) destabilization and oil separation is desired. Knowledge of the basic principles and mechanisms governing emulsion stability presents both academic and applied interest. Stability loss in emulsions may occur in four different ways [27,28]: • Creaming (or sedimentation) is due to density difference between the two immiscible liquids, the lighter phase will tend to go up while the heavier will move downwards. In the case of an O/W emulsion the oil droplets will accumulate at the top, forming a cream layer. • Flocculation occurs when the droplets stick to each other and form aggregates due to the presence of a minimum in the interaction energy. The aggregates could be compact or may form an expanded gel-like structure. However, the individual droplets remain separated by a layer of the continuous phase - thin liquid film (see Fig. 1). Flocculation may be weak and reversible or strong and irreversible. Flocculation enhances creaming since it forms large floes. It is also often a prerequisite step to coalescence.
Fig. 1. Sketch of two deformed droplets in the presence of long range repulsion (e.g., electrostatic). If no long range forces are present, the droplets deform upon contact of their surfaces. Otherwise a liquid film with thickness h is formed.
Theory of Emulsion Flocculation
315
• Coalescence is the process of the fusion of two or more droplets to form a larger droplet. In this case the droplets are no longer separated by the continuous phase and have become a single entity. The thin liquid film that may have separated them before in the flocculation stage has ruptured under the action of the attractive forces, or due to hydrodynamic instabilities [20,29-31]. • Ostwald ripening is a process of molecular diffusion transfer of oil from the smaller droplet to the larger droplet and is driven by chemical potential differences. The present chapter discusses only emulsion flocculation. Coalescence is outlined in Chapter 9 of this book. As mentioned above flocculation is a process during which the droplets aggregate, but remain separated by thin films of the continuous phase as shown schematically in Fig. 1. The formation of such configuration is due to an attractive interaction but the stability of the film between the interfaces proves the presence of repulsive force at shorter distances. Attractive and repulsive energy contributions have different dependence on the separation of the droplets and their superposition may exhibit a complex distance behavior.
Fig. 2. A sketch of DLVO interaction energy as a function of the separation (liquid film thickness) between two infinite flat surfaces. This picture is more relevant to solid surfaces. The droplets shown are to indicate qualitatively the effect of the attractive forces on the film formation (more detailed discussion is given in the text).
316
D.N. Petsev
In DLVO interaction, it is described by the combination of van der Waals attraction and electrostatic repulsion [1-4]. Van der Waals energy diverges as \lh at small separations, h, and decays as \/h6 at large distances. The electrostatic repulsion is finite at small distances and decays exponentially. Because of this, the overall interaction energy is dominated at small and large separations by attraction, while the intermediate region could be repulsive because of the electrostatic contribution, see Fig. 2. Fig. 2 suggests the presence of a strong and short-ranged repulsion also at very small separations. This repulsion is not included into the original version of DLVO theory [1-4]. Its possible origin is discussed briefly further in this chapter. The particular shape of the energy curve in Fig. 2, however implies that flocculation may take place in the far (secondary) minimum II or in the near (primary) minimum I. The attraction in the secondary minimum is much weaker than that in the primary and so is the flocculation. Droplets, flocculated in the secondary minimum would separate more easily than those in the primary. Flocculation in the primary minimum could have been preceded by such in the secondary minimum. Emulsions with an energy barrier for coalescence could still be amenable to flocculation. The kinetics of flocculation, as well as that for transition between secondary to primary minimum and film rupture depends on both direct (e.g., vane der Waals, electrostatic, etc.) and hydrodynamic interactions [15,20,26,29,31-35]. In the present chapter we discuss the droplet interaction energy and its relation to the kinetics of flocculation. A brief overview of the typical interactions that might be encountered in emulsion systems is given in Section 2. Section 3 outlines the mechanism and kinetics of droplet flocculation and film thickness transitions. Section 4 presents briefly some of the experimental methods for studying droplet interactions and flocculation and Section 5 contains the conclusions.
2. INTERACTION ENERGY BETWEEN TWO EMULSION DROPLETS 2.1. Energy density per unit area Let us consider the interaction between two infinite plane-parallel surfaces (representing the interacting droplets) separated by a layer of the different medium (representing the disperse phase, e.g., water for aqueous suspensions). This is a one-dimensional problem and its theoretical treatment is relatively simple. It also provides a foundation for further more elaborate treatment relevant to more realistic particle shapes (see below). The possible interactions between emulsion droplets may have a different physical origin [1-5,11-13]. Some of them are briefly outlined below.
Theory of Emulsion Flocculation
317
2.1.1. Electrical double layer (electrostatic) interactions Electrostatic interactions are due to the charge that many colloidal particles (or emulsion droplets) acquire when immersed in water. The reason for this charge is dissociation of surface ionic groups or adsorption of ions from the solution. In the particular case of emulsions, the surfaces charges are usually due to the use of ionic surfactants as a stabilizing additive. Electrostatic interactions are repulsive in general and play an important role in explaining colloid stability in DLVO theory [1-4]. In some special cases like interaction between surfaces with different charge (or potential) attraction at very short distances could be observed [1,2], see also [14]. Even the simple case of two plane parallel surfaces has no exact and general solution for the electrostatic interactions. There are, however, two reasonable approximations, briefly outlined below. Low surface potentials (weakly charged droplets). This approximation refers to the case (e$!olkT}<\ where e is the charge of the electron, AT is the thermal energy and ^Fo is the dimensional surface potential. For room temperature (T = 298 K) weakly charged droplets are those with ^Fo ^ 25 mV. The electrostatic energy density per unit area as a function of the separation, h, in this case is [1-4,12]
/J(/i) = 4 7 r e o £ « ^ l - t a n h ^ j .
(1)
The superscript *F means that the surface potential remains constant as the separation between the infinite plates varies. This assumption leads to significant mathematical simplifications but does not necessarily correspond to the real situation. It is a result of the boundary condition imposed at the droplet surface. An alternative is to assume that the potential may vary but the surface charge, a remains constant as the separation between the droplet surfaces, h, changes. Then [1-4,12]
/ ; lh) =
ZL_ i _ cothf—1 .
(2)
The parameter, K, takes into account the charge screening due to the presence of electrolyte (inverse Debye length), and for symmetric (z:z) electrolyte is
318
D.N. Petsev
2 2
1/2
(3)
«= ^ - 7 f Q
where Ce/ is the electrolyte number concentration and z is the charge number of the salt ions, s is the relative dielectric permittivity of the solvent and SQ is the dielectric constant of vacuum. Weakly overlapping double layers. In this case no constraints on the magnitude of the surface potential (or charge) are enforced. There is still an assumption, however, that the two interacting double layers are weakly overlapping, or nh > 1. This is a widely used approximation, known as the nonlinear superposition [12] and reads (see also[l-3,36]) fel{h)-
64CelkTK-1 tanh2 ^
exp(-^).
(4)
There is a number of other and more elaborate expressions for the electrostatic interactions available [1-3,36], and there are always the alternatives of numerical solutions, but the expressions shown above cover a very wide range of reasonable experimental conditions relevant to emulsions and offer rather accurate results. The surface charge (or potential), of emulsion droplets, could be conveniently controlled by using a mixture of ionic and nonionic surfactants [37]. 2.1.2. Van der Waals (dispersion) interactions Van der Waals interactions are due to the forces acting between the individual molecules in the macroscopic colloidal particles (e.g., droplets). These forces are between dipoles and induced dipoles and are typically short ranged. Integration of all these molecular interactions over the volumes of the interacting macroscopic bodies gives an overall energy that is considerably longer ranged. This was first done by de Boer [38] and Hamaker [39], assuming pairwise additivity of the interactions between the individual molecules in the colloidal particles. Such integration, sometimes referred to as Hamaker approach, could often be performed analytically for a number of particle shapes and geometries. Van der Waals interactions are always present between particles immersed in a liquid of a different refractive index and dielectric constant. Their magnitude may be negligible, especially when compared to other interactions that are present. However, van der Waals interactions diverge at contact as \lh and for sufficiently small particle separations they may easily become the dominant force.
Theory of Emulsion Flocculation
319
The van der Waals energy of interaction per unit area for two infinite parallel slabs is [1-3,12,13,36]
AH is the Hamaker constant and depends on the material properties of the droplets and the surrounding fluid. It was shown later [40,41] that due to phase shift between the interacting dipoles, AH might actually decay with the separation and hence, is not truly a constant. This effect is called electromagnetic retardation and may become substantial for h> C/LJ, where c is the speed of light and co is the frequency of dipole oscillation. Further developments of the theory of dispersion interactions [42-44] (see also [12]) lead to elaborate approaches for calculation of the Hamaker constant. These approaches usually require a substantial numerical effort but useful approximate analytical results are available [12] for the non-retarded
and fully retarded
AH =±kT 4
£L^2+iV e, - e 2
47T«2 nf + n{
«±ZA:'l
(7)
h
cases, where S\ and si are the dielectric permittivities, while n\ and n2 are the refractive indices for the droplets and the surrounding medium respectively. hp is Planck's constant, a> is the fluctuation frequency of the interacting molecular dipoles and c is the speed of light. Obviously, retardation affects only the second (frequency dependent) term of Hamaker constant [see Eqs. (6) and (7)]. The important difference between the two expressions above is that the nonretarded AH is a true constant with respect to the separation, h. The second term in the right hand side of the expression (7) for the fully retarded AH, however, decays as \lh [compare to (6)]. Hence, retardation leads to a faster decay of the van der Waals energy. A very useful expression that interpolates between (6) and (7) has been suggested {see [12]}
320
D.N. Petsev
where H = «2 njf + rq\ v '
— c
and
.
(8)
i~2/ 3
3 2
F(H) = ±H /•"C + ^)°iK-^) < f c M 1+ M '
Equation (8) covers the whole range from nonretarded to fully retarded van der Waals interactions. The presence of electrolytes in the solution containing the interacting colloids decreases the value of the Hamaker constant [45]. The first term in (6), (7) and (8) is the only one affected by electrolyte screening and decays exponentially with the separation h. While the electrolyte effect is substantial for high concentrations (e.g., about 1 M NaCl), it has virtually no effect when the electrolyte amount is lower [12]. Typical numerical values of Hamaker constants for oil/water/oil or water/oil/water systems (like most emulsions) are between 3 and 4xlO"21 J (see e.g., [13]). Dispersion interactions may also induce repulsion along with the attraction. It is due to the interactions between atoms and molecules at very small separations. Analysis of this repulsive contribution for colloidal particles has been done recently [46] and showed that the respective free energy of repulsion has the form
fZP{h) = Const
x^.
h
(9)
2.1.3. Steric interactions Steric interactions occur when there is an overlap of the adsorbed polymer layers that may cover the colloidal particles. Emulsion droplets covered by a nonionic surfactant are stable because of the steric repulsion between the hydrophilic headgroups of the amphiphile. Since the adsorbed polymers induce steric interactions, their interactions with the solvent are extremely important [5,12]. For good [47,48] and ©-solvents [49], the interaction is usually repulsive, see also [50], but for poor solvents it could be attractive [12,51]. The first two (repulsive) cases could be modeled theoretically relatively simply, the last one allows only for numerical treatment. Nonionic surfactants usually have
321
Theory of Emulsion Flocculation
polyoxyethylene headgroups, which are known to become more hydrophobic with increasing temperature. Hence, emulsion stabilized by such a surfactant may switch from a stable to an unstable (flocculated) state by varying the temperature. The free energy of interaction per unit area due to polymer steric repulsion in good solvent is approximately given by [47,48]
fsl(h) = 2kTI3/2Lg
h-5l4+HlIA-^
for h<2Lg
(10)
where Lg is the polymer layer thickness in a good solvent defined by Lg= N\Tls\
, hg=hl2Lg
and Y denote the polymer adsorption. For ©-
solvents the interaction is given by [49]
fAh) = FkT^-[^-ln
*l£\
for h<^3L0 J
fsl(h) = 4rkTexp
.
(11)
for h>^f3LQ 2 Lo
L0=ly/N is the mean-square end-to-end distance of the polymer molecule dissolved in the solvent layer separating the surfaces; / and N are the length of a fragment unit and their number in the polymer chain respectively. There are much more elaborate theories for the interaction between layers of adsorbed polymers (or nonionic surfactants) available based on single molecule mean field [52] or self-consistent field [53] approaches. None of them are analytical and they require numerical analysis 2.1.4. Depletion interactions Emulsion droplets often exist in the presence of much smaller species like polymer molecules, surfactant micelles, microemulsions droplets, etc. These species may not adsorb at the droplet interface but still affect the interactions and hence the stability of the emulsion. This is due to the depletion attraction [7,12]. It leads to destabilization and depletion flocculation [8,9,54]. According to the simple model of Asakura and Oosawa [7], the depletion attraction per unit area between two infinite and parallel surfaces is
322
UP(h)
D.N. Petsev
= -^(d-h)
fdep{h) = 0
for
for
h
(i
^
h>d
where d is the diameter of the small colloids and
(13)
where X is the decay length and/ 0 is the surface free energy at contact. Both are determined by fitting experimental data. In the case of hydration repulsion / 0 « 3 - ^ 30 mJ/m and A £»0.6-^ 1.1 nm, while for hydrophobic attraction / o = — 27 with 7 « 1 0 — 50 mJ/m2 being the interfacial tension for a single particle/water interface) and the decay length is A « 1 — 2 nm [11,13].
Theory of Emulsion Flocculation
2.2.
323
Interaction energy between curved fluid interfaces
2.2.1. Interfacial shape of interacting emulsion droplets The shape of a fluid interface is given by the Laplace equation of capillarity [68], see also Ref. [69] and the excellent review by Princen [70]. In the case of two interacting fluid surfaces (as between two emulsion droplets) the Laplace equation should be modified to account for the interaction. For not very small droplets, this could be accomplished by introducing a disjoining pressure [1,2] term in the Laplace equation [71-74], see also [75]. If the system has rotational symmetry, it obtains the form (see Fig. 3 and Ref. [18])
d\j(x)sma(x)} dx
7(x)sina(x) x
,N
where a(x) is the running slope angle of the interface, Pc is the capillary pressure and n(x) = II[//(JC)] is the disjoining pressure between the two droplet surfaces at point x.
Fig. 3. A qualitative illustration of the smooth shape of the transition region between the film and the surrounding curved droplet surfaces. H(x) is actually the shape of the droplet surfaces, a(x) is the running contact angle, ho is the distance in the middle of the film, Rc and a c are the effective film thickness and contact angle respectively, as is the radius of the spherical part of the droplet centered at yo. x\, is a computational parameter (see the text and Ref. [35]) Reproduced by permission of Academic Press, N.D. Denkov, D.N. Petsev, K.D. Danov, J. Colloid and Interface Sci. 176 (1995) 189.
324
D.N. Petsev
H(x) is the local film thickness and the y(x) is the local interfacial tension. To solve Eq. (14), one needs to know these two functions. The equations for them are dH(x) — = 2tancv(x) and
(15)
dx - ^
= -n(x)sina(x).
(16)
The boundary conditions for solving Eqs. (14-16) are (see Fig. 3 and Ref. [18]) l{*B) = l + \f{HB),
(17)
HB=H{xB) = 2 ^ - ^a] - x\) and
(18)
tan«(x g ) =
%B
(19)
^as - xB
The disjoining pressure, U(x) and surface free energy fix) are related by the expression
f{h)= [°°n{H)dh
(20)
*J h
and y is the interfacial tension of undeformed droplet [18]. Since the fluid in the droplet is incompressible, the volume, V, does not change upon deformation, or V = ^a3=\fH\x2{H)dH
+ \{3as~p),
p{as) = as + ^
- x\
(21)
where a is the radius of the undeformed droplet, as is the radius of the spherical part of a deformed droplet in a doublet (see Fig. 3) and x{H) is the function of the droplet shape in the region ^h$
Theory of Emulsion Flocculation
325
F = 2TrJX"ll(x)xdx = 0.
(22)
The dependence of the disjoining pressure Tl(x) on the radial coordinate x is shown in Fig. 4. The repulsion between the droplet surfaces in the thin film region is balanced by the attraction that is due to the surrounding spherical parts. The shape of the droplets in an equilibrium doublet is calculated using the following procedure [18]: equations (14-16) together with boundary conditions (17-19) are solved numerically to obtain the first approximation of the shape in the region 0 < x < xB. The remaining part of the droplet is assumed to be perfectly spherical with radius as. Then condition (22) is checked and if it is not fulfilled, a new starting value for y0 is chosen, which leads to a new value for as calculated by means of Eq. (21). This is repeated until y0 and as satisfy both (21) and (22) simultaneously. Extending the droplet interfacial area gives rise to the following contribution to the pair interaction energy Us = 4TT7 JX" x^l + tan2a(x)dx + asp(as)-
Fig. 4.
2a2 .
Disjoining pressure against the film radial coordinate (see Fig. 3).
parameters are droplet radius as = 2 p.m, Hamaker constant AH = l x l 0 ~
(23)
The 20
J,
interfacial tension y = 1 mN/m, surface potential \&0 = 1 0 0 m V , and monovalent electrolyte concentration Cei = 0.1 M. The vertical dotted line defines the thermodynamic film radius. Reproduced by permission of Academic Press, N.D. Denkov, D.N. Petsev, K.D. Danov, J. Colloid and Interface Sci. 176 (1995) 189.
326
D.N. Petsev
The value of as quantifies the extent of the droplet deformation. Note that this definition of the interfacial tension contribution to the overall droplet interaction does not account for the interaction between the surfaces which is discussed separately below. More details are given in Ref. [18]. 2.2.2. Derjaguin's approximation Calculating the energy of interaction between finite size colloidal particles (e.g., emulsion droplets) is not as simple as for the case of infinite parallel plates discussed above. This difficulty follows from the curvature of the interacting surfaces. This was partially resolved by Derjaguin [76] {see also [1,2,12,13,36]} who suggested that for small separations between the particle surfaces, hi a «; 1, the interaction energy could be given by the general formula
U(h) = 2nf™f{H{x)]xdx.
(24)
h is the shortest distance between the deformed droplet surfaces (see Fig. 3) and it may change as the droplets approach or separate. If the free energy of interaction per unit area is known (see Section 2.1 above) then calculation of its contribution to the pair energy can carried out using Eq. (24). For example, introducing Eq. (5) into (24) we obtain the van der Waals contribution to the droplet pair energy of interaction
Uvw = 27rj
xVl + tan 2 «(x) -
J 0
I
H[
J
\2ITH
>[ dx
(25)
[X)
while Eq. (4) provides the respective electrostatic term Uel = 2TTJX" x^l + tan2 a(x)
64CelkTn~] tanh2 ^ - exp[-K#(x)]Lt. (26)
A considerable simplification could be attained if the deformed droplet shape is represented by that of a truncated sphere, like in Fig. 1. This approximation may give very similar results to those that are obtained by the approach presented above [18]. For DLVO type of interactions (van der Waals, electrostatics and surface extension - see below) and wide range of parameters, the differences were s 7%. In this truncated sphere approximation, we distinguish a well-defined plane-parallel film surrounded by the spherical droplet surfaces (see Fig. 1). Then [14] {see also [15,17,18,21,25]}
Theory of Emulsion Flocculation
H(x) = h
forx
H(x)^h + -(x2-R2)
forx>/?"
327
h is again the closest distance between the surfaces and according to Fig. 1 it is the actual thin film thickness, R is the film radius and x is the running radial coordinate. Combining (24) and (27) one obtains for small deformations,
(rlaf <1 U(h,R) = 7rR2f(h) + nafC°dHf(H)
(28)
Jh
where f(h) is the energy density per unit area as defined in Section 2.1. The first term in (28) accounts for the interaction energy across the plane-parallel film while the second is due to the surrounding curved portions (see Fig. 1). If there is no deformation, r = 0, the first term in (28) disappears and Derjaguin's expression for interacting spheres is obtained [1,2,12,13]. Hence, Eq. (28) allows for an easy and accurate way to calculate the different terms in the total pair energy between deformed droplets which is discussed below in this chapter. An alternative method for calculating the interaction energy of curved interfaces is based on a perturbation approach [75]. Using Eq. (28) and assuming a model truncated sphere shape for the interacting droplets allows for obtaining very simple expressions for the different types of interactions that could be present - electrostatic, van der Waals, steric, depletion, salvation, etc., see Section 2.1. The van der Waals interaction, however, could be calculated explicitly for the model shape which is demonstrated below. 2.2.3. Van der Waals (dispersion) interactions between droplets Van der Waals interactions could be calculated exactly for some simple shapes of the interacting colloidal particles. This approach was developed by Hamaker [39], who calculated the dispersion energy between bodies of different shape including two spheres of different size. This result is particularly relevant to emulsions where the spherical shape of the droplets is governed by the interfacial tension. In some cases (low interfacial tension), the droplets deform and their shapes could be approximated with that of truncated spheres separated by a plane-parallel film. In this case (see Fig. 1), the van der Waals energy can be calculated exactly following the integration procedure of Hamaker and was
328
D.N. Petsev
done independently in Refs. [14] and [77]. The general result for two deformed spheres with equal radii is [14]
>K
12 \{L + h)2
' +
h(2L + h)
(L
+
hf
^ . _ _ ^ 2i?2 2 h h(L + h) (L + h)[2(L-a) + h] (h2+4R2)yh2+4R2-h)
2R2a(2L2+Lh + 2ah)
2h[2(L-a) + h]
h(L + h)2[2(L-a) + h\
+
where L = a + va 2 — R2 . Eq. (29) could be significantly simplified if one assumes that the deformations between the droplets is small, or (Vla) < 1. In this case, the van der Waals energy contribution becomes
rr /, N Uvw(h,r) =
A
4 2 H <3 4a2 , hUa + ti) !L\ J + TT, TT + 2 1 n ~ T 12 (2a + hf h{4a + h) (2a + hf LV
J
.
(30)
5 2
l28a R 2
h (2a + hf (4a + hf \ The first three terms on the right hand side of (30) are equivalent to the Hamaker result for spheres, while the last term is due to the formation of plane-parallel film between the droplets. 2.2.4. Surface extension and bending energies Emulsion droplets normally consist of incompressible fluid. Therefore any deformation is accompanied by extension of the droplet interface from the optimal (spherical) to the new shape (e.g., truncated sphere) [16]. The energy contribution related to this (again for small deformations, (r la) < 1) is [14]
E/,(r) = ™2 ^ 4 + — 4 V
^
2 a4
64
8 a
(31)
Theory of Emulsion Flocculation
329
where y0 is the interfacial tension of the spherical droplet in absence of
deformations
and
EG = (d-y/d\nS)s=s
= (dj/d\nr2)
is the Gibbs'
elasticity of the surfactant monolayer at the droplet interface. The second term on the right hand side of (31) is often negligible and only one proportional to the interfacial tension, y0, remains. For oil/water/surfactant systems with ultra low interfacial tension [78], the Gibbs' elasticity term may become important. Film formation between droplets also leads to energy contributions due to the bending of the interface [21], viz.
Ub = -2,r2B0H 1 - - £ - = *nke t f -^ 1 - f . 2HQ
[a) a0
(32)
2a
Bo is the bending moment, kc is the bending elasticity constant [79], H = — 1 / a and Ho = — 1 / a0 are the interfacial and spontaneous curvatures respectively (a and a0 are the droplet and the interfacial spontaneous curvature radii). It is important to emphasize that the surface extension and bending energy contributions depend only on the extent of deformation expressed by the film radius R and not on the separation (film thickness) h. 2.3. Structure of droplet aggregates 2.3.1. Doublet of two interacting droplets Colloid particles [ 1,2] as well as emulsion droplets may flocculate in the distant (secondary) or in the closer (primary) energy minimum - see Fig. 2. The overall energy of interaction between droplets is determined by the superposition of different energy contributions like those described above and depends on both the interdroplet distance and film radius (if any deformation is present)
U{h,r) = YJUi{Kr)
(33)
i
where the subscript / stands for the respective energy term (electrostatic, van der Waals, steric, depletion, etc.). If the droplets have low interfacial energy, they are deformable and the contributions of the interfacial deformation (surface extension and/or bending) have to be taken into account. Whether the droplets forming a doublet will deform or not is determined by a variety of factors including the interfacial tension, droplet size, charge, electrolyte concentration and magnitude of the attractive energy [14-18,21,23-25]. Hence, it is not
330
D.N. Petsev
possible to define a simple criterion for the occurrence of deformation. Usually, the total interaction energy has to be minimized with respect to the variables h and r (the film thickness and film radius) [17]. In cases when the parameters do not favor film formation the minimization predicts r = 0. As rules of thumb we find that, see also Ref. [80]: > Attractive interactions favor the deformation (film formation). Hence, strong van der Waals forces or depletion interactions, induced by smaller colloids, may lead to droplet deformation. > Repulsive interactions suppress film formation. A typical example is electrostatic repulsion: high surface potentials and/or low electrolyte concentrations do not favor the formation of plane a parallel film between the droplets in a doublet. Addition of electrolyte decreases the repulsion and therefore facilitates deformation. > The lower the interfacial tension the higher the droplet deformability. Interfacial tension resists any extensions of the droplet interface and thus the formation of film in the doublet. Similar is the role of high bending energy which increases the interfacial rigidity. ^ Larger droplets deform more readily than small ones if all the remaining parameters are the same. This could be attributed to the higher capillary pressure of small droplets that make them behave more like rigid spheres.
Fig. 5. Countour diagram of the interaction energy of two charge stabilized droplets. a= 1 urn, AH = l x l O " 2 0 J , y = l mN/m, tf0 = 100 mV, Ce, = 0.1 M.
Theory of Emulsion Flocculation
331
This is illustrated in Fig. 5, which presents a contour diagram of the interaction energy between two droplets as a function of the film thickness, h, and film radius, r. The droplet size chosen for this calculation is 1 \im, the electrolyte concentration corresponds to 0.1 M monovalent salt and the Hamaker constant is AH = l x l ( T 2 0 J, ^ 0 = 5 0 mV and y = 1 mN/m. A minimum in the energy surface is formed at r/a = 0.055 and h/a = 0.0069. The attraction energy is -29.1 kT. If the deformability is ignored, the energy vs. distance dependence runs along the ordinate axis and the energy minimum is -27.3 kT at h/a = 0.0058. From the ratio r/a (at the energy minimum), one may calculate the equilibrium contact angle, a using [18] a = sin—-. a
(34)
Increasing the electrolyte concentration or lowering the droplet surface potential leads to greater deformation. More details on the structure (values of the film radius and thickness) are given elsewhere [14,17,18,21,24,25]. These also include the effect of other types of colloidal forces like steric, depletion and structural. While steric repulsion acts similarly to the electrostatics, depletion may turn into oscillatory structural force at a high volume fraction of the small colloids [21,25]. These forces lead to metastable states of macroscopic emulsion films containing one or more layers of trapped particles [55,56,81,82] and their role for miniemulsion droplets (submicrometer sized and below) is usually destabilizing, leading to a deep depletion minimum without any small particles in it [21,25]. 3. KINETICS OF DROPLET FLOCCULATION The kinetics of flocculation of solid colloidal dispersions was first analyzed by Smoluchowski [83-85]. The theoretical model he developed was based on the notion that the time determining step is the approach of two colloidal particles due to diffusion. It has subsequently undergone numerous refinements and generalizations [2,86-96]. These include more realistic direct and hydrodynamic interactions and account for higher order particle correlations but mostly for solid colloidal particles. Still, the major differences between solid and fluid particles are the fluidity and flexibility of the droplet interfaces. 3.1. Droplet motion and hydrodynamic interactions The motion of a viscous droplet in unbounded viscous fluid without surfactants presents a relatively simple hydrodynamic problem and was solved
332
D.N. Petsev
by Rybczynski [97] and Hadamard [98]. However, in most real cases there are always surface-active substances present and they change the droplet movement considerably. Analyses of single droplet motion in the presence of surfactants were given by Levich [99]. The relative motion of two droplets presents a more difficult problem. The case of absence of surfactant and interfacial tension gradients has been considered by Davis et ah, who suggested numerical solutions for the mutual mobilities of non-deformable droplets [100] as well as accounting for the deformation in lubrication approximation [101,102] or using a boundary integral algorithm [103]. More detailed analysis of the hydrodynamic interactions between droplets is also given in Chapters 10 and 11 of this book. Emulsion droplets in absence of surfactants are usually unstable against coalescence and are unable to form doublets or higher order aggregates of individual droplets separated by thin films of the continuous phase. The presence of surfactant stabilizes the emulsion droplets and strongly affects the hydrodynamic interactions. When two droplets are close to each other, the main contribution to the hydrodynamic resistance (energy dissipation) comes from the thin gap between them, or the thin liquid film. The hydrodynamics of displacing solvent from the film during the approach of two surfactant stabilized emulsion droplets has been studied extensively [31,33,34,104-106]. These studies considered millimeter and sub millimeter sized droplets where the main force, bringing the two droplets together, is usually buoyancy. Knowing the magnitude of the buoyancy force and the hydrodynamic resistance, one may calculate the velocity, V{h,R) and hence the time of film thinning, r, viz. r{hm,hf,R) = ^ ^ - ^
V(h,R) = FBC(h,R)
(35)
where FB is the buoyancy force and <^h,R) is the hydrodynamic resistance that depends on the film thickness, h, and radius, R. For small (micrometer and submicrometer) sizes, the Brownian motion is bringing the two droplets together [15]. The kinetics of flocculation of Brownian droplets can be described following the Smoluchowski method modified to account for the droplet deformation if such occurs. Note that the hydrodynamic force between the droplets may induce deformation and it usually occurs at larger separations [15], than those that correspond to a minimum in the interaction energy. The relative diffusion approach of two droplets at small separations is related to the hydrodynamic drag force. The thin film formed between the droplets is characterized by two variables thickness, h, and radius, R and in Brownian systems they both fluctuate when subjected to the random
Theory of Emulsion Flocculation
333
Fig. 6. Random force acting on a thin liquid film between two droplets and the respective trajectory in the 2-dimensional space of the film thickness h and radius R. This is a qualitative illustration of film thickness transition due to Brownian fluctuating forces Fh and FR.
force, F, with normal, and radial components Fh and FR respectively, see Fig. 6. Hence,
\FR}=
^RR ^Rh\[VR
{Fh}
QhR ChhAVh '
(26)
For linear hydrodynamic flows one may apply the Onsager theorem [107] which implies that (Rh — QhR. The components of the friction tensor, £, are given by [26], see also [15,31] 6TTT]R2
h{\-£s) _ ShR—^Rh—
_W 2h
3nVR3 J2—
^'>
, , R2 , R4 ' ah
l l ah
where rj is the solvent viscosity and ss is a parameter that takes into account the interfacial mobility and droplet bulk viscosity - see [31,104]. For tangentially immobile droplet interfaces, ss = 1. This is the case when the interfaces are
334
D.N. Petsev
Fig. 7. Components of the hydrodynamic resistance tensor as defined by Eq. (37). These are valid for small separations between the droplets.
Theory of Emulsion Flocculation
335
completely saturated with adsorbed surfactants that are soluble in the continuous phase only. If there are no surfactants, or if they are soluble only in the droplet phase then ss decreases and may become about 0.001 [31,104,105]. In absence of deformation (R = 0) and tangentially immobile droplet surfaces, ChR =CRR = 0 a n d (hh =3-7rrya2/2/z, which corresponds to the result for two solid spheres of radius a [12]. Another limiting case is a very thin film with a large radius, Rlh^>\, when the contribution of the curved surfaces around the film could be ignored. In this case the Reynolds formula for the approach of two solid parallel disks [108], (hh = 3nrjR 12}?, is applicable. The component QhR increases as R3 while (RR becomes infinite for the case of tangentially immobile droplet surfaces and finite radius since es — 1, see Eq. (37). Fig. 7 represents the dependence of the different tensor components on the film thickness, h, and radius, R. 3.2. Brownian Flocculation 3.2.1. Effect of the droplet interactions on the flocculation kinetics If the direct and hydrodynamic interactions are known one may calculate the steady flux of droplets toward a given central droplet, J, and hence - the rate constant of pair flocculation
j
n^
WF
where nx is the droplet number concentration at infinite distance from the central droplet, Do is the Stokes-Einstein diffusion coefficient [109] and WF is the Fuchs factor defined by [2,12,86]
WF=4D0af dr Jo
[
\\ D{r)
2 2 r
J
-
(39)
The integral in (39) is taken over the distance between the flocculating droplets. U(r) is the interaction energy and D{r) = kTI(,(r) is the diffusion coefficient at small separations [15,26], also see below. The rate constant, kf, defined by (38), can be introduced in the kinetic equation for the formation rate of doublets of droplets [15]
336
D.N. Petsev
£ =*/-*,
CO)
«2 is the number of doublets formed and nx is the single droplet number at infinity. Detailed analysis of the above equation for the particular cases of droplet flocculation and coalescence is performed in Ref. [110]. The flocculation and coalescence kinetics of Brownian emulsions is not affected by deformability for diluted systems, in the presence of only attractive interactions [15]. In such systems the time determining step for a flocculation or coalescence of two droplets is the approach from large distances. On the other hand, the presence of strong repulsive interactions, e.g. high surface potentials, may prevent the droplet deformation for low electrolyte concentrations ^ 0.1 M of monovalent electrolyte. Therefore, the kinetics of flocculation and coalescence of very diluted or charged emulsions with low concentration of electrolyte present will be identical to solid dispersions with the same particle size, Hamaker constant, surface potential and other interaction determining parameters [15], see also [111,112]. In the case of charge-stabilized emulsions at high ionic strength (above 0.1 M monovalent electrolyte), the situation could be different. For such high salt concentrations droplet flocculation often occurs and is accompanied by droplet deformation [32,113-116]. The droplets may flocculate in the secondary minimum of the DLVO energy [117] and eventually jump into the primary similarly to macroscopic foam and emulsion films [29,118]. The primary doublets may further coalesce. This case was recently analyzed [26] and was shown that the mean time for transition between the secondary into the primary energy minimum, r depends on the droplet deformation and film formation. This mean transition time could be obtained by solving the following equation [26] (see also Fig. 6) —
^
V\Peq (R,h)D(R,h).Vr(R,h)\ = - 1 , where
611
Peg (/?,/,) = e x p - ^ P
X
'
(41)
and D = kT^(R,hf
Since the inverse friction matrix is [26] >•-! _
1 ( (>RR ~ C M | _
1
|
(M
the components of the diffusion tensor, D, become
~ChR\
,,2\
337
Theory of Emulsion Flocculation
kT
Chh
D K R =
CRRCM
kT
=
~ChR
+
,
ChR CRRCHH )
n
~kT<:hR -
- n -
U
U
hR — Rh — ~—";
~kT
T2~ —
r<m 2
"
'
'
SRRShh ~ ChR
CwChh j _ ChR ChR CRRChh kT > i ChR
kTQm = CRRChh ~ChR
CRRCM-I
The dependence of the diffusion tensor components on the film thickness, h, and radius, 7?, is given in Fig. 8. Note that DRR is infinite at R = 0 and becomes zero for es — 1. The transition (or coalescence) time equals twice the mean time necessary for a system (doublet of droplets) to move from the secondary minimum (Rmin,hmm) to the saddle point (RsaddAadd) (see Fig. 5). Hence, the boundary conditions for Eq. (41) are T R
{ min. Kin ) =T>
T R
(44)
{ sadd, Kadd ) = °
The multiplication by two is needed to account for the fact that a system at the saddle point can either cross it or return with equal probability [119,120]. In the case of tangentially immobile droplet surfaces DhR = DRh = DRR = 0 , Eq. (41) becomes [26]
"^^.^.ijiJlML-l. Peq{R,h)dh
H
,45)
dh
Equation (45) can be integrated directly and c Km,
dz
r°°
^ • ^ ^ ) = J L f ( > M ^ M J , dyPeq{R,y).
(46)
If hmin is at the secondary minimum and hsadd corresponds to the top of the barrier that has to be overtaken, Eq. (46) gives the mean transition time between the secondary and the saddle point (Fig. 6) [26]. It is interesting to compare (46)
338
D.N. Petsev
Fig. 8. Components of the diffusion tensor as defined by Eq. (43). These are valid for small separations between the droplets.
Theory of Emulsion Flocculation
339
to (35). The latter applies to film thinning and thickness transitions due to the action of a well-defined deterministic force (e.g., buoyancy), while the former is valid for processes that are driven by random Brownian forces. That is why the mean times given by (46) and (41) are statistical averages and depend on the probability distribution in the configuration space defined by Peq(i?,/?) = exp —U(R,h)I kT , as well as on the diffusion coefficient Dhh(R,h) = kT/(hh(R,h) and/or the diffusion tensor D(R,h) = kT[C)(R,h)\'1. Analysis of the dynamics of secondary-primary film transitions, based on (44) demonstrated that the deformation of the droplets may increase the mean transition time with orders of magnitudes when compared to referent nondeformable particles, for more details and examples see Ref. [26]. Eq. (46) and the underlying model do not take into account wave disturbances of the droplet (and more importantly, film) surfaces. Such disturbances are known to trigger film thickness transitions and breakups in macroscopic films (e.g. large millimeter sized droplets) [29-31]. According to experimental evidence [22,118,121] and some theoretical estimates [122], film surface corrugations and wavelike deformations do not occur for small films, like those that form between Brownian micrometer and submicrometer droplets. Still the analysis of some model systems based on Eq. (44) showed that the effect of the droplet deformability could be very substantial at high electrolyte concentrations [26]. 3.2.2. Flocculation and coalescence kinetics of dilute emulsions Calculating the flocculation and/or coalescence rate constant for diluted emulsions is relatively simple if only the Dhh(R,h) = kT /(,hh(R,h) term is taken into account and the other tensor terms are assumed to be zero, DhR = DRh = DRR = 0 [15]. In this case the problem could be solved by calculating the steady flux of droplets toward a single one as suggested by Smoluchowski for solid colloidal particles [83-85]. The important difference in case of droplets is the possibility for interfacial deformation due to hydrodynamic and direct forces. The droplet deformation in this case may start at distances that are substantially larger that those corresponding to the distant secondary minimum of the direct droplet interactions. The reason is the addition of an effective Brownian force to the action of the surface colloidal forces. The droplet flux is
J =4 ^
D W
* G +£ & * 4 2 L coml [ dr
kT
(47)
dr
where r is the distance from the central droplet, nx is the droplet number concentration at infinity, D(r) is the diffusion coefficient, P(r) is the probability
340
D.N. Petsev
to find a droplet at a distance r from the central droplet, and U(r) is the pair interaction energy. Eq. (47) can be solved for the probability function using the boundary conditions P -> 1, U -> 0 for r -»oo P^O, forr^O
(48)
and the result is [15,123]
pooexp[^(r)/^r] , . P(r) w = exp
U[r)\ —-1 itr
*r
D(r)r2 ^yi— . r ooexp U(r)lkT\ I ^—df Jo D(r)r 2
(49)
To determine the total force (Brownian and direct), it is convenient to rewrite Eq. (47) in the form [15] J<;(r) =kTd\nP(r) 2
47rr « 00
|
dU(r)
^
(5Q)
dr
where C,{r) — kTID[r) is the hydrodynamic resistance to the mutual droplet approach. The mean droplet velocity of motion toward the central droplet is [15] V r
() =
/
, ,-
(51)
4irr2nooP(r) This velocity is related to the mean total force, Fj{r), by the relationship FT(r) = t{r)V{r)
(52)
or [15]
FT(r) = tr*¥& dr
+
«W. dr
(53)
Theory of Emulsion Flocculation
341
The first term corresponds to the mean Brownian force, while the second is due to the direct interactions (e.g., van der Waals, electrostatic, etc.). Hence, the total force is
47rr2WnoP(r) For the special case of droplet approach with simultaneous deformation due to the force bringing them together, one may compute the hydrodynamic resistance coefficient using the fact that the radial, VR, and normal, Vh, film velocities are related by [15,26] Vr> R(\-e.\ -*- = - i *-!-. Vh 2h
(55)
Then Eq. (52) becomes Fr=Chh+ChR^^Vh
(56)
where Qhh + C,hRR{\ — es)/2h is the effective friction coefficient. Knowing the total force, one could easily obtain the deformation separation distance h, (where the surface curvature inverses its sign, and therefore the subscript /) from the relationship [31,34], see also [15] h -
FT
J
^rd)
p - p l
r
\
(
where rd = hf + 2a is the droplet mass center-to-center distance (see Fig. 9). The probability at the point of deformation is
Eq. (57) together with (56) gives [15]
342
D.N. Petsev
IrT
k=-Z^L7
(59)
where pr, rd ((r) Jo
l
r C,(rd)
\U(r)-U(rd)] kT
This integral has to be evaluated over the three regions Ac, ,4/and Ad (see Fig. 9).
Fig. 9. A qualitative illustration of the pair probability distribution function P(r) for deformable interacting droplets. Aa represents the region of droplet approach with no deformation, Aj correspond to the interdroplet distances where the droplets deform, A/ is the region of film thinning and Aa is the final region where coalescence may occur. rd is the distance between the droplet mass centers at which the deformation starts while rj is the value where the film starts to thin and rcr is the critical distance for film rupture. It is often convenient to work with film thicknesses instead of mass center distances. Thus ht is the film thickness where the deformation starts or the surface curvature inverts its sign (hence the subscript /). Sometimes it is assumed (for simplicity) that while the films forms and obtains radius rj the thickness remains constant and equal to h, (region Ad). Reproduced by permission of the American Chemical Society, K.D. Danov , N.D. Denkov, D.N. Petsev, I.B. Ivanov, R. Borwankar, Langmuir 9 (1993) 1731.
Theory of Emulsion Flocculation
343
If the emulsion is diluted, it can be assumed that the contribution from region Ac is much smaller than the contributions from the other two regions [15]. This assumption can be justified by the fact that the film rupture and droplet coalescence is usually accompanied by a sharp decrease in the interfacial energy and the exponential in (60) is small. Also, the rate of coalescence is much greater than the rate of approach and film thinning, which indicates that the ratio £(r)/£(r r f )
*=-«-, 4ir-yaf0
/0 =2/'<4^M*2L 1 Jra r C,[rd)
(61)
kT
Eq. (61) allows for determining ht if C,(f), U(r) and rcr {hcr = rcr — 2a), where hcr is the critical thickness of film rupture [31,124-126] 2
2
^,=0.243/*,.-^-,
1/7
.
(62)
2-7T7 hj
Eq. (61) is the last piece of information needed to obtain the Fuchs factor [see Eq. (39)], [15] and hence, the droplet flocculation kinetic coefficient [Eq. (38)]. Eq. (62) was derived for attractive van der Waals forces without any long range repulsion. Therefore its application to charged droplets is an approximation. Following this approach, one may also calculate the Wuchs factor and the kinetic constant, see equations (38) and (39) respectively. If the film between the droplets remains intact at the final stage, then such a calculation refers to the kinetics of flocculation. If, however, the film breaks (at distance hcr), it refers to the kinetics of coalescence. 4. EXPERIMENTAL STUDIES OF EMULSION FLOCCULATION AND INTERACTION BETWEEN FLUID INTERFACES This section presents a brief outline of some recently developed techniques for studying the interaction and flocculation between emulsion droplets and measuring the forces between fluid interfaces. 4.1. Direct measurement of the forces between emulsion droplets Performing well-defined model experiments with emulsions became possible due to the method for obtaining monodisperse sample developed by
344
D.N. Petsev
Bibette [54]. This method avoids all the ambiguities stemming to their polydispersity and allows for better comparison with theoretical models that are readily available for droplets of equal size. A method for directly studying droplet-droplet interaction was suggested recently, which is based on monitoring the force-distance relationship law between equally-sized magnetic emulsion droplets [127]. A monodisperse sample of octane ferrofluid (octane containing 10% Fe2C>3 grains) was prepared following [54]. Applying a magnetic field to the emulsion gives rise to a force pressing the droplets against each other, while monitoring the Bragg diffraction of visible light from the sample provides the necessary information about the distance between them. This powerful technique has a great potential in studying different types of interactions that might be present in emulsions [127,128]. The main inconvenience of this method is that Fe2O3 grains must always be present in the dispersed phase to provide the response to the externally applied magnetic field. 4.2. Liquid surface force apparatus Liquid surface force apparatus is a device for measuring the forcedistance profile between a droplet (e.g. oil) formed at the tip of a flexible micropipette and a macroscopic oil-water interface [129-132]. The force exerted on the droplet while pressed towards the interface is determined by the deflection of the pipette shaft. The disjoining pressure in the film between the droplet and the interface could be obtained by knowing the pressure inside the droplet. The latter is controlled throughout the experiment. This method allows for direct investigation of the important case of a droplet interacting with the macroscopic phase of the same material and extensive experimental studies have proven that it is very informative and reliable [129-131]. Its relevance to the interaction between two small droplets in an emulsion is a little less straightforward. 4.3. Experimental cells for studying emulsion films The formation and evolution of thin liquid films could be conveniently traced in specially designed experimental cells [29,133]. The cell is usually a capillary (with a diameter of the order of a few millimeters) in which the oil phase is separated in two parts by a thin layer of water to simulate an oil in water emulsion. The capillary is connected to a pressure controller and the film thickness could be changed as well as monitored by interferometry. The experimental setup allows for obtaining the forces acting across the thin water film between the oil phases. The main disadvantages of such a cell are [118]: (i) the typical film diameters formed are above 100 urn (much larger) and (ii) the capillary pressures are much lower than those in real emulsion systems. A step forward in the design of cells for thin film studies was the construction of a
Theory of Emulsion Flocculation
345
miniaturized version [118]. Using laser micro machinery a cell diameter of about 280 um was produced that allowed for the formation of films with radii in the range 20-220 urn. Although a significant improvement, these radii are still too large when compared to miniemulsions, where the film's radii could be in the nanometer range. 4.4. Video-enhanced microscopy This method was developed recently [134-136] and is based on direct observation of the kinetic processes in emulsions. The kinetics of flocculation (including reversible) [134,135] and flocculation combined with coalescence [136] were studied. The method is well designed for kinetic investigations of emulsions but it is also applicable to solid dispersions. 5. CONCLUDING REMARKS This chapter presents an overview of the droplet interactions in emulsions and their effect on the stability against flocculation. Surfactant stabilized emulsions are often treated theoretically in the same way as suspensions of solid particles. This is justified for small droplets (below one micrometer) with high interfacial tension because they are virtually undeformable. Large droplets (between tens of microns to millimeters in diameter) represent the opposite limiting case where all the important interactions are across the plane parallel liquid film that forms between them at small separations. There are many other situations, however, where the droplets are deformed and the overall interaction is due not only to the plane parallel liquid film, but also to the contribution from the curved surfaces surrounding the film. Examples for such emulsion systems are small droplets (a few hundreds of nanometers in diameter) with low interfacial tension. A special case are microemulsions where the interfacial tension could be extremely low and even droplets smaller that 100 nm in diameter are amenable to deformation. Such droplets exhibit Brownian motion and deform under the action of the hydrodynamic and surface forces present. Hence, all theories for the interactions and kinetics of flocculation have to be modified to account for the effect of the deformation. This chapter summarizes some of the work that we have done in this direction. We suggest expressions for the most common types colloidal interactions (electrostatic, van der Waals, etc.) that take into account the droplet deformation along with the contribution of the droplet surface extension. The theoretical analysis of the flocculation kinetics requires knowledge of the hydrodynamic interactions between the droplets. We have derived expressions for the friction tensor and developed a procedure for calculating the mean transition time over the energy barrier that might be present between two approaching droplets. More details and accurate numerical analysis of these interactions is presented in Chapter 10 of this book. For more
346
D.N. Petsev
information on coalescence and droplet breakup the reader may refer to Chapters 9 and 11. Complete understanding of the details of emulsion flocculation and coalescence is important because it helps in developing more accurate models describing emulsion systems. Such models would have better predicting power and can be tested with the constantly improving experimental methods for studying emulsions. REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II]
[12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24]
B.V. Derjaguin, N.V. Churaev and V.M. Muller, Surface Forces (Plenum, New York, 1987). B.V. Derjaguin, Theory of Stability of Colloids and Thin Films (Plenum, New York, 1989). E.J.W. Verwey and J.T.G. Overbeek, Theory and Stability of Lyophobic Colloids (Elsevier, Amsterdam, 1948). J.T.G. Overbeek, Stability of Hydrophobic Colloids and Emulsions, in: H.R. Kruyt (ed.), Colloid Science, Vol. 1 (Elsevier, Amsterdam, 1952) 302-341. D.H. Napper, Polymeric Stabilization of Colloidal Dispersions (Academic Press, New York, 1983). G. Hadziioannou, S. Patel, S. Cranick and M. Tirrell, J. Am. Chem. Soc, 108 (1986) 2869. S. Asakura and F. Oosawa, J. Chem. Phys., 22 (1954) 1255. M. Aronson, Langmuir, 5 (1989) 494. P. Richetti and P. Kekicheff, Phys. Rev. Lett., 68 (1992) 1951. J.L. Parker, P. Richetti, P. Kekicheff and S. Sarman, Phys. Rev. Lett, 68 (1992) 1955. P. Somasundaran, B. Markovic, S. Krishnakumar and X. Yu, Colloid Systems and Interfaces - Stability of Dispersions Through Polymer and Surfactant Adsorption, in: K. Birdi (ed.), Handbook of Surface and Colloid Science (CRC Press, New York, 1997). W.B. Russel, D.A. Saville and W.R. Schowalter, Colloidal Dispersions (Cambridge University Press, Cambridge, 1989). J.N. Israelachvili, Intermolecular and Surface Forces (Academic, New York, 1991). K.D. Danov, D.N. Petsev and N.D. Denkov, J. Chem. Phys., 99 (1993) 7179. K.D. Danov, N.D. Denkov, D.N. Petsev, I.B. Ivanov and R. Borwankar, Langmuir, 9 (1993) 1731. N.D. Denkov, P.A. Kralchevsky, C. Vassilieff and I.B. Ivanov, J. Colloid and Interface Sci., 143(1991)157. N.D. Denkov, D.N. Petsev and K.D. Danov, Phys. Rev. Lett., 71 (1993) 3226. N.D. Denkov, D.N. Petsev and K.D. Danov, J. Colloid and Interface Sci., 176 (1995) 189. J.A.M.H. Hofman and H.N. Stein, J. Colloid and Interface Sci., 147 (1991) 508. I.B. Ivanov and P.A. Kralchevsky, Colloids Surf, 128 (1997) 155-175. D.N. Petsev, N.D. Denkov and P.A. Kralchevsky, J. Colloid and Interface Sci., 176 (1995)201. D.N. Petsev and J. Bibette, Langmuir, 11 (1995) 1075. D.N. Petsev and P. Linse, Phys. Rev . E., 55 (1997) 586. D.N. Petsev, Physica A, 250 (1997) 115.
Theory of Emulsion Flocculation
[25]
[26] [27] [28] [29] [30] [31] [32] [33] [34] [35]
[36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60]
347
D.N. Petsev, Interactions and Macroscopic Properties of Emulsions and Microemulsions, in: B.P. Binks (ed.), Modern Aspects of Emulsion Science (RSC, London, 1998). D.N. Petsev, Langmuir, 16 (2000) 2093. B.P. Binks, Emulsions — Recent Advances in Understanding, in: B.P. Binks (ed.), Modern Aspects of Emulsion Science (Royal Society of Chemistry, London, 1998). B.P. Binks, Annu. Rep. Prog. Chem., Sect. C, 92 (1996) 97. Scheludko, Adv. Colloid Interface Sci., 1 (1967) 391. C. Maldarelli, R.K. Jain, I.B. Ivanov and E. Ruckenstein, J. Colloid Interface Sci., 78 (1980) 118. I.B. Ivanov and D.S. Dimitrov, Thin Film Drainage, Chap. 7, in: I.B. Ivanov (ed.), Thin Liquid Films (Dekker, New York, 1988). J. Bibette, Langmuir, 8 (1992) 3178. I.B. Ivanov, Pure Appl. Chem., 44 (1980) 1241. I.B. Ivanov, D.S. Dimitrov, P. Somasundaran and R.K. Jain, Chem. Eng. Sci., 40 (1985) 137. P.A. Kralchevsky, K.D. Danov and N.D. Denkov, Chemical Physics of Surface and Colloid Science, in: K. Birdi (ed.), Handbook of Surface and Colloid Science (CRC Press, New York, 1997). J.T.G. Overbeek, The interaction between colloidal particles, in: H.R. Kruyt (ed.), Colloid Science, Vol. 1 (Elsevier, Amsterdam, 1952)245-277. R. Aveyard, B.P. Binks, J. Esquena and P.D.I. Fletcher, Langmuir, 15 (1999) 970. J.H.d. Boer, Trans. Faraday Soc, 32 (1936) 10. H.C. Hamaker, Physica, 4 (1937) 1058-1072. H.B.G. Casimir and D. Polder, Nature, 158 (1946) 787. H.B.G. Casimir and D. Polder, Phys. Rev, 73 (1948) 360. E.M. Lifshitz, Soviet Physics JETP, 2 (1956) 73. I.E. Dzyaloshinskii, E.M. Lifshitz and L.P. Pitaevskii, Adv. Phys., 10 (1961) 165. N.G.v. Kampen, B.R.A. Nijboer and K. Schram, Phys. Lett, 26A (1968) 307. J. Mahanty and B.W. Ninham, Dispersion Forces (Academic Press, New York, 1976). D. Henderson, D.-M. Duh, X. Chu and D.T. Wasan, J. Colloid Interface Sci., 185 (1997)265. S.J. Alexander, Physique, 38 (1977) 983. P.G.d. Gennes, Adv. Colloid Interface Sci., 27 (1987) 189. A.K. Dolan and S.F. Edwards, Proc. R. Soc. London Ser. A, 337 (1974) 509. T.F. Tadros, Steric Interactions in Thin Liquid Films, in: I.B. Ivanov (ed.), Thin Liquid Films, Vol. 29 (Marcel Dekker, New York, 1988). H.J. Ploehn and W.B. Russel, Adv. Chem. Eng., 15 (1990) 137. I. Szleifer and M.A. Carignano, Tethered polymer layers, in: I.P.a.S.A. Rice (ed.), Advances in Chemical Physics, Vol. 94 (Wiley, New York, 1996). J.M.H.M. Scheutjens and G.J. Fleer, Macromolecules, 18 (1985) 1882. J. Bibette, J. Colloid and Interface Sci., 147 (1991) 474. X.L. Chu, A.D. Nikolov and D.T. Wasan, J. Chem. Phys., 103 (1995) 6653. D.T. Wasan, A.D. Nikolov, P.A. Kralchevsky and I.B. Ivanov, Colloids Surfaces, 67 (1992) 139. D. Henderson and M. Lozada-Cassou, J. Colloid Interface Sci., 114 (1986) 180. S. Basu and M.M. Sharma, J. Colloid Interface Sci., 165 (1994) 355. N.A.M. Besseling, Langmuir, 13 (1997) 2113. D. Henderson and M. Lozada-Cassou, J. Colloid Interface Sci., 162 (1994) 508.
348
[61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] [96] [97] [98]
D.N. Petsev
S. Marcelja and N. Radic, Chem. Phys. Lett., 42 (1976) 129. J. Forsman and C.E. Woodward, Langmuir, 13(1997) 5459. V.N. Paunov and B.P. Binks, Langmuir, 15 (1996) 2015. J.N. Israelachvili and H. Wennerstrom, Nature, 379 (1996) 219. J. Petkov, J. Senechal, F. Guimberteau and F.L. Calderon, Langmuir, 14 (1998) 4011. R.M. Pashley, J. Colloid Interface Sci., 80 (1981) 153. R.M. Pashley, Adv. Colloid Interface Sci., 16 (1982) 57. P.S. Laplace, Traite de Mecanique Celeste; Supplements au Livre X, Vol. 4 (GauthierVillars, Paris, 1806). R. Finn, Equilibrium Capillary Surfaces (Springer-Verlag, New York, 1986). H.M. Princen, Shape of Interfaces, Drops, and Bubbles, in: E. Matijevic (ed.), Surface and Colloid Science, Vol. 2 (J. Wiley, New York, 1969) 1. P.A. Kralchevsky and I.B. Ivanov, Chem. Phys. Lett, 121 (1985) 111. P.A. Kralchevsky and I.B. Ivanov, Chem. Phys. Lett, 121 (1985) 116. A.D. Nikolov, P.A. Kralchevsky, I.B. Ivanov and A.S. Dimitrov, AIChE Symp. Ser., 252(1986)82. I.B. Ivanov and P.A. Kralchevsky, in: I.B. Ivanov (ed.), Thin Liquid Films (Marcel Dekker, New York, 1988) Chapter 2. S.J. Miklavcic, Phys Rev. E, 57 (1998) 561. B.V. Derjaguin, Kolloid Z., 69 (1934) 155. J.K. Klahn, W.G.M. Agterof, F.v.V. Vader, R.D. Groot and F. Groenweg, Colloids Surf., 65(1992) 151. R. Aveyard, B.P. Binks, S. Clark and J. Mead, J. Chem. Soc. Faraday Trans. 1, 82 (1986) 125. W. Helfrich, Z. Naturforsch., 28c (1973) 693. D.N. Petsev, Mechanisms of Emulsion Flocculation, in: A. Hubbard (ed.), Encylcopedia of Surface and Colloid Science (Marcell Dekker, New York, 2002) 3192. K.G. Marinova, T.D. Gurkov, T.D. Dimitrova, R.G. Alargova and D. Smith, Langmuir, 14(1998)2011. A.D. Nikolov, D.T. Wasan, P.A. Kralchevsky and I.B. Ivanov, J. Colloid and Interface Sci., 133(1989) 1. M.v. Smoluchowski, Phys. Z., 17 (1916) 557. M.v. Smoluchowski, Phys. Z, 17(1916) 585. M.v. Smoluchowski, Z. Phys. Chem., 92 (1917) 129. N.A. Fuchs, Z. Phys., 89 (1934) 736. P.M. Debye, Trans. Electrochem. Soc, 82 (1942) 265. F.C. Collins and G.E. Kimbal, J. Colloid Sci., 4 (1949) 452. T.R. Waite, Phys. Rev., 107 (1957) 463. T.R. Waite, J. Chem. Phys., 28 (1958) 103. G. Wilemski and M. Fixman, J. Chem. Phys., 58 (1973) 4009. J.M. Deutch and B.U. Felderhof, J. Chem. Phys., 59(1973) 1669. D.F. Calef and J.M. Deutch, Annu. Rev. Phys. Chem., 34 (1983) 493. G.H. Weiss, J. Stat. Phys., 42 (1986) 3. J. Keizer, Chem. Rev., 87 (1987) 167. A.A. Ovchinnikov, S.F. Timasheff and A.A. Belyy, Kinetics of Diffusion Controlled Chemical Reactions (Nova Scientific, New York, 1989). W. Rybczynski, Bull. Intern. Acad. Sci. Cracovie Ser. A Sciences Mathematiques, 1 (1911)40. J.S. Hadamard, Compt. Rend. Acad. Sci. (Paris), 152 (1911) 1735.
Theory of Emulsion Flocculation
[99]
349
B.G. Levich, Physicochemical Hydrodynamics (Prentice-Hall, Englewood Cliffs, 1962). [100] X. Zhang and R.H. Davis, J. Fluid Mech., 230 (1991) 479. [101] R.H. Davis, J.A. Schonberg and J.M. Rallison, Phys. Fluids A, 1 (1989) 77. [102] S.G. Yiantsios and R.H. Davis, J. Colloid Interface Sci., 144 (1991) 412. [103] A.Z. Zinchenko, M.A. Rother and R.H. Davis, Phys. Fluids A, 9 (1997) 1493. [104] T.T. Traykov and I.B. Ivanov, Int. J. Multiphase Flow, 3 (1977) 471. [105] T.T. Traykov, E.D. Manev and I.B. Ivanov, Int.J. Multiphase Flow, 3 (1977) 485-494. [106] K.D. Danov, D.S. Vlahovska and I.B. Ivanov, Journal of Colloid and Interface Science, 211 (1999)291-303. [107] L. Onsager, Phys. Rev., 37 (1931) 405. [108] O. Reynolds, Phil. Trans. Roy. Soc. London A, 177 (1886) 157. [109] A. Einstein, Ann. Phys., 17 (1905) 549. [110] K.D. Danov, I.B. Ivanov, T.D. Gurkov and R.P. Borwankar, J. Colloid Interface Sci., 167(1994)8. [ I l l ] S. Dukhin and J. Sjoblom, Kinetics of Brownian and Gravitational Coalescence in Dilute Emulsions, in: J. Sjoblom (ed.), Emulsions and Emulsion Stability (Marcel Dekker, New York, 1996). [112] N.A. Mischuk, S.V. Verbich, S.S. Dukhin, O. Holt and J. Sjoblom, J. Disp. Sci & Technol., 18(1997)517. [113] M.P. Aronson and H. Princen, Nature, 286 (1980) 370. [114] M.P. Aronson and H. Princen, Colloids Surf., 4 (1982) 173. [115] J. Bibette, T.G. Mason, H. Gang and D.A. Weitz, Phys. Rev. Lett., 69 (1992) 981. [116] J. Bibette, Langmuir, 9 (1993) 3352. [117] T.G.M.V.d. Ven and S.G. Mason, J. Colloid Interface Sci., 57 (1976) 517. [118] O.D. Velev, G.N. Constantinides, D.G. Avraam, A.C. Payatakes and R.P. Borwankar, J. Colloid Interface Sci., 175 (1995) 68. [119] C.W. Gardiner, Handbook of Stochastic Methods (Springer, New York, 1985). [120] N.G.v. Kampen, Stochastic Methods in Physics and Chemistry (Elsevier, New York, 1992). [121] B. Radoev, A. Scheludko and E. Manev, J. Colloid Interface Sci., 95 (1983) 254. [122] R. Tsekov, Colloids Surfaces, 141 (1998) 161. [123] S.G. Yiantsios and R.H. Davis, J. Fluid Mech., 217 (1990) 547. [124] S.K. Chakarova, M. Dupeyrat, E. Nakache, C D . Dushkin and I.B. Ivanov, J. Surf. Sci. Technol., 6 (1990) 17. [125] C. Maldarelli and R.K. Jain, in: I.B. Ivanov (ed.), Thin Liquid Films (1988) 497. [126] A. Vrij, F. Hesselink, J. Lucassen and M.v.d. Tempel, Proc. K. Ned. Acad. Wet. B, 73 (1970) 124. [127] F.L. Calderon, T. Stora, O.M. Monval, P. Poulin and J. Bibette, Phys. Rev. Lett., 72 (1994)2959. [128] O.M. Monval, F.L. Calderon, J. Philip and J. Bibette, Phys. Rev. Lett., 75 (1995) 3364. [129] R. Aveyard, B.P. Binks, W.-G. Cho, L.R. Fisher, P.D.I. Fletcher and F. Klinkhammer, Langmuir, 12(1996)6561. [130] B.P. Binks, W.-G. Cho and P.D.I. Fletcher, Langmuir, 13 (1997) 7180. [131] W.-G. Cho and P.D.I. Fletcher, J. Chem. Soc. Faraday Trans., 93 (1997) 1389. [132] P.D.I. Fletcher, Interactions of Emulsion Drops, in: D. Mobius and R. Miller (eds.), Drops and Bubbles in Interfacial Research (Elsevier, New York, 1998). [133] B.V. Derjaguin, A.S. Titievskaya, I.I. Abrikosova and A.D. Malkina, Faraday Discuss. Chem. Soc, 18(1954)27.
350
D.N. Petsev
[134] O. Holt, O. Saether, J. Sjoblom, S.S. Dukhin and N.A. Mischuk, Colloids Surf., 123124(1997) 195. [135] O. Holt, O. Saether, J. Sjoblom, S.S. Dukhin and N.A. Mischuk, Colloids Surf, 141 (1998)269. [136] O. Saether, J. Sjoblom, S.V. Verbich, N.A. Mishchuk and S.S. Dukhin, Colloids Surfaces, 142(1998) 189.
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 9
Coalescence kinetics of Brownian emulsions N.O.Mishchuk Institute of Colloid Chemistry and Chemistry of Water, National Academy of Sciences of Ukraine, Vernadskogo avenue, 42, Kyiv 03142, Ukraine 1. INTRODUCTION Investigation of the reasons and regularities of emulsion destabilization, including droplet coalescence, is of considerable scientific and practical importance [1-3]. Despite the progress in theoretical [3-7] and experimental [4, 8-11] investigation of properties of thin films, there is not much information about coalescence rate in real emulsions. It is quite difficult to study the process of coalescence in emulsions directly and therefore coalescence between two droplets or between a droplet and a layer of liquid [12-14] or the general destabilization process are investigated. Although calculation of the total number of droplets [8] or use of dielectric spectroscopy [9] to establish coalescence rate are quite informative, these methods do not allow to follow the details of existent processes and, hence, to distinguish the rate of rupture of thin films between droplets from the rate of the other processes accounting for the approach and emergence of the films. Absence of industrial equipment for automatic measurement of the number of singlets, doublets and multiplets is one of the reasons for such unfavorable scientific and technological situation. Therefore, the characteristic time of coalescence r c in emulsions was established experimentally for the first time by quite laborious analysis of emulsion photos made at certain time intervals [15-20]. It should be also noted that, when developing a method of coalescence analysis, not only the technical support of investigations matters but the general methodological approach and availability of theoretic models ensuring the adequate consideration of the multiple factors influencing the investigated processes are of great importance as well. In particular, there are still no reasonably accurate theoretical models and reliable experimental verification of Van der Waals forces retardation and their screening with electrolyte [21, 22]. Practically, no theoretical models taking into account the variety of factors determining the stability of emulsions with medium and high concentration exist
352
N.O.Mishchuk
[23, 24]. Not enough attention is paid to investigation of coagulation reversibility. Emulsion destabilization is quite often considered as a combination of irreversible aggregation and coalescence (see, for instance, [8, 25]), although the aggregation reversibility may be an important factor influencing the rate of coalescence and destabilization regularities in whole [26- 29]. The role of aggregation reversibility diminishes or increases depending not only on the interaction energy of specific droplets in a doublet but also on the correlation of rates of the aggregate disintegration, emergence of a new aggregate and coalescence. Thus the aggregation irreversibility may be quite relevant for certain emulsions that are used for practical purposes. Yet, in many other cases, the aggregate reversibility in the secondary energy minimum - and sometimes in the primary minimum as well - may turn out to be the main factor determining emulsion properties. Hence, to develop general fundamental ideas of emulsion properties, it is imperative to select such conditions of experiment that would allow separating all the processes that are of importance for understanding of emulsion stability, including the aggregation reversibility. At the same time, it is necessary to ensure high accuracy of measurements and the possibility of univocal interpretation of obtained results, i.e., to conduct the experiment under conditions ensuring a theoretical description of the system with a minimum number of constants and arbitrary assumptions. For instance, the use of monodisperse emulsions or emulsions with inconsiderable scattering of droplet sizes allows not only to weaken the role of orthokinetic (gravitational) coagulation but also to avoid the difference in the value of energy of interaction between different droplets, which affects both aggregation kinetics and the coalescence process. Formation of multiplets is also one of the complicating factors. The energy of interaction between droplets in a multiplet and, therefore, coagulation reversibility and coalescence rate in these aggregates depend on the quantity of interacting particles and the character of their packing. Moreover, when multiplets are formed, initially monodisperse emulsion quickly changes into polydisperse and that leads to transition from purely Brownian coagulation to Brownian-gravitational one. Therefore Dukhin and Sjoblom recommended to use singlet-doublet emulsions, i.e., those in which forming of multiplets is unlikely, as a standard method for determining the characteristic time of an elementary coalescence act [26]. Such emulsions may be obtained at a certain combination of the time of rapid coagulation and the effective time of aggregate disintegration. Articles [28, 29] provide a thorough theoretical analysis of the possibility of obtaining singlet-doublet emulsions by selecting a droplet surface charge, electrolyte concentration and droplet volume fraction. The review [28] also analyzes the influence of various factors (retardation and screening of Van der Waals forces with electrolyte [30-32], presence of surfactants [33], etc.) on the energy of
Coalescence Kinetics ofBrownian Emulsions
353
interparticle interaction and, hence, the probability of aggregation in the primary and/or secondary minimums and, correspondingly, the probability of coalescence. The present chapter is based methodologically on the approach suggested in [26]. However, unlike [26], it explores not only the significance of coagulation reversibility but also investigates the role of transitions from the secondary minimum to the primary one. The theoretical analysis is limited to diluted monodisperse singlet-doublet Brownian emulsions that satisfy the aforementioned requirements. To provide a clearer view of the processes taking place and their influence on the kinetic dependencies that are measured, the investigation starts with simpler systems. Hence the first part of the chapter deals with stable thin films. This allows not only to carry out a qualitative and quantitative analysis of the processes taking place but also to dwell on certain moments of aggregation description that are of interest both for emulsions and dispersions. High emphasis is placed on discussion of the question of correct interpretation of obtained experimental data. 2.THEORY 2.1. Definition of Brownian emulsion and the main limitations on the investigated system The main property ofBrownian diffusion is thermal motion of droplets. If droplets are sufficiently small (with radius of micron order or smaller), it is the Brownian motion and not sedimentation that defines the coagulation behavior of emulsion. Emulsions with larger droplets may be referred to Brownian ones only in case that dispersion medium has almost the same density as emulsion droplets, which is almost impossible in real systems. Even if we suppose that there is ideal monodispersity and complete stabilization of the system, when large droplets with density p^ other than that of the medium pm are used, there will be complications with creaming of emulsion. At sedimentation or floating-up of droplets, they will concentrate in the lower or upper part of the cuvette. In the opposite part of the cuvette, volume fraction of droplets will decrease, which may be erroneously taken for coagulation. The investigated picture is also modified by the counter flow of dispersion medium, which compensates the sedimentation transfer of droplets from one part of the cuvette to another and distorts their Brownian motion. If coagulation takes place in the monodisperse system, a deviation from the regularities of Brownian coagulation may also emerge due to different velocity of sedimentation or floating-up of singlets and doublets (or enlarged droplets when they coalesce). The bigger is the difference between the density of droplets and the density of medium and the larger are the droplets, the more
354
N.O.Mishchuk
important is the gravitation effect. Significance of the effect of this process on Brownian coagulation may be defined using the model [34, 35], developed on the basis of the main ideas of gravitation coagulation [36]. Without dwelling on the details of these works, it should be noted that, when the superposition of Brownian and gravitation factors is assumed (which is admissible only in case of small effect of sedimentation or floating-up of droplets on Brownian coagulation), a change of the number of singlets could be described using rather simple formulas. In particular, the following expression is obtained for monodispersion [35]:
*.«).*, (.{•"^o[ 7 ^-'" ( ' + 2 ' /r - ) ]] {
(It™
(1)
2
8(l + l/r OT ) JJ
where N\{t) = N Q l(\. +111 Sm) is the change of the total number of particles at rapid purely Brownian coagulation [37], t is the time counted from the beginning of the process, tgm - n rj aQ / a kT is the time of rapid coagulation calculated according to Smoluchowski, Pe0-47taQAp/3kT is the Peclet number characterizing the relative role of gravitation and diffusion factor in coagulation and corresponding to the initial state of system with the droplet radius a0, TJ is the dynamical viscosity of the dispersion medium, a is the volume fraction of droplets, kT is the energy of Brownian motion, k is Bolzmann constant, T is the absolute temperature, Ap = \p^ - pm is the difference between the densities of droplets pj and medium pm, and g is gravity acceleration. According to Eq. (1), deviation of the number of singlets from the value calculated on the basis of Smoluchowski formula, increases in time. Let us assume that the accuracy of investigation may be considered acceptable when the error of determining the number of singlets caused by gravitation does not exceed 10%, i.e., N(t)/N{(t) = 0.9. The typical time t* that is necessary for appearance of such inaccuracy and which is calculated according to Eq. (1), is presented in Fig. 1. As the figure shows, the larger is the value Ap, the smaller is, at fixed radius a 0 , the value of t * /rSm, which leads to considerable deviation from the Brownian nature of coagulation. Dependence of t*lt^m on the radius of droplets is even sharper. It is not surprising taking into account that the coefficient of diffusion of Brownian particles according to Stokes-Einstein formula is inversely proportional to their radius D o -kT/(mr]a§, and the velocity of sedimentation of droplets is proportional to squared radius
Coalescence Kinetics of Brownian Emulsions
355
Used - 2aQ&pgl9ri [38]. The larger the value t*ltSm, the later, in comparison to Smoluchowski time, a deviation from the regularities of purely Brownian coagulation appears. Using the above figure, it is possible to determine the size of droplets for which, at the given Ap, the gravitation effect may be ignored. If the coagulation process within the Smoluchowski time is to be investigated, the intersection of the dotted line with the curves for relevant Ap (see fig.l) corresponds to the maximum acceptable droplet size. For instance, at Ap = 0.5, the radius of droplets should be less than 0.5 micron. Only in this case emulsion (or suspension) may be considered as Brownian. If emulsion is to be investigated within a longer period of time, the condition becomes even more rigid. Dependence of the critical droplets size on the difference between the densities of the medium and droplets is shown in Fig.2. As is seen from the figure, the gravitation effect may be ignored only at very small radiuses OQ , or only at very small differences of densities Ap. Similar calculations can be performed for non-monodisperse emulsion, showing that, at small scattering of droplets radiuses (for instance, when the volumes of the smallest and largest droplets differ twice), a considerable
Fig.l. Dependence of typical time /* , normalized to Smoluchowski time Tgm, on the radius of droplets. Numbers near curves refer to the difference of densities Ap. The dotted line shows equality t* = t^m. Fig.2. Critical radius of droplets fln^at gravitational coagulation as a function of the difference of densities Ap for investigations during / < T$m (curve 1) and / <\Qzgm (curve 2).
356
N.O.Mishchuk
distinction from the monodisperse system exists only in the beginning of the process, before multiplets are formed in the initial monodisperse emulsion, i.e., when the latter has not begun to behave as non-monodisperse emulsion. At A/? < 0.1, results of numerical calculations for such non-monodisperse system differ from the monodisperse one not more than for several percents. Thus, the above estimation of critical particle sizes is suitable also for moderate nonmonodispersity. Unfortunately, there is another restriction on droplet size as well. Droplets cannot be much smaller than a micron because, due to Ostwald ripening [39], nanosize droplets disappear quite quickly and therefore Brownian aggregation and coalescence are complicated by this process. Another important restriction is concerned with the fact that existing theoretical models calculate the time of rapid coagulation, assuming free Brownian motion of droplets, i.e. independence of particle motion from the neighboring particles. Such approach is correct when surface forces exist only at short distances, in other words, when droplets move independently of each other within the time interval between their collisions. This is true for aqueous dispersion mediums, where the radius of surface forces action is usually considerably smaller than droplets size and, correspondingly, considerably smaller than the distance between droplets even in rather concentrated emulsions. However, this restriction can prove to be important at investigation of "water-in-oil" emulsions, where, due to low ion concentration in non-aqueous medium, the extension of double electrical layer K~ can be not only commensurable to but even distinctly larger than droplet radius CQ, i.e., satisfying condition KOQ «1. Indeed, the volume fraction of droplets in emulsion a could be expressed through the diameter of droplets 2ag and the distance between their centers 2Z>Q as a ~ (OQ Ibo) . Let us suppose that droplet motion may be considered sufficiently free when a droplet gets in the field of neighboring droplets (roughly speaking, when double layers intersect) at the distance not exceeding 5% of the average distance between them b§ - OQ, i.e., 2K~ «(b§ - ao)l2§. Taking into account the expression for a, a restriction on the thickness of the double electrical layer is obtained: rcag »40\fa/\l-^faj. It follows from this that, at a = 0.001, 0.01 and 0.1, it is necessary that AT^Q were larger than 4, 10 and 35, correspondingly. Hence, the thickness of the double layer should not exceed 1/4, 1/10 and 1/35 of the droplet radius. These conditions may be fulfilled in real non-aqueous mediums only for large droplets OQ » l/^m • However, in this case even at small Ap gravitation plays an important part, i.e., the emulsion is not Brownian but Brownian-gravitational.
Coalescence Kinetics of Brownian Emulsions
357
2.2. Basic notions of reversible and irreversible coagulations in Brownian emulsion Kinetic behavior of miniemulsions is a manifestation of competition between surface forces and Brownian motion [2, 15, 16, 40]. According to DLVO theory, in the general case, energy of pair interaction of charged particles is characterized by the existence of two minimums with a barrier between them [2], which result from different amplitudes and characteristic lengths of action of electrostatic and Van der Waals forces (Fig.3). Energy of interaction at the presence of other forces (hydration, hydrophobic, steric etc.) has the same qualitative appearance. The scheme of theoretic and experimental investigation of coalescence kinetics presented in this chapter is based on the main ideas of Brownian emulsions. All processes taking place in emulsions, except for coalescence, are caused by Brownian motion of droplets. In dilute emulsions, droplets diffuse freely until they get into the area where surface forces of other droplets act.
Fig.3. Energy of interaction of two particles U as a function of distance between their centers r and scheme of fluxes in a system (see Section 2.3): J\ is the diffusion flux from infinity to droplet, J 2 t n e fux from secondary minimum to infinity, JT, the flux from secondary minimum to primary one, J4 the flux from primary minimum to secondary one, J5 the flux from primary minimum to coalesced state. R = 2OQ is the closest distance between the centers of particles with radiuses QQ . Barrier, secondary and primary energy pits are shown using letters B, S, P.
358
N.O.Mishchuk
Depending on the energy of interaction between droplets, the following scenarios of further developments may be distinguished [41, 42]. When getting into energy minimum (pit), droplets are attracted and, at sufficiently big depth of the minimum, they maintain this stable position, i.e. form a doublet. Since there are two minimums (a deeper, primary, minimum and a shallower, secondary, one), there may be two types of doublets, which, following the established terminology [41, 29], are called here as primary (PD) and secondary (SD). When droplets approach, they first find themselves in the secondary minimum (S) and hence form a secondary doublet (SD). If the minimum is very shallow, i.e., attraction energy is small, the doublet is quickly disintegrated. This is reversible aggregation (or fiocculation). When the barrier is not very high (B), transition from the secondary (S) to the primary (P) minimum is possible. Since the latter is usually quite deep, the aggregate that is formed (PD) is stable and thus irreversible coagulation takes place. At the distances between the surfaces of particles having order of several nanometers, due to surface hydrophylicity additional repulsion may also appear, hindering further approach in the primary energy minimum. Adsorption layers, creating a steric barrier between droplets, act in a similar way. Sometimes the length of the layer of adsorbed macromolecules or macroions may be big enough to overlap the primary minimum, either in part or in full, and thereby to prevent irreversible coagulation. An opposite situation is possible when droplets are weakly charged and the barrier is negligible or inexistent. In that case two minimums become confluent. Not only the possibility of formation of primary and secondary aggregates but also the probability of their disintegration considerably depends on the change of interaction energy at the approach or moving away of droplets. For each pair of approached droplets, the probability and, consequently, the time of change of their aggregate state (aggregate disintegration or transition from one minimum to another one) depends on the thermal motion of surrounding molecules and the motion of the Brownian droplets themselves. For the same period of time, some doublets may disaggregate; some may transfer from one minimum to another, and others may retain their state. However, in general, transition from one state to another in the system has certain statistic character. Therefore it is convenient to consider transition processes as droplet fluxs Jj from one state to another (Fig.3) and, by analogy with multi-stage chemical reactions, to introduce characteristic times corresponding to each stage of such "reaction" (Fig.4). Unlike disperse particles, deformation is of importance for droplets. Since, in the absence of other forces except surface ones, deformation may take place only at attraction between droplets, it is possible in the energy minimum only. Apparently, in most cases, it may be accomplished only in the primary minimum, where interaction energy is sufficient for noticeable deformation.
Coalescence Kinetics of Brownian Emulsions
359
This means that no characteristic time, except for the period of disintegration of the primary doublet r pcj S(j, depends on droplet deformation. Since surface flattening intensifies interaction between droplets, their deformation leads to deepening of the primary minimum and increase of coagulation irreversibility. Thus, the process of coalescence depends on droplet deformation for two reasons: due to increase of stability of the primary doublet and because both the form of film and its thickness and area change as the gap flattens. Taking into consideration that the process of deformation depends not only on the interaction energy of approached droplets (which decreases as the droplets become smaller, according to the DLFO theory) but also on Laplace pressure [43] (which increases as droplet size decreases), deformation is more peculiar to quite large droplets, with the radius exceeding a micron. Moreover, as the study [25] shows, for micron droplets, change of interaction energy due to droplet deformation is not too significant and therefore can be ignored. Since the chapter is focused on rather small droplets, the role of deformation in aggregation will not be analyzed here in detail. Furthermore, to simplify theoretic models, we will assume that coagulation is irreversible in the first minimum, i.e., r p j sc[ -> oo even without droplet deformation. As regards coalescence, even though it depends on droplet deformation, the chapter is aimed at determining the time of coalescence irrespectively of the details of this process and therefore the mentioned issue will not be considered here. Investigation of small droplets also allows to consider them as solid particles, i.e., to neglect the effect of the surface and internal mobility of droplets on their sedimentation and on interparticle interaction. Since usually droplets of emulsion are covered with a layer of adsorbed maeromolecules, their mobility, especially for micron-sized droplets, diminishes to zero.
Fig.4. Scheme of aggregation, disintegration and coalescence. Characteristic times of different stages: TS SC{- collision of two singlets (formation of a secondary doublet), TSC]^Sdisintegration of a secondary doublet, rsd^pd ~ transformation of a secondary doublet into a primary one, rpd sd~ transformation of a primary doublet into a secondary one, TC coalescence time.
360
N.O.Mishchuk
2.3. Method of calculation of characteristic times The most consecutive model of aggregation and disaggregation in dilute suspensions, including calculation of characteristic times of subprocesses, is the model of Muller [41, 42]. This model could be directly used for description of non-coalescing droplets and expanded with account of coalescence [26-29]. If emulsion is rather dilute, multiplets practically do not exist and therefore it is enough to take into account the fluxes of single particles (singlets) Jj directed to or from a certain arbitrarily chosen particle (fig.3). Let us suppose that a system has reached a quasi-equilibrium, in other words, the aggregation and disaggregation are mutually equilibrated and the concentrations of singlets «1, primary rip and secondary ns doublets are constant. Naturally, at such quasi-stationary regime, diffusion fluxes are constant and can be described by the equation [41, 42]
L dr
kT dr J
y }
where v(r) is the concentration of singlets at the distance r from a chosen particle, U{r) is the energy of interaction of two particles, D = 2DQ is the coefficient of reciprocal diffusion of two singlets, defined through the diffusion coefficient of single sphere Do = kT 167rr/a0. At short distances D(r) = 2DQ /p(r), where p(r) is the factor taking into account hydrodynamic resistance between the particles approaching each other (for example,
p{r) = l +
ao{r-2ao)[44]).
To define fluxes J\ -J4, presented in Fig.3, Muller proposed to solve Eq.(2) with the following boundary conditions. The flux Jx leads to appearance of secondary doublets and satisfies the boundary conditions of Smoluchowski [45] - the absence of particles in the secondary pit and constant concentration of particles at the infinity v(r e S) - 0
and v(r = <x>) = NQ = const
(3)
The flux J2 leads to disintegration of a secondary doublet and satisfies the conditions |V(r)4;zr2dr = 1
and v(r = <x>) = 0
(4)
i.e., at the moment of doublet disintegration there is only one particle in the
Coalescence Kinetics of Brownian Emulsions
361
secondary pit and, during the process of disintegration, the particle that moves away interacts only with the second particle from the doublet. The flux J 3 defines the probability of transformation of secondary doublets into primary ones and satisfies the condition v(r e P) = 0
(5)
and the first condition in (4), that is, the absence of a particle in the primary pit and the presence of a single particle in the secondary pit. Finally, the flux J 4 defines the probability of transformation of primary doublets into secondary ones and satisfies the conditions
^v(r)4nr2dr = \
H
v(r€S) = 0
(6)
(rsP)
i.e., at the moment of transition there is a single particle in the primary pit and no particle in the secondary pit. According to Eq.(2) and conditions (3-6), the expressions for fluxes can be presented in the following way [42]:
R2TS
R2WTS
R2WTP
where 1^,1^,1^, are the "powers" of primary and secondary minimums, Wis the retardation factor [46, 47] which can be presented as
^ = \r2
exp(- U{r)) \P}Q txV{U(r'))dr' dr « r^
(8)
(r'Y
W=
f (reB)
P(r
'hxp{O(r'))dr'
fcxp(-O(r))r2dr
Fs= (reS)
(9)
(r'Y >
v
I> =
Jexp(-£/(r)/ kT)r2dr (reP)
(10)
362
N.O.Mishchuk
Here the distance r is normalized on diameter of particle 2a 0 U(r) = U(r)/kT. Taking into account the achievement of a local equilibrium J] n\ = 2 J2 %
H J^fip-J^ng
and
(11)
and existence only of singlets and doublets, the characteristic times of subprocesses have been expressed [42] as
r
s sd
'
-
l
1
-
J,
sd s
T
J3
-J_-^l
> ~J2~2D0'
R2WT^
1 l
sd,pd
~
SnD0RN0'
—
„ _. ^-^0
i '
l
pd,sd~
j
R2WTP
~~ ~ r. *M ^^O T
m\
(
v1J/
Although Eqs.(7-13) are received for a quasi-equilibrium state, which can be reached only at t » Ts sd'^sd,s'^sd pd>^pd,sd> they c a n t>e a l s o u s e c l when analyzing transition to an equilibrium state [42]. It is interesting to note that the time of transition from a secondary to a primary pit depends not only on the barrier height but also on the depth of the secondary pit, and the characteristic time for inverse transition depends on the depth of the primary pit. The larger the depth of the primary and secondary pits and, correspondingly, the larger the values of FS,TS and Tp, the less "readily" the particle leaves its place and the larger is the time of transition from the secondary pit into unconstrained state TS(J}S or into primary pit fsdnd • As a consequence, the flux J\ is divided into two fluxes J2 and J 3 , the intensities of which, according to Eq.(7), correlate as J2 UT, = W^s l^s ^W . This means that, due to the barrier, the probability of formation of a primary doublet is W times smaller than the probability of disintegration of an aggregate. It should be stressed that, according to the model presented above, even if a secondary pit is very shallow, i.e. the characteristic times Tsj s and/or ^sd,pdare small the transition from two singlets to a primary doublet formally has two stages: transition from singlets to a secondary doublet and transition through the barrier from a secondary doublet to a primary one. Although coalescence was not investigated in [42], formally it is possible to introduce the notion of the flux J 5 and the characteristic time of rupture of
Coalescence Kinetics ofBrownian Emulsions
363
thin film between two droplets (or the time of coalescence) Tc = 1/ J 5 , as it was done, inter alia, in [19, 26]. Some problems of correct calculation of interaction energy and characteristic times, necessary to ensure the proper description and understanding of existent processes, were discussed in the review [29]. 2.4. System of differential equations for description of coagulation and coalescence According to papers [42] and [19, 26], the change of concentrations of singlets (n s ), primary ( n p d ) and secondary ( n s d ) doublets can be described by interconnected differential equations
dns^_nj T
d?
+2nsd
+npd
r
s,sd
(i4)
f
sd,s
c
^ = -3L-»J^- + - U+ -^L 2r
d?
s,sd
dn
pd
{Fsd,.!
nsd
=
npd
r
&
r
pd,sd
npd
r
sd,pd
sd,pd)
(15)
r
pd,sd
r
c
or, with introduction of dimensionless values, by the equations djh^_n2+2^L dt
T
+^ L
(1?)
T
sd,s
c
d
-^=i-J—+-^1
dt dn
pd dt
2
T
y sd,s nsd
= T
sd,pd
as)
T
sd,pd)
npd T
c
where the concentrations of singlets and doublets are normalized on the initial droplet concentration NQ: ns=nsl NQ, npd =rtpdl NQ and nsd =nsd INQ and all characteristic times are normalized on the time of formation of secondary doublets Tj = T; I Ts sd .
364
N.O.Mishchuk
Since the size of the chapter is limited, the possibility of transformation of a primary doublet into a secondary one and correspondent terms npd I Tpd scj at transition from Eqs. (14-16) to Eqs. (17-19) are neglected. Eqs.(17-19) take into account the process of formation of secondary doublets, their possible disintegration into two singlets or transformation into primary doublets. Coefficient 2 in Eq. (17) shows that disintegration of a doublet results in two singlets and coefficient 1/2 in Eq.(18) shows that two droplets form a doublet. Eqs. (17-19) assume that formation of triplets is negligible. This is possible only in the case when the number of doublets in the system is not too high. The condition necessary for this will be discussed in Section 2.5. The terms npci I rc describe coalescence. The number of coalesced droplets is directly proportional to the number of formed thin films, i.e., the number of primary doublets, and inversely proportional to the characteristic time of their rupture r c . Since existent experimental methods do not allow to measure relatively small changes of droplet size, the coalesced droplets are included into Eqs. (17) and (19) as initial singlets. Taking into account that it is practically impossible to create an absolutely monodisperse emulsion, the 26% difference between the radiuses of initial (<3Q) a n d coalesced (a\ — 42CIQ) droplets cannot strongly reduce its non-monodispersity. Eqs.(17-19) should be solved together with initial conditions ns(t = 0) = l
nsd(t = 0) = 0
npd(t = 0) = 0
(20)
To better understand the role of coalescence and the way of its experimental definition, first, the process of coagulation will be investigated for the case of stable thin layer between droplets. Such analysis is important both for emulsions with thick adsorbed layers of macroions or macromolecules preventing coalescence and for suspensions, whose aggregation in primary and secondary energy minimums is underinvestigated. In this particular case, conditions (20) should be supplemented with the additional condition of the constant total number of droplets n
s + 2nsd + 2nPd =!
(21)
2.5. Single energy minimum, non-coalescing droplets The simplest case to be considered is the existence of a single energy minimum combined with high stability of thin film. Such situation takes place when the energy barrier is very high or the first minimum is inaccessible due to steric repulsion and, as a result, the primary doublets cannot form. Another case is when the surface charge is very low and, thus, a barrier is absent and only one
Coalescence Kinetics ofBrownian Emulsions
365
minimum exists, and thin film is stabilized with a dense layer of macromolecules. In this case Eqs.(17-19) can be reduced to two differential equations ^ .
=
_
n
2
dt
dnsd
+ 2 ^ T
_ sd,s
_ ns
dt
2
(22)
nsd T
sd,s
where the indexes corresponding to the secondary minimum are preserved, although, as was written above, the nature of a minimum could vary. With account of two first conditions (20) and condition (21), the solution of this system of equations can be presented as „ _ q O ~ c 2 ) - c 2 (l - q)exp(- (q - c2)t)m 0 " c2) ~ (1 - c, )exp(- (c, - c 2 ) 0
_^~ns 2
where c
i = ^
2r
;
C
2=^-^
M^
( 25 )
lT
sd,s
At ? -> oo, the quasi-equilibrium concentrations of singlets and doublets are established 1-Ci
ns (t -> co) = d
« ^ (/ -> oo) = - ^ - i -
(26)
Solution (24) is obtained at an arbitrary ratio of characteristic times. However, since the focus is made here on singlet-doublet emulsions, the following limitation for the characteristic times can be obtained from Eq.(25) nsd «1
or
^sd,s « l
(27)
With account of these conditions, the expressions (24) can be simplified: ns » l - ( l - q X l - e x p ( - ( q -c2)t)) , nsd * ^f>(l
-exp(-(q -c2)t))
(28)
366
N.O.Mishchuk
The accuracy of the obtained approximation depends on the correlation between time t and the multiplier in the exponent (q -c-i). Comparison of (28) and (24) shows that at t > (q -c 2 )~ the deviation of (28) from the exact solution is about 0.1%. At t<(c\-C2)~ and strict observance of condition (27), the deviation is less than 0.25%. When condition (27) is not fulfilled, the deviation of approximation (25) from exact solution increases (at TS(J S ~ 1, it is less than 2% and, at Tsds ~ 5, it reaches 15%). Analysis of numerical and analytical solutions of differential equations shows that a quasi-stationary state is reached during the time tst that is a few times larger than tsci s. For example, if it is assumed that doublets have reached the quasi-stationary state when their concentration is equal to 90% from possible maximum value, the critical time of stationarity can be obtained from (28) as _ ln(10) _, 2.3 r ^ , S t
~~
/1
C]-c2
A
*•
'
Jl + 4TsdjS
With account of (27), quasi-stationary concentrations of singlets and doublets can be evaluated as ns (t -> oo) = 1 - rsdtS + 2x\s
;
nsd (;->«,) = ^
~^sd's
(30)
The larger the time of doublet existence, the higher is their concentration and the lower is the concentration of singlets. Under condition (27), nscj «1/8. At such small number of doublets, the appearance of triplets is very unlikely, i.e., this is really singlet-doublet emulsion. Examples of numerical calculations according to Eqs.(24) are shown in fig. 5. Since the quasi-stationary state could be reached very quickly, the best way to investigate experimentally the aggregation in the secondary minimum with account of possible disintegration of doublets is to measure the stationary concentration of doublets or the total concentration of "particles" ns + n^ using light absorption. The values should be stable during a very long time. Modern equipment allows to measure light absorption with very high accuracy during very short time t « TSCJ S , i.e., long before the quasi-stationary state is reached. However, the interpretation of such results could be incorrect. Indeed, as could be seen in Fig. 5, the beginnings of curves at considerably different values of TSCJ s are quite close to each another. Therefore accidental
Coalescence Kinetics of Brownian Emulsions
367
Fig.5. Dependence of concentrations of singlets ns (curves 1-6) and doublets nsci (curve 1'6') on time t. Disintegration timeTscj^s equals to : 1000 (curve 0) , 1 (1,1',1"); 0.8 (2,2',2"); 0.6 (3,3',3"); 0.4 (4,4',4"); 0.2 (5,5',5"); 0.1 (6,6',6"). Curves 0*-6* - sum of singlets and doublets ns + nsci, presented in Figs. 5c and 5d in different scales.
deviations of concentrations can lead to inaccuracy of measurement comparable to the distances between the curves. It is also necessary to stress that the decrease of slope of beginnings of the curves representing reversible coagulation in comparison to the slope of the curve of rapid coagulation (see fig. 5) can be
368
N.O.Mishchuk
wrongfully considered as manifestation of slow coagulation related with the overcoming of energy barrier between particles or droplets. However, in reality, as can be seen from curves in fig.5, at high energy barrier that prevents transformation of secondary doublets into primary ones, the slope of curves changes owing to aggregation reversibility. In contrast, at large depth of a secondary minimum (curves 0, 0*), aggregation looks as rapid coagulation in a primary minimum, although droplets do not overcome the energy barrier. 2.6. Secondary minimum of limited depth and infinitely deep primary minimum. Non-coalescing droplets The problem with account of transition into a primary minimum without coalescence can be described by three differential equations
* ' = - „ ? +2 "** dt
(31)
T
sd,s
dn
n
sd
S „ f
dt
2
dn
Pl dt
=
i
l
l
m
x
T
\Jsd,s
sd,pd )
^cL_ T sd,pd
(33)
Differentiation of Eq.(32) with account of Eqs. (31), (33) and (21) leads to the differential equation of the second order
d\£L 2
dt
= _2rhn^_dn^(2ns T
sd,pd
dt
{
+
J_
+
T
sd,s
_L_ > |
(34)
T
sd,pd)
the solution of which could be received according to the scheme proposed in [19, 26], where the linearization of a similar equation was performed by replacing function ns (t) with its initial value ns (t = 0)« 1. The exactness of solution at t > tst (29) could be higher when ns is replaced with its quasiequilibrium value ns « ns (t -> oo) - c\ (26), which was obtained for simpler emulsion without transition into a primary minimum. After presentation of Eq.(34) as
^V^^L dt2
dt
\
]-2c,-^+ T- ^ + ^— T T sd,s
sd,pd)
sd,pd
(35)
369
Coalescence Kinetics ofBrownian Emulsions
its solution is expressed very simply: a f
e-
_e~a2'
\
2(af2-a,) where
1 1 2ci ^,=20,+-!-+—!-;?,=—*T
sd,s
T
(38)
T
sd,pd
sd,pd
Concentrations of primary doublets and singlets are obtained from Eqs.(33, 36, 21) as
a2(l-e-^)~a,(l-e-a^) "/«/= o ^"7°
H;
. . . 2 «5=l- n^-2«pd
..„ (39)
Emulsion will remain singlet-doublet if, in addition to condition (27), it satisfies the condition of slow transition into the primary minimum, which is possible if hd,Pd»hd,s
and ^d,Pd>1
( 4 °)
In the opposite case, the number of primary doublets increases considerably, i.e., the total number of doublets rises and the possibility of triplet formation appears. Under conditions (40) q\ <1, p\ » 1 and a\ « 1 a2 » 1 and, therefore, at t »1 /«2 solutions (36, 39) can be simplified:
2«2
2*sd,pda\a2
{ a2
T
sd,pd^\«2)
The deviation of these simple expressions from exact numerical solutions during the period tst
370
N.O.Mishchuk
The rate of transformation of secondary doublets into primary ones can be found by differentiation of obtained functions (41): dn
dnscLa_a}_e-aV^ dt 2a2
vd ._, e~a* dt 2Tsd^pda2
'
(42)
These expressions become more obvious after expansion of parameters a, (37) in series dn
sd dt
T
., 2r
2 sd,s
sd,pd
T
c
sd,s Tsd^pd
T
dnpd dt
Tsd^
^
sd,s Tsd^pd
2t
sd,pd
The rate of formation of primary doublets is not only inversely proportional to the time of overcoming of barrier Tsdtpd but also proportional to the time of existence of secondary doublets rsd^s. The longer is the existence of secondary doublets, the larger is the probability that disintegration of secondary doublets will be replaced with their transformation into primary ones. The decrease of the number of secondary doublets caused by their transformation into primary ones is also inversely proportional to the characteristic time rsd pd. However, the obtained expression contains, instead of the primary degree of disintegration time, the secondary one r , . This fact, SCI jS
strange at the first glance, is actually quite correct. On the one hand, the rate of decrease of the number of secondary doublets is proportional to the probability of their transformation into primary doublets rsd s I rsd pd; on the other hand, it is proportional to the number of secondary doublets nsd(see Eq.(30)) which grows as the time of existence of secondary doublets r ^ 5 increases. Taking into account that in the investigated case of short disintegration time Tsds «1 (27), the rate of change of the number of secondary doublets dnsd I dt is lower than the rate of formation of primary ones dnpdl dt (see Eqs.(43)), the total concentration of doublets nd - nsd +npd grows. It is reasonable that under condition (40) the intense transition from a secondary to a primary pit takes place when the number of secondary doublets is sufficiently high. Joint analysis of the numerical solution of Eqs.(31-33) and the analytical solution (24) of the problem of aggregation without transition into a primary minimum showed that during the time / < tst (26) they practically
Coalescence Kinetics of Brownian Emulsions
371
coincide. Thus, at the short time of investigation t
, m= ]5C«iZ«a) a
2
<44)
~a\
Obtained value of rmax is very close to the time when the kinetic curves for a single minimum reach their quasi-stationary values (Fig.5,6). However, at ? m a x , bends of the curves representing the number of singlets and total number of doublets appear. The values of the maximum concentration of secondary doublets and concentration of singlets in the point of the bend can be found by substitution of Eq.(44) into Eqs.(36) and (39). Comparing the theoretical and experimental maximums of doublets concentrations and bends of curves for singlets can give a lot of information as that allows to determine the region of transition to quasi-stationary state and the value of TSCJS . The slope of curves after the transition region allows to define the value of rsd,pd • Let us compare the rates of processes at t < tst (29) and at t > tmax(44). In both cases the rate of investigated processes is defined by exponential function of time. In the first case, according to Eq.(28), the characteristic time is l/(cj -C2)- ?sd,s I ^+ ^Tsd,sx
T
sds~^T
d
while in the second case,
according to Eq.(41), it equals \l ct\ « rsc( ^ ( 3 + l/TS(jfS). On the basis of these expressions one can conclude that at t > tmax the change of singlets concentration caused by transformation of secondary doublets into primary ones is a i / ( q -c2)={r2sds-2T3sdJ/(xsdpd(\ + 3zsd^)) time slower than the change at t
372
N.O.Mishchuk
Fig.6. Dependence of concentrations of singlets ns (curves 1-6), secondary nsj(\'-6') and primary « p j (4"-6") doublets on time t. tsd,pd =1000 and Tsd^= '000 (curves 0,0') ; 0.03 (1,1'); 0.1 (2,2'); 0.3 (3,3'); r ^ ^ ^ = 1 0 and r^ ;iS =0.03 (curves 4,4',4"); 0.1 (5,5',5"); 0.3 (6,6',6"). Curves 0*, 4*-6* represent the sum of singlets and all doublets ns +npd +nsci; curves 4**-6** show the sums of primary and secondary doublets npd+nsdzero and the concentrations of singlets and secondary doublets coincide with the concentrations presented in Fig.5. A decrease of transition time Tsd^pd creates conditions for transformation of secondary doublets into primary ones (see
Coalescence Kinetics ofBrownian Emulsions
373
curves 4"-6"). As a result, the concentrations both of singlets (curves 4-6) and secondary doublets (curves 4'-6') decrease. However, since characteristic times for all the processes strongly differ (rsej>pci » TS^SCJ » TSCI^ ), at first, the faster processes equilibrate and that leads to coincidence of the beginnings of curves for singlets (1-3 and 4-6) and doublets (l'-3' and 4'-6'). A difference between the curves at the same value of TS(Jp(j but different values of TSCI s is clearly seen in Fig.6. Therefore, both the barrier that should be overcome by droplets and their location and interaction energy are of importance. The lower the interaction energy in the secondary minimum and, correspondingly, the smaller the time of secondary doublet disintegration, the lower is the probability of transition into the primary minimum at the same height of the barrier. Since secondary doublets transform into primary ones slower than they disintegrate, the concentration of primary doublets in the beginning of the process is low and does not affect the concentration of singlets and secondary doublets. However, as the number of primary doublets increases, the number of droplets participating in aggregation and disaggregation in the secondary minimum decreases and that leads to a decrease of singlet concentration. As a result, maximums of the curves 4'-6' and bends in the curves 4-6, 4*-6* appear demonstrating the qualitative change of the investigated process. Similarly to aggregation in the secondary minimum (Section 2.5), in the present case, not only the time of transition to a quasi-stationary state but also the values of corresponding concentrations depend on the characteristic times of each subprocess. Both in case of insurmountable and surmountable barriers, the concentration of secondary doublets grows and the concentration of singlets decreases as the time of doublet disintegration TS(JS increases (see transition from curves 1, 1' to 3, 3' and from curves 4, 4',4" to 6, 6', 6"). At very low TSCJs, the probability of secondary doublets disintegration is to such extent higher than the probability of their transformation into primary doublets that the concentration of secondary doublets becomes independent of the latter process. This conclusion is proved by small difference between curves 1' and 4', which coincide at the given scale of the figure. It is useful to compare the kinetic curves at fixed TSJS and different T
sd pd ( s e e Fig-7)- The decrease of the barrier leads to acceleration of
transitions into the primary minimum and affects the number of secondary doublets and residuary singlets. However, even at the transition times that are commensurable with the time of rapid coagulation (TSJ nd=^> curves 5, 5',5") or even smaller ( r 5 j n ( i < l, see curves 6,6",6"), the general rate of coagulation is considerably lower than the rate of rapid coagulation. This means that at a very
374
N.O.Mishchuk
Fig.7. Dependence of concentrations of singlets ns (curves 1-6), secondary nscj (V6') and primary rip^ (l'-6') doublets and the sum of doublets and singlets (l*-6*) on time / . The calculations are performed at TSCJ s =0.1 and
TSCJ n ^ = 1 0
(curves 1,1',1",1*); 5
(2,2',2",2*); 3 (3,3',3",3*); 2 (4,4',4",4*); 1 (5,5',5",5*) and 0.3 (6,6',6",6*), correspondingly. Curve 0* is calculated at zsc[ s =1000 and TSCJ pj =1000. Figs.c and d present the same results on different scales.
shallow secondary minimum and, consequently, low xs£s, droplets scatter sooner than they receive an opportunity to overcome the barrier. The analyzed peculiarities of primary doublet formation could be used for
Coalescence Kinetics of Brownian Emulsions
375
theoretical explanation of the well-known experimental fact - at reduction of electrolyte concentration, the deceleration of coagulation in comparison with rapid coagulation is smoother than according to the calculation of the retardation factor (see, for example, [46, 47]). Let us compare Figs.7d and 5d. Curve 1 * in Fig.7d coincides with curve 6* in Fig. 5d and the fan of curves 2*-6* in Fig.7d is the branching of the same curve 1 * that appears owing to the change of transition time TSCJpcj. Since in Fig.7 all values of x^
pcj
are larger than TscjfS , at short time t < TSCJs curves
l*-6* practically coincide. This means that if the measurement is performed within the time interval 0 < t < TS^^S , there will be a difference between curve 0* for rapid coagulation and curves l*-6*, caused not by transition into the primary minimum but by disintegration of secondary doublets. In other words, this will be "slow" coagulation, although not in its classical meaning. It will be slow due to reversibility of aggregation in the secondary minimum. The qualitative illustration of two different types of "retardation" is shown in Fig.8, where the factor of retardation at transition into the primary minimum without disintegration of doublets (curve 1) and the factor of the "retardation" caused by the change of the slope of kinetic curves related with the reversibility of aggregation in the secondary minimum (curves 2,2',2") are presented.
Fig.8. Factor of retardation W as a function of electrolyte concentration. Curve 1- traditional factor of retardation (see Eq.(9)), curves 2,2',2' - "retardation" factor caused by reversibility —21 of aggregation in the secondary minimum. Hamaker constant is ^ = 2 1 0 J, surface potential y/ = 25mV, radius OQ =0.5/jm. Curve 2" is calculated at the linear change of surface potential (// from 25mV at C=\molll to 50mV at C = 10~ moll I. Volume fractions of droplets are 1% (curve 2) and 10 % (curves 2',2").
376
N.O.Mishchuk
The numerical calculations were performed for arbitrary chosen parameters with account of retardation and screening of van der Waals interaction according to the models [30, 31] that particularly were used in [28, 29]. It was also assumed that experimental measurements of the number of droplets were performed at t - 0.2, which corresponds to the practically linear section of the kinetic curves in Fig.7d. As seen from the curves in Fig.8, at reversible aggregation, theoretical "retardation" of the process at reduction of electrolyte concentration could be even slower than that obtained experimentally, for example, in [48, 49]. The degree of "retardation" depends both on the change of volume fraction of droplets (2, 2') and on possible change of surface forces related with the change of surface potential (curves 2', 2"). The dependence of the position of curve 2 on volume fraction could be used to prove the "non-traditional" mechanism of coagulation retardation and to separate it from other mechanisms, for example, those proposed in [48] and other papers. Choice of the time of measurement is also important. For example, in Fig. 7d, the kinetic curves at /<0.2 look linear, i.e., seemingly it is enough to do the measurements at t - 0.2 and use them to define the retardation factor. However, curve 6* in Fig.5d (the same time interval but different scaling) shows the evident deviation from linearity. Thus, measurement of droplet concentration at smaller t will produce another value. That is why it is very important to study aggregation not only in the beginning of the process but within the large time interval giving more information about the investigated process. It is necessary to stress that, when investigating "water in oil" emulsion, one should take into account the hydrophobicity of droplet surface and, correspondingly, the possibility of hydrophobic attraction between droplets, which can considerably decrease the height of energy barrier and accelerate both aggregation in the primary minimum and rupture of thin films between droplets [50]. The dissolution of air in dispersion medium could also affect the rate of aggregation [49] since it changes the van der Waals interaction [51]. Essential influence on the rate of aggregation can be made by the mobility of droplet surface [52] and presence and behavior of adsorbed layers [3, 6]. Since the above examination of aggregation was carried out without account of coalescence, all obtained results are also valid for solid particles. The performed analysis of the role of TSJ>S in formation of primary doublets is especially important for "water in oil" emulsion, where the thickness of the double layer is very large and, consequently, TSCJ^ diminishes to zero. 2.7. Single energy minimum. Coalescing droplets The simplest version of coalescence takes place after aggregation directly in the primary minimum, i.e., without complicative transition from the
311
Coalescence Kinetics ofBrownian Emulsions
secondary minimum to the primary one. This is possible at low surface charge when two minimums merge in one. The system of three differential equations (17-19) in this case is reduced to two: ^
= -,,2 + 2 - ^ + ^
dt
T
(45)
x
pd,s
c
dn
pd n2s f 1 l) -jr = ^-npd\ +— dt 2 \rpd,s Tc)
where zpds
( 46 )
has the same meaning as rsds
in Section 2.5. The system of
equations can be solved using the scheme of linearization presented in Section 2.6. The solution for the number of doublets is e-ft'-e-/ht nPd rnd=—-. 2{/32-px)
(47)
where
A , 2 = " f [-1± P f ]; P2=2cXp + ^ ^ ; 9 2 = ^ 2
1
i
T d s
p \ \
P>
Tc
Tc
(48)
and quasi-equilibrium concentration of doublets c\p is expressed by Eq.(25) with the replace of ts^^s by rp^s. Since the total number of droplets at coalescence is not preserved, the concentration of singlets can be found from Eqs.(45, 46) as
(ns + 2npd)=-
y
or
ns=\-2npd
-nc
where
is the decrease of the total number of droplets caused by coalescence.
(49)
378
N.O.Mishchuk
The location of the maximum for doublet concentration can be found similarly to Eq.(44) as
_ln(/V/?2) 'max
KJ1J
n n
P2 ~ P\
However, contrary to the previous case, where the similar expressions defined the location of maximum number of secondary doublets and the bend of kinetic curve for the sum of primary and secondary doublets, in the given case, this is the location of the maximum concentration of one type of doublets. In Fig.9 presented below, calculations are performed specially for shallow minimums and, correspondingly, quick disintegration of doublets, in order to make the role of doublets disintegration more illustrative. As seen from the figures, the rate of coalescence depends on aggregation reversibility. The larger the doublet lifetime r ^ , the higher is the probability of coalescence. The increase of
TSC[S
(compare Figs. 9a,b and 9 c,d) leads not only to
stronger decrease of ns and ns +np(^ but also to more pronounced maximums of npci that demonstrate faster coalescence. It should be underlined that the intensification of coalescence is caused not by the decrease of the time of film rupture r c , but by the increase of the number of droplets that are in the energy minimum and wait for the rupture of the film. Attention should be also paid to the fact that although the total number of droplets decreases, according to the curves in Fig.9, the emulsion remains to be singlet-doublet, since npci lns < 0.15 when the time of investigation is not too long. To show the role of doublet disintegration for a wider range of parameters, the decrease of the total number of droplets is calculated for a few sets of characteristic times of disaggregation and coalescence (see Fig. 10). The influence of rp^s depends on its correlation with the time of coalescence r c . If these values are of the same order (see curves 1, 1' or 2, 2'), the number of coalesced droplets changes weakly. However, when they differ to the great extent (compare curves 1 and 1" or 2 and 2"), the number of coalesced droplets decreases considerably. It is interesting to note that, excluding the beginning of the curves, i.e. the time interval t
Coalescence Kinetics of Brownian Emulsions
379
exponent is a slow function of time and even at t » 1 can be approximated using two first terms of expansion into series as e~^]t ~ 1 - f5\t, reflecting the linear dependence of nc on t and, correspondingly, the linear dependence of the total number of droplets on time.
Fig.9. Dependence of concentrations of singlets ns (curves 1-4), doublets tip^ (l'-4') and their sum ns + npd (l'-4') on time t. tpd,s
=
^ ' (Figs, a, b), Tpd>s = 0.3 (Figs, c, d) and
TC = 1000 (curves 1,1', 1"); 10 (2, 2', 2"); 3 (3, 3', 3"); 1 (4, 4',4"). Curve 0* is calculated for T
pd
5
= 1000 and rc = 1000.
380
N.O.Mishchuk
Fig. 10. Dependence of the decrease of total number of droplets nc on time at tp^s = 100 (1 3); Tpd,s=™ (1'- 2') and
tpdjS=\
(l"-2") for r c =30 (1,1',1"); r c =100 (2,2',2");
r c =1000(3).
Taking into account that, at numerical calculations for Fig. 10, rather large values of coalescence time r c were used, the coefficient fi\ for this limiting case can be written in a simpler view:
A=^ Tc
!
(52)
2c]p+l/TpdfS
Thus, it is inversely proportional to the coalescence time. This is not surprising since it is coalescence that is the slowest process limiting the rate of emulsion destabilization in the given system. 2.8. Two energy minimums. Coalescing droplets Let us analyze a more general case of coalescence with the possibility of transition from the secondary minimum to the primary one that can be described by three equations (17-19). Since the solution of these equations is more complicated than in the previous cases, fist of all the numerical solution will be analyzed (see Figs. 11 and 12). The comparison of Figs. 11 and 12 clearly points to the influence of characteristic time isd,pd o n m e formation of primary doublets and, consequently, on the rate of coalescence at a fixed time of rupture of thin films rc. The lower energy barrier leads to quicker appearance of primary doublets
381
Coalescence Kinetics ofBrownian Emulsions
and rupture of thin layers and that, in its turn, causes quicker decrease of numbers of singlets and secondary doublets. Similarly to the case of noncoalescing droplets, at the fixed value of rsd S , the slope of curves in the
Fig.ll. Dependence of concentrations of singlets ns (curves 1-5), secondary « 5 ^(l'-5') and primary tip^ (2"-5") doublets, sum of doublets n^ (l*-5*) and sum of singlets and doublets n +n
s d
Tsd,pd
(curves ]**-5**)
On
time t.
Curves 0,0** are received at r ^ ^ =1000,
=1000, r c = 1 0 0 0 ; curves l,l',l", 1*,1** at 7 ^ = 0 . 3 ,
TC =1000; the others at r ^ ^ =0.3, rsd,pd
=1°
and 7
r^^=1000
and
c =1000 (2, 2', 2", 2*, 2**); 10(3,
3', 3", 3*, 3**); 3 (4, 4', 4", 4*, 4**); 0.1 (5, 5', 5", 5*, 5**), correspondingly.
N.O.Mishchuk
382
beginning of aggregation is independent of TSC[p j , which was taken considerably larger than r ^ s. The initial slope of curves is also independent of rc,
the value of which is comparable with TSCJS and is considerably smaller
than TSC{pc[. The latter result is caused by the need to overcome a barrier before coalescence starts. This means that instead of comparing TSCI s and r c , it is necessary to compare the values of TS(jtS and T
T
T
sd sp + c >> sd s'
me
T
sd,sp
+T
c • Since
initial slopes of curves are defined by a faster
process- the disintegration of secondary doublets.
Fig. 12. The same as Fig. 11, but rsd^s = 0.3 is replaced with tgd^pd =
Coalescence Kinetics of Brownian Emulsions
383
As seen from numerical analysis, slow rupture of thin film in case of quite high barrier (large TSCJpj) leads to a slow change of the total number of droplets. This means that the system should be examined for a very long time and therefore the exactness of the investigation can be low owing to different accompanying processes, the influence of which increases with time. The decrease of the total number of droplets calculated for a few sets of parameters is shown in Figs. 13 a,b. The curves in Fig. 13a are similar to curve 2" in Fig. 10, which is also shown in Fig. 13a to make comparison easier. Indeed, the calculations for curve 2" were performed at xc =100, r ^ =1 (since in Fig.10 a single minimum was investigated, rpci s is the analogue of ts^^s), therefore the curves in Fig. 13a can be considered as the complication of the process shown in Fig. 10. The droplets represented by curve 2" participate in three processes (appearance of a doublet, its disintegration or coalescence). In the case shown in Fig. 13a, the additional process appears, that is the overcoming of a barrier. It might seem that, when a barrier exists, coalescence should be slower. However this is true only for a short period of time when the number of primary doublets is very small (see beginning of curves 2,3 and 2" in Fig. 13a). Later, at a low barrier (curves 1-3), coalescence occurs quicker than in a single minimum case (curve 2"). This non-trivial result has a very simple explanation. The disintegration of a doublet from a single shallow minimum occurs quicker than its possible coalescence. In case of two minimums, after formation of secondary doublets, some of them disintegrate into singlets, while others transform into primary doublets, which wait until thin films rupture. For a single energy pit, the number of coalesced droplets is proportional to the ratio of times iS(j s lxc . For two pits, the TSC[ S/TSCJ sp part of secondary doublets transforms into primary ones, and the l / r c part of transformed doublets coalesces. Thus, the rate of coalescence is proportional to Tsd,s/[rsd,spTc)- This very rough qualitative evaluation of the role of two energy pits is reflected by kinetic curves for the number of coalesced droplets. At a high barrier (TSCJ^P = 10, curve 4), when the probability of formation of primary doublets is low, at fixed tc, the rate of coalescence for two minimums is lower than for a single minimum. In contrast, at low barrier (and small TSCJsp, see curves 1,2), the probability of formation of primary doublets is high and therefore the rate of coalescence for two minimums is higher than for a single one. It is also worthwhile to analyze results of numerical calculations at fixed r sd,pd' a s presented in Fig. 13b, in which curve 3 coincides with the same curve in Fig.l3a and curve 2" from Fig. 10 is also shown repeatedly. The shorter the
384
N.O.Mishchuk
Fig.13. Dependence of the decrease of the total number of droplets nc on time a) TC = 100, TSCJ^S = 1 and Tsd,pd = °-' (curve 1), 1 (curve 2), 3 (curve 3), 10 (curve 4); 6) TSCI pd =0.1 and r c = 1 0 0 (curves 1-3) and r c = 1 0 (curves l'-3'); xscjs
equals O.I
(curves 1,1'), 0.3 (curves 2,2') and 1 (curves 3,3'). Curve 2" is copied from Fig. 10.
time TSCI s, the lower is the number of coalesced droplets. Droplets from secondary doublets, instead of transformation into primary doublets, "prefer" to leave aggregates and therefore the probability of coalescence decreases. The decrease of r c , when other parameters are fixed, leads to quicker coalescence, as shown by curves l'-3' inFig.l3b. Unfortunately, the analytical solution of differential equations for a low barrier and slow rupture of thin layer is rather difficult. Therefore we will analyze analytically only the opposite limiting case of a relatively high barrier and quick rupture of thin films. This case is more interesting from the experimental point of view than the quick rupture of film and a low barrier, since the latter process would look as rapid coagulation and on its background it would be difficult to define the rate of coalescence. When condition T
r
sd,pd »
(53)
c
is fulfilled, it may be assumed that each primary doublet merges into a single droplet immediately after it is formed, i.e., as may be described by the condition dn
pd dt
nsd
=
T
sd,pd
_npd=Q T
c
( 5 4 )
385
Coalescence Kinetics ofBrownian Emulsions
In the framework of the above scheme, the solution of differential equations (17-19) with account of (54) takes the following view: e-nt
n
_e-r2l
e-r]t_e-nt
T
sd =
"pd = —
(55)
T
n-n
sd,Pd
n-n
and d /
n
\
—\ns+2nsd) at
pd
=
or
n
=1 -2nsd
TC
- In
-nc
(56)
ft=-S_
(58)
d F
where
*Tsd,Pd
nnTsd,pd
ri,2=^f-l± P f 1 P3=2c 1+ -U^- ; 2
I
\|
P3 I
T
sd,s
T
sd,pd
T
sd,pd
The maximum of the kinetic curves for secondary doublets is reached after the time _ Hn 172)
(59)
\jy;
' m a x ~~
72-71 In the limiting case when xc I tsdnd ~^> 0, the same time corresponds to the maximum of primary doublets (curve 5"in Fig. 12b) and, consequently, to the maximum of the total number of doublets (curve 5* in Fig.l2d). However, when ratio ?clTsd,pd'v& n o t t 0 ° l ° w ' condition (53) is violated and the maximum of curves for primary doublets and the total number of doublets shift to larger times (see, for instance, curves 3", 4" in Fig.12b and curves 3*, 4* in Fig.l2d). 3. DESIGN OF EXPERIMENTAL INVESTIGATIONS Comparison of kinetic curves for total concentrations of singlets and doublets (see Figs. 7c, 9c, l i e and 12c) shows that trends of curves seem to be very similar at considerably different subproccesses of emulsion destabilization.
386
N.O.Mishchuk
This means that there may be different interpretation of the experimental data obtained by absorption or scattering of light. In particular, absorption of light does not allow discrimination between two aggregated droplets and two droplets that have already coalesced. Moreover, taking into account that absorption of light depends on the length of the used wave, droplet size, size and shape of the aggregate and total concentration of droplets, the applicability of the used method and equipment for the specific investigated system should be thoroughly controlled. It is clear that for this aim it is necessary to develop a special technique that allows to count not only the general number of droplets and aggregates but also singlets, doublets and multiplets separately. Such apparatus based on the combination of streaming ultramicroscopy and single particle light scattering is used for a long time to investigate aggregation in suspensions [53, 54]. As distinct from streaming ultramicroscopy that owing to hydrodynamic flow could affect the stability of secondary doublets, transition into primary minimum and the process of coalescence, videomicroscopy [15-20] is a method that, in principle, does not influence the state of emulsion. Moreover, videomicroscopy allows not only to count the number of doublets and multiplets but also to observe the behavior of chosen doublets (see, for instance, Fig.3 in paper [17]) and follow general changes in emulsion. Therefore it seems to be the only reliable way to define the rate of rupture of thin films against a background of all subprocesses that lead to drop coalescence . Owing to the fact that the analysis of the videopictures [15-19] was not automated, the information about the rupture or coalescence of doublets was not sufficiently complete. Development of special computer programs for videoimage processing could allow to analyze large data files. Only in this case, careful study would allow separation of primary (irreversible) and secondary (reversible) doublets and direct definition of the mechanism of slow coagulation. Beside the development of special videotechnique and corresponding software tools, details of experiment arrangement also play an important part. As it was repeatedly written above, at a given ratio between the density of droplets and medium, droplet size should be limited to avoid the influence of gravitation. The scattering of emulsion size should be minimal, as that could allow not only to weaken the gravitational component of coagulation but also to avoid the change of interaction energy and, correspondingly, characteristic times for different pairs of droplets. One can easily see that the strict monodispersity is less important for rapid coagulation and more important for all other subprocesses of aggregation. Indeed, the most important factor for rapid coagulation is the efficiency of collision between droplets. The analysis of collision efficiencies [55] for identical 4DQQQ and different (Z)o + Z)jX^o +a\) droplets with radiuses QQ a n ^
Coalescence Kinetics of Brownian Emulsions
387
a\ shows that, owing to inverse proportion between diffusion coefficients and radiuses Dq\~\l a§\, t n e efficiencies differ weaker than the radiuses. For example, the efficiencies of collision between initial (a 0 ) droplets only and between initial ( CIQ ) and coalesced (a\ = iflciQ = 1.26ao) droplets differ for 1.3%. When two types of quite different droplets (ag ar>d a\ - 2«o) a r e present, the efficiencies differ for 12.5% only. Emulsion with such scattering of droplet sizes, which does not considerably affect Smoluchowski time, can be easily obtained using even very old methods (see, for instance, [56]). Unfortunately, all other characteristic times depend on radiuses considerably stronger and therefore the aforementioned radius scattering is too large to provide a small difference between the values of TS^JS, tsci,pd a n d Tc for different pairs of droplets. This is caused both by the dependence of characteristic times r,- on droplet radiuses f,- ~ R l2D§~a^ (12-13) and by their exponential relation (see Eqs. (12, 13) and (8-10)) with the energy of interaction U{r), which is a function of particle radiuses 2a o «i /(ao + a\) [2]Indeed, calculation of the characteristic time of doublet disintegration [28, 29] showed that a relatively small change of droplets radius leads to a sharp change of its value. It is clear that the larger the scattering of radiuses and of characteristic times, the more inaccurate is the definition of the time of thin film rupture. Thus, the design of the experiment should provide for the preparation of emulsion with maximum monodispersity. The experiment also requires to select such electrolyte concentration and volume fraction of droplets that would allow forming of singlet-doublet emulsion with prevalence of singlets, relatively low number of doublets and almost total lack of triplets and other multiplets. It is this type of emulsion that makes it possible to observe kinetics of singlet and doublet concentrations, define the characteristic time of disaggregation and calculate the time of thin film rupture. For simplification of theoretical analysis all times were normalized on the time of formation of secondary doublets Ts>scj. However the choice of the optimal value of this time is very important. It should be about several minutes; otherwise bends in kinetic curves, which are of significance for interpretation of investigated processes, could be lost in the time interval between preparation and investigation of emulsion. Since, at the given difference of droplets and medium densities, the choice of droplet size is limited because of gravitation influence, it is necessary to change both the size and volume fraction of droplets. Thus, for example, for one-percent emulsion (a = 0.01) and radius a0 = 0.5jam, Ts sci is approximately equal to 10 sec, but for OQ =\/MI it is already 1.5 min.
388
N.O.Mishchuk
Decreasing the volume fraction in 10 times results increases tgm to 15min. With account of condition (26), the dimensionless disintegration time should be T sd s K< 0-25, therefore the bend in kinetic curves should appear within a few seconds, a few dozens of seconds or a few minutes, correspondingly. It is also necessary to draw attention to the fact that while at rapid coagulation of emulsion the main developments take place in the bulk of emulsion, at slow coagulation the key role could be played by wall effects. First of all, this refers to the abovementioned emulsion concentration or dilution near the upper or bottom wall of the cuvette caused by gravitation. This means that behavior of droplets should be investigated not near the wall but in the bulk of emulsion. For this aim the cuvette should satisfy a few conditions. It should be vertical with the height about 5-10 cm so that the state of emulsion in the central part of cuvette was independent of creaming. With account of the droplet sedimentation velocity Usecj =2aoApg/9r/,
for example at Ap =
O.Olg/cm , such height of the cuvette allows to investigate the properties of emulsion with one micron droplets during a week, with 3-micron droplets during two days and so on. The thickness of the cuvette should be about several hundreds of microns to avoid natural convection. It is also necessary to stress that investigations may not be carried out just near the vertical wall of the cuvette since, owing to hydrodynamic resistance between a droplet and the wall, the latter hampers diffusion [57] and, correspondingly, slows down the rate of droplet collision. The requirement of studying emulsion at a certain distance from the wall does not cause any considerable complications, since investigations should be performed at strongly diluted emulsions, which are sufficiently transparent to allow video filming in the depth of the cuvette. Another important fact to be considered is the attraction of droplets to walls, in result of which the droplets are either attached to the walls or spread on their surface. Therefore the walls should have a strong charge of the same sign as the droplets to create a high energy barrier between them. They can also be covered with non-charged macromolecules, overflowing primary pits in the gap between walls and droplets and in this way preventing aggregation in the primary minimum. For example, in [28] it was proposed to use for this aim agarose gel, which immobilizes electric double layer hydrodynamically [58] and, hence, is supposed to overflow the primary minimum. In addition, the adsorption of macromolecules should be irreversible, since even their small admixture in emulsion changes its behavior. In conclusion, it should be said that the theoretical investigation presented in [26-29] and, for a more general case, in the given chapter shows the role of different factors in destabilization of emulsions and observable kinetic of coalescence. The comprehensive experimental study and theoretical analysis of
Coalescence Kinetics ofBrownian Emulsions
389
kinetic behavior of singlets and doublets allow to discriminate the role of reversibility of aggregation, transformation of secondary doublets into primary ones and coalescence in the general process of emulsion destabilization. The characteristic time of rupture of thin film between two droplets regardless of the nature of this rupture could be determined. Although the experience of application of videomicroscopy in such research is not quite wide, the accomplished experimental investigations [15-20] confirm the possibility of designing a device to measure coalescence and disintegration time of doublets, which, in combination with emulsion dynamics modeling, could be a useful tool in analyzing emulsion properties. Development of theoretical modeling and experimental investigation that takes into account overcoming of energy barrier and provides a thorough analysis of all existing subprocesses would refine the insight on emulsion properties and allow to optimize emulsion technologies with respect to stabilization and destabilization. REFERENCES [I] J.A. Kitchener and P.R. Musselwhite, in: "Emulsion Science", I.P. Sherman (Ed.), Academic Press, New York, 1968. [2] B.V. Derjaguin, Theory of Stability of Colloid and Thin Films, Plenum, New York, 1989. [3] I.B. Ivanov and P.A. Kralchewski, Colloids Surf. A, 128 (1997) 155. [4] I.B.Ivanov. (ed.), Thin Liquid Films. Marcel Dekker, New York, 1988. [5] K.D. Danov, I.B.Ivanov, T.D.Gurkov and R. Borwankar, J. Colloid Interface Sci., 167,
(1994)8. [6] [7] [8] [9]
I.B.Ivanov, K.D. Danov and P.A. Kralchewski, Colloids Surf. A, 152 (1999) 161. D.N.Petsev, Langmuir 2000, 16, 2093-2100. R.G.P. Borwankar, L.A. Lobo and D.T. Wasan, Colloids Surf. A, 69 (1992) 135. J. Sjoblom, in: J. Sjoblom (ed.), Emulsion and Emulsion Stability, Marcel Decker, 1996, p. 393. [10] V.Mishra, S.M.Kresta and J.Masliyah, J. Colloid Interface Sci., 197, (1998) 57. [II] K. Khristov, S.E. Taylor, J. Czarnecki and J. Masliyah, Colloids Surf. A, 174 (2000) 183. [12] T.Gilespie and E.K.Rideal, Trans.Faraday Soc, 52 (1956) 173. [13] E.G.Cockbain and T.S.McRoberts J. Colloid Sci. 8 (1953) 440 [14] H.Sonntag and H.Klare. Z.Phys. Chem. 223 (1963) 8. [15] S.S. Dukhin, O. Saether and J. Sjoblom, in: K.L. Mittal and P. Kumar (eds.), Emulsions, Foams, and Thin Films, N.Y., Basel, 2000. [16] S.S. Dukhin, O. Saether and J. Sjoblom, in: J. Sjoblom (ed.), Encyclopedic Book of Emulsion Technology, Marcel Decker, 2001, 71. [17] O. Holt, O. Saether, J. Sjoblom, S.S. Dukhin and N.A. Mishchuk, Colloids Surf. A, 123124(1997) 195. [18] O. Saether, J. Sjoblom, S.V. Verbich, N.A. Mishchuk and S.S. Dukhin, Colloids Surf. A, 142(1998) 189. [19] O. Holt, O. Saether, J. Sjoblom, S.S. Dukhin and N.A. Mishchuk, Colloids Surf. A, 141 (1998)269. [20] O. Saether, J. SjQblom, S.V.Verbich and S.S.Dukhin, J. Disp. Sci. Technol., 20 (1999) 295.
390
N.O.Mishchuk
[21] W.B.Russel, D.A.Saville and W.R.Schowalter, Colloidal Dispersions, Cambridge University Press, New York, 1989. [22] J.N. Israelachvili, Intermolecular and Surface Forces, 2 nd Edition, Academic Press, London, 1991. [23] D.N.Petsev, V.M.Starov and I.B.Ivanov, Colloids and Surf. A, 1 (1993) 65-81. [24] W.Albers and J.Th.G. Overbeek, J. Colloid Interface Sci., 14 (1959) 518. [25] K.D. Danov, N.D. Denkov, D.N. Petsev and R. Borwankar, Langmuir, 9 (1993) 1731. [26] S.S. Dukhin and J. Sjoblom, J. Dispers. Sci. Technol., 19 (1998) 311. [27] S.S. Dukhin, J. Sjoblom, D.T. Wasan and O. Saether, Colloids Surf. A, 180 (2001) 223. [28] N.A.Mishchuk, R.Miller, A.Steinchen and A.Sanfeld, J.Colloid Interface Sci., 256 (2002) 435. [29] S.S.Dukhin, N.A.Mishchuk, G.Loglio, L.Liggieri and Miller R., Adv. Colloid Interface Sci., 100-102(2003)47. [30] Ya.I. Rabinovich and N.V. Churaev, Kolloid Zh., 52 (1990) 309. [31] V.N. Gorelkin and V.P. Smilga, Kolloid Zh., 34 (1972) 685. [32] N.A. Mishchuk, J. Sjoblom and S.S. Dukhin, Kolloid Zh., 57 (1995) 785. [33] E.D. Shchukin, E.A. Amelina and A.M. Parfenova, Colloids Surf. A, 176 (2001) 35. [34] N.A. Mishchuk, S.V. Verbich, S.S. Dukhin, O. Holt and J. Sjoblom, J. Disp. Sci. Technol., 18(5) (1997) 517. [35] A.S.Dukhin, Kolloid Zh., 49 (1987) 630. [36] V.M.Voloschuk and Y.S. Sedunov, Coagulation Processes in Disperse Systems. Hydrometeoizdat: Leningrad, 1975 (in Russian) [37] M.Smoluchowsi, Z.Phys.Chem. 92 (1918) 129. [38] V.G.Levich, Physico-Chemical Hydrodynamics, Prentice Hall, New York, 1962. [39] A.S.Kabalnov, A.V.Pertsov, Yu.D. Aprosin and E.D.Shchukin, Kolloidn.Zh. 47 (1985) 1048. [40] S.S. Dukhin and J. Sjoblom, in: Emulsion and Emulsion Stability, J. Sjoblom (Ed.), Marcel Decker, 1996,41. [41] V.M. Muller, Kolloid. Zh., 40 (1978) 885. [42] V.M. Muller, in: 'Surface Forces in Thin Films", B.V. Derjaguin (Ed.), Nauka, Moscow, 1979,30. [43] S.Ljunggren, J.C.Eriksson and P.A.Kralchevsky, J.Colloid Interface Sci. 191 (1997) 424. [44] V.M. Muller, Colloid J., 58 (1996) 598. [45] M.Smoluchowski, Phys.Z., 17 (1916) 557, 585. [46] N.A. Fucks, Z.Phys., 89 (1934) 736. [47] B.V. Derjagin and V.M. Muller, Doklady Acad./Nauk SSSR, 176 (1967) 869. [48] H.Kihira, N.Ryde and E. Matijevic, J.Chem.Soc.Faraday Trans., 88 (16) (1992) 2379. [49] D.Snoswell, J.Duan, D.Fornasiero and J.Ralston, J.Phys.Chem., B, 107 (2003) 2986. [50] RJ.Pugh, Adv.Colloid Interface Sci., 64(1996)67. [51] N.Mishchuk, D.Fornasiero and J.Ralston, J. Phys. Chem., A. 106 (2002) 689-696. [52] S.A.K. Jeelani and S.Hartland, J.Colloid Interface Sci., 206 (1998) 83. [53] N.Buske, H.Hedan, H.Lichtenfeld, W.Katz and H.Sonntag. CoI.Polym.Sci. 258 (1980) 1303.
[54] [55] [56] [57] [58]
H.Sonntag, V.Shilov, H.Gedan, H.Lichtenfeld and C.Durr. Colloids Surf., 20 (1986) 303. N.A.Fuchs, The Mechanics of Aerosols, Pergamon, Oxford, 1964. M.A. Nawab and S.G. Mason, J.Colloid Sci.13 , 179, 1958. J.Goldman, R.G. Cox and H.Brennet, Chem.Eng. Sci., 22 (1967) 637. C.J. van Oss, R.M. Fike, R.J. Good and J.M. Reinig, Analyt. Biochem., 60 (1974) 242.
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 10
Hydrodynamical interaction of deformable drops A.Z. Zinchenko and R.H. Davis Department of Chemical and Biological Engineering, University of Colorado, Boulder, CO 80309-0424 1. INTRODUCTION For small emulsion drops with diameters in the range of 1-100 /im, the surface deformation is typically small, and much theoretical work has been done to date to investigate drop interactions with negligible shape distortion. Another relevant assumption for small emulsion drops is the neglect of fluid inertia on the microscale, making it possible to use simplified (Stokes) equations for the fluid motion. Exact semi-analytical solutions have been constructed to describe hydrodynamical interaction of two surfactant-free spherical drops at arbitrary surface separations and orientations, both in a quiescent liquid and shear flows [1-8]. These results allowed us to study the effects of drop interactions on the collision efficiency of Brownian and non-Brownian drops in gravity-induced and shear-induced motion, and on the non-Newtonian rheology of semi-dilute emulsions [4,8,911]. For 'well-mixed' concentrated emulsions of spherical drops, numerical multipole techniques were used to evaluate the effective properties (viscosity, sedimentation rate) [12]. These studies have been complemented by asymptotic analyses for two-drop interactions at small surface separations [4-5,13-15]. In particular, it was found that the lubrication resistance of the liquid film between two non-deformed drops with arbitrary viscosities is inversely proportional to the square root of the gap, and therefore this resistance does not preclude drops from coming into contact (in a finite time) even without van der Waals attraction. This finding is in stark contrast to solid-particle interaction; it is well known (e.g., [16]) for the latter case that the lubrication resistance of the thin film is inversely proportional to the gap, and so collisions of small solid particles are not possible
392
A.Z. Zinchenko and R.H. Davis
without attractive molecular forces, non-continuum effects, or microscale surface roughness. The present review focuses on the effects of surface deformation on drop interactions. While deformation is relatively unimportant for describing the behaviour of solitary and well-separated drops of small size (1-100 lira) in dilute emulsions, it may have a crucial effect, as it turns out, on the efficiency of drop coalescence and on the collective behaviour of highlyconcentrated emulsions. The reason is that the deformation of the thin film between two surfaces approaching each other is known to preclude drops from coming into contact, unless van der Waals forces become important [17-18], and so the coalescence efficiency of slightly deformable drops is a result of subtle interplay between the resistance of the thin film and short-range molecular attractions. The degree of deformation on the drop lengthscale is measured by the capillary number Ca ~ /ief//cr
,
where /j,e is the dynamical fluid viscosity outside the drops, U is the characteristic velocity of the flow relative to the drops, and a is the constant surface tension (the effects of surface tension gradients due to surfactants or temperature variations are not included here). When Ca
Hydrodynamical Interaction of Deformable Drops
393
useful for two or more drops with interactions. A relatively new element is multipole acceleration (Sec. 3.2), which, in particular, extends the capabilities of dynamical boundary-integral simulations to very large systems of drops (on the order of a thousand drops) with large deformations [23-25]. In Sec. 4, multipole-accelerated, boundary-integral calculations are used to obtain exact results for two drops in very close approach at Ch -C 1, and thereby to verify the asymptotic theory of drop coalescence of Sec. 2 and find its range of validity. In Sec. 5, we review some results of the boundaryintegral analysis [26] for two drops or bubbles with large deformations in buoyancy/gravity-driven motion. It is demonstrated how interactions give rise to cusped shapes, drop breakup and capture. The capture phenomenon is of particular interest; it was first discovered experimentally by Manga and Stone [27-28] for two bubbles in a highly-viscous corn syrup. In Sec. 6, multidrop simulations in concentrated emulsions with large interface deformations are discussed. Of particular interest here is the phenomenon of sedimentational instability due to deformations [23,25], which, we believe, is akin to Koch-Shaqfeh type of instability predicted previously for a dilute suspension of solid, rod-like particles [29]. Multidrop rheological simulations in a shear flow [24] are also discussed in Sec. 6; even a small amount of deformation is shown to make an emulsion non-Newtonian (at high drop volume fractions). As another application of the multipole-accelerated boundary-integral method, the problem of a large bubble or drop motion through a concentrated emulsion of deformable freely-suspended drops [25] is discussed in Sec. 6. It is shown how to calculate the steady-state settling velocity of the bubble/drop from first principles, by rigorous multidrop simulations, and thus to avoid difficulties with the constitutive equation for the emulsion. In Sec. 7, we discuss some unresolved issues and open areas for future research. 2. HYDRODYNAMICAL THEORY OF COALESCENCE OF SLIGHTLY DEFORMABLE DROPS Consider two Newtonian drops of non-deformed radii a\ and a,2, viscosity fjf (assumed to be the same for both drops, just for simplicity) and density p' moving in an unbounded Newtonian liquid of viscosity /ie and density pe. The Reynolds number, Re = peUai//j,e, for the flow relative to the drops is assumed to be small (which is the most practical case for emulsion drops) and so the Stokes equations for the fluid motion inside and outside the drops apply [30]: ^V 2 u = Vp ,
V •u = 0
,
(1)
394
A.Z. Zinchenko and R.H. Davis
where u and p are the fluid velocity and pressure, respectively. Two cases are of specific interest: (i) relative drop motion in a quiescent liquid under gravity/buoyancy (Ap = p' — pe ^ 0) and (ii) shear-induced motion of two freely-suspended drops (p' — pe), where the unperturbed flow around the drops has the form u^x) = (—7x2, 0, 0), with 7 > 0 being a shear rate. The capillary number is precisely defined as
for shear-induced motion, and a=^\vr-VT\
2(A+l)a,V~P.)g
p)
for gravity-induced motion. In Eq. (3), V?° is the settling velocity of an isolated spherical drop calculated by Hadamard-Rybchinski's formula [30], g is the gravity acceleration, and A = n'/ne
(4)
is the viscosity ratio. For Gi < 1, the problem of drop coalescence is best handled by matched asymptotic expansions. On the lengthscale of the drop size, the drops are considered as non-deformed fluid spheres, and they can reach contact in a finite time because the hydrodynamic resistance coefficient for mutual approach has an integrable singularity ~ /i~1//2 [4,13,15], as the surface clearance h -» 0. On the deformation lengthscale, however, there is a small non-uniform gap between the drops that evolves under the action of hydrodynamic and short-range, singular molecular forces; inclusion of molecular forces such as van der Waals attractions is imperative to make two deformable surfaces come into contact and, thereby, coalesce. It was shown in Ref. [20] that, at least, at not too high viscosity ratios A, relative tangential motion of drops in close approach is practically unaffected by small deformation, which allows us to calculate the time-dependent 'contact' force acting along the line-of-centres, from the 'outer' solution for spherical drops in apparent contact. This 'contact' force is equal to the lubrication force, which closes the local integro-differential equations for the thin film in the gap. Solving these thin film equations (see below) allows us to determine whether the physical contact of the deformed surfaces (which is identified with the onset of coalescence in the present approach) is reached for given initial conditions, or if the drops instead move past each other and separate without coalescing.
Hydrodynamical Interaction of Deformable Drops
395
To demonstrate this methodology, consider first gravity/buoyancydriven motion. Due to linearity of the Stokes equations (1), hydrodynamic forces acting on spherical drops 1 and 2 in apparent contact can be written as
F i = - 6 7 ^ 1 [A[2V + Ttn{Vl - V 2 ) x + T{2V^\ + F F2 = - 6 7 ^ 0 3 [A22V + Tt2l{V2 - V^
+ Tl2Vi] - F
,
(5)
where Vi and V2 are the velocities of geometric centres of the spheres, V is the common velocity of the drops along the line of centres, and F and — F are contact forces acting along the line of centres; _l_ denotes the vector components perpendicular to the line of centres. The non-dimensional hydrodynamic resistance coefficients, A\2 and A22, are related to the forces acting on drops in their motion as an aggregate along the line of centres; these coefficients depend on size ratio k = a\/a2 and viscosity ratio A, and are known from the exact solution of Ref. [31]. Likewise, the nondimensional coefficients T^Ak, A) are known from the exact solution [3] for the two-drop motion perpendicular to the line of centres in the limit of touching. Projecting the force balances Fi + |?raf Apg = 0 on the line of centres and excluding V yields the expression for the contact force: 4 (fc + 1)3
,
A.U-k2AL
where /3 is the angle between the centreline and the vertical (Fig. 1), and R = a\a2/(ai + a2) is the so-called 'reduced radius.' Similarly, projecting the force balances on the plane orthogonal to the centreline gives the equation for the relative tangential motion in apparent contact: M-K1
sm fJ ,
- - ^
+ a 2
^ ,
-TtiTt2
+ TtiTt2
•
W
Integrating Eq. (7) yields the contact force F(/3(t)) as an explicit function of time, given the initial conditions for (3{t) (see below). The values of the non-dimensional coefficients a(k, A) and Tl(k, A) are given in Table 1. For two freely-suspended drops in a shear flow uoo(x) = (—7x2,0,0), the centre-to-centre unit director p = {pi,p2,Ps) from sphere 2 to sphere 1 is characterized by two angles /3 and 9 (Fig. 2): Pi = sin 9 cos (3 , p2 = sin 9 sin (i , Pz = cos 9 .
(8)
Similar algebra gives the contact force: F(/3,9) = 37r/ueai(oi + a2)D*j sin2 9 sin/? cos /3
,
(9)
396
A.Z. Zinchenko and R.H. Davis
Fig. 1. Gravity-induced motion of two drops in apparent contact. The close-up shows the inner region with the deformed thin film.
Table 1
The values of a and T ( for gravity-induced motion k 0.10
A = 0.1 a = 72.870 T* = 0.713 0.15 a = 50.291 Tl = 0.728 0.25 a = 31.323 Tl = 0.733 0.35 a = 22.246 T* = 0.709 0.50 a = 14.162 T ( = 0.623 0.75 a = 5.792 T ( = 0.371 0.90 a = 2.118 Tl = 0.162
1 2 0.5 0.25 69.241 64.300 57.158 48.460 0.616 0.509 0.390 0.282 47.989 44.861 40.348 34.874 0.316 0.635 0.533 0.419 30.098 28.434 26.039 23.148 0.357 0.648 0.555 0.451 21.490 20.463 18.984 17.201 0.632 0.547 0.453 0.369 13.758 13.207 12.412 11.453 0.560 0.490 0.412 0.343 5.653 5.463 5.188 4.856 0.212 0.335 0.295 0.251 2.069 2.002 1.905 1.788 0.146 0.129 0.110 0.093
where g . = A + | D{K\2 + D\K\2 A+ 1
A22 + &A'i2
4 39.590 0.201 29.323 0.237 20.239 0.285 15.415 0.304 10.495 0.291 4.524 0.183 1.671 0.081
10 29.720 0.133 23.188 0.168 17.059 0.220 13.477 0.247 9.462 0.245 4.167 0.158 1.546 0.070
(10)
Hydrodynamical Interaction of Deformable Drops
397
Fig. 2. Shear-induced motion of two drops in apparent contact.
and the hydrodynamic coefficients D\ and D\ are obtained from the exact solutions [5,8] in the limit of touching drops. The dynamics of contact motion for two touching spherical drops in simple shear flow is described by
— = - 7 ( 1 - £ f )sin0cos0sin/3cos/3 at
,
where B* is the value of hydro dynamic relative mobility function in the zero-gap limit, found from the solutions [5,8]- Again, solving Eq. (11) with given initial conditions (see below) gives the contact force F({3(t),6(t)) as a function of time. The values of the non-dimensional coefficients D*(k, A) and Bt(k, A) are given in Table 2. Even though relative drop motion is three-dimensional for both gravitational and shear flows (Figs. 1,2), it was proved in Ref. 20 that the near-contact deformation (shown as a close-up in Fig. 1) is mainly axisymmetrical for Ca
398
A.Z. Zinchenko and R.H. Davis
Table 2
The values of D* and Bl 0.5 A = 0.1 0.25 k 0.538 0.545 0.547 0.1 0.057 0.127 0.213 0.15 0.734 0.749 0.762 0.047 0.103 0.174 1.050 1.088 0.25 1.015 0.032 0.072 0.122 1.237 1.299 0.35 1.185 0.024 0.053 0.090 1.368 1.456 0.50 1.299 0.037 0.064 0.016 1.280 0.75 1.358 1.459 0.011 0.026 0.046 1.227 1.321 1.154 1.00 0.010 0.024 0.042
1 0.539 0.325 0.767 0.267 1.129 0.189 1.376 0.141 1.570 0.101 1.596 0.074 1.450 0.069
Normal stress balance A (I 1\ p
-6^
=a
+
{^
2 0.515 0.445 0.756 0.369 1.160 0.265 1.450 0.200 1.694 0.146 1.750 0.109 1.596 0.101
4 0.476 0.553 0.728 0.466 1.174 0.341 1.509 0.262 1.803 0.194 1.894 0.147 1.733 0.137
10 0.421 0.664 0.682 0.571 1.172 0.432 1.555 0.339 1.906 0.256 2.035 0.197 1.870 0.184
a d ( dh\
,1rt.
a-2)-Yrdr- V Yr)
(12)
'
Momentum balance k f-~2dr~
}
9P
(13)
{6)
'
Local boundary integral u(r,t) = -, /o°° >(r',r)f(tS)dr'
,
Mass continuity dh 1 d r , 1 —+ - — 77m=0
h2
dt r dr L J Integral force balance
, '
r(p-fi^i)2'rr*=F
u=u
•
(14) dp -£-
12/xe dr
,
(15)
y
'
(i6)
In Eqs. (12) and (16), A is the Hamaker constant, and — A/(6irh3) represents the disjoining pressure term for unretarded van der Waals attraction (other colloidal forces, such as electrostatic, are not considered herein). The analysis can be easily extended to retarded van der Waals force expressions, if needed. The boundary integral (14) can be written in an equivalent form:
f(t, r) = V f
(17)
Hydrodynamical Interaction of Deformable Drops
where ,, , s 1 ^(r',r =
r'
/-Trr r« / 1
2rr'cos6)i-i/2 n n = 5cos9d9
399
. . 18
is the elliptic-type Green function. Relation (15) accounts for the parabolic velocity profile in the film and generalizes the well-known lubrication equation of Hocking [32] for mobile films (u ^ 0); u(t,r) is the average fluid velocity across the film. Infinite limits of integration in Eqs. (14) and (16) are justified, since the film radius becomes infinitesimally small, as Ca —> 0. Most importantly, the contact force (6) or (9) from the outer solution enters the integral balance (16) for the film as the lubrication force. The thin-film system of equations (12)-(17) (or its simplifications for immobile or non-deformable films, constant driving force F, or no van der Waals attractions) have been derived and used by a number of authors [33,15,17,18,34] to study head-on collisions. In Refs. [20,21], film drainage equations (12)-(13), (15)-(17) were combined with the outer solution equations (6)-(7) (or (9), (11)) to describe three-dimensional coalescence in gravity- or shear-induced motions, but without the pressure-gradient term in Eq. (15), which has limited the theory in Refs. [20,21] to moderate viscosity ratios A
•
exp /
^—
'- dr\
,
(19)
where r is the centre-to-centre distance, and the mobility functions L and M, for instantaneous relative motion along and normal to the line of centres [4, 35], respectively, can be taken from exact two-sphere solutions [4]. For a trajectory analysis in shear-induced motion, see Refs. [8, 9, 21].
400
A.Z. Zinchenko and R.H. Davis
Given (3C (or (3C and 9C), the simplest matching strategy [20], suitable for moderate viscosity ratios A, is based on calculating the 'collision' time required for a spherical drop trajectory to reach contact from an initial configuration where the drops are slightly separated. The hydrodynamical lubrication force between two non-deformed drops in close contact has the form [4, 13]
F « - | f f V2V#3/2%^S
,
(20)
where /i min is the minimum gap between the drop surfaces; for nondeformed drops the minimum separation is located along their line of centres. On the other hand, this force is approximately equal to the contact force (6) or (9), and (3 K, (3C (Q J=S QC) during the mutual approach of spherical drops from a nearly touching configuration. These arguments give the collision time for spherical drops by integrating Eq. (20): tcoll
= I 7T3V2 n'R^ (h°mmf2/F((3c)
(21)
(for shear-induced motion, F(f3c) is replaced with F(/3C,8C)), where h°min is the initial gap at t = 0 between the spherical drops. In all cases of interest, short-range van der Waals attraction has no appreciable effect on matching (see Ref. [20]) and it was neglected in Eq. (21). In the spirit of the asymptotic method, h°min is subject to the limitations h°min
,
t =0
,
(22)
which corresponds to local parabolic approximation of nondeformed shapes. Accordingly, the thin-film equations (12)-(17) are solved first from t — 0 to t — tcoii with the initial condition (22) and a constant driving force F((3C) (or F((3C,8C)), and then, for t > tcou, the time-dependence of the contact force through Eq. (7) or (11) is taken into account. The exact choice of the initial separation h°min does not affect the overall solution in the range of interest t > tcou (since, for most of the time during t < tcou, deformation is unimportant). This simplest matching strategy based on Eq. (21) is asymptotically correct for Ca -> 0 (except for the special case of sliding collisions, when F(fic) m 0 or F((3C,8C) & 0, which have a negligible probability in random emulsions), but it is limited to moderate viscosity ratios A
Hydrodynamical Interaction of Deformable Drops
401
Refs. [20, 21]. A more general matching condition [19] not subject to this restriction takes into account that, during an initial approach of highly viscous drops, the angles (3 and 9 and the contact force may have substantial variations due to slow film drainage. In addition, a more general lubrication form [14-15], suitable for A > 1, is used instead of Eq. (20): w -67r/i e - — $(p) — — , p = AJ ——, (23) h-min at N where $ ( p ) has an exact expression [14]:
H
^ ^ f ^ n + i ^ i + i ^ + e]
(24)
'
The lubrication form (20) is the limiting case of (24) for p -C 1. Equating (23) to the contact force (6) or (9), and integrating backwards from Kiin = 0 to hmin = h°min, yields
<25>
*""*'=£ IT* • where
^
*(p)
-
n(n + 1)
F
(2n + l)Po]
and po = X(2h^nin/R)1^2\ the contact force F and d^/di in Eq. (25) are taken from equations (6)-(7) or (9)-(ll). Equation (25) relates the collision angles (3C (or f3c and 0c) to the initial angles (30 (or (30 and 90) at t = 0. In particular, for gravity-induced motion,
sin/30 = s i n / 3 c e x p [ - ^ ^ l [ aApgR J
.
(27)
For relative centre-to-centre trajectories in the plane 9 = n/2 in shearinduced motion, Eq. (25) yields ! - ( ! - £ ' ) cos 2/?e = 8fcI(l-g') 1 - (1-Bt)co82/30 (l + k)3D*
, , '
{
(for general trajectories, 9 ^ TT/2, it is also possible to relate 9O, (30 to 9C, (3C through the second equation (11), but the algebra becomes more involved). Knowing /3O (or (30,90) and the initial film profile (22) allows us to start integrating the thin-film equations together with the contact-motion equations. Again, the choice of initial separation h°min has no effect on the solution in the range of substantial film deformations (for /i/a8- < O(Ca)).
402
A.Z. Zinchenko and R.H. Davis
For AGz1/2 < 1, the matching strategy (25) and that of Refs. [20, 21] yield practically identical results, and both do not contain any adjustable parameters; for \Call2 > 0(1), however, the matching rule (25) is the one to use. The numerical solution of thin-film equations (12)-(17) is greatly complicated by numerical stiffness, so that the simplest, explicit algorithms require extremely small time steps (especially at the initial stage of film deformation), and are very impractical for systematic coalescence efficiency calculations (when these equations need to be solved many thousand times). Far more efficient, semi-implicit, absolutely stable algorithms have been developed recently; see the original works [19, 20] for details. To demonstrate how the theory is used to calculate the coalescence efficiency, consider the case of two drops in gravity-induced motion with size ratio k — a\/a,2 = 0.5, viscosity ratio A = 1, Ca = 2.2 x 10~3, and the non-dimensional Hamaker parameter 5 = 1.1 x 10~3, where
Fig. 3. Dynamics of the minimum surface separation for two trajectories close to critical.
Fig. 3 presents the minimum surface separation, hmin/a2, scaled with the larger radius 02, versus the angle (3, for two different collision angles /3C = 1.035 rad (solid line) and 1.040 rad (dashed line). The minimum separation is located along the line of centres, or at the rim when a dimple forms. When (3C — 1.035 rad, the van der Waals attraction has enough time to pull the surfaces together, leading to coalescence. In contrast, for a slightly
Hydrodynamical Interaction of Deformable Drops
403
higher value of j3c = 1.040 rad, the strength of the van der Waals attraction is insufficient for coalescence, the drops reach minimum separation of hmin/o-2 — 7.6 x 10~6 at (3 « TT/2 and eventually separate. The critical value (3fu in the range 1.035 - 1.040 is slightly different from f3?it m 0.99 found in Ref. [20] for the same conditions, simply because we used in the present calculations much larger cutoff radii r max in the numerical solution of thin-film equations (12)-(16) for full numerical convergence. Fig. 4 shows the film profile evolution for a slightly supercritical value of pc = 1.040. The profiles are plotted in non-dimensional variables [20]: h = hR/b2
,
r = r/b
,
(30)
where b is defined by irb2a/R = ApgR3 and is a measure of the film radial extent. At /3 — 1.101, the film just starts to dimple. Subsequently, the dimple radius decreases with increasing (3. At (3 = 1.712, when the surface clearance hmin has reached its last local minimum (Fig. 3), the dimple disappears altogether, causing very rapid drop separation thereafter.
Fig. 4. Dynamics of the film profile for f3c = 1.04 rad.
Once the critical collision angle f3^u is found by trial-and-error (which typically requires extensive calculations and an efficient thin-film algorithm), Eq. (19) may be used to determine the corresponding critical offset parameter d^\ separating coalescence (c?oo < d^11) and noncoalescence (doo > d^f). The coalescence efficiency is defined as / dcrit •^12 =
\2 •
(<Jl)
404
A.Z. Zinchenko and R.H. Davis
Most importantly, knowing E\i allows us to solve the population dynamics equation (e.g. [36-38]) and predict the evolution of the drop-size distribution in dilute emulsions due to coalescence. It was found [20] that there is a narrow 'second coalescence zone' for collision angles j3c close to n/2, but its contribution to the coalescence efficiency £12 is typically very small, so Eq. (31), assuming a single coalescence zone d^ < d^lt, is quite adequate for practical purposes. Figure 5 shows the coalescence efficiency for slightly deformable (solid curves) and spherical (dashed lines) drops with Ca — 0.001 and A = 1 as a function of the non-dimensional Hamaker parameter S. Figure 6 presents E12 as a function of the capillary number at k = 0.5 and 8 = 4 x 10~4. It can be predicted from Fig. 6, for example, that deformation decreases the coalescence efficiency of two drops with a\ = 262 /xm, a-i = 525 /im, a = 10 dyn cm"1, A = 0.5, Ap = 0.1 g cm"3, and A = 10~14 erg (5 = 4 x 10"4, Ca — 5.8 x 10~3) by about five fold. The sharp dependence of S on the drop size (see Eq. (29)) is the reason for steep behaviour of Eyi in the transition range from spherical to deformed drops. Similar calculations can be performed for shear flow [21], although they are technically more involved, since the upstream interception area for deformable drops is no longer a circle, and two collision angles, j3^%t and #£"*, have to be found by trial-and-error. The collision efficiency, E\%, is defined as En = Jn/J°12
,
(32)
Fig. 5. Coalescence efficiency for gravity-induced motion at Ca = 0.001 and A = 1. Reproduced by permission from Ref. [20].
Hydrodynamical Interaction of Deformable Drops
405
Fig. 6. Coalescence efficiency for gravity-induced motion at k = 0.5 and S = 4 x 10 4 . Reproduced by permission from Ref. [20].
where J\2 is the flux of pairs through the upstream interception area, and Jf2 is the corresponding Smoluchowski's value [40] assuming the neglect of drop interactions, shape distortions and molecular attraction. Figure 10 from Ref. [21] shows, for example, that the deformation effect on the coalescence efficiency of two drops of ethyl salicylate in diethylene glycol for shear flow with k = 0.5, 7 = 1 s"1, a = 1.9 dyn cm"1, A = 0.1, /j,e = 0.35 g cm"1 s"1, A = 5 x 10~14 erg becomes significant when the average radius a = (ai + 02)/2 exceeds 300 /zm. At a = 300 /xm, the capillary number (2) is 0.005, which is expected to be within the range of validity of the asymptotic theory, even for this small A = 0.1 (see Sec. 4). For A = 0(1), it is also possible to perform scaling for the range of drop sizes where deformations have a significant effect on shear-induced coalescence. For two drops with A = 0(1), coming into 'apparent contact' without van der Waals attraction in a general 3D collision, the minimum surface separation h along the trajectory scales like h ~ RCa4^ (Sec. 4). Roughly, if the molecular force Fw ~ AR/h2 becomes competitive with hydrodynamic forces FH ~ HeiR2 at this separation, the drops will coalesce. Substituting the definition (2) of the capillary number Ca into the criterion Fw ~ FH yields r A5rr8 1 1/17
*~&]
(33)
In deriving this scaling, we neglected all non-dimensional numerical factors that depend only on k and A.
406
A.Z. Zinchenko and R.H. Davis
3. COMPUTATIONAL METHODS FOR INTERACTING DROPS WITH ARBITRARY DEFORMATIONS 3.1 Boundary-integral Method For drops with finite deformations (Ca = 0(1)), the asymptotic method of Sec. 2 is obviously not valid, and so theoretical analyses in this case rely more heavily upon numerical calculations. If the Reynolds number is still small, typically requiring a highly viscous matrix fluid, the boundaryintegral method is most appropriate. This method was pioneered by Rallison and Acrivos [22] for one-drop calculations, and since then it has received considerable development, including binary interactions [19, 26-28, 41-42], and multiple drops in a periodic box [23-25, 43-44] to simulate locally homogeneous emulsions. The essence of this method is to reduce the Stokes equations (1) in the fluid domains inside and outside the drops to a system of integral equations for the fluid velocity u on drop surfaces only, using the reciprocal theorem for Stokes flows [30] and the Green function (which corresponds to the Stokes flow generated in an unbounded fluid by a point force). In particular, for two drops moving in an unbounded liquid, this system of boundary-integral equations takes the form u(y) = 2K £ / ti(a:) • T(x - y) • n(x)dSx + F(y) 0=1
,
(34)
•/6<3
where K — (A — 1)/(A + 1), Sp (/? = 1, 2) is a drop surface,
T(r) = £ - ^
(r = x - y)
(35)
is the stresslet third-rank tensor corresponding to the free-space Green second-rank tensor (with unit viscosity):
and n(x) is the outward unit normal at x G Sp. The inhomogeneous term F can be written as Oil
ltl\
F{y) = ^ff 1+ A
9
+ -jf-VS
2
r
£ /, /(»)»(*) • G(x - y)dSx
, (37)
He{± + A) p=iJSl3
where u^iy) is the unperturbed flow (which would exist at point y in the absence of drops), f{x) - 2ak(x) + (Pe - p')gz
,
(38)
Hydrodynamical Interaction of Deformable Drops
407
k(x) = \{k\ + k,2) is the mean surface curvature at x (ki, ki being the two principal curvatures of the surface), g is the gravity acceleration, and z is the Cartesian coordinate in the direction of gravity. The first term on the RHS of Eq. (38) is related to the stress jump across the interface, since 2ak is nothing but the Laplace pressure caused by the surface tension. Knowing the fluid velocity u on drop surfaces from the solution of Eq. (34) at each time step allows us to advance the drop shapes (see below). For matching viscosities A = 1, Eq. (34) yields an explicit expression u(y) = F(y), which greatly facilitates calculation; otherwise, Eq. (34) must be solved by iterations. The most popular way is that of 'successive substitutions' of the previous iterate u^ into the right-hand side of (34) to get a new approximation «("+1) until such iterations converge. However, K = ±1 are known to be the spectral values for Eq. (34) (see, e.g., [45-46]), so that the homogeneous system (34) (i.e., with F = 0) has non-zero solutions u at K = ±1. As a result, for extreme viscosity ratios A < 1 or A > 1, when \K\ « 1, the system (34) is ill-conditioned and successive substitutions are very poorly convergent. A procedure called 'Wielandt's deflation' (e.g., [4546]) eliminates K = ±1 from the spectral values at the expense of slightly modifying the boundary-integral equation (34), which greatly accelerates iterations for well-separated drops with A < 1 or A > 1. For drops in close approach, however, the spectrum of K-values for Eq. (34) becomes nearly continuous, and so elimination of just marginal values n = ±1 from the spectrum by deflation does not alleviate all difficulties for extreme A; it is better to use in this case more sophisticated, biconjugate-gradient iterations of Lanczos [42], after deflation. For numerical solution of Eq. (34) (or its deflated version), the drop surfaces must be discretized. There are two general approaches to surface discretization: (i) global parametrization (e.g., [46-47]) and (ii) unstructured triangulation. The first approach (with a clear analogy to globe parametrization by a system of longitude and latitude lines) is losing popularity and is now considered generally inferior to the second method, primarily due to the presence of coordinate singularities (poles) in the global parametrization approach, with a negative effect on numerical stability. In dynamical simulations, drops often start from spherical shapes. Unstructured, almost uniform triangulations of a sphere can be constructed by simple techniques. In the first scheme [45], each face of a regular icosaedron inscribed into the sphere is divided into four triangles, the new vertices are projected radially on the sphere, and the process is repeated as many times as necessary, giving discretizations with N& = 20 x Ak (k = 0,1,2,...) triangles;
408
A.Z. Zinchenko and R.H. Davis
Fig. 7 shows an example for N& = 1280. The second scheme [42] differs only in that we start from a regular dodecaedron, project each pentagon face centre radially on the circumscribed sphere, connect the projection with the face vertices, and then proceed as in the first method; this scheme produces triangulations with N& — 60 x 4k [k = 0,1,2,...). Subdivision of each triangle face into m2 smaller triangles [43] may also be practiced; when the latter procedure for small m < 5 is combined with the first two schemes, it gives additional possibilities NA = 720, 1500, 2160, 2880, 6000, 6480, etc., but still with highly uniform unstructured triangulations (the maximum-to-minimum mesh edge ratio being within 1.19-1.22).
Fig. 7. An example of surface triangulation into 1280 elements starting from a regular icosaedron.
When drops move (and deform), one can advect the mesh nodes either with the interfacial fluid velocity u or with the normal velocity (u • n)n (both found from the solution of Eq. (34)), to update the drop shapes. A familiar difficulty with both strategies, making them unsuitable, is that the internode distances become highly irregular, invalidating the mesh after a short simulation time. One remedy [43] is to add an artificial tangential velocity to the node motion (which has no effect on shape evolution, for sufficiently small time steps); this additional velocity is constructed by some local rules to prevent mesh degradation for a long simulation time, based on the idea of internode 'springs'. Unfortunately, in the original form, this idea of 'grid tension' leads to numerical stiffness, with tight stability limitations on the time step. A more advanced, adaptive restructuring method [48] combines a dynamical spring-like mesh relaxation (performed iteratively) with topological mesh transformations (node addition/subtraction/reconnection). The latter approach is very flexible and was used, in particular, in drop pinch-off simulations [48]. With isotropic mesh restructuring into compact elements, however, it is not easy to reach convergence (i.e., independence of the global results from triangulation) for elongated shapes, because adequate azimuthal resolution requires a
Hydrodynamical Interaction of Deformable Drops
409
very large total number of boundary elements as a drop stretches. A very different approach to mesh control called 'passive mesh stabilization' [42, 26, 23-25] uses fixed topology triangulations (i.e., with a fixed number of nodes and fixed connections) and seeks to prevent mesh degradation by constructing an additional global tangential field on each surface Sa separately from the solution of a variational problem. These additional node velocities are required to minimize, in some sense, the 'kinetic energy' of disordered mesh motion (as opposed to 'potential energy' in the simplest grid tension method), thus avoiding excessive numerical stiffness. Suitable forms of the kinetic energy function were found both for compact [42] and highly stretched [26] shapes. This method provides less control over the mesh than does adaptive restructuring [48]. However, it was found to be surprisingly flexible in a variety of problems, including dimpled shapes and highly stretched drops closely approaching breakup [26]; most importantly, it was possible to systematically demonstrate, with passive mesh stabilization, the independence of the results (global shapes, dynamics of thin neck thinning, etc.) from the degree of triangulation even for a modest number of boundary elements [26]. The results for finite deformations discussed below are all obtained using passive mesh stabilization. Local surface fitting by a paraboloid has become a popular method to calculate the curvatures k(x) and normals n(x) in the nodes of an unstructured method. The first parabolic fitting algorithm is, probably, due to Rallison [49]; a much simpler, more general and efficient version was subsequently developed [42]. Numerical discretization of the singular integrals (34) and (37) by a trapezoidal rule [49] (see [42] for details) is preceded by singularity and 'near-singularity' subtractions [43] (with u(x*) and f(x*) subtracted from u(x) and /(as), where a;* is the mesh node on Sp nearest to y); the latter greatly improves the quality of the numerical solution. For gas bubbles or low-viscosity drops (A
410
A.Z. Zinchenko and R.H. Davis
A remedy allowing us to continue the calculation beyond the singularity formation is the 'curvatureless' form [26] of the boundary integral (37):
fSfik{xMx)-G{x-y)dSx = ±fSfl{l-*^^}dSx
. (39)
The right-hand side of Eq. (39) (understood as a principal-value integral) contains only normal vectors n(x) much less sensitive to discretization errors than is k{x), and eliminates the difficulty with the concentrated capillary force. However, the curvatureless form (39) is more singular at r —>• 0, and requires special desingularization techniques [26]. Using the curvatureless form is also crucial for successful drop breakup simulations, when a fixed-topology mesh is used (allowing boundary elements to stretch); examples are given in Sec. 5. Besides binary interactions in an unbounded medium (relevant to dilute emulsions), as described by Eq. (34), it is of great interest to simulate multidrop interactions with periodic boundaries. Namely, a random system of TV ^> 1 drops with centres in a box V is assumed to be continued triply-periodically into the whole space. The idea of periodic continuation is borrowed from statistical physics (e.g. [51]) and allows us to accurately simulate the effective properties of locally-homogeneous concentrated emulsions away from the boundaries with a limited number of drops N. For emulsion sedimentation (Sec. 6.2), the periodic box V is stationary, and can be taken as a cube (Fig. 8). For emulsion shear flow (Sec. 6.1), the box V is initially a cube, but it is then skewed by the flow, until the whole system repeats itself in a cyclic manner [43] (Fig. 9). The main modification to the boundary-integral equation (34) is in using Hasimoto's [52] periodic Green function. The details are involved, however, and the reader is referred to the original papers [23-25]. 3.2 Multipole Acceleration of Boundary-integral Calculations Direct point-to-point calculation of all boundary integrals (34) and (37) has 0(N2N^) computational cost per time step, which severely limits the number of drops N and/or the number of triangular elements N& per drop possible in dynamical simulations, especially for contrast viscosities (A
Hydrodynamical Interaction of Deformable Drops
411
Fig. 8. A schematic for multidrop sedimentation with periodic boundaries (a 2D sketch, not to scale). Surfaces S'p and S'p are periodic images. Reproduced by permission from Ref. [23].
Fig. 9. A schematic for multidrop simulation in shear flow with periodic boundaries. The non-dimensional periodic box V (contoured bold) is based on the lattice vectors ei, e^, ez\ initially V is a unit cube based on ei, e\, e%. When the strain 7 = 7* reaches 1/2, the periodic boxes are restructured for the first time, to avoid excessive skew. The third dimension is not shown. Reproduced by permission from Ref. [24].
412
A.Z. Zinchenko andR.H. Davis
and resolution N& ~ 103. In this scheme, elongated drops are first sliced into compact blocks (if a drop is compact, it coincides with its only block), and minimal spherical shells (V) are constructed around each block (Fig. 10).
Fig. 10. Multipole-accelerated calculation of near-field interactions. The contributions of block By (surrounded by shell Z>7) to the boundary integrals for y, y' and y" are evaluated, respectively, by (i) reexpansion of Lamb's singular series from x° to x°s, (ii) pointwise calculation of Lamb's singular series, and (iii) direct point-to-point summation. Reproduced by permission from Ref. [23].
The periodic Green function is partitioned into the free-space and 'farfield' part. Each block contribution to the free-space part of the boundary integrals is expanded as Lamb's singular series [30] (which may be viewed as an expansion in inverse powers of the distance from the block centre). Interactions between well-separated blocks (e.g., B1 and Bs, with shells P 7 and V$, respectively) are handled by Lamb's series reexpansions from singular to a regular form (i.e., in positive powers of the distance from the centre x°6). If the shells V^ and V1 overlap, but mesh node y' of block B% is 'well outside' shell X>7, Lamb's singular series for block B1 is used directly to calculate this block's contribution to the boundary integrals for node y' (Fig. 10). Only if node y" of block B( is inside shell £>7 (or is outside, but
Hydrodynamical Interaction of Deformable Drops
413
too close to X>7), should we use direct summations to calculate the contribution of block B7 to the boundary integrals for node y" (Fig. 10). The far-field contributions are handled by Taylor expansions about block centres. Thus, the essence of the algorithm is to use efficient multipole/Taylor expansions as much as possible, and employ direct summations only as the last resort. Technical details are quite involved, though, and the interested reader is referred to the original papers [23-25]. For N — O(102 — 103) and NA ~ 103, the computational gain of this algorithm over the standard summation is 2-3 orders of magnitude at each time step, which has made such large long-time simulations feasible [23-25]. Perhaps contrary to expectations, 3D boundary-integral simulations for only two drops in close approach is a very difficult problem, when the capillary number is small and the drops are nearly non-deformed. One reason is a very tight stability limitation on the time step stemming from the Courant condition. Another difficulty is localization of stress, requiring very high resolution in the narrow gap between two drops. Less obviously, the outer region (away from the gap) also needs unlimited resolution as Ca —> 0; the reason is that, in the non-dimensional form of (37), the curvature deviation (of order O(Ca)) from a uniform value is divided by Ca, to produce an 0(1) effect. As a result, for Ca
414
A.Z. Zinchenko and R.H. Davis
Fig. 11. Node partition into non-overlapping 'patches' in multipole-accelerated calculations for two drops with high resolution.
4. VALIDATION OF THE SMALL-DEFORMATION THEORY OF COALESCENCE The asymptotic theory of coalescence (Sec. 2), which is based on matching the thin-film solution in the gap with the outer solution for spherical drops, avoids formidable numerical difficulties inherent in boundaryintegral simulations for Ca
Hydrodynamical Interaction of Deformable Drops
415
an asymptotic film thinning to arbitrarily small separations. It is therefore of great interest to study if the small-deformation theory of Sec. 2 is indeed a correct approach to non-axisymmetrical interactions and to find its range of validity; in particular, it is of interest to see if (and when) the neglect of the pumping flow is an accurate assumption. Multipole-accelerated 3D boundary-integral calculations for two drops with very high resolution [19], albeit costly, are able to provide exact results for comparisons. Examples below are for two equal drops of radius a freely suspended in a shear flow (gravity-induced motion of unequal drops is additionally discussed in Ref. [19]). For simplicity, van der Waals attraction is neglected, and only centre-to-centre trajectories in the plane 6 = n/2 (Sec. 2) are considered.
Fig. 12. Shear-induced motion of two drops at A = 1 and Ca = 0.01. Close-ups at t = 5.5 and 6.5 show the gaps of 3.2 x 10~3a and 7.3 x 10~4a, respectively.
Fig. 12 presents snapshots of relative motion for A = 1, Ca = /j,eja/a = 0.01, and initial centre-to-centre offsets Ay0 = 0.7a (across the flow) and Ax 0 = 5a (along the flow), using a gap-adaptive 'projective' mesh [19] with N& = 34560 elements per drop; time is scaled with j 1. Spherical drops
416
A.Z. Zinchenko and R.H. Davis
would collide at (3C « 0.55 rad. In contrast, there remains a very small, but non-zero gap (reaching 7 X 10"4a) during the apparent 'contact' motion, and the drops eventually separate. This simulation was repeated for varying capillary numbers, to compare the surface clearance hmin{(3) with the asymptotic theories. In Fig. 13a,b, for Ca = 0.04 and 0.01, respectively,
Fig. 13a,b. Comparison of exact and asymptotic results for the surface clearance at A = 1.
the solid lines are from fully-convergent boundary-integral calculations (for Ca = 0.01, we had to use N& — 138240). Short-dashed lines are from the asymptotic theory [20] (which neglects the pressure gradient term in Eq. (15), assuming A = 0{Call2\)
Hydrodynamical Interaction of Deformable Drops
All
Fig. 14. Comparison between the exact (squares) and asymptotic (crosses, A = 0 theory) values of the separation angle for A = 1.
„_ r
~ _ hR
r
~n
'
p=—p
- _ Qa(t - t0)
TJ2 ' *~
2i?y
, u = —f— , f = -^—
,
40
where
Then, the thin-film equations (12)-(16) (without van der Waals attraction and the pressure gradient term in (15)) can be written as P
r dr {dr)
'
J
u(i,r) = f§°
r')df'
dh
r°°
1 d , " .
2 dr
,
-
/^ \
^ - + - -W2 (rhu) = 0 , / prdr = -t . 42 Equations (42) are universal and do not contain any parameters. The separation time isep must be therefore 0(1). Using Eqs. (9), (11) to evaluate F'(t0), together with Eqs. (40)-(41) and (2), one can find that isep = 0(1) corresponds to f3sep — TT/2 = O(QJ 1 / / 3 ), which is in agreement with the asymptotic curve in Fig. 14 and shows, indeed, an extremely slow approach
418
A.Z. Zinchenko andR.H. Davis
Psep —>• vr/2, as Ca —> 0. The same universality arguments can be used to find the scaling [19] 0{a Ca4^) for the minimum surface separation along the trajectory, which was used in Sec. 2. While, for A = 1, the A = 0 theory is preferred, the situation is reversed for highly-viscous drops. Figure 15a,b presents the non-dimensional
Fig. 15a,b. Comparison of exact and asymptotic results for the surface clearance at A = 4.
surface separation hmin/a for A = 4, Ay0 = 0.7a, Ax0 = 5a, and two capillary numbers Ca = 0.02 and 0.01. Fully convergent boundary-integral calculations (solid lines) are much closer to the predictions of the A ^ 0 theory (long-dashed lines) than to the results by the A = 0 theory (shortdashed lines). Both theories, however, give similar errors for (3sep (but of opposite signs), which slowly disappear, as Ca —> 0; for Ca = 0.005, the exact result (3sep = 2.06 rad is only slightly different from the prediction (3sep = 2.03 rad by the A / 0 theory. Similar calculations (Fig. 16a,b) for A = 10, Ax 0 = 5a, and Ay0 — 0.5a show that, in this case, the A ^ 0 theory becomes accurate for hmin{f5) at Ca < 0.01, and gives only a modest error for the separation angle (this error in (3sep further decreases to 1.2% at Ca = 0.005); the A = 0 theory was totally unsuccessful for A = 10, for obvious reasons. Interestingly, the behaviour of the separation angle at A = 10 is even farther from the naive expectation (3sep m ?r/2 than for A = 1: (3sep varies only from 2.40 to 2.16, as the capillary number is reduced from 0.04 to 0.005; to observe f3sep —>• TT/2, much smaller Ca would be required. Loewenberg and Hindi [41] predicted, by qualitative arguments without a detailed solution, that the minimum separation (min hmin) that two highly-viscous drops can reach along a trajectory scales like O{a Ca), as Ca -» 0 (unlike O(a Ca4/3) for A = 0(1), see above). Our results for A = 10 in Fig. 16 closely follow this predicted scaling.
Hydrodynamical Interaction of Deformable Drops
419
Fig. 16a,b. Comparison of exact and asymptotic results for the surface clearance at A = 10.
Overall, the above comparisons with exact boundary-integral results demonstrate the validity of the asymptotic method of Sec. 2 for generic, non-axisymmetrical collisions with offsets Ayo ~ a, for a wide range of Ca
420
A.Z. Zinchenko and R.H. Davis
Fig. 17. Comparison of boundary-integral and asymptotic results for the surface clearance at A = 0.25 and Ca = 0.04.
A
Hydrodynamical Interaction of Deformable Drops
421
5. BUOYANCY OR GRAVITY-INDUCED MOTION OF TWO BUBBLES/DROPS WITH LARGE DEFORMATIONS Interaction of two strongly deformable drops or bubbles in the gravitational field offers a surprising variety of different types of behaviour. In the examples below, calculated by the curvatureless algorithm [26], the characteristic velocity V and time for nondimensionalization are \/\p\ga\/{&irjie) and CI2/V, respectively, where ai is the largest of the two non-deformed radii. Drops/bubbles start from spherical shapes. Each relative trajectory is determined by A, k = a\jai < 1, initial non-dimensional horizontal A i o = Axo/a2 and vertical Azo = Azo/a2 offsets of the drop centres, and the Bond number B — \Ap\gal/a (which is used in place of Ca to characterize deformation). Figure 18 shows relative motion of two bubbles for A = 10~3, k = 0.9, B = 14.3, Ax0 = 4.75, Ay0 = 0 and Azo = 15. Even though the initial shapes were smooth, interaction induces apparent surface singularities. Most interestingly, the side view in the (y, z)-plane (Fig. 18d) reveals that the larger, bottom bubble tends to develop a point-singularity, while the singularity formed on the smaller bubble by t = 8 is scallop-like, which is locally a 2D cusp. The occurrence of such sharp edges was also observed in related experiments [27,28] and is not a numerical artifact. If untreated, this singularity causes the calculation to stall. A special smoothing technique was developed [26], allowing for this and similar calculations to continue beyond the development of apparent singularities. The key idea is to add an artificial normal velocity to that obtained from the boundaryintegral solution at each time step. This additional velocity is on the order of e
422
A.Z. Zinchenko andR.H. Davis
Fig. 18. Relative buoyancy-driven motion of two bubbles with a\/a2 = 0.9, A = 10~3, B = 14.3 and 3840 triangular elements on each surface; no smoothing: (a-c) the bubble shapes in the (x, z)-plane; (d) in the (y, z)-plane. The calculation could not be continued due to the 2D cusp formation on the upper bubble at t w 8, without smoothing. Reproduced by permission from Ref. [26].
procedure with the use of conventional boundary integral (37) is generally successful and allows us to continue a simulation well beyond the singularity formation. When line singularities (2D cusps) form, however, it is crucial to use the curvatureless form (39) instead, in addition to smoothing, in order to eliminate a numerically unresolvable 0(1) cusp contribution. The above arguments, showing that detailed cusp resolution (which would be numerically impossible due to molecular scale of the cusp [50]) can always be avoided in the calculation of global motion when the curvatureless form is used, also indicate that additional physical mechanisms should not have a global effect, if they invalidate the classical model in
Hydrodynamical Interaction of Deformable Drops
423
the small cusp region only. For example, in practice, tip streaming often occurs for A
< q >a 11 - exp L ( < ^ i >
- l) I
,
(43)
where ea -C 1 is the smoothing parameter for surface Sa,
K={{k\ + kl)/2\1'2
(44)
is a measure of the local curvature (note K = k for a sphere), and < q >a and < K >a are the average values of |g| and K over Sa. In addition, an appropriate small constant is added to (43) to make zero the total flux through Sa (to conserve mass). On the main, smooth part of the drop or bubble, the modification to the normal velocity u, • rii is very small, of order ea < q >a (or even smaller, if the surface is nearly spherical, as is often the case in bubble cusping). In the cusp region, however, where K may attain high values, (43) gives a stronger negative correction to the normal velocity, and the exponential dependence keeps the cusp curvature under control. The principal curvatures k\ and ki for Eq. (44) do not have to be accurate, and so the method remains effectively curvatureless. Repeating the simulation in Fig. 18 with a small cusp-smoothing (ei = 62 = e = 6.67 x 10~3) allows us to proceed much farther and predict the further motion (Fig. 19). Namely, a larger bubble is being sucked into the dimple formed on a smaller one, resulting in bubble capture. This interesting phenomenon of deformation-induced capture was first discovered experimentally by Manga and Stone [27-28] for two bubbles in corn syrup, although adequate computer simulations were not available at that time. Manga and Stone [27-28], however, offered a far-field asymptotic analysis for well-separated drops with small deformations by the 'method
424
A.Z. Zinchenko andR.H. Davis
Fig. 19. The same simulation as in Fig. 18 in the (x, z)-plane, repeated with smoothing (e = 6.67 x 10~3) and leading to bubble capture. Reproduced by permission from Ref. [26].
of reflections', to qualitatively explain the onset of bubble capture. Most strikingly, simulations in Figs. 18-19 show that the critical offset Ax% for bubble capture can be much larger than the geometrical Smoluchowski's value 1 + k, which is in accord with experiments [27-28]. In practice, this capture phenomenon is followed by bubble coalescence when the van der Waals attraction drains the thin film, but the actual coalescence time (which may be quite large [27-28]) depends on the details of near-contact physics. Additional calculations [26] confirm global numerical convergence of the results in Fig. 19 in the limit N& —> oo, e —»• 0, and their independence of the relation between e and N& in this limit, thus making the choice of smoothing (43) purely technical. The cusp-smoothing, in combination with the curvatureless algorithm, is therefore a correct method to predict the global shape evolution both with point and line 'singularities' (limitations are discussed in Ref. [26]). The local structure of cusps,
Hydrodynamical Interaction of Deformable Drops
425
however interesting, is irrelevant to predicting global dynamics of cusped shapes. A different mode of bubble capture is shown in Fig. 20. The parameters k = 0.7, A = 10"3, -6 = 7, Axo = 2.3 and Az0 - 10 are the same as in Fig. 9 of Ref. [26], but we used here much finer triangulations (N& = 15360 triangles per bubble, instead of N& = 3840 in Ref. [26]), and, accordingly, a four times smaller smoothing parameter ei = 8.2 x 10~4 fa — 0). Only a point singularity forms on the smaller bubble, while the other bubble remains smooth. Unlike in Fig. 19, the smaller bubble is swept around the larger one and eventually becomes sucked into the dimple formed at the rear of the larger bubble. This later mode of bubble capture is known experimentally [27-28] as 'entrainment.' The present results are close to those in Fig. 9 of Ref. [26] and show just a slightly faster capture; again, a good global accuracy is achieved without fully resolving the cusp. Calculations similar to those in Figs. 19-20 (although, for more limited resolutions) were used systematically [26] to find the critical offsets (Axo)cr for capture by trial-and-error (Fig. 21). The critical offset increases with B due to increased deformation-induced alignment but becomes only a weak function of B at large Bond numbers. Also, the critical offset is very sensitive to the size ratio and is the largest for bubbles of nearly the same size (although the mutual approach is slow in this case). Additionally, a comparison of the dark squares and the crosses in Fig. 21 at a\/a2 = 0.7 shows that the capture efficiency is slightly reduced when the bubbles (A = 10~3, which is representative of A < 0(1O~2)) are replaced with drops having A = 0.1. Further increase in A to 0(1) values leads to interaction-induced breakup, instead of capture. For ai/02 = 0.7, A = 1, B = 5.31, Ax0 = 1, and Azo = 5.09 (Fig. 22), the smaller drop is swept around the larger one, stretches due to hydrodynamical interaction, and starts necking. The reason for stretching is that the hydrodynamical field in the wake of the larger drop is mainly an extensional flow. The simulation in Fig. 22 was repeated for four different numbers of triangles iVAl = 2160, 3840, 6000 and 8640 on the smaller drop [26], to assess accuracy. After t = 6, the drops continue to separate, and the smaller drop experiences neck pinchoff, while the other drop remains compact. In Fig. 23, only the evolution of the breaking drop is shown. Despite some local mesh imperfections inherent in our fixed-mesh topology breakup simulations, the global convergence is quite good (Fig. 24), especially in the top bulbous area; in particular, the radius 03 = 0.920ai of the main (top) fragment after breakup is accurately predicted [26]. Moreover, these simulations accurately describe the temporal
426
A.Z. Zinchenko and R.H. Davis
Fig. 20. Bubble capture through 'entrapment' at ai/a2 = 0.7, A = 10~"3, B = 7, NA = 15360, ei = 8.2 x 10" 4 .
Fig. 21. The non-dimensional critical capture offset far upstream for bubbles and low-viscosity drops. Dark squares are for A = 10~3, crosses are for A = 0.1 (at a.i/a.2 = 0.7 only). Reproduced by permission from Ref. [26].
Hydrodynamical Interaction of Deformable Drops
427
Fig. 22. Relative buoyancy-driven motion of two drops with 0,1/0,2 = 0.7, A = 1, B = 5.31; 3840 elements on the smaller and 1280 elements on the larger drop. Reproduced by permission from Ref. [26].
dynamics of the neck thinning. As breakup is approached, the local neck shape becomes axisymmetrical (even though the whole problem is 3D), thus allowing us to introduce an equivalent neck radius r = (<STOm/7r)1/'2, where Smin is the minimum area of cross-sections orthogonal to the line of maximum elongation. This observation supports, incidentally, the body of local axisymmetrical studies on viscous pinchoff (e.g., [59-62]). In particular, for A = 1, a local self-similar axisymmetrical solution [62] predicts a linear dependence r(t) at pinchoff, with the dimensional slope of —0.034cr/^e- For our simulations in Fig. 24, the dynamics of the equivalent neck radius is given in Fig. 25. The small lack of accuracy at the bottom of the drop at large times (Fig. 24) does not seem to appreciably affect the dynamics of the neck thinning, and excellent numerical convergence with respect to triangulations is observed in Fig. 25; this agreement may be due to the universality of self-similar thinning at pinchoff. Our dimensional slope dr/dt at the end of the simulation is about —0.029cr//xe, which is close to the theoretical result [62] —0.034a//ie; a plausible reason for small discrepancy is that the global drop shape at the end of our simulation
428
A.Z. Zinchenko and R.H. Davis
Fig. 23. The stretching and breakup of the smaller drop. The parameters are the same as in Fig. 22, except that 8640 triangular elements are used for the smaller drop.
Fig. 24. Comparison of the absolute positions and shapes of the smaller drop in the (x, z)-plane for the simulation of Fig. 23 using JVAl = 3840 (dashed lines), 6000 (solid lines), and 8640 (dotted lines). Reproduced by permission from Ref. [26].
Hydrodynamical Interaction of Deformable Drops
429
Fig. 25. The non-dimensional neck radius vs. time for the simulations in Figs. 23-24 with different triangulations of the smaller drop. The results for N^i = 8640 and 6000 are practically indistinguishable. Reproduced by permission from Ref. [26].
is not quite axisymmetrical, causing a small error in the direction of the minimal cross-section. Another possible reason is that the ultimate slope dr/dt is approached only for very thin necks, which was seen previously in axisymmetrical simulations [62-63]. In any case, using dr/dt between —0.029cr//ie and — 0.034a/fie to extrapolate the neck radius in Fig. 25 to zero yields tight bounds on the non-dimensional breakup time: t^ = 7.95 — 8.02. Breakup simulations similar to those in Figs. 22-24 were also made for contrast viscosities (A ^ 1), and critical horizontal offsets (Axo)^. for breakup were also calculated [26]. An additional mode of 3D two-drop interaction in gravity-induced motion predicted theoretically [26] is 'combined capture and breakup.' Namely, after the smaller drop is swept around the large one, it becomes sucked into the dimple formed on the larger drop, like in the pure capture phenomenon (Fig. 20), but simultaneously undergoes considerable elongation and starts necking. This behaviour strongly suggests that the smaller drop will break without being released from the dimple, as has also been observed in axisymmetric simulations [63]. Another, and very interesting phenomenon, discovered experimentally [64] for moderate A and moderate-to-large B is 'pass-through', in which the leading drop forms a torus and the smaller drop passes through its centre, sometimes in multiple cycles reminiscent of leap-frogging. It can be noted that a gravity-induced breakup simulation for a set of parameters close to ours in Figs. 22-23 was also performed by Cristini
430
A.Z. Zinchenko and R.H. Davis
et al. [65] using isotropic mesh restructuring [48] (with a growing number of boundary elements as a drop stretches). Their mesh algorithm is very versatile, with an excellent local control over the mesh, and even allows the calculations to proceed after primary breakup by using mesh splicing [48]. With isotropic restructuring, however, it is often more difficult to obtain convergent results (in particular, an accurate temporal dynamics of drop elongation in primary breakup), and a very large total number of elements may be needed for this purpose, compared to 'passive' mesh stabilization [26]6. MULTIDROP SIMULATIONS 6.1 Shear Flow of Concentrated Emulsions One of the useful applications [24] of multipole-accelerated, boundaryintegral simulations in a periodic box is to study non-Newtonian rheology of a concentrated emulsion of deformable drops in a steady shear (7x2, 0, 0) (where 7 > 0, without a loss of generality). For simplicity, monodispersity is assumed. A large number N of spherical drops of radius a with centres in a periodic box V (initially, a cube) starts from a random non-overlapping configuration (prepared by a standard Monte-Carlo method, e.g. [66]), and the system is subject to shear, which causes the drops to deform and the periodic cell V to skew (excessive cell skew is simply avoided [43, 24] by restructuring periodic cells V at half-integer strains jt). At every time moment t > 0, the quantity of interest is the space-average stress tensor Eij, which can be expressed in terms of surface integrals (e.g., Ref. [67]) over interfaces only. For N ^> 1, the only essential components are [68] the shear stress £12, the first £11 — £22 and the second £22 — £33 normal stress differences. It is convenient to introduce non-dimensional quantities M* = E12 /(/vy), M = (En - £22)/(/ie|7l) and N2 = (£ 22 - E M ) / ^ ! ) Fig. 26 presents a typical snapshot [24] of the simulations, at drop volume fraction c = 0.5, N = 200 drops in a periodic cell, N& = 1280 triangular elements per drop, and capillary number Ca — fieja/a = 0.1; the whole simulation was continued to strains jt ss 28 with about 3000 time steps. For the number of drops N attainable in present-day dynamical simulations (roughly, N = O(102 — 103) with multipole-accelerated codes [23-25]), a single configuration is not representative, and time averaging of the stress Ey is essential. Fig. 27(a-c) presents the trajectories of the dimensionless effective viscosity //* and normal stress differences JVi, N2 vs. strain 7^ at c = 0.5, A = 1, N = 100 and capillary numbers Ca = 0.025, 0.1 and 0.2. These results were produced by the algorithm
Hydrodynamical Interaction of Deformable Drops
431
Fig. 26. A snapshot of an emulsion shear flow simulation at c = 0.5, A = 1, Ca = 0.1, and strain -yt = 4.45. The centres of 200 independent drops have been mapped into the unit cube, an initial periodic cell. Reproduced by permission from Ref. [24].
[24], but we used here much higher resolution N& = 8640 at Ca = 0.025 essential for Ca
432
A.Z. Zinchenko andR.H. Davis
Fig. 27. The trajectories of the effective viscosity (a), and first (b) and second (c) normal stress differences for c = 0.5, A = 1, NA = 1280 - 8640 and N = 100, with Ca = 0.025 (thick lines), 0.1 (dashed lines) and 0.2 (thin solid lines). The initial configuration for Ca = 0.1 calculations (shown in a,b only) was taken from a steady state with Ca = 0.2
was shown [24] to overestimate the average of N\ almost 1.5-fold (note that the first normal stress difference in this case is essential, comparable to the shear stress). In contrast, the average results for iV = 100 and 200 showed very good convergence [24]. The time-averaged results for the effective shear viscosity (< fi* >) and normal stress differences (< Ni >, < iV2 >) at A = 1, N = 100 and different concentrations c (from 0.3 to 0.5) are shown in Fig. 28(a-c); at
Hydrodynamical Interaction of Deformable Drops
433
c = 0.55, the behaviour is complicated by "phase transition" [24] and will not be discussed here. For moderate capillary numbers {Ca > 0.1), the results are identical to those in Ref. [24], but for Ca = 0.025 and 0.05, we used much higher resolutions (N& = 8640 and 6000, respectively) in the present calculations to improve the accuracy. Several trends in Fig. 28(a-c) are interesting to discuss. The first is a sharp dependence of the emulsion viscosity on Ca at high concentrations (Fig. 28a), so that most of the shear-thinning occurs for drops with only small deformation. This shear-thinning occurs because the drops slide more easily past each other when the shear rate and, hence, capillary number are increased, due to drop deformation.
Fig. 28. Time-averaged (a) effective viscosity and (b,c) normal stress differences at A = 1.
Unlike for the viscosity < / / >, there are no general mechanical principles to predict the signs of < N\ > and < N2 >, but our calculations always yield positive < N\ > and negative < N2 > (which is in accord with
434
A.Z. Zinchenko and R.H. Davis
more limited simulations [43-44J for small systems (N = 12) at 30% concentration, and with single-drop calculations [69]). However, our results for c = 0.3, 0.4 and 0.45 indicate that the first normal stress difference, < N\ >, extrapolates to small but negative values at Ca = 0 (Fig. 28b). The second normal stress difference, < N2 >, although inevitably subject to some statistical errors, is seen to remain negative as Ca —>• 0 (Fig. 28c). Negative signs of both < N\ > and < N2 > at Ca = 0 are also in agreement with the asymptotic rheological calculations [9] for semi-dilute emulsions of spherical drops on the level of pairwise interactions. The physical reason for predicted sign change of < N\ > at small Ca (Fig. 28b) is that, for moderate-to-large Ca, drop deformation has the primary effect on N\ while, at Ca
Hydrodynamical Interaction of Deformable Drops
435
Fig. 29. A snapshot of the emulsion shear flow simulation for c = 0.55, A = 3, Ca = 0.1 and jt = 6.38. The centres of 100 independent drops have been mapped into the unit cube, an initial periodic cell. Reproduced by permission from Ref. [24].
Fig. 30. The trajectories of the effective viscosity for c = 0.55, A = 3, and N = 100. Reproduced by permission from Ref. [24].
436
A.Z. Zinchenko andR.H. Davis
6.2 Sedimentation of Concentrated Emulsions of Deformable Drops In this section, we discuss sedimentation (or, equivalently, creaming under buoyancy) of concentrated emulsions of deformable drops starting from a homogeneous initial state with no deformation. Although this problem did not attract attention in the literature until recently [23, 25], sedimentation of interacting drops with moderate-to-large deformations may have interesting applications in phase separations. Under normal gravity, the Bond number for emulsion drops is typically small, so the sedimentation would occur only with small drop deformation. If such an emulsion, however, is placed in a centrifuge (thereby increasing the acceleration due to gravity by an order of magnitude, or more), the Bond number may become 0(1), causing strong deformation (and drastically changing the sedimentation regime!). Another example is emulsion sedimentation in the miscibility gap, with low surface tension. The behaviour of interacting deformable drops settling under gravity is strikingly different from that for freely suspended drops in shear flow. Figure 31 presents snapshots of the simulation at Bond number B = Apga2/a = 1.75, drop volume fraction c = 0.35, viscosity ratio A = 1, N = 600 drops in a periodic cell and N& = 960 triangular elements per drop, for different non-dimensional times t (scaled with fj,e/(Apga)). The initial 'well-mixed' state of non-overlapping spherical drops of radius a (not shown) was prepared by the standard Monte-Carlo method. The main quantity of interest is the sedimentation rate, i.e., the volume-averaged fluid velocity over the drop phase, in the reference frame where the emulsion, on average, is at rest (only the vertical component, U, is essential). The non-dimensional rate U/Uo (where Uo is the settling velocity of an isolated spherical drop) for this simulation is presented in Fig. 32a by the thick line. An apparent non-existence of the steady state is due to drop clustering, and is akin to the instability of a dilute suspension of sedimenting solid rod-like particles predicted analytically by Koch and Shaqfeh [29] by far-field analysis of the pair distribution function (for more recent studies on rod-like particles, see Ref. [73-74]). Indeed, most drops in our simulation (Fig. 31) acquire prolate shapes and become partly oriented along the vertical, and there is some qualitative analogy between the two systems. The main difficulty of the emulsion sedimentation study lies in a strong dependence of the instability growth on the initial random configuration; thin lines in Fig. 32a are for nine other random initial conditions showing considerable dispersion, as time proceeds. Probably, to make one realization representative in a wide time range, systems with N > O(105)
Hydrodynamical Interaction of Deformable Drops
437
Fig. 31. Snapshots of the emulsion sedimentation simulation from a homogeneous initial state of spherical drops for B = 1.75, c = 0.35, A = 1 and TV = 600. The drops sediment downward. Reproduced by permission from Ref. [25].
Fig. 32. The dependence of the sedimentation rate on random initial configurations at B = 1.75, c = 0.35, A = 1 and NA = 960: (a) 10 realizations with N = 600; (b) 10 realizations with N = 300. Reproduced with permission from Ref. [25].
438
A.Z. Zinchenko andR.H. Davis
would need to be considered (since statistical fluctuations decay slowly, ~ N~ll2), which is far beyond the present-day capabilities; ensemble averaging over many realizations must be performed instead at each time moment, with assumptions made about the probability density of the initial states. In what follows, we assume [25] that all initial configurations of non-overlapping spheres have equal probability; a standard Monte Carlo method for 'hard spheres' (e.g. [66]) is known to generate realizations satisfying this condition. To study the effect of N, similar calculations were done for 25 uncorrelated initial realizations with N = 300; the first 10 are given in Fig. 32b showing even larger dispersion. Remarkably, however, ensemble-averaged results for N = 300 and 600 are in excellent agreement over a wide time range (Fig. 33) and are thus representative of the macroscopic behaviour; in contrast, a smaller system N = 100 greatly underestimates the average sedimentation rate (Fig. 33), except for very small times. The physical reason why very large systems are imperative in this problem is due to clustering; as time grows, so does the cluster size, and more drops are required to make the results box-size independent. In addition to the effect of N, the numerical effect of drop triangulation on the results in Fig. 33 was also studied and found to be small in a wide time range [25]. Calculations were also performed at the same A and B to study the effect of drop volume fraction (from 0.15 to 0.4) on the ensemble-average
Fig. 33. The ensemble-averaged sedimentation rate for B = 1.75, c = 0.35, A = 1 and N& = 960, with 14, 25 and 20 realizations for TV = 600, 300 and 100, respectively. Vertical bars show statistical errors (with 67% confidence level) for N = 100 and 600; statistical errors for N = 300 are similar to those for N = 600. Reproduced by permission from Ref. [25].
Hydrodynamical Interaction of Deformable Drops
439
sedimentation rate (Fig. 34); at c = 0.15, quite large systems N — 1200 were used (Fig. 35). At c — 0.4, the instability and clustering grow slowly (Fig. 34). This slower growth is likely due to geometrical constraints at high concentrations hampering drop deformation. The instability growth rate is higher at c = 0.35. Emulsions with 25% drop volume fraction show the strongest instability, with the sedimentation rate increasing 2.15 times from its minimum by t = 150 (by this time, a single drop would have fallen a distance equivalent to 40 radii). At c = 0.15, the initial sedimentation rate is, of course, higher than at c = 0.25 (due to less backfiow and hindrance at lower concentration), but it grows somewhat more slowly as time proceeds. The average drop length and deformation at c = 0.15 are also somewhat smaller than at c = 0.25, thus helping to explain why the emulsion instability at c = 0.15 is not as strong as at c = 0.25. The existence of the optimum volume fraction for emulsion instability and clustering (about 0.15—0.25 for A = 1 and B = 1.75) simply follows from the fact that, as c —>• 0, there are no interactions, and, hence, no deformation of sedimenting drops - a primary cause of clustering and instability. It remains an open question, however, if the graphs of the average sedimentation rate for different drop volume fractions can intersect (so that a more concentrated system would start sedimenting faster than a less concentrated one). We
Fig. 34. The ensemble-averaged sedimentation rate for B = 1.75, A = 1, N& = 960 and different drop volume fractions. Twelve realizations with TV = 400, 14 realizations with N = 600, 14 realizations with N = 800 and 10 realizations with N = 1200 were used for c = 0.4, 0.35, 0.25, and 0.15, respectively. Reproduced by permission from Ref. [25].
440
A.Z. Zinchenko andR.H. Davis
Fig. 35. A typical snapshot of a clustered configuration with B = 1.75, c = 0.15, A = 1, N = 1200 and N& = 960. Drops sediment downward, t = 140. Reproduced by permission from Ref. [25].
cannot rule out that further evolution would lead to massive, statistically significant drop break-up changing the trends in Fig. 34. Surprisingly, however, in none of the calculations for Fig. 34 have we seen individual drop break-up; the time-scales where drop break-up might have a significant effect on the sedimentation rate are much larger than those in Fig. 34. The phenomenon of instability of sedimenting emulsions is strongly sensitive to the Bond number: When B is decreased, the instability growth slows down dramatically, due to less deformation and clustering, which is demonstrated in Fig. 36 for two random realizations. It is expected, however, in the spirit of the Koch-Shaqfeh [29] theory for a dilute suspension of rod-like particles, that there is no critical Bond number for the instability. Similar calculations for A ^ 1 are much more expensive and could be done only for individual realizations (e.g., Fig. 37), without ensemble averaging. Even under this limitation, it was possible to study the effect of A on the sedimentation rate starting from the same random initial configuration (Fig. 38). The most interesting observation from Fig. 38 is that the instability growth accelerates dramatically, when A is decreased from 1 to 0.25, and the results for A = 0.1 show the same trend; in the latter case, however, we could not proceed to large times due to numerical difficulties.
Hydrodynamical Interaction of Deformable Drops
441
Fig. 36. The effect of the Bond number on the instability growth for two random realizations (1 and 2) with c = 0.35, A = 1, N = 300 and JVA = 960. Solid lines, B = 1.75; dashed lines, B = 0.9. Reproduced by permission from Ref. [25].
Fig. 37. Snapshots of the emulsion sedimentation simulation from a homogeneous initial state of spherical drops for B = 2.5, c = 0.4, A = 0.25, N = 125 and 7VA = 1500. Drops sediment downward. Reproduced by permission from Ref. [25].
442
A.Z. Zinchenko and R.H. Davis
Fig. 38. The effect of the viscosity ratio A on the instability growth for an individual realization with B = 2.5, c = 0.4, N = 125. Reproduced by permission from Ref. [25].
6.3 Flow of a concentrated emulsion past a deformable drop or bubble Another problem [25] which can be studied by large-scale numerical simulations is the steady gravity-induced motion of a deformable bubble (or drop still called a 'bubble' in what follows) through a concentrated emulsion at low Reynolds numbers. When the bubble is much larger than the emulsion drops, the emulsion can be treated, in principle, as an effective medium with the macroscopic boundary conditions (no flux, continuity of velocity and stress) on the bubble surface. For a Newtonian form of the effective stress tensor and a spherical bubble, the classical solution of Hadamard and Rybchinski [75] shows that no deformation will occur, irrespective of the nonzero surface tension value; surface tension, however, is required for the stability of the drop. Unfortunately, a Newtonian form of the effective stress tensor is not an accurate approximation for concentrated emulsions of deformable drops (Sec. 6.1), and a more complicated constitutive equation valid for arbitrary kinematics would be required, which is problematic and could be done, at present, only in an ad hoc manner. For a bubble comparable in size with the emulsion drops, the problem can be studied instead by rigorous, first-principle numerical simulations, without a constitutive equation. A large bubble of non-deformed radius a& and surface tension G\, is placed in a cubic box, together with N 2> 1 drops of non-deformed radius a^ forming a random emulsion. The whole system is continued triply periodically into the whole space. The emulsion drops are deformable (with surface tension aa) and neutrally buoyant, so that the
Hydrodynamical Interaction of Deformable Drops
443
motion in the system is entirely due to density difference pb — pe between the bubble and the surrounding medium. The bubble and the emulsion drops are assumed to start with spherical shapes. The main quantity of interest is the settling bubble velocity (in the reference frame where the whole system is at rest) as a function of emulsion concentration c, bubble Bond number Bb = \pb- Pe\alg/ab
,
(46)
bubble-to-emulsion-drop size ratio ab/ad and surface tension ratio o^/orf, bubble-to-medium-viscosity ratio A& = w,//i e , and emulsion-drop viscosity ratio Xd = p-d/p-e- The limit N —> oo must be taken to approach the solution for a single bubble in an unbounded medium. The solid line in Fig. 39a shows the non-dimensional settling bubble velocity Ub/U0 (where
Fig. 39. The non-dimensional settling velocity of a bubble through a concentrated emulsion at c = 0.45, ab/ad = 2, \b = 0.05, ab/ad = 2, and \d = 1, with (a) N = 800, (b) N = 200. The bubble and each emulsion drop are discretized by 2160 and 1280 elements, respectively. The horizontal lines are the stationary levels for Bb = 4. Reproduced by permission from Ref. [25].
Uo is the isolated bubble velocity, in the absence of emulsion drops) in a 45% concentrated emulsion for N = 800 and Bb — 4; time is scaled with Me/(|/°6 — Pe\ga>b)- The initial decline in Ub/U0 is due to emulsion-drop interactions increasing the effective viscosity around the bubble; a statistical steady state is reached at t ~ 50. Despite a relatively large Bb = 4, the bubble deformation was found to be small (although an emulsion at c = 0.45 and Ad = 1 is noticeably non-Newtonian, see Sec. 6.1). In contrast, the emulsion drops experience large deformations in the vicinity of the bubble, with typical oblate and prolate shapes upstream and downstream, respectively (Fig. 40); the other drops (away from the bubble) are only slightly
444
A.Z. Zinchenko and R.H. Davis
Fig. 40. A typical snapshot of emulsion drops in the vicinity of the bubble, from the simulation in Fig. 39 with N = 800 and -B& = 4 at t = 31. The bubble rises upward. Reproduced by permission from Ref. [25].
deformed, but they create a necessary background for the simulation. A similar simulation with N = 200 (solid curve in Fig. 39b) shows strong sensitivity of the initial velocity Ub/Uo (at t = 0) to N (0.68 for iV = 200 vs. 0.80 for N = 800; the limit must be 1 at N —>• oo). This dependence on N merely reflects strong artifact interactions between single bubbles in different periodic cells at t = 0, since strictly spherical emulsion drops with \d — 1 have no effect at all. Fortunately, at finite 5&, the necessary steady-state results for Ub/U0 are less sensitive to N (0.578 at N = 200 vs. 0.619 at N — 800), and thus the limit N —>• oo can be practically achieved in this problem. Similar calculations at Bb = 2 (dashed lines in Fig. 39a,b) give slightly worse convergence of the steady-state results (0.566 for N = 800 vs. 0.521 for N = 200). It appears that the screening effect of emulsion-drop deformation weakens as Bb is decreased, and convergent steady-state simulations for smaller Bond number may require N > O(103); the same is obviously true for larger size ratios ab/ad- Nevertheless, our simulations demonstrate the possibility of solving the problem without an ad hoc constitutive equation for an emulsion.
Hydrodynamical Interaction of Deformable Drops
445
7. SOME PROSPECTS FOR FUTURE RESEARCH In this chapter, we have demonstrated some contemporary progress made in the area of deformable drop interactions and coalescence, and concentrated emulsion flows by rigorous theories and first-principle numerical simulations. Still, there remains a large number of issues to be resolved, and a large room for further development. The hydrodynamical theory of coalescence (Sec. 2), based on matching the thin-film 'inner' solution with the 'outer' solution was shown to be a correct approach for arbitrary 3D interaction of slightly deformable drops. Applications to coalescence efficiency calculations, however, have been made so far [20, 21] only for classic, unretarded van der Waals attractions. To make this theory more realistic, additional account for electromagnetic retardation and other colloidal forces may be necessary. Besides, real emulsions are rarely uncontaminated, and it would be practically important to include surfactant effects on interaction and coalescence, which opens a wide area of research; some initial studies have already been made [76]. For multidrop interactions, it appears possible to further improve multipole-accelerated boundary-integral algorithms and make them suitable for accurate manythousand drop simulations with large deformations, with more powerful computer resources available in the near future. For solid particle simulations, the lattice-Boltzmann (LB) method was found to be a very promising tool [77-79]. It would be useful to explore if LB can become a competitive method for deformable drops, in terms of accuracy and speed. When fluid inertia cannot be neglected (which is often the case for large drops), the problems become much more involved; a front-tracking method [80] appears to be a suitable tool for deformable drops, although much remains to be done to improve the accuracy of this method. In the area of computational rheology of concentrated emulsions (Sec. 6.1), generalization for time-dependent shear flows would be straightforward; a major challenge, however, is to develop a simulation method suitable for an arbitrary history of macroscopic deformation (different from shear flow).
446
A.Z. Zinchenko and R.H. Davis
REFERENCES [1] E. Rushton and G.A. Davis, Appl. Sci. Res., 28 (1973) 37. [2] S. Haber, G. Hetsroni and A. Solan, Int. J. Multiphase Flow, 1 (1973) 57. [3] A.Z. Zinchenko, Prikl. Mat. Mekhan., 44 (1981) 30. [4] A.Z. Zinchenko, Prikl. Mat. Mekhan., 46 (1983) 58. [5] A.Z. Zinchenko, Prik. Mat. Mekhan., 47 (1984) 37. [6] Y.O. Fuentes, S. Kim and D.J. Jeffrey, Phys. Fluids, 31 (1988) 2445. [7] Y.O. Fuentes, S. Kim and D.J. Jeffrey, Phys. Fluids A, 1 (1989) 61. [8] H. Wang, A.Z. Zinchenko and R.H. Davis, J. Fluid Mech., 265 (1994) 161. [9] A.Z. Zinchenko, Prikl. Mat. Mekhan., 48 (1984) 198. [10] A.Z. Zinchenko and R.H. Davis, J. Fluid Mech., 280 (1994) 119. [11] A.Z. Zinchenko and R.H. Davis, Phys. Fluids, 7 (1995) 2310. [12] G. Mo and A.S. Sangani, Phys. Fluids, 6 (1994), 1637. [13] A.Z. Zinchenko, Prikl. Mat. Mekhan., 42 (1979) 1046. [14] A.Z. Zinchenko, Prikl. Mat. Mekhan., 45 (1982) 564. [15] R.H. Davis, J.A. Schonberg and J.M. Rallison, Phys. Fluids A, 1 (1989) 77. [16] M.D. Cooley and M.E. O'Neill, Mathematika, 16 (1969) 37. [17] S.G. Yiantsios and R.H. Davis, J. Fluid Mech., 217 (1990) 547. [18] S.G. Yiantsios and R.H. Davis, J. Colloid Interface Sci., 144 (1991) 412. [19] A.Z. Zinchenko and R.H. Davis, J. Comput. Phys. (2004, submitted). [20] M.A. Rother, A.Z. Zinchenko and R.H. Davis, J. Fluid Mech., 346 (1997) 117. [21] M.A. Rother and R.H. Davis, Phys. Fluids, 13 (2001) 1178. [22] J.M. Rallison and A. Acrivos, J. Fluid Mech., 89 (1978) 191. [23] A.Z. Zinchenko and R.H. Davis, J. Comput. Phys., 157 (2000) 539. [24] A.Z. Zinchenko and R.H. Davis, J. Fluid Mech., 455 (2002) 21. [25] A.Z. Zinchenko and R.H. Davis, Phil. Trans R. Soc. Lond. A, 361 (2003) 813. [26] A.Z. Zinchenko, M.A. Rother and R.H. Davis, J. Fluid Mech., 391 (1999) 249. [27] M. Manga and H.A. Stone, J. Fluid Mech., 256 (1993) 647. [28] M. Manga and H.A. Stone, J. Fluid Mech., 300 (1995) 231. [29] D.L. Koch and E.S.G. Shaqfeh, J. Fluid Mech., 209 (1989) 521. [30] J. Happel and H. Brenner, Low Reynolds Number Hydrodynamics, Nijhoff, Dordrecht, 1973. [31] L.D. Reed and F.A. Morrison, Int. J. Multiphase Flow, 1 (1974) 573. [32] L.M. Hocking, J. Engng Maths, 7 (1973) 207. [33] A.F. Jones and S.D.R. Wilson, J. Fluid Mech., 87 (1978) 263. [34] I. B. Bazhlekov, A. K. Chesters and F. N. van de Vosse, Int. J. Multiphase Flow, 26 (2000) 445. [35] X. Zhang and R.H. Davis, J. Fluid Mech., 230 (1991) 479. [36] S.R. Reddy, D.H. Melik and H.S. Fogler, J. Colloid Interface Sci., 82(1981) 116. [37] J.R. Rogers and R.H. Davis, Metall. Trans., 21A (1990) 59. [38] H. Wang and R.H. Davis, J. Fluid Mech., 295 (1995) 247. [39] K. Barton and R. Subramanian, J. Colloid Interface Sci., 133 (1989) 211. [40] V. Smoluchowski, Z. Phys. Chem. (Leipzig, 92 (1917) 129. [41] M. Loewenberg and E.J. Hinch, J. Fluid Mech., 338 (1997) 299. [42] A.Z. Zinchenko, M.A. Rother and R.H. Davis, Phys. Fluids, 9 (1997) 1493. [43] M. Loewenberg and E.J. Hinch, J. Fluid Mech., 321 (1996) 395.
Hydrodynamical Interaction of Deformable Drops
447
[44] M. Loewenberg, Trans. ASME: J. Fluids Engng, 120 (1998) 824. [45] S. Kim and S. Karilla, Microhydrodynamics: Principles and Selected Applications, Butterworth-Heinemann, 1991. [46] C. Pozrikidis, Boundary Integral and Singularity Methods for Linearized Viscous flow, Cambridge University Press, 1992. [47] G.P. Muldowney and J.J.L. Higdon, J. Fluid Mech., 298 (1995) 167. [48] V. Cristini, Blawzdziewicz and M. Loewenberg, J. Comput. Phys., 168 (2001) 445. [49] J.M. Rallison, J. Fluid Mech., 109 (1981) 465. [50] J.-T. Jeong and H.K. Moffatt, J. Fluid Mech., 241 (1992) 1. [51] J.P. Hansen and I.R. McDonald, Theory of Simple Liquids, Academic Press, 1976. [52] H. Hasimoto, J. Fluid Mech., 5 (1959) 317. [53] H. Yang, C.C. Park, Y.T. Hu and L.G. Leal, Phys. Fluids, 13 (2001) 1087. [54] M. B. Nemer, X.Chen, D. H. Papadopoulos, J. Blawzdziewicz and M. Loewenberg, Phys. Rev. Lett. 92 (2004) 114501. [55] R.A. de Bruijn, Chem. Engng Sci., 48 (1993) 277. [56] H.A. Stone, Ann. Rev. Fluid Mech, 26 (1994) 65. [57] W.J. Milliken, H.A. Stone and L.G. Leal, Phys. Fluids A, 5 (1993) 69. [58] J.D. Sherwood, J. Fluid Mech, 144 (1984) 281. [59] J. Eggers, Phys. Rev. Lett, 71 (1993) 3458. [60] J. Eggers, Rev. Mod. Phys, 69 (1997), 865. [61] M.P. Brenner, J.R. Lister and H.A. Stone, Phys. Fluids, 8 (1996) 2827. [62] J.R. Lister and H.A. Stone, Phys. Fluids, 10 (1998) 2758. [63] R. H. Davis, Phys. Fluids, 11 (1999) 1016. [64] J. Kushner IV, M.A. Rother and R.H. Davis, J. Fluid Mech, 446 (2001) 253. [65] V. Cristini, J. Blawzdziewicz and M. Loewenberg, Phys. Fluids, 10 1781. [66] P.K. MacKeown, Stochastic Simulation in Physics, Springer, 1997. [67] C. Pozrikidis, J. Comput. Phys, 169 (2001) 250. [68] G. Astarita and G. Marrucci, Principles of Non-Newtonian Fluid Mechanics, McGrawHill, 1974. [69] M.R. Kennedy, C. Pozrikidis and R. Skalak, Comput. Fluids, 23 (1994) 251. [70] R.S. Rivlin and J.L. Ericksen, Arch. Rat. Mech. Anal, 4 (1955) 323. [71] N. A. Frankel and A. Acrivos, J. Fluid Mech, 44 (1970) 65. [72] D. Barthes-Biesel and A. Acrivos, Int. J. Multiphase Flow, 1 (1973) 1. [73] M.B. Mackaplow and E.S.G. Shaqfeh, J. Fluid Mech, 376 (1998) 1493 [74] J.E. Butler and E.S.G. Shaqfeh, J. Fluid Mech, 468 (2002) 205. [75] G.K. Batchelor, An Introduction to Fluid Dynamics, Cambridge University Press, 1967. [76] A. K. Chesters and I. B. Bazhlekov, J. Colloid Interface Sci, 230 (2000) 229. [77] A.J.C. Ladd, J. Fluid Mech, 271 (1994) 285. [78] A.J.C. Ladd, J. Fluid Mech, 271 (1994) 311. [79] A.J.C. Ladd, Phys. Fluids, 9 (1997) 491. [80] S. O. Unverdi and G. Tryggvason, J. Comput. Phys, 100 (1992) 25.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 11
The role of inertial effects and conical flows in breakup of liquid threads Vakhtang Putkaradze Department of Mathematics and Statistics, University of New Mexico, Albuquerque, NM 87131-1141, USA email: [email protected]
We study the appearance and relevance of conical flows in the problem of droplet breakup. First, we give the theory of break-up of a slender jet of fluid in air [1]. We then proceed to the two-fluid break-up when a thread of one fluid is rupturing up while being surrounded by another fluid. The two-fluid break-up shows the appearance of conical shapes [2]. Inspired by these experiments, we develop a theory of exact solutions of Navier-Stokes equations which describe the flow of two fluids fluid separated by an interface in the shape of a single or double cone. We also describe the extension of these solutions to two dimensions. 1. INTRODUCTION When observing a dripping faucet in the kitchen, one can notice that the process of forming an individual droplet can be separated into two parts. The first part is the slow accumulation of water in the end of the faucet so that the gravitational force pulling the droplet down can overcome the surface tension holding the droplet inside the faucet. Once enough mass is accumulated, the droplet falls from the faucet with ever-increasing speed. The "slow" phase finishes when substantial volume of the droplet appears from the faucet; from that time on, the surface tension's role becomes destructive, snapping the liquid thread extending from the faucet to the droplet.
450
Vakhtang Putkaradze
The process of snapping the liquid thread and forming the droplet is very fast, so inertial effects are important for understanding this process. Formation of droplets and the role of inertial effects when air is the surrounding substance is understood quite well. Less well understood is the role of inertial effects when a thread of one fluid snaps while being surrounded by another fluid. The major difficulty of describing the fluid motion at the very moment of droplet formation lies in the fact that at this point in time and space Navier-Stokes equations exhibit a singularity. This singularity is physical, as at the moment when the droplet is born, the thread becomes infinitely thin in finite time. An amazingly beautiful and complex dynamic during droplet break-up has been revealed in great detail in [3]. We shall start by giving the reader an introduction into the theory of droplet formation in air, where a rigorous theory balancing surface tension, viscosity and inertia can be derived. 2. 2.1.
DROPLET FORMATION IN AIR Derivation of equations of motion
In this section, we follow the derivation of equations for fluid break-up in [1] without getting into too many technical details. The analysis of this problem has been mainly completed, and we will concentrate on giving the reader enough information to see the difference between this case and the thread break-up in case the ambient fluid is present. The reader is advised to refer to the original article in order to fill in the gaps in the derivation. Let us start with the derivation of the mass conservation at the point of breakup. We assume that the fluid is incompressible and the motion has radial symmetry with the axis of symmetry being z. By assumption, the velocity has only two components: vr, at the direction of the polar radius r (perpendicular to the thread's axis) and vz at the direction of z (along the thread's axis). To make further progress, let us identify the small parameter e inherent for this problem as the ratio of the typical scales in r and z direction. Thus, we re-scale r as r —> er, whereas the z coordinate remains unchanged: z —> z. Let us therefore assume the following expansion for the z component of the velocity: vz = v0(z,t)
+ e2r2v2{z,t)
+ ...
(1)
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
451
The incompressibility requires that WV
Id ~ rdr ^Vr'
+
dvz ~dz ~
'
^
which in turn yields an expansion for the r component of the velocity: Id
loot?
Suppose the equation for the thread is F(r, z, t) = h(z, t) — r = 0. The kinematic condition on the free surface requires that molecules which are on the surface remain on the surface, or DF/Dt = 0, where D/Dt is a full (material) derivative. This leads to
dh dt
dh dz
^7 + vz-
vr \z=h = 0
(4)
Integrating (2) with respect to rdr from r — 0 to r = h(z, t) and using (4) gives the equation of mass conservation -7T- + w (voh2) = 0.
(5)
Equation (5) is easy to understand, as the area of the radially symmetric thread at point z is given by S(z,t) = nh(z,t)2, so (5) describes simply the advection of volume of a radially-symmetric fluid element. The velocity of fluid VQ is still unknown. To complete the system, we must consider the z-component of Navier-Stokes equation, which connects VQ and V2- The full Navier-Stokes for three-dimensional velocity v in vector form reads: dv 1 — + (v • V)v = — V p + i/Av,
(6)
with p being pressure and v being kinematic viscosity. Let us consider the z-component of this equation and do appropriate rescaling of spatial coordinates (r —> er , z —> z), and velocities (1,3). Keeping only the terms of the lowest order in e we obtain dvo_ dt
dvo__ _ldp dz pdz
(. \
dhjo\ dz2 J
where p is pressure and v is kinematic viscosity. To connect pressure with the velocity components, we utilize the boundary conditions at the free
452
Vakhtang Putkaradze
surface. The normal dynamic boundary condition (vanishing of the stress normal to the surface) gives p = 7/c, where 7 is the surface tension coefficient and K is local curvature of the surface. Finally, the tangential dynamic boundary condition (vanishing of the tangential stress) yields the connection between VQ and v
3 dv0 dh
.
2
4 dz 2h dz dz Combining all these equations in one system, we obtain dv0 at
dv0 dz
7 d , , . ~v d fdv0 pdz h2dz \oz
f+!(**')=0.
2\
J
do.
For a thin slender thread the curvature is approximated as K = 1/h. However, it was demonstrated that because of numerical stability and other reasons, it is advantageous to keep higher order terms in curvature [4, 5]. The full expansion for curvature is 1 K = -
1 .
hy/l + e2h2z
2.2.
hzz
2
—g —.
7 1 + 62^
(11)
Analysis of t h e equations
We shall refer to (9,10) as Cosserat equations, which were first obtained by Green in 1976, with an inviscid version given by Lee in 1974, see [1] for references and history of these equations. (Technically speaking, equations (9,10) are only leading-order approximations to the Cosserat equations. However, we shall call them Cosserat equations for simplicity). The inviscid equations can approximate the solution quite well for some time and have interesting mathematical properties, such as linearization under Hodograph transformation and infinite hierarchy of conservation laws, see [6, 7] and references therein. To make quantitative analysis of the Cosserat equations, let us first introduce a length scale, constructed out of the physical parameters of the problem lv = ^ 2 p/7, and time scale tv = uzp2/j2. The ratio of a typical length scale L to the viscous length scale /„, F = L/lv, determines how important the viscosity is on different regimes. Numerical simulations of equations (9,10) reveal that there is a singularity forming in finite time. The exact timing of singularity, to, depends
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
453
on the initial conditions, but the structure of the solution on approach to the singularity has a universal self-similar pattern which we will briefly outline here [1]. The behavior of solutions observed in simulations and experiments is quite different for small and large values of F, and we will describe the difference below. Of course, as we have mentioned earlier, the viscosity is important no matter how small it is, but the effect of viscosity for a very large F shows only very close to the singularity. For large viscosities, F < 1 or even F ~ 1, the equations show excellent agreement with the experiment. Not only do the equations capture the results of experiments qualitatively, but excellent agreement in quantitative details is also achieved. Equations (9,10) correctly predict the length of the thin necking region and the shape and size of the drop [4]. These equations also show that the break-up of liquid threads close to the singularity is well approximated by a universal self-similar solution. To see how the self-similar scenario comes about mathematically, we suggest that, close to the singularity, all dimensional parameters of the problem may enter the solution only through the time and length scales tv and lu. Thus, we assume the following self-similar ansatz for radius and velocity describing the evolution of the thread close to the point of singularity z$ and timing of singularity to (here we use t' = (to — t)/tv and z! = (z-zo)/lv):
h{z,t) = t'
(12)
= (t')-V2z'
We shall only consider the solution for t < to (before breakup). The selfsimilar functions 4>(r]), ip{rj) satisfy
I»
+
#') + W = |
+
3 < ^
i ( * 2 + ^(*2)') + ( « 2 ) ' = o
(13) (I")
To specify the boundary conditions on the solutions of (13,14), let us notice that far away from the singularity in non-rescaled variables, both height and velocity should be finite. Non-rescaled variables z',t', which are close to singularity, correspond to very large values of 77. This is referred to as matching of inner solution (12) and outer solution, describing the evolution of fluid far from the singularity. The inner and outer solution must match,
454
Vakhtang Putkaradze
in particular, both the height and velocity of the inner solution should remain finite. Technically speaking, it is an assumption that both height and velocity should tend to a constant value (as this constant value can be zero). However, this assumption is well-justified by extensive comparison between the theory, numerical simulation and experiment. Thus, we posit >(v) -> A±V2,
77->±oo,
n
rx
The boundary conditions (47) close the system and allow one to find a solution of (13,14). Since the conditions on the solution are asymptotical, the solution is not unique and additional solutions satisfying (13,14) and boundary conditions (15) were found [8]. The universal law of scaling was later studied in more details in [9]. Notice that the inner solution satisfying (15) grows quadratically away from the singularity. Let us now investigate how the inner solution described by the self-similar ansatz merges with the outer solution. Close to the singularity, when the length scale of the self-similar solution becomes small, we can see that when 77 —> ±00, | ^ | ^ . /
W
±
(
z
_
2
o
)
,
(16)
so the outer solution must be quadratic close to the singularity. More precisely, Viter-^-^o)2-
(17)
Thus, the outer solution for the neck's width h(z, t) consists of two parabolas with (very) different curvatures, joined smoothly at the singularity. It is interesting that in this particular case the microscopic processes driving the inner solution specify the asymptotics of the outer solution (17), whereas in most standard cases, it is the large-scale processes at the outer solution which are driving the asymptotic of the inner solution. The process we have described is universal and is observed in any fluid rupturing in air, provided that the process of rupture does not develop an instability. However, for very small viscosities, this universal self-similar process of rupture can be hard to observe, as the viscous length scale lv can become extremely small. On the scales much larger that lu, the outer solution forms conical shapes. To describe the flow of fluid in conical geometry, we shall later introduce a self-similar ansatz separating radial and
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
455
angular variables. While that description lacks the power of Cosserat equations describing the dynamics of rupture, it will introduce the contribution of nonlinear terms in a consistent way. However, before we do that, we will describe another set of experiments showing rupture of a thread of one fluid inside another fluid, when the appearance of conical interfaces is more pronounced.
3. 3.1.
PHYSICAL EXPERIMENTS SHOWING CONICAL INTERFACES Lava lamps: break-up of a thread of one fluid in another fluid
We see that in the case of a liquid bridge rupturing in air close to the singularity, the slender-body approximation works excellently. We shall now turn our attention to the case when a liquid bridge of one fluid is rupturing while being surrounded by another fluid, which is immiscible with the first. A very beautiful demonstration of this process, familiar to everybody, is the lava lamp. In the experiment (or the lava lamp), a lighter fluid is slowly rising through heavier highly viscous fluid. At some point, the rising bubble of lighter fluid separates so far away from the bulk of it that a neck is formed, which is then ruptured by the surface tension. Watching a lava lamp, we notice that the process of forming a neck is extremely slow, yet the process of rupturing is extremely fast, even when the viscosity of surrounding fluid is high. This makes recording the exact moment of break-up challenging; but these difficulties were successfully resolved in [2]. The photographs taken at the very moment of break-up show the formation of two cones with a common tip and axis of symmetry. Let us suppose that fluid i, i = 1,2 has kinematic viscosity ?].;. The angle of the cones depends on the ratio of viscosities A = v\jvi) but is practically independent of other parameters of the experiment. However, a recent study [10] showed that the break-down of a droplet is not self-similar for certain fluids (water and silicone oil) and remembers the initial as well as boundary conditions. Careful experiments and numerical simulations showed formation of exceptionally thin and slender threads of fluid (with the diameter of the order of several nano-meters with length up to a few millimeters) bridging the gap between the two conical sections drifting apart.
456
Vakhtang Putkaradze
The theory of the phenomenon of two-cone formation was developed by [2] and [11] (see also [12] for the numerical analysis of the process of break-up of a thin liquid thread). The theory and simulations are based on assuming Stokes' flow (low Reynolds number) on both inner and outer fluid. Based on the integral re-formulation of the motion of the interface, the theory provides a dynamic way of predicting the interface position, along with velocity field in space. The formation of two cones at the instant of the drop pinch-off is confirmed by the numerical simulation. The cones' angles also compare favorably with experimental data. In what follows, we shall develop an exact solution of two fluids separated by an interface in the shape of a cone (or two cones) without assuming that Reynolds number is small. We are motivated by the following argument: if the shape of the interface is exactly conical very close to the origin, there is no typical length associated with it, and, as a consequence, no Reynolds number can be defined. We show that it is possible to find exactly self-similar (in space) stationary solutions respecting the conical geometry and taking into account both linear and nonlinear terms. However, in searching for the stationary solutions of this type we sacrifice the dynamic description of the interface afforded by the Stokes flow. Thus, our solutions should be considered as supplementary to those found in [2] and [11]. In our development of exact conical solution, we shall first start with the case when only a single conical interface was present. While we show that mathematically, no exact solution of this problem satisfying the requirements of regularity at the cone's axis and all boundary conditions at the interface can be found, such flows are also of interest as the interface in a shape of a single cone is observed in the experiment of two-fluid selective withdrawal [13, 14]. In the selective withdrawal experiment, a lighter fluid (oil) is withdrawn through a tube which is positioned at a certain elevation from the interface between the lighter fluid and heavier fluid (water). The interface between two fluids is deformed by the flow. For small values of withdrawal rate, only the lighter fluid enters the tube; whereas for high enough values of withdrawal rate, both fluids are withdrawn. At some critical value of the flux, stationary flows are observed for which the interface assumes the shape of an almost perfect cone. The rounding of the cone tip is much smaller than the capillary length, and the appearance of such singular conical shapes makes the phenomenon interesting for our studies. It has also been demonstrated that it is possible to use the thin slender
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
457
threads observed close to the bifurcation in two-fluid withdrawal regime to coat microscopical particles used in medical applications [15]. The theory of this phenomenon in the regime of two-fluid withdrawal based on the assumption of slender thread entering the tube is also in preparation [16]. 4. 4.1.
FORMAL DESCRIPTION OF CONICAL FLOWS: SELFSIMILAR SOLUTIONS General considerations
Motivated by appearance of conical interfaces in the two-fluid flow experiments, we shall derive exact solutions of Navier-Stokes equations describing the flow of two immiscible fluids, with the interface separating the fluids being a cone. Solutions of this type have velocities growing without bound at the conical tip as 1/r, where r is the spherical radius from the tip of the cone. Therefore, stress tensor components diverge as 1/r3 at the tip of the cone. Since the curvature of the cone is growing as 1/r at the tip, forces induced by surface tension diverge as 1/r2, i.e., slower than viscous forces. Thus, this solution persists close to the tip of the cone even if the surface tension is introduced. Note that far way from the tip the conical interface will be deformed by surface tension (and gravity in the case of density difference). If r = 0 were the only singularity of the flow, such two-fluid conical solution would be very interesting, not only as a model of what happens in selective withdrawal close to the tip of the cone, but also as an example of a physically realizable solution with finite energy exhibiting singularity in space. Indeed, if the singularity would be present only at r = 0, one could in principle set up an experimental apparatus with prescribed flow at the boundaries, say r = R leading to the singular solution. Unfortunately, as we show below, our exact solution must exhibit infinite velocity not only at the tip of the cone, but also at the all of the cone's axis. We show that the solution without singularities at the axis can only be achieved with unphysical fluids having negative viscosity or density ratio. The singularity at the axis is typical of the three-dimensional solutions of single fluid of this type. A swirling motion can be successfully incorporated into this ansatz as well; however, we shall not use it here, since the formulas become extremely complex and analytical treatment of two-fluid flow is almost impossible. The situation is similar for the case of two-fluid two-cone flows. We
458
Vakhtang Putkaradze
assume that the interface is located at two positions 9 = B\ and 9 = 02, which describes two coaxial conical surfaces. In this case, it is still impossible to achieve a solution with no singularity at the axis 9 = O,vr. Nevertheless, the two-cone flows are still of interest and we shall outline their structure as. The flows described in this section are natural extensions of the singlefluid flows bounded by conical interfaces, which are a class of analytical solutions of Navier-Stokes equations [17] - [21]. One of the important features of the conical flows of this type is that a singularity on the axis of the cone is ubiquitous. Much work has been devoted to the explanation of this singularity. In particular, it was recently proved rigorously [22] that specifying either tangential stress or pressure on the interface makes the regular solution unique. However, it seems to be impossible to satisfy all boundary conditions and still obtain a regular solution. No such result was available for the case of two-fluid flow until now. We will also show that it is possible to derive an exact two-fluid conical solution of Navier-Stokes equation which does not have a singularity on the axis if one of the boundary conditions on the interface is violated. For example, a solution can be achieved if a slip boundary condition on the interface between two fluids is assumed, and we will describe such a solution below for the case of two cones. A solution for a single cone is possible as well, but will not presented here, as it is a simple technical extension of the two-cone case. Instead, we shall choose a more pedagogical approach. First, we discuss the impossibility of a regular solution for the single-cone case satisfying all the regularity requirements. Then, we show that if a slip condition on the interface is assumed, a two-cone solution is possible. We shall just mention that a single-cone regular solution with a slip on the interface is possible (the proof of this fact is relatively easy), but a twofluid conical solution satisfying all the regularity requirements is impossible (which is rather tedious to show) [23]. We shall begin with a derivation of equations, common for both single - and two-cone solutions. 4.2.
Derivation of equations and boundary conditions
We begin by introducing the spherical system of coordinates (r, 9, 0), with the origin being at the cone's tip, and 9 = 0, IT axis aligned with the axis of the cone. Two fluids denoted by a subscript j — 1,2 are separated by
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
459
the interface at 6 — #*. Let us assume that fluid with subscript 1 occupies the area above the interface, i.e., fi < /z*, and fluid 2 is in the area /z > /z*. We assume the velocities of each fluid using the ansatz vr, = VjFjW/r
vej = VjfjW/r
v^ = T^/r,
(18)
where Uj are kinematic viscosities of each fluid and Fj, fj are dimensionless. In this section, we shall put Tj = 0 (no swirling), greatly simplifying the formulas. This ansatz leads to a first order equation of Riccati type, first obtained by Slezkin [24]. We shall repeat the derivation of the Slezkin equation briefly, since the presence of the interface modifies the final result and requires the expressions for stress tensor components in order to match them at the boundary. Incompressibility condition 1 9 , , 1 d . . . x r2Qr
\
>J> rsmQdO
J
leads to a relationship between F.j{9) and fj(6):
which can be simplified introducing the new coordinate /J, = cos 9 and
The incompressibility condition is then simply
If we now define pressure in each fluid as 2
pj(r,n) = ?i£-Pi(li),
(20)
from the Navier-Stokes equations (^-component) we obtain the pressure in the form
Pl
= f-\JLd[i
2 1 - [il
+ Aj,
(21)
where Aj is an integration constant. Substituting (20, 21) in the r-component of Navier-Stokes equations and integrating twice, we obtain the full version of the Slezkin equation for g(fi): (1 - /z 2 )^ + 1
m
+l-g]= -AjU2 + 2BjH + Cr
(22)
460
Vakhtang Putkaradze
Two new integration constants Bj,Cj are introduced in addition to Aj from (21). To solve (22) in each fluid, we have to specify g.j at the interface for each fluid, as well as three unknown constants Aj, Bj, Cj, which gives a total of eight unknowns. In addition, the position of the interface /z* is unknown. In order to complete the system, we have to find the corresponding number of boundary conditions. The kinematic boundary conditions are the continuity of the r and 9 components of the velocity. Moreover, since the flow is stationary, vgj must vanish at the interface 9 — 6>*, or ji — /z*. Thus, the kinematic boundary conditions on the 9 components of velocity are
(23)
The continuity of the r-components of velocity gives Vl(M*) = 02 (/•*»)> which, in view of (22) and (23), can be written as
A(-A lt 4 + 2Blfi* + d) - {-A2/4 + 252/u* + C2).
(24)
Here, we introduced the viscosity ratio A = V\jv2. Let us now compute the dynamic boundary conditions, which are the continuity of tangential arg and normal ogg components of stress tensors. Using the velocity ansatz (18), expression for the pressure (21), equation for g (22) and vanishing of g on the interface (23) after some algebra leads to the following answer: _
P]v)
/ - 2 % i . + 2£?A
a
«* - ~ ^
\—I-M2
2 P
-^{-g'^)
) ' + A3).
(
}
(26)
If we introduce 5 — {viIv%){p\/Pi), the dynamic boundary conditions at the interface become 8(2Axn* - 2Bi) = 2A2li, - 2B2
(27)
for the rB component and
KJM
~ M) = 92(»*) ~ M
for the 99 component. Again, the 90 stress equation can be simplified using (22), becoming 5(Ai - 2J3i^ - Cx) = A2-
2B2[i* - C2.
(28)
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
4.3.
461
Single cone: regularity requirements and singular solutions
Equations (22) together with the boundary conditions (23, 24, 27, 28) specify five equations for nine unknowns. Additional equations can be obtained if certain conditions on regularity of solution are imposed at /j, = ± 1 , the axis of symmetry of the cone. It is clear that vg must be zero on the axis [i = ± 1 . In addition, we require that vr is bounded when \i —> ± 1 . Thus, we infer regularity conditions on g.y. gi{n) -» 0
g[(fi) bounded
/j, ->• - 1
(29)
g2(/i) -> 0
g'2(n) bounded
/i-»l.
(30)
We shall now show that enforcement of these regularity conditions leads to the unphysical value for the interface position //», and so regular solution to this problem cannot exist. From Slezkin's equation, regularity conditions on g\ (29) is achieved only if the right-hand side of (22) is proportional to (1 + //) 2 , i.e., Bx = -A,
d = -A,.
(31)
Similarly, regularity conditions on g2 require that B2 = A2
C2 = -A2.
(32)
Then boundary conditions (24) and (28) become \Ax{\ + ii\) = A2{\-[il)
(33)
<JA1(l + //,) = A 2 ( ^ - l ) ,
(34)
and boundary condition (27) becomes identical with (28). Dividing (34) by (33), we find the following expression for the position of the interface
/x* =
T
r.
(35)
0 — A
Clearly, for A > 0 and 5 > 0 we have |/i*| > 1. But since /i* = cos#*, I/i* I < 1 must be satisfied. This shows that we cannot find a solution to a single-cone problem satisfying regularity requirements at /x = ± 1 . On the other hand, if we impose weaker regularity conditions, removing the requirement on the boundedness of g'j(fi) at ji = ± 1 , we can now find
462
Vakhtang Putkaradze
Figure 1: An example of single-cone solution with g'(fi) having logarithmic singularity on the axis /j, = ±1.
a solution. Instead of four conditions imposed by (31) and (32) we have just two conditions: at \i = 1. -Ax
+ 2Si + Cx = 0.
and at /i = — 1. - 4 2 - 2 5 2 + C2 - 0. Since we have dropped the number of constraints by two, we can choose the value of the interface angle /z* arbitrarily. We see that a solution is possible for any /i*, and is parameterized by two additional parameters Ai and ^42. Both vr and v$ components diverge as 1/r as r —> 0, and vr diverges when //—>±1, or#—>-0,7r. The total energy is finite in any ball of finite radius. Thus, we see that it is impossible to have a stationary solution with exact conical singularity at the origin, while remaining regular everywhere else. However, we can still interpret the solution described in this section as the outer solution of the problem, which requires regularization close to the origin, presumably at length scale less than lv. The regularization can also smoothen out the apparent singularity at the symmetry axis \x = ± 1 . Let us note that there is no condition on the critical angle /i*. This, perhaps, is physical as the inner solution at the origin determines this angle. An example of solution g{\i) with the singularity (/•*) ~ log(|/i. ± 1|) when /i —> ± 1 is shown on Fig. 1. Another way to obtain a meaningful solution is to drop one of the boundary conditions at the interface. We shall derive equivalent solutions for the
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
463
two-cone case under the assumption that the continuity of tangential velocity is violated. The single-cone case can be solved in a similar fashion. In this case, the regularization in the form of a thin shear layer at the interface is required.
4.4.
Two cones: Regularity requirements and singular solutions
The theory of self-similar conical solution can also be applied to the case when the interface consists of two cones having a common axis. In this case, the interface is positioned at \i = ji\ and [i = ^2, so we must consider three domains. Fluid 1 is at —1 < // < n\ (domain 1) and \ii < \i < 1 (domain 3), whereas fluid 2 is ii\ < {i < ^ (domain 2). In each domain, equation (22) holds. As we mentioned before, a regular solution satisfying all the boundary conditions at the interface is impossible. In our further study of twocone flows, we shall assume that the fluids can slip at the interface. The continuity of velocity does not have to be enforced then. Being forced to drop one of the boundary condition, we have decided to admit the discontinuity of tangential velocity at the interface as it is perhaps the least dangerous sacrifice, at least from the physical point of view. First, one can assume that the addition of certain surfactants can allow the fluids to slip. Second, even if the true boundary condition is no-slip, we can still imagine a thin boundary layer between the fluids which accommodates the difference of tangential velocities. In this boundary layer, the flow is not self-similar, but outside this thin boundary layer the flow is well described by our self-similar ansatz. The problem of the flow of a plane jet consisting of two immiscible fluids was considered recently by Herczynski et al [25], and an extension of this work to conical geometry is possible. On the other hand, if we drop continuity in tangential or normal stress at the interface, we see no physical way to explain that a finite force is acting on an infinitely thin (or even finite, but very thin) layer of fluid adjacent to the interface. Thus, we shall assume that the continuity of tangential velocity is violated and proceed with the calculation. Notice that fluid 2 can now have singularities at the axis \i = ±1, as it does not extend there. Thus, for fluid 2 there is no condition on A^^B^ and Ci- Let us first seek solutions which are regular at /i = ± 1 . In domains 1 and 3, having regularity condition at the axis requires that Bx = -Ax
Ci = - A i
(36)
464
Vakhtang Putkaradze
at fx = — 1 (domain 1) and B3 = A3
C3 = -As
(37)
at fj, = 1 (domain 3). Similar to the one-cone case, on each interface H = nu i = 1, 2, we enforce the continuity of normal and tangential stress as follows. For the r6 component we have now 5(2Alfjll - 2Bi) = 2Am - 2B2 [M>
5{2A3ii2 - 2B3) = 2Am - 2B2. The continuity of 69 component requires
5{Al - 2Bm - d) = A2- 2B2JM - C2 [ 5(A3 - 2B3fi2 - C3) =A2- 2B2[i2 - C2. ' Equations (37,36,38,39) specify eight equations for eleven parameters A t , Bi and C% (i = 1,2,3) and the angles fj,\^2. Additional constraint comes from the fact that there is a solution of (22) in domain 2, which we call g2(/j,), must solve boundary-value problem 92^ = Mi) = 0
92^ = to) = 0.
(40)
Thus, we expect the final solution to depend on two parameters, let us say Ai and n\. We have also assumed that the fluids are given, which fixes the value of 8. Analytical progress can be achieved under the assumption that in domain 2 (i.e., in ambient fluid), the flow is sufficiently slow, so nonlinear terms g2/2 in (22) are small. Then, general analytical solution of this linearized version of (22) is given as
g2(ri = C(l-ii2)
+ A15F(ri,
(41)
where C is arbitrary for now and F
^)
=
I ^U(^ TX ( ^ - J ) - ^ v (Mi + 1)0^1 -to)
+ ^ + 2^2M+
(l + Mi)(M2-l)(l-/i 2 )log[^]).
(42)
Then, the condition (40) yields a homogeneous linear system for C and Ai, which is solvable if and only if the determinant A of this system is zero, i.e., A( M l , fi2) = 6 [(1 - f4)F(to)
- (1 - vl)F(vi)}
=0
(43)
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
465
Figure 2: fi2 versus n\ for linearized flow in ambient fluid (domain 2).
with F(fi) given by (42). It is important to keep in mind that if second fluid is absent, S = 0 so (43) does not pose any condition connecting fi\ and /i2-
Notice that the solution of (43), defining an implicit relationship between (i\ and [i2 is no longer dependent on the parameter 5 for 5 ^ 0 (and, naturally, on the amplitude A\), making it very convenient for analysis. Equation (43) has a trivial root ^2 = A*i, and another non-trivial root which is of interest to us. We present the dependence H^i^i) of Fig- 2. As is clear from this figure, (43) can only be solved for a limited range of — 1 < Hi < 0.30. Note that a solution described here was obtained for linearized flow only in ambient fluid (domain 2). In domain 1 and 3, the flow remains fully nonlinear. Let us finish our analysis by showing that a fully nonlinear solution can be obtained as well. If we do not assume linearity of (22) in domain 2, we cannot connect boundary conditions at fi = HI and /i = [ii easily and a fully nonlinear numerical solution must be sought. (We shall mention here that a solution of (22) can be sometimes expressed in terms of hypergeometric functions, but such analysis is more complex than the one presented here and does not add in either clarity or simplicity). Given 5 and for any A\. Hi and /J,2, we can find the rest of the coefficients A t , B{ and Q and proceed by shooting from ji = Hi to /i = ^i- Then we numerically find Hi such that 52(^2) = 0. Notice that we have to repeat this process iteratively, as
466
Vakhtang Putkaradze
Figure 3: Nonlinear solution g{ji) satisfying regularity conditions at fi = ±1.
the coefficients A,, B{ and d depend on fi2. It is interesting that while the solution of the iterative process does depend on A\, /ii and 5, the main part of the solution is already correctly predicted by the linearized result in Fig. 2. The difference between the linearized and fully nonlinear solution for (i2 never exceeds 1 %. The limit of 5 —> 0 is of interest as well, as it describes the conical flow with free surface when the second fluid is inviscid and not moving, or simply absent. This limit also give us a physical insight into the region of validity of the linearized solution. In Fig. 3 we present the result of fully nonlinear solution for A\ = 1, 5 = 0.2,/xi = —0.3. The oil-water mixture frequently used in the experiment has 5 ~ 10~2, however we have used a somewhat larger value of 8 in Fig. 3 for demonstration purposes. As S —> 0, the solution in domain 2 converges to 0 uniformly. If we assume that A\ remains of order one, it can be seen from the boundary conditions that A2, Bi and C2 are of order 5, Then, necessarily, g2 = 5G2(fx) + O{52) where G2 is of order one and satisfies the linearized equation (22). We then proceed analogously to our discussion of the linearized solution in the domain 2: an exact solution of this linearized Slezkin equation can be found, yielding the same connection between \x\ and \i2 as shown on Fig. 2. The behavior of solutions for large values of Ai, A\ ~ 1/5, remains unclear but will be addressed in future studies. Presumably, this question is of mostly academic interest, as large values of A\ will presumably correspond to unstable flows. However, since solutions described here are exact solutions of Navier-Stokes equations, a detailed information about the behavior in all regimes may be of importance for checking advanced numerical codes for simulation of fluid flows with an interface. Having completed the study of conical flows, let us now turn our at-
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
467
tention to a similar problem of radial two-fluid flow in two dimensions [26]. Also in this case, a complete analytical solution of the problem can be found. Such a solution can be realized when a point where two cones merge is artificially stretched out in space (say, by the effect of external forcing) so the interface between two fluids will now consist of two sheets joining along a line. 5. 5.1.
TWO-DIMENSIONAL TWO-FLUID FLOWS General Considerations
The physical picture considered here is the following: one fluid is being injected through a line source into another fluid. Such flow may be realized when a cracked oil pipe leaks oil when being submerged in water. We shall consider a slightly more general case: two viscous, incompressible fluids being ejected from the same point source in a plane. Two fluids are assumed immiscible and are separated by an interface, which is denoted by the two straight dashed lines originating at the source. We assume that the fluid j , j = 1, 2 is flowing with the flow rate Qj and has density pj and kinematic viscosity Vj. The effect of gravity is neglected. The only case relevant for practical applications is Q\ ^ 0, Q2 = 0. It appears, however, that arbitrary Q2 ^ 0 can be considered without extra effort, and it introduces valuable insight into the structure of the solutions. Thus, in the rest of this section we shall consider arbitrary values of Q\ and Q2The flows developed here are based on the classical Jeffery-Hamel solutions, which describe the outflow and inflow of viscous incompressible fluid in a linearly expanding channel (wedge) with a given angle between solid walls, as formulated by [27] and [28]. One dimensionless parameter of the problem is R = Q/v, Q being total flux and v being kinematic viscosity, and another dimensionless parameter is the angle of the aperture, a. Behavior of the solutions is very interesting. It was shown in [29] that there is an infinity of solutions and each solution undergoes a bifurcation at some R = Rc, where the critical Reynolds number Rc depends on the number of oscillations in the velocity profile of the solution and the angle of the wedge. In spite of a long history, many important results concerning the structure of Jeffery-Hamel solutions and their stability are quite recent, see [30]-[39].
468
Vakhtang Putkaradze
Several generalizations of Jeffery-Hamel flows have been made. An extension describing the flow of fluid in free space without boundaries was first considered in [40] but the analysis was not complete. Goldshtik and Shtern [32] were the first to point out the exact bifurcation values: it was shown there that there is an infinity of solutions with fc-fold symmetry, k = 1,2,... and each such solution exists if R < vr(fc2 — 4). This type of flows was subsequently analyzed in [33]. Unaware of that earlier work, Putkaradze and Dimon [41] re-derived the same results using analytic structure of solutions (Weierstrass P-functions). The analytical structure of solutions described in [41] will be crucial for exhaustive description of two-fluid case, as we demonstrate below. Another generalization of JefferyHamel solutions which includes a vortex and a sink at the origin was first done by Goldshtik, Hussain and Shtern [33]. Voropayev and co-authors extended Jeffery-Hamel flows to include a quadrupole at the origin, see [42] [45]. Bourgin and Tahiri [46] studied Jeffery-Hamel type flows for the case when velocities on the interface are prescribed. Moffatt [31] posed different boundary conditions at the surface of the wedge and solved Jeffery-Hamel flow in the limit of small Reynolds numbers. Jeffery-Hamel flows were also used as a basis for the study of fluid flows in channels with curved walls in [47] and [48].
5.2.
Equations and Boundary Conditions
The derivation of governing equations in this section is similar to that for Jeffery-Hamel flows. We shall therefore only go through the derivation briefly. For details, see [24]. We assume that two immiscible, incompressible fluids are ejected from a point source in two dimensions. The fluids have kinematic viscosities v\ and V2 and densities p\ and pi. If we assume that the fluids are flowing purely radially, i.e., ty, = 0 in each fluid, the incompressibility condition dr(rvr) = 0 gives vrJ = u ^ .
(44)
Here and below, the subscript j = 1,2 indexes the fluid. The prefactor v:l in (44) makes the velocity function Uj((f>) dimensionless. For convenience, substitute
uj(<£) = - 6 / j ( 4 0 - 2 ,
(45)
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
469
then each of the functions fj(
ff = Aft-arfi-bj,
(46)
where the constants a,j and bj, j — 1,2 are to be determined to satisfy boundary conditions at the interfaces. The prime denotes the derivative with respect to 0. We assume that the interface is located at (p = —a/2 and (p = a/2. If we denote A = vi/1/2 the viscosity ratio and 5 — A2pi/p2, the boundary conditions at the interface become: A(6/i(-a/2) + 2) - 6/ 2 (-a/2) + 2 A(6/i(a/2) + 2) = 6/ 2 (a/2) + 2 <J/{(-a/2) = / 2 ( - a / 2 ) 5f[(a/2) = /2(a/2) <J(3ai - 4) = 3a 2 - 4
(47)
/_°52(-6/i( = —«o/2 and 4> = OLQ/2. We have discovered, however, that the answer strongly depends on the values of a 0 and chosen initial profile, showing that several solutions of this problem are possible. Thus, even though the continuation method is a possible way to obtain a two-fluid Jeffery-Hamel solution, it does not necessarily converge to the desired solution. In the remainder of this section, we present an alternative method of studying the solutions, based on the Weierstrass Elliptic
470
Vakhtang Putkaradze
Functions. The method allows the complete classification of the hierarchy of solutions. In addition, our method yields a type of solution which is impossible to obtain by Newton's method with continuation (SE solutions, see below). The idea of our method consists of finding the solutions satisfying the boundary conditions at (j) — Oi/2 automatically provided they satisfy the boundary conditions at
5.3.
Analytic Properties of the Solutions
Equation (46) can be interpreted as a description of motion of a particle in a cubic potential U(f) = — 4 / 3 + a,jf + bj, which shows that the solutions of (46) are periodic with some period, let us say pj, with pj being real. Upon the change of variables 4> —> i<j), equation (46) describes motion of a particle in a cubic potential —[/(/), which shows that the solution fj(
j = 6 0 E' n , m ( n Pj + m 2iT3y4 bj = i4OTltn,m(npj+m2iTj)-6,
(AS) K }
where the prime denotes that the summation is taken over n, m such that n2 + m2 > 0. In general, the periods may be two arbitrary linearly independent complex numbers. The solution of (46) can be represented either in terms of invariants, i.e., a.j and bj, or in terms of periods, i.e., pj, 2iTj. These representations are completely equivalent from the mathematical point of view. Dependent on the case, however, it is advantageous to use either the invariants or the periods representation. To distinguish between these representations, we shall denote f(z) = V{z; a,j, bj) if f(z) is represented in terms of invariants a,j and bj, and use curly brackets after the semicolon f(z) = V(z; {pj, 2ITJ}) if f(z) is represented in terms of periods pj, 2%rr Also, we always explicitly mention which representation is used in each particular case to avoid confusion.
471
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
Figure 4: Velocity function u(
5.4.
Satisfying boundary conditions at 0 = —a/2 and (j) = a/2
We are now ready to describe different types of solutions obeying boundary conditions (47). As was mentioned earlier, we seek solutions satisfying boundary conditions at (p — a / 2 provided that boundary conditions at 0 = —a/2 are satisfied. The first possibility is to seek solutions f\{4>) which are periodic with the real period p\ = a/k\ and f2{4>) with the real period p2 — (2TT — a)/k2 for some integer numbers k\ and k2. Using our mechanical analogy, solutions of this type represent a particle oscillating in a cubic potential [/(/) = - 4 / 3 + a,/ + bj. The velocity function u(>) = - 6 / ( 0 ) - 2 for such solution is illustrated on Fig. 5.4. We distinguish different periodic solutions by their periodicity. We denote solutions with fi(4>) being k\ periodic and f2(4>) being A;2 periodic as P(ki)P(k2) (periodic-periodic). In the analysis of P{k{)P{ki) solutions we will use the period representation. The P{k\)P{k2) solutions are given by fcW)
= V{<\> + ir, + 9h {PJ,2IT3}).
(49)
The unknowns are 6j (phase-shifts) and Tj (imaginary periods). Thus, it is advantageous to use the period representation for P(k\)P{k,2) solution. The second possibility to satisfy boundary conditions at 0 = ± a / 2 is to require the flow to be symmetric with respect to the reflection around t h e line (f> = 0, more precisely, f\(-
- 4>) = h{^
+ 4>)-
Since there is no condition on the periods, the representation of solutions in terms of invariants a,j,bj is advantageous. Solutions of this kind are given by: fx{4>) ="P(0 + ZTi(ai,6i);ai,&i) / 2 (0) = p( 7 r + 0 + ? r 2 (a 2 ,6 2 );a 2 ,6 2 ).
l
'
472
Vakhtang Putkaradze
The unknowns are the invariants a-j and bj. The imaginary half-periods Tj(a,j,bj) can be expressed in terms of the invariants (a,j,bj) by inverting (48). Even though no requirements on real periods are made, we shall also enumerate these solutions in terms of two integer numbers, k\ and &;2. This implies that the real periods pj satisfy fci -
^! —
fci+1
/t;i\
We shall denote such solutions E(k\)E{k2) ("even-even"). In the study of E(h)E(k2) solutions, it is best to use invariants as unknowns. The third possibility is to have one of the solutions fj{4>) be singular. From the properties of equation (46) we conclude that singularities of /,(>) are periodically spaced second-order poles: f3 ~ (z — npj)~2, where n is any integer number. These solutions are physical, if the singularity occurs outside the domain of the solution, i.e., in another fluid. For example, if f\{4>) is such singular solution, we need the singularities of f\(4>) to lie outside the region —a/2 <
, , l
;
Again, the unknowns are the invariants dj, bj and a. There are two more possibilities to satisfy the boundary conditions automatically, which correspond to degenerate types of solutions (in the sense that they are realized for the set of parameters (R\, i?2) of measure zero). The fourth possibility is to have one of the functions fj constant, and another function oscillatory. One can see that if, for example, / 2 = const,
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
473
Figure 5: Velocity function u(4>) in the singular domain of SE(k2) solution. The poles of velocity are occuring outside the physical domain
/i(0) must be periodic with the period a/k\. We shall call these solutions P{k\)C solutions if f2 = const and CP{k2) if A = const. Finally, the most trivial solution possible is the one with f\ = const and f2 = const. We shall refer to these solutions as CC solutions. As was demonstrated in [26], three cases P{k\)P{k2) , E{k\)E(k2) , SE(k2) are realized for the set of parameters (i?i,i?2) of positive measure. Other types of solutions, involving at least one C in their notation, can only be realized for sets of codimension at least one in the parameter space (i?i,jR2)- Therefore, the analysis of P(ki)P(k2) , E{ki)E{k2) and SE(k2) type of solutions is physically most relevant. However, we shall mention P(k\)C and CP{k2) solutions later as they play an important role in bifurcations. To specify a two-fluid Jeffery-Hamel solution, we must first determine the type of solution {P{k\)P{k2) , E{k\)E{k2) or SE(k2) ) as well as periodicity of the velocity profiles k\ and k2. Once these are chosen, we can analyze the behavior of the system on the parameter plane (Ri, R2). The most interesting bifurcation diagram is exhibited by the P{ki)P{k2) solutions. The introduction of an extra Reynolds number as a free parameter yields a highly complex bifurcation picture, typical of bifurcations of degeneracy two [50]. In particular, if we take a line through the parameter space (Pti, R2) (denoted by Reff) and compute the value of u'(a/2) (proportional to the shear stress on the interface), we obtain a boomerang-shape diagram typical for the umbilical catastrophies [50], as shown on Fig. 6. Such a bifurcation diagram is only understandable if two parameters Ri and R2 are considered. The boomerang structure persists even if we take the line R2 = 0 in the parameter plane, corresponding to no flux in the second fluid (which is the physical regime). Such 'ghost entrance'
474
Vakhtang Putkaradze
Figure 6: The value of non-dimensional shear stress on the interface u'(aj2) taken along a line on the parameter plane Ri,R2.
of the second parameter R2 into this problem is highly non-trivial and unexpected, and shows that the free surface affects the flow beyond the apparent complexity of the formulas. As we have seen, we must continue the study into the physically unrealizable regime R2 ^ 0 in order to fully understand both the physics and mathematics behind of the flow. The bifurcations of E{k\)E(k,2) and SE(k2) solutions, while not showing the same umbilic character of the bifurcation as the P{k\)P{k2) , are still sufficiently complicated. It was shown in [26] that E(k\)E{k,2) solutions, as well as P{k\)P{k2) solutions, exist on large areas of the parameter plane (Ri,R2), whereas SE(k2) type can only be found in relatively small (in measure) areas. The boundaries of the existence domains are formed by the P{k\)C (or CP{k2)) solutions, or, alternatively, by merging of a solution with its mirror image. Finally, one can demonstrate that the SE(k2) solutions cannot be obtained by the method of continuation in principle, as the singular part exhibits a pole-type singularity on real line. While this singularity is hidden by the fact that it appears in the domain of the second fluid, it must appear on real line if the starting point of the continuation is the two fluid being identical (i.e., no free surface). It is essential therefore to employ the method of P-functions described above to achieve a complete classification of solutions. We shall also note that this method includes, as a particular case, the solutions when second fluid is inviscid and not moving or simply absent. In
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
475
this case, only solutions of the periodic or even types are possible (the singular solution must necessarily have poles in the physical domain, yielding infinite flux, and must therefore be ruled out). Such solutions (fan jets) may play important role in some applications. The question of the selection mechanism for the two-fluid solutions is still unresolved. It may be argued that the E(1)E(1) solution looks like two parabolic profiles joined together and therefore is the most probable one to be chosen by the flow. It is also conceivable to conjecture that the profiles with high number of oscillations and alternating regions of strong inflow and outflow are unstable due to Kelvin-Helmholtz-type instability. While these arguments are plausible, a convincing argument for selecting one of the solution is still missing. Moreover, it has been demonstrated [51] that the stability of radial flows strongly depends on the upstream and downstream boundary conditions on velocity perturbations. Thus, perhaps it is possible that several solutions can be realized by altering the upstream and downstream boundary conditions in numerical simulations or experiments. 6.
CONCLUSIONS
In this chapter, we have outlined the role conical flows may play in the process of droplet break-up in two and three dimensions. While most of the previous work in droplet fission is based on the Stokes approximation, which allows treatment of moving interface, our approach utilizes the Landau-Squire or Jeffery-Hamel exact stationary solutions of Navier^-Stokes equations. Thus, our approach cannot treat moving interface and should be seen as complimentary to the previous studies using Stokes methods. Nevertheless, some of the interesting effects caused by the combination of inertia and free surface (multiplicity of solutions, coexistence of inflows and outflows, intriguing bifurcation diagrams, etc) should persist when time dependence is taken into account. The exact introduction of time dependence is impossible in the purely analytical framework outlined here. Thus, numerical simulations of the full Navier-Stokes equations and detailed measurements of velocity profiles in experiments are of paramount importance for future progress.
476
Vakhtang Putkaradze
7. ACKNOWLEDGEMENTS The author acknowledges fruitful and inspiring discussions with W. Zhang and J. Eggers. The idea of using two-fluid radial flow (originally in Stokes approximation) was suggested to the author in a discussion by H. Stone and M. Brenner. This work was partially supported by Petroleum Research Grant (Type AC). The hospitality of the Department of Mathematics, Ibaraki University (Mito, Japan), where this work was completed, is greatly acknowledged, as well as financial support of the Japanese Society for Promotion of Science (JSPS). REFERENCES [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33]
J. Eggers, Rev. Mod. Phys., 69 (3) (1997) 865. I. Cohen, M. Brenner, J. Eggers, and S. Nagel, Phys. Rev. Lett., 83 (6) (1999) 1147. X. Shi, M. Brenner and S. Nagel, Science, 265, (1994) 219. J. Eggers and T. Dupont, J. Fluid Mech., 262 (1994) 205. F. J. Garcia and A. Castellanos, Phys. Fluids, 6 (1994) 2676. J. Hoppe Lectures on integrable systems Springer, Berlin 1992. J. Keller and M. Miksis, J. Fluid Mech, 232, (1991) 191. M. Brenner, J. Lister and H. Stone, Phys. Fluids, 8(11) (1996) 2827. A. U. Chen, P. K. Notz and O. A. Basaran, Phys. Fluids, 8(11) (2002) 2827. P. Doshi, I. Cohen, W. Zhang, M. Seigel, P. Howell, O. Basaran and S. Nagel, Science, 302 (2003) 1185. W. Zhang and J. Lister, Phys. Rev. Lett., 83 (6) (1999) 1151. J. Lister and H. Stone, Phys. Fluids, 10, (1998) 2759. I. Cohen and S. Nagel, Phys. Fluids, 13 (12) (2001) 3533. I. Cohen and S. Nagel, Phys. Rev. Lett., 88(7) (2002) 074501. I. Cohen, H. Li, J. Hougland, M. Mrksich and S. Nagel, Science, 292 (2001) 265. W. Zhang, In preparation. B. Squire, Phil. Mag., 43 (1952) 942. A. F. Pillow and R. Paull, J. Fluid Mech., 155(1985) 327. M. A. Goldshtik, M and V. N. Shtern, J. Fluid Mech., 218 (1990) 483. V. N. Shtern and F. Hussain, J. Fluid Mech., 309 (1996) 1. V. N. Shtern and F. Hussain, Ann. Rev. Fluid Mech., 31 (1999)537. C. F. Stein, J. Fluid Mech., 438 (2001) 159. V. Putkaradze, in preparation. L. D. Landau and E. M. Lifshitz, Fluid Mechanics, Pergamon Press, Oxford, 1991. A. Herczynski, P. D. Weidman and G. I. Burde, Phys. Fluids, 16 (4) (2004) 137. V. Putkaradze, J. Fluid Mech., 483 (2003) 1. G. B. Jeffery, Phil. Mag 29, (1915) 455. G. Hamel, Jahresbericht Deutsch. Math. Verein, 25 (1916) 34. L. Rosenhead, Proc. Roy. Soc. A, 175 (1940) 436. W. H. H. Banks,P. G. Drazin and M. A. Zaturska, J. Fluid Mech., 186, (1988) 559. H. K. Moffatt and B. R. Duffy, J. Fluid Mech., 96 (2), (1980) 299. M. A. Goldshtik and V. N. Shtern, Fluid Dynamics, 24 (1989) 191. M. A. Goldshtik, F. Hussain and V. N. Shtern, J. Fluid Mech., 232 (1991) 521.
The Role ofInertial Effects and Conical Flows in Breakup of Liquid Threads
[34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51]
477
L. E. Fraenkel, Proc. Royal Soc. Lond. A, 267 (1962) 119. G. A. Georgiou and P. M. Eagles, J. Fluid Mech., 14 (1988) 259. P. M. Eagles and G. A. Georgiou, Phys. Fluids, 30 (12) (1987) 3838. P.M. Eagles, J. Fluid Mech., 24(1) (1966) 191. S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska, J. Fluid Mech., 336 (1997) 183. F. J. Uribe, E. Diaz-Herrera, A. Bravo and R. Peralta-Fabi, Phys. Fluids, 9 (9) (1997) 2798. R. Berker, Handbuch Der Physik, VI1I/2, (1963) 20. V. Putkaradze and P. Dimon, Phys. Fluids, 12 (1), (2002) 66. S. I. Voropayev and Ya. D. Afanasyev, J. Fluid Mech., 236 (1992) 665. S. I. Voropayev and Ya. D. Afanasyev, Phys. Fluids, 6 (1992) 2032. S. I. Voropayev, H. J. S. Fernando and P. C. Wu, Phys. Fluids, 8 (2) (1996) 384. S. I. Voropayev and H. J. S. Fernando, Phys. Fluids, 8 (9) (1996) 2435. P. Bourgin and N. Tahiri, in P. H. Gaskell, M. D. Savage and J. L. Summers (eds.), Proc. of the First European Coating Symposium, University of Leeds, (1995) 23. L. E. Fraenkel, Proc. Royal Soc. Lond. A, 272 (1963) 406. I. J. Sobey and P. G. Drazin, J. Fluid Mech., 171, (1986) 263. K. Chandrasecharan, Elliptic Functions, Springer-Verlag, 1984. R. Thorn, Structural Stability and Morphogenesis, Benjamin Inc, 1975. V. Putkaradze and L. Romero, in preparation.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 12
Statistical mechanics of dense emulsions D. N. Petsev Department of Chemical and Nuclear Engineering, University of New Mexico, Albuquerque, New Mexico 87131 1. INTRODUCTION Emulsions are a specific type of colloidal systems where both the disperse media and disperse phase are liquids. Typical examples are dispersions of oil in water or water in oil. In most cases the dispersed phase (oil or water) is in the form of spherical droplets of different sizes. The shape (usually spherical) is enforced by the interfacial tension while the size distribution is usually dependent on the emulsion preparation protocol and, as time progresses, on the dynamics of flocculation and coalescence. The latter, on the other hand, is a function of the droplet direct and hydrodynamic interactions. The focus of this review will be mostly on ensembles of submicrometer sized monodisperse emulsion droplets often referred to as miniemulsions. An important property of miniemulsions is that they exhibit Brownian motion, which has a strong impact on their dynamic and thermodynamic properties. Therefore, such droplet ensemble is amenable to statistical mechanical analysis and this is the purpose of this article. Emulsions are usually polydisperse but taking into account this fact will complicate the analysis too much without contributing to the main results and conclusions. However, a procedure for preparation of monodisperse emulsions is available [1], which means that the theoretical treatment outlined here could be always related to a particular experimental system. When comparing emulsions to solid liquid dispersions (e.g. hard spheres, charged hard spheres, etc.) it becomes evident that there are some important differences following from the fact that the droplet interfaces are fluid and flexible. Often the interfacial fluidity is suppressed by surfactant adsorption and the interfacial flexibility might be unimportant because the interfacial tension renders the droplets effectively rigid. However, there are many other cases where one or both of these conditions do not apply. In general, for droplets with size of the order of a few hundreds of nanometers and interfacial
480
D.N. Petsev
tension of about a few dynes/cm (usually below 10), one may expect that droplet deformability will impact both the direct (e.g., electrostatic, van der Waals) and hydrodynamic interactions. This impact is due to the fact that all these interactions are sensitive to the specific shape of the interacting surface. Besides, interfacial deformation itself contributes to the interaction energy since work is done against the forces governing the droplet shape (interfacial tension, bending energy). As a result two droplets will deform upon approach and acquire a shape that could be reasonably approximated as truncated spheres, see Fig. 1. For large macroemulsion drops (between a few micrometers to a few millimeters) the effect of the deformation becomes so great that one may consider only the thermodynamic and hydrodynamic interactions within the plane parallel film that forms between the surfaces and neglect the contribution of the surrounding regions. Particular cases of oil-water-surfactant mixtures are microemulsions, which are formed when the interfacial tension is low [2-4]. Contrary to ordinary emulsions, microemulsions are thermodynamically stable. The interfacial deformability of the microemulsion droplets is governed primarily by the bending rigidity rather than by the interfacial tension. Microemuslions can be sorted in three groups based on the phase state [2,3]. When the surfactant in the solution has high hydrophilic-lipophilic balance the microemulsion exists in the form of small oil droplets (normally with diameter below a 100 nm) in a continuous aqueous phase. If excess oil is present in the
Fig. 1. Sketch of two deformed droplets (a) in absence of long ranged repulsive forces and (b) in the presence of long range repulsion (e.g., electrostatic). If not long range forces are present the droplets deform upon contact of their surfaces. Otherwise a liquid film with thickness h is formed.
Statistical Mechanics of Dense Emulsions
481
system, it resolves as a separated macroscopic phase. This type of microemulsion is often referred to Winsor I type. For surfactants with low hydrophilic-lipophilic balance, the opposite situation could be observed when water droplets form in the oil phase and the excess water (if any) exists in separate macroscopic phase. This is Winsor II type of microemulsion. By varying some experimental parameters (temperature for nonionic and electrolyte concentration for ionic surfactants) one may switch from Winsor I to Winsor II and vice versa. By doing that, the microemulsion goes through a region of extremely low interfacial tension [4]. Based on the composition oil/water/surfactant a third phase may resolve between the oil and water where both the oil and water are interconnected forming a bicontinuous network [2-4]. This type of microemulsion is called Winsor III. A thermodynamically stable microemulsion could be mechanically stirred to form a macroemulsion which may exhibit very interesting properties in terms of stability due to the coexistence of small microemulsion droplets (below a 100 nm) and large macroemulsion drops, which could about micrometer size [5]. In the case of sub-micrometer miniemulsion droplets both the plane parallel film region and the surrounding curved regions contribute to the droplet energy of interaction and it is often incorrect to ignore one in comparison to the other. Hence, for such emulsions one has to be able to account properly for the energy of interaction of deformable droplets in order to be able to model and analyze their macroscopic properties and stability. The same argument is valid for the interactions between microemulsion droplets. The main difference is that the contribution of the interfacial tension is often negligible while one should properly take into account the bending rigidity. Recently, extensive studies have been undertaken to derive expressions for the pair energy, [6-12] see also Chapter 8. This chapter is organized as follows: the next Section presents a brief overview of the statistical mechanical treatment of dense fluid systems based on the application of the integral equation theory. Section 3 elucidates the implications in the approach due to the fluidity and deformability of the droplets, Section 4 presents some computation results for the macroscopic structure and properties of emulsions. The particular case of extremely dilute emulsions is very important since it provides a reference state for normalizing the radial distribution function. That is why it is discussed in Section 5. Section 6 summarizes the concluding remarks. 2. STATISTICAL MECHANICS OF COLLOIDAL DISPERSIONS A BRIEF OVERVIEW Statistical mechanics is a valuable tool, which relates microscopic parameters to the macroscopic properties of a system that consists of a large number of
482
D.N. Petsev
molecules or colloidal particles. It is important and useful to be able to deduce the structure and the thermodynamic parameters of such a system from the interaction energy between the molecules or particles that can be often determined or calculated from elsewhere. A convenient approach to achieve this goal is using integral equation theories [13-17]. Integral equation theories in general aim to relate the pair interaction energy between two molecules (also colloidal particles or Brownian emulsion droplets) to the macroscopic thermodynamic properties of the whole system as osmotic pressure or compressibility. There are a number of integral equation based approaches that employ different approximations and are therefore applicable to different particular cases. Central concepts in all versions of the integral equation theories are the pair correlation and/or radial distribution functions, which are discussed briefly in this section. 2.1. Correlation functions, structure and thermodynamics Our consideration of the topic of correlation functions, structure and their relation to thermodynamics will be restricted to briefly recalling some of the basic ideas and definitions since there are many excellent texts that deal with this matter in detail [13-17], see also Chapter 1. When considering colloidal systems and emulsions in particular it is important to emphasize the fact that there are two levels at which the distribution and correlation function formalism could be applied. The first level is molecular - the pertinent statistical ensemble consists of the molecules of the droplets, the solvent surrounding them and all species that might be present (electrolyte, surfactants, etc.). This level provides valuable information about the structure of the droplets, their interfaces and the solvent in close vicinity to the interfaces at a molecular scale and is discussed in greater detail in Chapter 1 of this book. The second level is particle where the ensemble consists of the dispersed colloids, or droplets in the case of emulsions. The molecular structure of the drops and the disperse medium that surrounds them are not taken into account explicitly but indirectly by means of macroscopic parameters that modify the particle interactions like dielectric permittivity, ionic strength, Hamaker constants, etc. Although less detailed than the molecular level, this approach allows successfully treating complex colloidal systems and offers some useful insights about their structure at the particle length scale, as well as their overall macroscopic thermodynamic properties. Let us consider a fluid of N particles that occupies a volume V and has temperature T. The probability of having any two of these particles to be in volumes dv\ around rj and dr2 around r2 while the remaining N-2 particles are allowed to be anywhere in the rest of the volume Fis
Statistical Mechanics of Dense Emulsions
g
483
( r " r ») = V(jV_2)i;Jo -Jo exp[-#/,(rp...,rN)]*,...*N J^
poo
poo
~ — J o ...Jo exp[-/3t/w(r1,...,rN)jJr3...JrN where o
...J o exp[-/?t//v(r1,...,rN)]
(2)
is the configuration integral, [/^(r,,...,!^) being the TV-particle potential energy. The multiplier (3 = \lkBT is the inverse thermal energy, where kB is Boltzmann constant and T is temperature. The transition between the first and second line in Eq. (1) is justified for large number of particles, N. The quantity g (r,,r 2 ) is the two particle distribution function and in case of an isotropic system (e.g., fluids) depends only on the distance between the particles r = r, — r2 and g{2)(rt,r2) = g(r)
(3)
where g[r) is the radial distribution function. Its meaning is probability to have one particle at the origin and another at a distance r from it. As it could be seen from Eq. (1), the two particle distribution function and hence the radial distribution function depends on the //-particle energy UN(r{,...,rN). This quantity is very difficult to obtain but for particles, interacting via central forces, and assuming pairwise additivity the TV-particle energy becomes
l/ff(r,,...,rN) = i $ > f o )
(4)
where M ( ^ ) is the pair interaction energy between particles i andy, separated by a distance rtj. Knowing the radial distribution function, one may calculate the macroscopic properties of the system of dissolved molecules or suspended colloidal particles (including droplets) as the osmotic pressure, Pos, or the isothermal osmotic compressibility, Xr respectively
484
D.N. Petsev
^ . = 1-4/30 P ^ r ) — f d ? , J
p
(3
(3 dpos
o
r = r/2a
(5)
^
^ >
df
^ oy
T
>\
where a is the particle radius, p is the particle number density and <> / = 4iraip/3 is the volume fraction. The latter quantity is very convenient when dealing with colloidal dispersions, includingly emulsions. The radial distribution function, g(r), is directly related to the structure factor of the system, S(q) by a Fourier transform (q is the wave vector and q is its magnitude)
S(q) = l + 6^ f"g(f)exp(-iq-r)df,
r = rl2a,
q = 2qo
(7)
7j- JO
or for isotropic systems
S{q) = \-mr\\-g{r)p^r2df, Jo
l
J
q = 2qa.
(8)
qr
The structure factor can be determined experimentally by light or neutron scattering experiments. Knowing the radial distribution function allows for calculation of all thermodynamic quantities of interest (for details one may refer to Refs.) [1317]. Therefore this approach allows for theoretical analysis of phase eqilibria. The conditions for "gas-liquid" phase equilibrium, when the system separates into two regions with low and high density, are given by 7} =7},, P,=Pn,
n,=nu
(9)
where T, P and u, are the temperatures, pressures and chemical potentials in the low density (subscript L) and high density (subscript H) phases respectively. Recently it was shown that the radial distribution function could also be used for computation of the freezing boundary in the phase diagram [18-20], see Ref. [16]. These authors have shown that the freezing point can be determined by the following equation se-s2
=0
(10)
Statistical Mechanics of Dense Emulsions
485
where sex is the excess entropy multiparticle correlation expansion of the entropy of liquids [21]
«=2
sn is the partial entropy contribution of n particles and sex has the meaning of excess entropy per particle due to multiparticle interactions and correlations. The pair term is given by
s2=-12> J o ~[g(r)lng(r)- g(r) + \]r2dr .
(12)
This simple freezing criterion has been tested against molecular dynamics simulations for a variety of potentials and proved to be remarkably accurate [16,18-20,22-25]. 2.2. Ornstein-Zernike integral equation for the radial distribution function Obtaining the radial distribution function for a multiparticle system (like colloidal dispersion) is not simple. The reason is that the interaction between each two particles is affected by the presence of all the others in the ensemble. That is why all equations for determining the radial distribution function introduce assumptions and approximations. There are a number of excellent reviews where the equations for the radial distribution function are discussed and compared in detail [13-17]. Below we present a very brief description of the Ornstein and Zernike [26] approach and some of the pertinent approximations that are often applied. This equation has to be solved numerically and the best current algorithm was suggested by Labik et al. [27]. The Ornstein-Zernike equation reads [26], see also Refs. [13,14,16,17]
h{r) = c(F) + ^JQnJc(\f-r'\)h(f')dr'
(13)
where /z(r) = g(r) —1 is the total correlation function and c(r) is the direct correlation function. Eq. (13) that means the total correlation between two particles could be separated into a direct, c(f), and an indirect term given by the integral. The indirect term on the other hand is also expressed by direct correlations, averaged over the positions of the remaining particles. Hence, the Ornstein-Zernike equation has a very simple and clear physical meaning. Still, it is not sufficient because in order to obtain the radial distribution function
486
g(r)
D.N. Petsev
[or the total correlation function h(f), which is equivalent] one must
know the direct correlation function c(r). Unfortunately it is not known and one has to make an assumption about its form. There are different models for the direct correlation function, also called closure approximations. A very good discussion of the different closures is given in Ref. [16]. Below we will consider only a few that are most relevant to colloidal dispersions and emulsion systems. 2.2.7. Percus-Yevick approximation The Percus-Yevick approximation for the direct correlation function reads [13,14,16,17,28]
c(r) = {l- exp[-/3«(r)]}g(r).
(14)
This approximation allows for obtaining an analytical solution of the OrnsteinZernike equation for hard sphere fluid where u(r) = oo for r<\,
u(r) = 0 for r>\.
(15)
In some cases the hard sphere model represents very well emulsion systems. For example, oil in water emulsion with high interfacial tension. In general, the Percus-Yevick approximation works well when the pair interaction energy is short ranged although it tends to overestimate the first peak in the radial distribution function. Besides the hard sphere model it gives an analytical solution for the sticky hard sphere model where the attraction is infinitely short ranged [29]. However, the Percus-Yevick approximation suffers from thermodynamic inconsistency [13,16,17]. The hard sphere equation of state obtained from the pressure virial expression (5) is different from the one derived using (6). Similar problems occur for the sticky sphere model [29-32]. In this chapter, however, we will concentrate on emulsions where the interactions are more complex than in the simple hard or sticky sphere cases. 2.2.2. Hypernetted chain approximation The hypernetted chain closure approximation [13,14,16,17,33-37] for the direct correlation function is given by c(r) = h{r)-f3u(r)-\ng{r).
(16)
The hypernetted chain approximation is known to work better for long ranged interactions. Contrary to the Percus-Yevick model, the hypernetted chain
Statistical Mechanics of Dense Emulsions
487
approximation underestimates the height of the first peak in the radial distribution function. There are a number of modifications of the hypernetted chain approximation [38-41], see also Ref. [16]. 2.2.3. Rogers-Young approximation Rogers and Young [42] suggested a closure approximation that imposes thermodynamic consistency. To achieve this, they interpolated between the Percus-Yevick and the hypernetted chain approximation using
where f(r) = \-exp(-ar)
(18)
with a being an adjustable parameter. The Rogers-Young approximation is currently the best closure for systems with repulsive interactions. Its application, however, for attractive particles is questionable [43]. Still, our analysis of moderately attractive emulsions [11,44] showed that this closure gives reasonable results. 2.2.4. Other closure approximations There are other closure approximations that have been suggested in the literature. For more details the reader may refer to the excellent reviews of Caccamo [16] and Nagele [53]. As discussed above the choice of closure approximations should be based on the specific system under consideration and taking into account the specific assumptions involved in each of them. A good practice is to test the solution against Monte Carlo or Brownian dynamics simulations. 3. DISTRIBUTION AND CORRELATION FUNCTIONS IN ENSEMBLES OF EMULSION DROPLETS: FEATURES DUE TO THE INTERFACIAL DEFORMABILITY This Section presents analysis of the possible impact that the interfacial flexibility and deformability may have on some of the basic statistical mechanical quantities. The effect on the pair interaction energy is also briefly discussed (see Chapter 8).
488
D.N. Petsev
3.1. Configurational intergral, pair correlation and radial distribution functions Similarly to hard particle dispersions, emulsions could be represented by a system of N oil droplets suspended in water. The droplets may be stabilized by surfactants (ionic and/or nonionic) while the continuous phase may contain some background electrolyte and is considered as structureless. Below we present a brief outline of the impact the droplets deformability has on the statistical mechanical analysis. This subject has been presented in more details in Refs. [44,45] and [11]. The main difference between solid dispersions and emulsions (with low interfacial tension) is in the fact that droplets may deform upon approaching each other [6-11,46]. The deformation is due to the pair interaction energy between the droplets but at the same time the pair energy itself depends on the deformation, see Chapter 8. As a consequence, the positions of the mass centers of the droplets are not the only integration variables in the configuration integral (2). For each given distance between two droplets there could be a number of different configurations corresponding to different droplet deformations, (or formation of films with different radii - see Fig. 2).
Fig. 2. A pair of interacting droplets may exhibit different deformations at the same mass center to center distance. Therefore for a given configurational variable r there are multiple states corresponding to different film radii.
Statistical Mechanics of Dense Emulsions
489
Therefore, the film radii, R formed between the droplets are additional degrees of freedom that have to be accounted for when defining the configuration integral. It becomes [44]
o -Jo Jo "Jo ex P[-^(^'-^M;^-^)]^,-^^...Jr N This is the configuration integral for a system of N droplets that may form M films. If two droplets are far and do not form a plane parallel film, its radius is essentially zero and the configuration is still properly counted. The pair correlation function for such a system is g{2)(Ri;rl,r2)=constx—-
V2 Z
»
.../ Q
%J 0
/ ... *J 0
*J 0
. (20) exp[-(3U{Rl,...,RM;rl,...rN)]dR2...dRMdrr..drN
As mentioned before, the energy of droplet interaction,L^(/?,,...,i?M;rl,...rjV), depends on the deformation. The "const" multiplier has to be determined by the requirement that the correlation function becomes equal to one for infinite separation of the droplets, |r, — r2| —> oo. As we will discuss later, it depends on the configurations generated by droplet deformations at infinite distances, i.e. without interactions. The number of films (including those with zero radius) is difficult to determine. However, this number is not important [44]. In fact, one can integrate (20) over the film radius R\ and derive an analog of Eq. (1)
g
(2)
V2 (r,,r 2 )=co^x — Z
. (21)
N
.../
I ...I
exV[-pU(Rv...,RM;rl,...rN)]dRr..dRMdrr..drN
There is an important difference between Eq. (1) and Eq. (21) and there are additional configurations and variables, which contribute to the averaging. This is actually an application of the potential distribution theorem [47]. The constant multipliers in Eqs. (20) and (21) are not the same. Similarly to the case of solid particle dispersion, we can simplify the problem further by assuming a pairwise additivity of the interactions, including the effects of the droplet deformation
490
D.N. Petsev
U{Rl,...,Ru;r[,...,rN) = ±-J2u(R
(22)
where Ry is the film radius, formed between droplets / and/, separated by a distance ry. Expressions for uyR^r^ are available in the literature [6-11], see also Chapter 8. Using approximation (22) is most helpful for obtaining the averaged pair energy at infinite dilution as well as the normalization constant in Eq. (21). Thus for vanishing volume fraction <j> —> 0 we define g°(R;F)~exp[-0u(R;r)},
R = R/2a
(23)
and g°(r) = constx f """g°(R;r}dR^ constx Pexp\-/3u[R;r)\dR.
(24)
Eq. (24) gives the radial distribution function for an infinitely diluted system of droplets, averaged over all possible deformations. Note that in the second integral the integration limit Rmax is replaced by infinity. This is justified since the interaction energy ui,R;r\ strongly increases with the film radius R and hence the integrand in (24) rapidly decreases and tends to zero. Therefore, large values of R have negligible contributions to the configurations of the droplet pair and to the radial distribution function. Eq. (24) allows defining a pair potential of mean force (averaged over all possible film radii), viz.
/?w(r) = ln[g°(r)].
(25)
This pair energy, w(r), is the one to be used in the closure approximations (14,16,17) instead of simply u(r) when deformable liquid droplets are considered and not solid dispersions of spheres. The normalization constant in Eq. (24) can be determined by the condition that the radial distribution function becomes equal to one at infinite separation of the droplets,
const = { r~exp[-/?«(£;/:r-»oo)ldKl . [Jo
L
^
More details are given in Section 5.
J
(26)
Statistical Mechanics of Dense Emulsions
491
3.2. Pair interaction energy between droplets Emulsion droplets with high interfacial tension behave very much like a solid colloidal dispersion. The interaction energy and stability of solid dispersions has been studied for a long time and there are excellent sources on the subject, for example see Refs. [48-53], which by no means represent a full list. In this review, however, we are interested mostly in emulsion systems with low interfacial tension (usually below 5 mN/m) where the approach and interaction of two droplets may lead to deformation and film formation due to both thermodynamic and hydrodynamic forces. 3.2.1. Interfacial extension and bending contributions Deformation modifies the colloidal interactions like electrostatic, van der Waals, steric, depletion, hydration, etc. It also gives rise to additional contribution due to the deformation of the interface itself, which is not present in solid dispersions. This contribution is due to the extension of the droplet surface and/or change of the local curvature upon deformation [6-11,44-46,5456], see also Chapter 8. For small deformations the surface extension energy is given by «J(/?) = 47ra 2 (2 7 ( ) ^ 4 +£ G ^ 8 ),
R = R/2a
(27)
where 70 is the interfacial tension of a spherical, undeformed droplet and EG=(dj/d\nS)s=s
=(d"ild\nR2}
_ is the Gibbs elasticity of the surfactant
monolayer stabilizing the droplet. Often the second term in Eq. (27) could be ignored when compared to the first one, and this is the case for most emulsion systems. For systems with extremely low interfacial tension, however, the Gibbs elasticity term becomes dominant. Such systems often tend to form microemulsions [4,5,57] in which the bending rigidity of the interface also becomes important u, (R) = -8ira2B0H v
;
1
— R2 = 32irkc —f 1 - ^ U 2H0 ao{ 2a)
2
(28)
where Bo is the bending moment and kc is the bending elasticity constant [58], H = — 1 / a and H0 = — \/a0 are the local interfacial and spontaneous curvatures respectively while a and a0 are the droplet and the spontaneous curvature radii. The energy contributions given by Eqs. (27) and (28) do not depend on the distance between the droplets explicitly. The deformation (given by the radius R) is determined by the colloidal interactions that are present and they are distance dependent (see the discussion below, Chapter 8 and Refs. [6-11,44-
492
D.N. Petsev
46,54,55]). However, all colloidal type of interactions like van der Waals, electrostatic, depletion, steric, salvation, etc., decay with the droplet separation and become zero for infinite distance. The interfacial deformation related contributions do not necessarily disappear at infinite droplet separations and therefore may still generate statistical configurations that need to be properly taken into account. That is accomplished by the condition (26), see also the discussion in Section 4. 3.2.2. Colloidal interactions and droplet deformation The typical colloidal forces such as electrostatic, van der Waals and others induce the droplet deformation and are modified by it but at the same time. That is why the shape of a droplet doublet is determined by minimizing the free energy of interaction while taking into account all possible contributions [6,8-11]. Since detailed analysis of many particular cases has been presented elsewhere [6-11,44-46,54-56], in this review we will restrict ourselves to a few examples based on the Derjaguin-Landau-Vervey-Overbeek (DLVO) theory [48,50,51], which includes electrostatic repulsion and van der Waals attraction. The latter is accounted for in the framework of the model of Hamaker [59]. The electrostatic repulsion between two deformed droplets is given by the following expression u(h,R) v
'
= ^^tanh>[P^}eW(-2Kah)LRi K (3K
{
4
)
'\
+±-\
naj
h = h/2a
(29)
where h and R are the film thickness and radius respectively (see Fig. 1), Cel is the number electrolyte concentration, e is the elementary charge, \t 0 is the droplet surface potential and K is the Debye screening parameter defined by J =2 ^
C
"
.
(30)
z is the background electrolyte charge number and eoe is the dielectric permittivity of the solvent. In absence of any deformation, the film radius is zero and Eq. (29) reduces to the well-known nonlinear superposition approximation for the electrostatic repulsion between two colloidal solid spheres [48,50,51,53]. The van der Waals attraction between two deformed spheres with shapes approximated by those of truncated spheres could be determined by the following expression
Statistical Mechanics of Dense Emulsions
/ - ~\ A u (h,R) = -^L
493
1 1 hil + h) —+„, l ,,+21n-^ / L
J
(31)
2
4R
h2(l + h)\2 + h)2 where, again for no deformation (R = 0), the expression for the van der Waals energy reduces to that for two undeformed solid spheres [59], see also [48,50,51,53]. The coefficient AH is known as the Hamaker constant although it is not truly a constant and may depend on the separation distance between the droplet surface h due to the electromagnetic retardation [60-63]. If retardation is unimportant (like for small particles below one micrometer [53])
4=A^'A'"''l. 413 £ , - £ ,
(32,
16^2 ( „ » + „ > ) "
ex and e2 are the dielectric permittivities, while n\ and n2 are the refractive indices for the drops and the surrounding medium respectively. hp is Planck's constant, co is the fluctuation frequency of the interacting molecular dipoles and c is the speed of light. Hence, in absence of retardation, the Hamaker coefficient is a true constant. For larger particles however electromagnetic retardation becomes substantial and the Hamaker coefficient becomes
^=^-^^ 2 + ^4^4 2 i. 4/3 ex — e2
4im2
(33)
n, + n2 h
Eqs. (32) and (33) are approximations and there is a more complete theory [63]. However, it is too complicated to be introduced into a statistical mechanical model. A relatively simple interpolation formula between the non-retarded and fully retarded cases has also been suggested (see for example [53], and Chapter 8 of this book). The retarded Hamaker coefficient decays with the distance between the interacting surface and therefore the van der Waals energy decreases even faster with the separation. In addition to the retardation, the Hamaker coefficient could decrease even further if there is a presence of extremely high (~1M) electrolyte concentrations [53,62]. For lower
494
D.N. Petsev
concentrations, the electrolyte effect on the Hamaker constant could be safely ignored. The typical numerical values of Hamaker constants for the cases of oil/water/oil or water/oil/water systems (like most emulsions) are between 3 and 4xlO"21 J[52]. 3.2.3. Total energy of pair interaction We assume that the total energy of interaction between two droplets is a sum of all the contributions described above u(h,R) = umv(h,R) + uel(h,R) + us(R) + ub{R).
(34)
This is the energy to be introduced in Eq. (24) to obtain the pair energy averaged over all possible deformations. It is convenient, therefore, to rearrange the separation variable from film thickness, h (distance between the drop walls, see Fig. 1), to center to center distance, r using the simple geometrical relationship
I H* r = h + 2a.\\ V
r,
or
- /
r = h + yll-4R2
a
.... •
(•)•>)
r = r/2a, h = h/2a, R = R/2a Using (35) we can transform u\h,R\ into
u(f,R\.
4. EXAMPLES In this section we present some computation results for the radial distribution function, the structure and macroscopic properties of emulsion and microemulsions. The model system that is considered is a monodisperse ensemble of Brownian droplets suspended in quiescent fluid. It is important to remind at this point that monodisperse emulsion systems can be obtained [1] and have been for extensive experimental studies [54,64-72]. 4.1. Radial distribution function, structure factor and macroscopic properties of monodisperse droplets with low interfacial tension In this section we present some results for radial distribution functions and structure factors of emulsions and microemulsions. The computations were performed using the Rogers-Young closure approximation [42] as implemented in the numerical algorithm developed in Ref. [27]. Although Rogers-Young closure works best for repulsive potentials, our comparison to Brownian simulations shows that it is also applicable to systems where there might be an
Statistical Mechanics of Dense Emulsions
495
attractive component in the interaction (e.g., van der Waals attraction). An obvious effect of the droplet deformation is that the radial distribution function will exhibit non-zero values for interdroplet distances less than their diameter. This is shown in Fig. 3a where the radial distribution function for emulsion droplets (full line) is plotted together with that for hard spheres with the same droplet diameter (dotted line). The interaction between the droplets includes only the surface extension energy governed by the interfacial tension and infinitely short length surface repulsion that prevents the droplets from coalescing. This type of interaction implies that there is only one value of the film radius (if film is formed) for each center to center separation, see Fig. 1 a. Using Eq. (35) we obtain [11,45] us(F) = 27ra2lo(l-F)2
for
r <\
= 0 for r > 1 This form of the interaction potential provides that g(r —»oo) = 1 without the explicit application of the normalization procedure described above, see Eqs, (23-26). The volume fraction used in this calculation is 0 = 0.42 and ira2j/kT = 760. Besides the effect of non-zero values at separation less than the droplet diameter, it can be seen that the peaks are lower and less developed than the referent hard sphere case. This means that the local ordering decays more rapidly with the distance between the droplets compared to hard sphere suspension with the same volume fraction. This is also evident from the plot of the structure factor of the same systems, see Fig. 3b. Hence, the fluidity of the droplet interface decreases the propagation of the structure induced by the excluded volume interactions. It is instructive to compare the radial distribution function for emulsions (where the surface deformation is controlled by the interfacial tension) and microemulsions in which the interfacial properties are governed by the bending elasticity. Such and example is shown in Fig. 4. The energy of interaction is [11,45] u* r) = — ^ 1 - ^ 1-r, aQ 2j = 0,
f>\
r<\,
ao=a0/a .
(i/)
496
D.N. Petsev
Fig. 3. Radial distribution functions (a) and structure factors (b) for emulsion droplets (solid curves) and hard spheres (dotted curves). The parameters for the calculation are Pna2^ — 760 and (/> = 0.42. Reproduced by persmission of the American Institute of Physics, D.N. Petsev and P. Linse, Phys. Rev. E 55 (1997) 586.
Statistical Mechanics of Dense Emulsions
497
Fig. 4. Radial distribution functions (a) and structure factors (b) for microemulsion droplets (solid curves) and hard spheres (dotted curves). The parameters for the calculation are kc=\l(3, a0 = 1 and ^ = 0.42. Reproduced by permission of the American Institute of Physics, D.N. Petsev and P. Linse, Phys. Rev. E 55 (1997) 586.
498
D.N. Petsev
Eq. (37) was derived using similar geometric arguments as for (36). Again the normalization at infinite separation is fulfilled. The parameters for this calculation are> = 0.42, the bending constant is kc =kT and the radius of spontaneous curvature a0 = a . The fluidity of the droplets, although governed by different parameters, has the same effect on the structure of the single drop phase microemulsion. This is evident by comparing the radial distribution function, Fig. 4a, and the structure factor, Fig. 4b, to the respective plots for hard sphere suspension. The effect of the droplet deformability for these particular examples (emulsion and microemulsion) is strong enough to be detected by experiments like measurements of the structure factor using light scattering [43].
Fig. 5. Radial distribution functions for uncharged (solid curve) and charged (dotted curve) emulsion droplets. The volume fraction in both cases is
Statistical Mechanics of Dense Emulsions
499
Long range interactions (i.e., electrostatic and van der Waals) lead to a shift in the radial distribution function peaks to slightly greater droplet separations and at the same time increase their height, see Fig. 5. This is due to the fact that for most typical parameters for emulsions, the electrostatic repulsion dominates the van der Waals attraction. The increase in the repulsion between the droplets enhances the structuring colloidal systems [43] and that apparently applies also to emulsions and microemulsions. The full line in Fig. 5 is for uncharged emulsion droplets and the dotted line is for charged droplets in the presence of 0.001 M monovalent electrolyte. The remaining parameters are: droplet radius a = 0.4 u.m, surface potential % = 100 mV, interfacial tension y = 0.1 mN/m, Hamaker constant AH = 5xlO"21 J and the volume fraction of the droplets is § = 0.42. Similar is the situation with charged microemulsions where the surface tension is negligible and the interfacial properties are determined by the bending rigidity, see Fig. 6. The full line is for uncharged droplets. The other lines plotted are for charged droplets in the presence of different concentrations of background electrolyte: dotted line - 0.01 M, dashed - 0.05 M and dashed dotted - 0.1 M.
Fig. 6. Radial distribution functions for uncharged (solid line) and charged microemulsion droplets at different electrolyte concentrations. The dotted line is for 0.01 M, the dashed line for 0.05 M and the dashed-dotted for 0.10 ML The volume fraction in both cases is
500
D.N. Petsev
The radial distribution function allows for computation of the macroscopic thermodynamic properties of the system. For example, using Eq. (5) one can obtain the osmotic pressure of the emulsion or microemulsion. Fig. 7 shows the osmotic pressure against the droplet volume fraction dependence for two types of uncharged microemulsions with different interfacial rigidity determined by the radii of spontaneous curvature, ao=a (solid curve) and a0 = 1.2a (dashed curve). The greater the radius of spontaneous curvature, the more deformable the droplet interface [10,11,45,73]. These two cases are compared to hard sphere dispersion (dotted curve). The droplet deformability has a considerable effect on the osmotic pressure. The pressure decrease is about 30% for a0 = a and 40% fora0 = 1.2a. Hence, interfacial deformability should be taken into account when considering the macroscopic properties of microemulsions and emulsions with low interfacial tension.
Fig. 7. Osmotic pressure vs. volume fraction for nonionic microemulsions with a0 12 = 1.0 (solid curve) and a0 12 = 1.2 (dashed curve). kc = kT for both cases. The dotted curve is for hard spheres. Reproduced by permission of the American Institute of Physics, D.N. Petsev and P. Linse, Phys. Rev. E 55 (1997) 586.
Statistical Mechanics of Dense Emulsions
501
Fig. 8. Osmotic pressure vs. volume fraction for nonionic microemulsions with ao/2 = \.O (solid curve) and a0 12 = 1.2 (dashed curve). kc = kT for both cases. The dotted curve is for hard spheres. Reproduced by permission of the American Institute of Physics, D.N. Petsev and P. Linse, Phys. Rev. E 55 (1997) 586.
Fig. 9. Osmotic pressure vs volume fraction obtained using the integral equation approach (Ornstein-Zernike, in Percus-Yevick approximation, full curve) and Brownian Dynamics simulation (dots). The parameters are kc = kT and <50 12 = 1.0. Reproduced by permission of the Royal Society of Chemistry, D.N. Petsev in "Modern Aspect of Emulsion Science, B.P. Binks Ed., RSC, 1998.
502
D.N. Petsev
The application of the Ornstein-Zernike integral equation approach in conjunction with the Rogers-Young closure approximation has been tested against Brownian dynamic simulations [11]. The results for the radial distribution functions for uncharged microemulsions at different volume fractions are in excellent agreement - see Fig. 8. The same statement could be asserted for the pressure - comparison with simulations is shown in Fig. 9. Hence the calculations based on the Ornstein-Zernike integral equation are reliable and readily applicable to emulsion and microemulsion systems. 5. DROPLETS AT INFINITE SEPARATION The statistical mechanical analysis of the properties of a single droplet is fundamentally important. It elucidates the effect of the interfacial curvature on the interfacial energy and also the contribution of the deformability to the overall properties of the whole ensemble of droplets. It also helps defining the proper normalization constant for calculating the radial distribution function, as used above. A more detailed discussion of physical foundations of the normalization is given below following [11,44,74]. The general approach to the statistical mechanical analysis of deformable droplets [67,68,75-80] is also briefly outlined. 5.1. Normalization constants for the radial distribution functions for deformable emulsion and microemulsion droplets The normalization condition of the radial distribution function at infinity [see Eq. (26)] depends on the droplet states at infinite separation. In general, even without any droplet interactions one may expect that the interfacial deformability will contribute to the thermodynamic state of a single droplet. Our model accounts for this contribution (in an approximate way) since the expressions for the droplet interfacial extension and deformation, Eqs. (27) and (28), do not depend on the droplet distance. Therefore, they do not decrease to zero at infinite separation. Infinitely separated droplets do not form films with defined radii. However, we calculate the normalization constant for the radial distribution function using Eq. (26) in conjunction with (27) or (28). The physical meaning of the result, which we will discuss in more detail below, is that the average single droplet surface area increase is represented by a quantity TTR2 where R is not related to an actual film but is a formal quantity that characterizes the extent of the deformation and area increase. For emulsion droplets, the interfacial properties are governed by the tension, y0, and Eq. (26) becomes
503
Statistical Mechanics of Dense Emulsions
const'1 = f"" exp\-(3us(R;r ^ o o ) dk - f™ exp(y-$f3ira2'y0R4)dR
° r(|) ^ (8/W 7o )
r(i) ^
=
1M
4(8/W 7o )"
°
•
(38)
4
Hence, const xexp(—/387ra27o.R4) defines a distribution function for the possible deformation that may occur in a single droplet. The radius is simply a formal measure for the deformation and one may calculate its statistical moments using the distribution function. Thus, the n-th moment is .. .
(8/37ra270)'/4 rco „
/Rn\\^__JU— ^
^
r
(D
J
_
f R"eXp(-P$na2%)R4)dR = °
[
^
r(^)
^^
^.
(39)
r(i)(8/3vra 2 7o )" /4
The first and second moments present a special interest. The first moment is given by 1/4
®=W)«k
<40)
and the second
(R2\
£(|)
(41)
The second moment quantifies the average interfacial increase of a single droplet due to shape fluctuations, AS, which is
-^HiP) =
r({)
„,=
" • ° 6 7 : 2 , AS = ^ ^ - .
(42)
For the parameters of Fig. 3a, the area increase of single droplet is about 0.43%. In microemulsions the interfacial tension is low [2-5,57]. That is why the interfacial properties are often dependent entirely on the bending elasticity, see Eq. (28). Introducing the latter into Eq. (26), one obtains
504
D.N. Petsev
exp — (3ub (i?;F —> ooj dR
exp -32Pirkc — 1 - ^- \R2 dR. 1
ao{
(43)
2a)
1
8{(2f3kca/a0)(\-a0/2a)\m Similarly to Eq. (39), above we define the n-th moment of the effective radius (that measures the deformation of a single droplet) r
/
.
Ni"2
•
(Rn) = %2l3kc—\\-^-\ rR"exp-32(37rkc^~l-^\R2 dR 1 [ ' ao{ 2a)\ ^o [ a0 2a) \ 11 \ l / 2
^^"/2
/-, ,
«c 2a)\ { 2
Ui I The first moment is (R)
[77
=
(45)
4ir[2f3kca/a0(\-a0/2a)] and the second
\
'
64(3irkc(a/a0)(l-a0/2a)'
For kc =1/(3 and ao = a, the second moment for the radius becomes
47ra2
\
7
32TT
8
The relative increase of the area of such microemulsion droplet is about 1%. Now we can trace the change in the average deformation for two droplets that are initially in close contact and then separate to infinite distance. At small separations, the deformation is governed mainly by the interactions and by increasing the distance their importance decreases. Then a distance dependent film radius could be defined by [11,44]
Statistical Mechanics of Dense Emulsions
(R(f)) = const x J™Rexp\-/3u(r,R)}.
505
(48)
The constant multiplier is given by Eq. (38) for emulsions and Eq. (43) for microemulsions, see also Eq. (26). A calculation for the distance dependence of the film radius for two emulsion droplets is shown in Fig. 10. The two curves in the figure correspond to different electrolyte concentrations. The higher concentration (0.01 M) leads to greater deformation of the interacting droplets while decreasing the amount of salt to 0.005 M commands a smaller film radius. As the separation between the droplets increases the interaction energy becomes unimportant and the deformation of the droplet (film radius) is given by Eq. (40) for emulsions and Eq. (45) for microemulsions. Again, the value of the radius R at infinite separations does not mean that a plane parallel film is formed but rather represents the degree of droplet deformation. The reason that the average film radii, given by Eqs. (40) or (45), are not zero is because any deviation from the spherical shape (R = 0) always provides a positive contribution to the effective droplet surface increase and hence to the effective radius. Eq. (40) gives the value of (R(f —»• oo n and for a = 0.4 um and y0 = 0.1 mN/m one obtains \R) = 0.0277. The dependence of the effective film radius on the droplet separation for charged microemulsions is shown in Fig. 11. Again, the effect of droplet interactions disappears with the distance between the droplets and the surface area increase of a single droplet is given by the value of (R). The repulsion for microemulsions is greater for the given choice of parameters and therefore a greater amount of electrolyte is needed. The solid curve is for 0.005 M while the dotted is for 0.01 M. The change of the film radius with the distance is much smoother than for emulsions where the surface properties are dictated mostly by the interfacial tension, see Fig. 10. Also, the relative increase in the single droplet surface due to fluctuations is greater for microemulsions than for emulsions. The value of (i?(r —>oo)) is calculated by means of Eq. (45). For a o / a = l.O and kc=l//3 we obtain (j?) = 0.0796. 5.2. Shape fluctuations of droplets Thermal perturbations of the droplet surface lead to variations of its shape that lead to surface increase [81]. Hence the droplet itself is a statistical mechanical system with many degrees of freedom - the different deformation modes. A very detailed analysis has been given by Henderson and Schofield
506
D.N. Petsev
[75,76] and further elaborated and applied to different systems by other authors [67,68,77-80,82,83]. The surface of such droplet is given by [81]
p r a2 (da? ^(dafV Jo J o
v
{d9j
sin 2 eldip)
v
sin0dd
J
•
(49)
The variable droplet radius a is a function that depends on the angular coordinates 9 and cp. The second line in (49) is true for small deviations from spherical shape where av{6',?) = const — a. The time and angular dependent radius can be expressed in a series of spherical harmonics /=/ m a x m=I
(5°)
av(0,c^) = a 1+ £ £ 4 * ( 0 U M 1=2 m=-l
where
UM^cos^exp^),
ir(cosg) = Sin"/Mf(c°^)
(51)
dycosO) and /}(cos#) is the Legendre polynomial of order /. Note that / = 0 and / = 1 are not included in the summation because they correspond to droplet dilation and translation respectively. The value of the maximum mode number is /max -aid, where d ~ 0.5 nm, is the molecular diameter. The interfacial energy of the droplet is then
I
, /=/„„ m=l
]
i+^rEE^+O-^Kf • OTT ,=2
m=-i
(52)
J
The ensemble average of (52) in conjunction with the equipartition theorem (for general discussion of this theorem see for example Ref. [17]) leads to
(K| 2 ) = {/37oa2[/(/ + l ) - 2 f ' .
(53)
Statistical Mechanics of Dense Emulsions
507
An important quantity is the mean square difference between the variable radius and that of the undisturbed sphere given by the expression [68]
The angular brackets indicate ensemble averaging. This expression gives an alternative and more rigorous opportunity to estimate the droplet surface area increase due to thermal fluctuations. The problem is that /max is generally unknown. Although it is usually a very large number [80], the series in Eq. (54) do not converge for /max —> oo and it is not possible to circumvent the problem by simply taking the upper summation limit to infinity. The shape fluctuations of droplets have been studied experimentally by difusing wave spectroscopy [67,68] and has been shown that this is a detectable phenomenon. The authors found experimentally that /max « 2 0 , but this is more of an empirical observation from their data than a rigorous result. The brief outline above is mostly for emulsions systems where the interfacial tension is the important parameter governing the interfacial fluctuations. A very detailed and rigorous analysis of microemulsion systems (with the bending rigidity being the important interfacial parameter) is given in Ref. [83]. 6. CONCLUSIONS In this chapter, we have presented an outline of the statistical mechanical analysis of ensembles of droplets, emulsions and microemulsions. In many cases, the approach and methodology are identical to solid colloidal dispersions, which are studied in detail. However, if the interfacial tension is low enough, the droplet interfaces become amenable to deformation and this may have an important effect on their structure and macroscopic thermodynamic properties. The main reasons for this are two: (i) the droplets deform at small separations due to the colloidal interactions (van der Waals, electrostatic, depletion, steric, solvation, etc.) and (ii) each droplet offers degrees of freedom related to the different modes of shape fluctuations, thus being a statistical system itself. The droplet deformation, induced by the pair interactions, leads to an additional energy contribution that is due to the surface extension and local curvature changes. The van der Waals, electrostatic and other possible colloidal forces on the other hand depend on the geometry of the interacting surfaces and hence, they also change with the droplet deformation that they have incited. Therefore, the droplet interactions and deformations are strongly coupled. The deformation of a single droplet requires a procedure for normalizing the radial distribution function and in this way contributes to the structure and
508
D.N. Petsev
macroscopic properties of the whole ensemble. It takes into account the fact that each droplet has a many degrees of freedom related to the shape fluctuations that are absent in a dispersion consisting of solid spheres. The deformability of droplets affects not only the equilibrium thermodynamic properties of emulsions and microemulsions but also their dynamics: flocculation and coalescence. These problems are also discussed in Chapters 8, 10-13, 16 and 17. Emulsions and microemulsions have a great practical importance for a wide variety of industries. That is why fundamental studies of these liquidliquid mixtures have a direct and applied importance for a broad range of human activities. REFERENCES [I] J. Bibette, J. Colloid and Interface Sci., 147 (1991) 474. [2] P.A. Winsor, Trans. Farady Soc, 44 (1948) 376. [3] P.A. Winsor, Solvent Properties of Amphiphilic Compounds (Butterworth, London, 1954). [4] R. Aveyard, B.P. Binks, S. Clark and J. Mead, J. Chem. Soc. Faraday Trans. 1, 82 (1986) 125. [5] B.P. Binks, W.-G. Cho, P.D.I. Fletcher and D.N. Petsev, Langmuir, 16 (2000) 10251034. [6] K.D. Danov, D.N. Petsev and N.D. Denkov, J. Chem. Phys., 99 (1993) 7179. [7] K.D. Danov, N.D. Denkov, D.N. Petsev, I.B. Ivanov and R. Borwankar, Langmuir, 9 (1993) 1731. [8] N.D. Denkov, D.N. Petsev and K.D. Danov, Phys. Rev. Lett., 71 (1993) 3226. [9] N.D. Denkov, D.N. Petsev and K.D. Danov, J. Colloid and Interface Sci., 176 (1995) 189. [10] D.N. Petsev, N.D. Denkov and P.A. Kralchevsky, J. Colloid and Interface Sci., 176 (1995)201. [II] D.N. Petsev, Interactions and Macroscopic Properties of Emulsions and Microemulsions, in: B.P. Binks (ed.), Modern Aspects of Emulsion Science (RSC, London, 1998). [12] J.K. Klahn, W.G.M. Agterof, F.v.V. Vader, R.D. Groot and F. Groenweg, Colloids Surf., 65(1992)151. [13] R. Balescu, Equilibrium and Nonequilibrium Statistical Mechanics (John Wiley & Sons, New York, 1975). [14] J.P. Hansen and I.R. McDonald, Theory of Simple Liquids (Academic Press, London, 1976). [15] T.L. Hill, Statistical Mechanics Principles and Selected Application (Dover, New York, 1987). [16] C. Caccamo, Phys. Rep., 274 (1996) 1-105. [17] D.A. McQuarrie, Statistical Mechanics (University Science Books, Sausalito, 2000). [18] P. Giaquinta and G. Giunta, Physica A, 187 (1992) 145. [19] P. Giaquinta, G. Giunta and S.P. Giarritta, Phys. Rev. A, 45 (1992) 6966. [20] P.V. Giaquinta, G. Giunta and G. Malescio, Physica A, 250 (1998) 91. [21] R.E. Nettleton and M.S. Green, J. Chem. Phys., 29 (1958) 1365.
Statistical Mechanics of Dense Emulsions
509
[22] C. Caccamo, D. Costa and G. Pellicane, J. Chem. Phys., 109 (1998) 4498. [23] C. Caccamo, P.V. Giaquinta and G. Giunta, J. Phys. Condens. Matter, 5 (1993) B75. [24] C. Caccamo, M.A. Varisco, M.A. Floriano, E. Caponetti, R. Triolo and G. Lucido, J. Chem. Phys., 98(1993) 1579. [25] E. Lomba, J.L. Lopez-Martin, H.M. Cataldo and C.F. Tejero, Phys. Rev. E, 49 (1994) 5164. [26] L.S. Ornstein and F. Zernike, Proc. Acad. Sci. (Amsterdam), 17 (1914) 293. [27] S. Labik, A. Malijevsky and P. Vonka, Mol. Phys., 56 (1985) 709. [28] J.K. Percus and G.J. Yevick, Phys. Rev., 110 (1958) 1. [29] R.J. Baxter, J. Chem. Phys., 49 (1968) 2770. [30] B. Barboy, J. Chem. Phys., 61 (1974) 3194. [31] G. Stell, J. Stat. Phys., 63 (1991) 1203. [32] R.O. Watts, D. Henderson and R.J. Baxter, Hard Spheres with Surface Adhesion: The Percus-Yevick Approximation and Energy Equation, in: I. Prigogine (ed.), Advances in Chemical Physics, Vol. 21 (Wiley, New York, 1971) 421. [33] J.M.J.v. Leeuwen, J. Groenveld and J.D. Boer, Physica, 25 (1959) 792. [34] L. Verlet, Nuovo Cimento, 18 (1960) 77. [35] G.S. Rushbrook, Physica, 26 (1960) 259. [36] T. Morita, Progr. Theor. Phys., 23 (1960) 829. [37] E. Meeron, J. Math. Phys, 1 (1960) 192. [38] F. Lado, Phys. Rev. A, 135 (1964) 1013. [39] F. Lado, S.M. Foiles and N.W. Ashcroft, Phys. Rev. A, 28 (1983) 2374. [40] G. Stell (ed.), Phase Transitions and Critical Phenomena, Vol. 5b, Academic Press, New York, 1976, Ch.3 pp. [41] Y. Rosenfeld and N.W. Ashcroft, Phys. Rev. A, 20 (1979) 1208. [42] F.J. Rogers and D.A. Young, Phys. Rev. A, 30 (1984) 999. [43] G. Nagele, Phys. Rep., 272 (1996) 215. [44] D.N. Petsev, Physica A, 250 (1997) 115. [45] D.N. Petsev and P. Linse, Phys. Rev . E., 55 (1997) 586. [46] N.D. Denkov, P.A. Kralchevsky, C. Vassilieff and I.B. Ivanov, J. Colloid and Interface Sci., 143(1991)157. [47] G.S. Rushbrook, Trans. Faraday Soc, 36 (1940) 1055. [48] E.J.W. Verwey and J.T.G. Overbeek, Theory and Stability of Lyophobic Colloids (Elsevier, Amsterdam, 1948). [49] J.T.G. Overbeek, The interaction between colloidal particles, in: H.R. Kruyt (ed.), Colloid Science, Vol. 1 (Elsevier, Amsterdam, 1952) 245-277. [50] B.V. Derjaguin, Theory of Stability of Colloids and Thin Films (Plenum, New York, 1989). [51] B.V. Derjaguin, N.V. Churaev and V.M. Muller, Surface Forces (Plenum, New York, 1987). [52] J.N. Israelachvili, Intermolecular and Surface Forces (Academic, New York, 1991). [53] W.B. Russel, D.A. Saville and W.R. Schowalter, Colloidal Dispersions (Cambridge University Press, Cambridge, 1989). [54] D.N. Petsev and J. Bibette, Langmuir, 11 (1995) 1075. [55] D.N. Petsev, Langmuir, 16 (2000) 2093. [56] D.N. Petsev, Mechanisms of Emulsion Flocculation, in: A. Hubbard (ed.), Encylcopedia of Surface and Colloid Science (Marcell Dekker, New York, 2002) 3192. [57] B.P. Binks, J. Meunier, O. Abilon and D. Langevin, Langmuir, 5 (1986) 1755. [58] W. Helfrich, Z. Naturforsch., 28c (1973) 693.
510
[59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83]
D.N. Petsev
H.C. Hamaker, Physica, 4 (1937) 1058-1072. H.B.G. Casimir and D. Polder, Nature, 158 (1946) 787. H.B.G. Casimir and D. Polder, Phys. Rev, 73 (1948) 360. J. Mahanty and B.W. Ninham, Dispersion Forces (Academic Press, New York, 1976). E.M. Lifshitz, Soviet Physics JETP, 2 (1956) 73. J. Bibette, Langmuir, 8 (1992) 3178. J. Bibette, Langmuir, 9 (1993) 3352. J. Bibette, T.G. Mason, H. Gang and D.A. Weitz, Phys. Rev. Lett., 69 (1992) 981. H. Gang, A.H. Krall and D. Weitz, Phys. Rev. Lett., 73 (1994) 3435. H. Gang, A.H. Krall and D. Weitz, Phys. Rev. E, 52 (1995) 6289. P. Poulin and J. Bibette, Phys. Rev. Lett., 79 (1997) 3290. P. Poulin, F. Nallet, B. Cabane and J. Bibette, Phys. Rev. Lett., 77 (1996) 3248. O.M. Monval, F.L. Calderon, J. Philip and J. Bibette, Phys. Rev. Lett., 75 (1995) 3364. O.M.-. Monval, A. Espert, P. Omarjee, J. Bibette, F.L.-. Calderon, J. Philip and J.-F. Joanny, Phys. Rev. Lett, 80 (1998) 1778. P.D.I. Fletcher and D.N. Petsev, J. Chem. Soc.Faraday Trans., 93 (1997) 1383. D.N. Petsev, unpublished. P. Schofield and J.R. Henderson, Proc. Roy. Soc. London A, 379 (1981) 231. J.R. Henderson and P. Schofield, Proc. Roy. Soc. London A, 380 (1982) 211. J.S. Huang, S.T. Milner, B. Farago and D. Richter, Phys. Rev. Lett., 59 (1987) 2600. S.T. Milner and S.A. Safran, 36 (1987) 4317. S.A. Safran, J. Chem. Phys., 78 (1983) 2073. L.C. Sparling and J.E. Sedlak, Phys. Rev. E, 39 (1989) 1351. L.D. Landau and E.M. Lifshitz, Fluid Mechanics (Nauka (in Russian), Moscow, 1988). B. Farago, D. Richter, J.S. Huang, S.A. Safran and S.T. Milner, Phys. Rev. Lett., 65 (1990) 3348. K.M. Palmer and D.C. Morse, J. Chem. Phys., 105 (1996) 11147.
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 13
Highly concentrated (gel) emulsions: formation properties
and
C. Solans", J. Esquena", N. Azemar", C. Rodriguezb, H. Kuniedab a
Departamento de Tecnologia de Tensioactivos, Instituto de Investigaciones Quimicas y Ambientales de Barcelona (IIQAB), Consejo Superior de Invetigaciones Cientificas (CSIC), Jordi Girona, 18-26, 08034-Barcelona, SPAIN
Graduate School of Environment and Information Sciences, Yokohama National University,Tokiwadai 79-7, Hodogaya-ku, Yokohama 240-8501, JAPAN 1. INTRODUCTION Highly concentrated or high-internal-phase-ratio emulsions are characterized by disperse phase volume fractions, <j), higher than the critical value (j)c = 0.74, for the most compact arrangement of uniform spherical droplets [1,2]. At this critical volume fraction an emulsion is expected to invert. However, emulsions can retain the original oil-in-water (O/W) or water-in-oil (W/O) structure at volume fractions exceeding the critical value because of two phenomena: polydispersity and droplet deformation. Consequently, the structure of this type of emulsions consists of polyhedrical droplets, with typical radii of a few microns, separated by thin films of continuous phase, a structure resembling gas-liquid foams. Their rheological properties [2-4] range from elastic (solidlike) to viscoelastic depending on the system components, composition variables and temperature, having a gel appearance. Fig. la shows a micrograph of a highly concentrated emulsion with a dispersed phase volume fraction close to unity and Fig.lb shows that the emulsion does not flow if the container is turned upside down.
512
C. Solans et al.
Fig. 1. a) Micrograph of a highly concentrated emulsion with a dispersed phase volume fraction close to unity, b) Macroscopic aspect of the emulsion, it does not flow if the container is turned upside down.
Highly concentrated emulsions are also referred to in the literature as gel emulsions [5-8], hydrocarbon gels [9], biliquid foams [10], etc. The term highly concentrated emulsion will be used throughout this text. They are classified, as ordinary emulsions, in two categories, water-in-oil (W/O) and oil-in-water (O/W). They can also be classified according to the interaction forces between droplets as adhesive and non-adhesive [11]. Adhesive emulsions, in the presence of attractive forces between droplets, do not relax and maintain a high degree of packing. Non-adhesive emulsions, brought into contact with its continuous phase, relax in a way that the emulsion dilates to the state of spherical (nondeformed droplets). Another classification is based on the microstructure of the continuous phase. Although it is generally assumed that emulsions consist of "simple" liquid phases being the droplets stabilised by a surfactant monolayer, the organisation of the different components in an emulsion can be more complex. Indeed, it has been proved that the microstructure of the phases present in the system plays a key role in emulsion formation and properties [5,12-15]. In this chapter, attention is focussed to highly concentrated emulsions in which the continuos phase is either a microemulsion or a micellar cubic liquid crystalline phase.
Highly Concentrated (gel) Emulsions: Formation and Properties
513
The characteristic properties of highly concentrated emulsions are of particular interest for theoretical studies and for applications. Their use for practical applications is well know since long ago. They are widely used as formulations in cosmetic, food, pharmaceutical, etc. fields. One of the most promising applications is their use as reaction media. In this context, they have received a great deal of attention for the preparation of low-density organic and inorganic materials (solid foams, aerogels) [10,16-19] and in chemical and enzyme-catalysed reactions [20,21] as alternative to conventional solvent media. In this chapter, highly concentrated emulsions constituted by a continuous phase based on microemulsion (Section 2) and micellar cubic (Section 3) phases will be discussed first. It will be followed by a description of adhesion properties (Section 4).
2. MICROEMULSION-BASED HIGHLY CONCENTRATED EMULSIONS 2.1. Phase behavior of water/polyoxyethylene nonionic surfactant/oil systems: general aspects Polyoxyethylene nonionic surfactants become lipophilic with increasing temperature because of dehydration of the polyoxyethylene chains. At low temperature, the surfactant monolayer has a large positive spontaneous curvature. At high temperatures, the spontaneous curvature becomes negative while at intermediate temperatures, the HLB temperature, the spontaneous curvature becomes close to zero and a bicontinuous, D phase, microemulsion containing comparable amounts of water and oil phases may be formed [22,23]. Fig. 2 is a schematic representation of the phase behavior of a ternary water/polyoxyethylene nonionic surfactant/oil system at increasing tenperatures from Ti to T3. At low temperatures, Ti, the two-phase region (extending from the water to the oil axis) consists of aqueous micellar (or O/W microemulsion) and excess oil phases, as indicated by the tie lines that converge near the oil apex. At higher temperatures, T3, the surfactant is mainly dissolved in the oil and the constituting phases are a reverse micellar solution (or W/O microemulsion) and excess water phase. At intermediate temperatures, T2 or (THLB), the surfactant does not show preferential solubility for oil or water and a three-phase region consisting of water, surfactant (middle phase or bicontinuous D microemulsion) and oil phases appears. This intermediate HLB temperature, is also called the Phase Inversion Temperature (PIT) because inversion from oil in water (O/W) to water in oil (W/O) emulsions or vice versa is produced [24].
514
C. Solans et al.
Fig. 2. Schematic phase diagrams of water (W) / polyoxyethylene nonionic surfactant (S) / oil (O) systems at various temperatures. I, II and III designate one-, two- and three-phase regions. uE: microemulsion.
Examples of the phase behavior of two ternary water/polyoxyethylene nonionic surfactant/oil systems are given in Fig. 3 and 4. Fig. 3 shows the phase behavior of water/R12EO4/decane system at 25°C and 50°C [5]. The HLB temperature of this system is 19°C and at the temperatures shown in Fig. 3, the surfactant is relatively lipophilic and acts as a W/O emulsifier. In the two-phase region (II), one of the phases is almost pure water. The other phase is a nonaqueous reverse micellar solution or microemulsion phase, indicated by (I). A lamellar liquid crystalline phase appears at higher surfactant concentrations above the isotropic phase (I). With a rise in temperature, the one-phase (I) region shifts towards the highly concentrated (in surfactant) area, an indication that the solubilization of water in the oil phase decreases with increase in temperature. The phase behavior of water/RnECVdecane system at 30°C [25] is shown in Fig. 4.
Highly Concentrated (gel) Emulsions: Formation and Properties
515
Fig. 3. Phase diagram of water/R,2EO4/decane system at: a) 25°C and b) 50°C. I and II are one and two-phase regions. L.C. present indicates the region in which a liquid crystalline phase appears. The dotted lines are tie lines (Reproduced by permission from Ref. 5).
The surfactant R)2EO6 is mainly water soluble and in the surfactant-water binary system forms a micellar solution (Wm) up to 40 wt% surfactant. In the presence of oil the Wm phase coexists with an excess oil phase (O). At higher surfactant concentrations, there are several regions with liquid crystalline phases of different structure.
516
C. Solans et al.
Fig. 4. Phase diagram of water/R^ECVhexadecane system at 30°C. I and II are one and two isotropic liquid phase regions. Wm and Om are aqueous and reversed micellar solution phases.L.C. present is a region, which contains a liquid crystalline phase (Lam.: lamellar type. Hex.: hexagonal type). V.I. means a viscous isotropic solution phase. Broken lines are tie lines. (Reproduced by permission from Ref. 25)
2.2.
Formation of highly concentrated microemulsion-phase-based emulsions and phase behavior Highly concentrated W/O emulsions form at temperatures above the hydrophile-lipophile balance (HLB) temperature of the corresponding system while highly concentrated O/W emulsions form below this temperature [5]. Those with W/O structure separate into two isotropic liquid phases at equilibrium: one phase is a submicellar surfactant solution in water and the other phase is a swollen reverse mieellar solution (or W/O microemulsion). This phase equilibrium is represented in Fig. 2 by the ternary diagram at T3 and in Fig. 3. Similarly, highly concentrated O/W emulsions separate into two isotropic liquid phases at equilibrium: an oil phase and an aqueous mieellar solution or O/W microemulsion. The phase equilibrium is represented by the ternary diagrams at T, of Fig. 2 and in Fig. 4. Highly concentrated emulsions exist only in limited regions of the miscibility gap (the two-phase region). The boundaries of the emulsion regions depend on the system and also on the method of preparation. The highly concentrated W/O emulsion regions in the partial phase diagrams of the water/R^EOVdecane system at 25°C and 50°C [5] are shown in Fig. 5. The notation A, A' and A" indicates their optical appearance as translucent, bluish-white and milky, respectively. The other regions, denoted as
Highly Concentrated (gel) Emulsions: Formation and Properties
517
Fig. 5. The highly concentrated W/O emulsion region in partial phase diagrams of a water/Ri2EO4/decane systems at: (a) 25°C and (b) 50°C. Method (2) was used for producing gel. I and II are one and two-phase regions. L.C. present indicates the region in which a liquid crystalline phase appears. The dotted lines are tie lines. (Reproduced by permission from Ref. 5).
B, C and D do not correspond to highly concentrated emulsions. It should be noted that in the vicinity of the one-phase region, close to the water-surfactant axis (indicated by D) emulsions are very unstable consist of very unstable emulsions. The boundary of this region is along a tie line and the change in
518
C. Solans et al.
coalescence rate along the border is very drasctic. It means that the phase volume ratio is not related to this phenomenon, instead it could be related to the change in structure of the oil phase [5]. With increase in temperature, the highly concentrated emulsions in this system are no longer translucent, the whole region shrinks. Fig. 6 is a detail of the oil-rich region of the phase diagram corresponding to the water/R12EO6/decane system at 30°C [25]. The O/W highly concentrated emulsion region is formed in the Wm+O region of the phase diagram. Since the highly concentrated emulsion region lies just below the constant surfactantwater composition line, which extends from the upper limit of the Wm phase (40 wt% surfactant) to the oil corner, the aqueous film surrounding the oil droplets is a highly concentrated aqueous micellar solution [25]. The usual preparation method for highly concentrated emulsion consist on dissolving a suitable emulsifier in the component that will constitute the continuous phase followed by addition of the component which will constitute the dispersed phase, with continuous stirring. They can be also prepared by weighing all the components at the final composition, followed by shaking or stirring the sample [5-7]. At a certain step of this emulsification process, the mixture consists of a multiple W/O/W emulsion, for this reason it was called the multiple emulsification method. Emulsification by the usual method and by this method results in rather polydisperse emulsions, like that shown in Fig. la.
Fig.6. Detail of the oil-rich region in Fig. 4. The dotted area is the O/W-type gel region. (Reproduced by permission from Ref. 25).
Highly Concentrated (gel) Emulsions: Formation and Properties
519
Another method of preparation, which takes advantages of the phase transitions produced during the emulsification process, is the so-called spontaneous formation method [26-28], which is based on the PIT emulsification method [24]. Emulsification is achieved by a rapid temperature change of a micellar solution or microemulsion without the need for mechanical stirring. For highly concentrated W/O emulsions, the system is below the HLB temperature at the start and emulsification takes place when an O/W microemulsion is rapidly heated from temperatures below to above the HLB temperature [26,27]. Formation of highly concentrated O/W emulsions, by this method, is achieved by quickly cooling a water-in-oil microemulsion from a temperature higher than the HLB temperature of the system to a temperature below it, at which two phases, an aqueous micellar solution phase and an oil phase, appear [28]. The emulsions produced by this method possess smaller droplet size with narrower size distributions than those obtained by the other two methods described above. Fig. 7 shows the phase behavior of the system 0.1 M aqueous NaCl//Ri2EO4/decane as a function of decane concentration and temperature [27]. The detailed phase behavior is depicted in Fig. 7b in a narrower range of temperatures than in Fig. 7a. When a composition corresponding to, for instance, point B is heated from 9 °C to higher temperatures, the phase changes are the following: initially an O/W mocroemulsion (Wm) forms and as the temperature increases a narrow two-phase region consisting of bicontinuous microemulsion and aqueous phases (D + W) is crossed and it is followed by another two-phase region composed of aqueous phase and lamellar liquid crystalline phases (LC + W). With increasing temperature, a single lamellar liquid crystalline phase is formed (LC). With further increase of temperature, the lamellar liquid crystal is melted and an isotropic phase L3 appears in which the surfactant molecules form a bicontinuous network [29]. Above the single-phase L3 region, excess water (or brine) is separated (region L3 + W). With increase in temperature, this two-phase region is continuously changed to reverse micellar solution (or W/O microemulsion) and aqueous phase (Om + W). Therefore, the spontaneous curvature or the surfactant aggregates change from convex to concave towards water. Phase inversion from water continuous O/W microemulsion to W/O highly concentrated emulsion occurs through a lamellar liquid crystalline phase and a bicontinuous surfactant L3 phase. The change in self-organising structure during spontaneous formation is schematically represented in Fig. 8. When the temperature change is slow the water droplet size in the gel emulsion is large because of coalescence of water droplets in the (L3 + W) region. In order to obtain fine droplet size emulsions.the temperature change should be very quick. The mechanism of highly concentrated O/W emulsion formation was determined by studying the phase behaviour of the 0.1 M NaCI aqueous
520
C. Solans et al.
solution/ Ri2E06/monolaurin/n-decane system as a function of temperature and brine concentration is shown in Fig. 9 [28].
Fig. 7. Phase diagram of the 0.1M aqueous NaCI/C^ECVdecane system as a function of temperature. Decane was added to 3wt% C12EO4 aqueous solution, and the weight percent of decane in the system is plotted horizontally, (a) Phase diagram over a wide range of temperature, (b) Detailed phase diagram around the HLB temperature. Wm oil-swollen micellar solution (O/W-type microemulsion); Om' water-swollen reverse micellar solution (W/O-type microemulsion); D, surfactant phase (middle-phase microemulsion); L3, bicontinuous surfactant phase; LC, lamellar liquid crystal; W and O, excess water and oil phases, respectively. I, II, and III indicate one-, two- and three-phase regions (Reproduced by permission from Ref. 27)
Highly Concentrated (gel) Emulsions: Formation and Properties
521
Fig. 8. Schematic change in the spontaneous curvature of surfactant layers in the process of spontaneous formation of W/O gel emulsions (Reproduced by permission from Ref. 27) The surfactant changes from hydrophilic to lipophilic with an increase in temperature, as illustrated by following the changes experienced by composition A in Fig. 9. At low temperatures, the surfactant dissolves as micelles in water and this phase coexists with excess oil (Wm + O). The excess oil phase is completely solubilised in the lower phase within a narrow temperature range at which the single isotropic L3 phase appears. In contrast to the L3 phase that appears in the water-rich region, the hydrocarbon parts of the surfactant layer are separated [29]. Above the L3 region (Fig. 9), a multiphase region with lamellar liquid crystal (LLC) is formed. With further increase of temperature a two-phase isotropic region (D + O) composed of a bicontinuous microemulsion phase (D) and an oil phase (O) is observed. In this two-phase region, emulsions are highly unstable and reach equilibrium within several minutes. At higher temperatures a single water-in-oil-microemulsion phase (I) appears and with further increase of the temperature an excess water phase is separated from the single-phase microemulsion. A W/O emulsion consisting of two liquid phases, a W/O microemulsion (Om) and an aqueous phase, exists in this (Om + W) region.
522
C. Solans et al.
Fig. 9. Phase diagram of the 0.1M aq. NaCl/Ci2EO6/monolaurin/n-decane system as a function of temperature. The Ci2E06:monolaurin:n-decane ratio is kept constant at 3.5:1.5:95. The weight percentage of 0.1 M aq. NaCI in the system is plotted horizontal. I, II and III indicate one-, two- and three-phase regions. W excess water phase; O, excess oil phase; Wm oil-in-water microemulsion; Om, water-in-oil microemulsion; L3, reverse bicontinuous phase; D, surfactant phase; LLC, multiphase region including lamellar liquid crystal. (Reproduced by permission from Ref. 28).
The phase behaviour studies represented in Fig. 9 suggest that at composition A the spontaneous curvature of surfactant molecular layers continuously changes from convex to oil to convex to water while cooling. The change in self-organising structures during emulsification is schematically shown in Fig. 10 [28]. The self-organising structures change from reverse micelles or water-in-oil (W/O) microemulsions to an oil-in-water (O/W) emulsion via the bicontinuous microemulsion D phase, the lamellar liquid crystal and the L3 phase with a decrease in temperature. The D phase and LLC phase coexist with the oil (O) phase, but oil and surfactant form a bicontinuous structure in the L3 phase. The existence of a single L3 phase region was shown to be very important for spontaneous emulsification [28]. Similar to the process of emulsification in W/O emulsions, an unstable emulsion region is crossed during the process. Therefore the more rapidly the temperature is lowered, the better the emulsification.
Highly Concentrated (gel) Emulsions: Formation and Properties
523
Fig. 10. Schematic change in self-organising structures during the spontaneous formation of highly concentrated emulsions. (Reproduced by permission from Ref. 28).
In systems with aliphatic hydrocarbons, the stability is maximum at about 25-30°C above or below the corresponding HLB temperature for W/O or O/W emulsions, respectively, due to the change in structure of the continuous phase as the system moves away from the balanced temperature [5]. 2.3. Structure The structure of highly concentrated emulsions was inferred from phase behaviour studies. Both O/W and W/O separate into two isotropic liquid phases [5,25]. O/W highly concentrated emulsions separate in an oil phase and a swollen aqueous micellar solution (O/W microemulsion droplets). Therefore, a O/W highly concentrated emulsion can be considered as an oil phase dispersed in a O/W microemulsion, structure similar to conventional emulsions. In the other hand, W/O highly concentrated emulsions were found to separate in an aqueous phase and a swollen reverse micellar solutions (or W/O microemulsion). The structure of equilibrium phases of W/O emulsions was studied by Fourrier Transform Pulse Gradient Spin-Echo (FT-PGSE) NMR [13], determining the self-diffusion coefficients of water, surfactant and oil, for the different phases. The results confirmed that the external continuous phase of the highly concentrated emulsions is a discrete microemulsion. Equilibrium separated phases and non-equilibrium W/O highly concentrated emulsions were studied by small angle X-ray scattering (SAXS).
524
C. Solans et al.
No significant differences were detected between the structures of the respective microemulsions [30]. The specific surface area and mean droplet size of highly concentrated emulsions was determined, from SAXS absolute intensity, according to the Porod-Auvray equation [31]. The results showed that the average droplet size increases on increasing the volume fraction of the dispersed phase, oil/surfactant ratio, salinity and temperature. These results were interpreted in terms of surfactant availability for surface coverage, and of interfacial tension [30]. At very high volume fractions of dispersed phase, close to unity, most of the surfactant is adsorbed on droplet surface and therefore the number of microemulsion droplets, contained in the continuous external phase, is small. Confirmation of these results was obtained by electron spin resonance (ESR) [32,33]. The variations in apparent order parameter and the isotropic hyperfine splitting constant, were determined in W/O emulsions, as a function of the volume fraction of the dispersed phase [32]. The results indicated that increasing the dispersed phase volume fraction, the water-oil interfacial area increases and the surfactant molecules are preferentially adsorbed at droplet surfaces. Therefore, the number of microemulsion droplets decreases with increasing water content and finally no microemulsion droplets are present in the continuous media at very high water content, when the fraction of water in the system is around 0.99. 2.4. Rheological Properties Rheology studies of highly concentrated emulsion have shown that these systems produce a viscoelastic response which can be described in terms of Maxwell model [34]. At low frequency the viscous modulus G" generally is bigger than the elastic modulus G'. However, at high frequencies, the elastic modulus can be the predominant, being the main contribution to the shear modulus. Experimental results have shown that both the elastic and the viscous modulus produce a good fit to Maxwell equations [34], described as follows: G>,G.=
^
i+(cody
G .- / G o =-^V
0)
(2)
l + (eo0)2 where to is the pulsation frequency and 6 is the system relaxation time, which corresponds to: O = rjl Go where r\ is the viscosity and Go the static shear modulus.
(3)
Highly Concentrated (gel) Emulsions: Formation and Properties
525
The normalized elastic modulus G7G0 and the normalized viscous modulus G'VGo are function of the pulsation frequency as shown in Fig. 11. The rheology of highly concentrated emulsions was well described by Princen et al., [2-4] who built a model based on assuming infinitely long cylindrical drops and interfacial tension independent on strain. The variables considered were the volume fraction of the dispersed phase, the radius of monodispersed droplets, the interfacial tension, the thickness of the films between droplets and the contact angle separating adjacent droplets.
Fig. 11. Maxwell theoretical elastic modulus G' and viscous modulus G", as a function of the adimensional time co9. The values of G' and G" have been normalized dividing by the static shear modulus Go.
Fig. 12. Droplet deformation and reorientation. 1) Initial unstrained configuration, 2) Strained, but below droplet reorientation, 3) Strained to maximum, 4) Unstrained, after reorientation. (Reproduced by permission from Ref. 3).
526
C. Solans et al.
Princen was able to derive the relationship between stress and strain, as a function of the above variables. The theoretical analysis was based on reorientations induced by strain, as shown in Fig. 12. According to the scheme shown in Fig. 12, droplets can be strained up to a maximum, after which droplet reorientation occurs. A semiempirical equation was derived, with two fitting parameters that could be determined by studying the static shear modulus, Go, as a function of average droplet radius, R32, interfacial tension, y, and volume fraction, (|). G o = « ^ f^-b)
(4)
This equation describes properly the strong dependence of rheological properties on the volume fraction of the dispersed phase (j>. The static shear modulus greatly increases as the volume fraction tends to unity, as it can be seen in the example in Fig. 13. According to the Maxwell liquid model, it can be considered that the system has two components, an elastic solid and a liquid of viscosity r\. In the analysis of highly concentrated emulsions, the elasticity is due to the structure of elastic films and the viscosity is due to the shearing of the continuous phase Considering that only a part of the system is responsible for the modulus, the viscosity of the continuous phase should be divided by the fraction of the system responsible for the relaxation time [34], as indicated in the following equation: T, = K T , C / ( 1 - 0 )
(5)
And using equations (3) and (5), the relaxation time 9 is: e = *Tic/(Go(i-<|0)
(6)
where k is an adjustable parameter, r\c is the viscosity of the continuous phase and (j> is the volume fraction of the dispersed phase. The relaxation time, as calculated by this equation, is shown in Fig. 14. The experimental relaxation times, obtained from the same system as before, are also represented. The agreement between the experimental and the theoretical relaxation time is fairly good. The best fit is obtained at high volume fractions of dispersed phase (<])>0.9). Therefore, it can be inferred that the main assumptions are basically correct: the viscous modulus arises from the viscosity of the continuous phase and the elastic modulus from the structure of elastic films. Consequently, the decrease in the fraction of continuous phase produces a decrease in the viscous component and an increase in the elastic component.
Highly Concentrated (gel) Emulsions: Formation and Properties
527
Fig. 13. Static shear modulus, Go, as a function of volume fraction of dispersed phase. Experimental and theoretical values, according to equation (4), are shown. The system is H2O/Ci6E4/decane, with oil-to-surfactant ratio equal to 1.5 at 40 °C. The lines are guides to the eye. (Reproduced by permission from Ref. 34).
Fig. 14. Experimental and theoretical values of relaxation time. The system is H2O/Ci6E4/decane, with oil-to-surfactant ratio equal to 1.5 at 40 °C. The lines are guides to the eye. (Reproduced by permission from Ref. 34).
528
C. Solans et al.
3. HIGHLY CONCENTRATED EMULSIONS BASED ON MICELLAR CUBIC PHASES 3.1. Micellar cubic phases: general aspects Micellar cubic phases are isotropic, transparent and highly viscous liquid crystals, composed of discrete aggregates (micelles) ordered in a threedimensional array. The normal discontinuous cubic phase (Ii) is built up with discontinuous hydrocarbon regions embedded in continuous aqueous medium, namely, normal micellar aggregates in water. The reverse discontinuous cubic phase (I2) consists of discontinuous aqueous regions embedded in a continuous lipid medium, namely, reversed micellar aggregates in oil. Ii phases show highly positive surface curvature and they have been found in hydrophilic systems containing anionic [35], cationic [36,37], non-ionic surfactants [38-40] and phospholipids [41], The I2 phase is less common but it has been found in lipids [42,43] and block-copolymer [44,45] systems. Some cubic phases in a binary system may incorporate a hydrocarbon [46-53], or some other organic compound [54,55], while other cubic phases lack this property and deteriorate. The interplay of two or more compounds may promote the formation of the cubic phase. This increases the possibility of finding cubic phases at various concentrations in amphiphilic systems. A spectacular feature of some cubic phases is their ringing property: they sound when struck [56]. However, this property is not characteristic only for cubic liquid crystalline structures, since it is connected with the threedimensional structure. The knowledge of the phase equilibria is important to understand the structure of the cubic phase as thermodynamically stable equilibrium systems. A cubic phase, when formed may be pseudostable for a very long time at nonequilibrium temperatures. One general location where normal micellar cubic phases have been found is that between the aqueous micellar solution phase and the normal hexagonal phase [38,41]. The structure for the cubic phase in this location has been debated. Early proposals were based upon the concept of closed globular aggregates. Among the various models proposed for the Pm3n cubic phase [5765], two of the more plausible are depicted in Fig. 15. In Fontell's model (Fig. 15a), the cubic phase is composed of short rod-like aggregates; two of them in the unit cell have complete rotational freedom while the six other only can rotate laterally. This structure has obtained support from NMR [58-60] and fluorescence quenching [62] studies. Low temperature and transmission electron microscopy [63,64] have provided more evidence about the presence of globular aggregates. The structure proposed by Charvolin and Sadoc (Fig. 15b) is analogue to that present in foams, namely, one consisting of a unit cell that contains a dodecahedron and a tetrakaidecahedron. This proposal is in agreement with X-
Highly Concentrated (gel) Emulsions: Formation and Properties
529
ray and NMR measurements, although in some cases the estimated micellar parameters are not as reasonable as expected. Although the previous models are still on debate, it can be said that among the four discontinuous cubic phases clearly identified [66] in amphiphiliccontaining systems, three are of the normal I, type (type I), which are assigned to the simple cubic space group Pm3n (8 micelles) [61], the face centred Fm3m space group (4 micelles)[43], and the body-centred Im3m space group [43]. Concerning the reverse I2 type (type II), only a structure containing 24 micelles and assigned to the Fd3m space group has been reported [66].
Fig. 15. Models for the cubic phase of space group Pm3n (a) Fontell [61] (b) Charvolin and Sadoc [65].
530
C. Solans et al.
There is a great deal of experimental evidence showing that liquid crystalline structures in lip ids participate in many important functions of the cells, and some of them might show three-dimensional symmetry, similar to the cubic phase. Cubic phases can also serve as stable and versatile matrices in a variety of fields ranging from material sciences, in which the may be used as reaction vessels or templates for the synthesis of structures with novel properties, to biology, where eompartmentation and enzyme immobilization play a significant role in metabolic regulation in vivo [67]. There might also be chemical uses such as the fabrication of semiconductor "quantum dot" superlattice arrays [68]. Moreover, it is thought that the cubic phase is related to many of the so-called : 'gel" materials [69-71], which constitute the basis of many formulations of creams, transdermal therapeutic systems and topical preparations. 3.2. Normal micellar cubic (Ii) phases with added oil As already mentioned, the normal micellar cubic phase is formed in highly hydrophilic surfactants, such as poly(oxyethylene) alkyl ethers with long ethylene oxide (EO) chains. The phase behavior of C^ECWwater/oil systems as a function of temperature is shown in Fig. 16. The single-phase cubic liquid crystal (Ii) is formed in a narrow range of oil concentration. A two-phase region (Ii plus excess oil) phase appears beyond the solubilization limit. At high temperatures, the Ii phase melts to an isotropic solution.
Fig.16. Phase behavior of CnECWwater/n-decane system as a function of temperature. The CnECWwater mass ratios are 50/50. Tmax is the maximum melting temperature of cubic phase. Ii = normal micellar cubic phase; Wm =micellar solution phase; O = excess oil.
Highly Concentrated (gel) Emulsions: Formation and Properties
531
The melting point of the cubic phase is plotted in Fig. 17 as a function of oil concentration for different oils. As described later, the solubilization of oil in the cubic phase and in the fluid micellar phase is low in this system. Hence, in most regions of Fig. 3 an excess-oil phase is present. It has been found [73-75] that in some systems there is a close structural relation between the cubic phase and the adjacent micellar solution, and that the transition between these two phases looks very much like a melting process in a certain range of surfactant concentration. In fact, the non-fluid I, phase suddenly starts to flow at the boundary of the 11-phase region. In the case of n-decane, the thermal stability of the cubic phase increases upon addition of a very small amount of oil. The melting point of the cubic phase increases sharply with the incorporation of oil in the micelles forming the cubic structure and then becomes constant when phase separation occurs. The thermal stability of the cubic phase may be expressed by Tmax, the maximum melting temperature of the cubic phase. It seems that both the thermal-stability-enhancing effect and the solubility of oil in the cubic phase decrease as the molar volume of the oil increases, from decane to squalane. Oil has an effect in Tmax as long as it is incorporated in the micelles forming the cubic structure, therefore its solubility seems to determine the magnitude of such an effect.
Fig. 17. Effect of added oils on the melting temperature of cubic phases in C^ECWwater systems. The surfactant/water weight ratios are 50/50. (B) n-decane; (J) n-hexadecane; (G) squalane (il) triglyceride (Reproduced by permission from Ref. 72)
532
C. Solans et al.
Fig. 18 shows the effect of surfactant concentration in water on the melting temperature of the cubic phase in the presence of excess oil. With decreasing the molecular weight of oil, the maximum melting temperature increases and its composition tends to shift to higher surfactant concentration, except in the heptane system, where, instead of an aqueous micellar phase (Wm) for other oils, a birefringent liquid crystal (probably a hexagonal phase) appears at temperatures above the melting temperature of the cubic phase in the concentrated region. Since short-chain hydrocarbons like heptane have a strong tendency to penetrate in the surfactant palisade layer, its curvature might be changed to be less positive. As a result, the maximum temperature of the cubic phase in the heptane system is not really a melting temperature but a transition temperature to the hexagonal phase. Oils stabilize the cubic phase by inducing the formation of micelles with a moderate axial ratio, and therefore, a more positive curvature [39, 46, 66-70]. This might improve the packing and order of micelles in the cubic structure, so that the I,-Wm phase transition temperature increases. Accordingly, the effect of oil is more pronounced in the surfactant concentrated region, where micelles tend to deviate from the spherical shape. This is supported by the tendency observed in Fig. 18 and by the fact that in some systems the cubic phase only appear when oil is added [39,46,56,74].
Fig. 18. Effect of added oils on the maximum melting temperature of cubic phases 7 ^ in Ci2EO25/water systems. Ws is the surfactant/(surfactant+water) weight ratio. The total oil weight fraction in the system is fixed at 0.1. In all systems there is an excess-oil phase and micelles are considered to be "saturated" with oil. (O) n-heptane ( • ) n-decane; ( • ) nhexadecane;(D) squalane; (O) no oil added (Reproduced by permission of Ref. 72).
Highly Concentrated (gel) Emulsions: Formation and Properties
533
In Ci2EO25/water/oil systems, the I] phase is organized in a body-centered cubic structure, as deduced from SAXS results (Fig. 19).
Fig. 19. SAXS data for the cubic phase in C^ECfe/water/decane systems. Composition (wt%): 2% n-decane, 49% C|2EO2s, 49% water, q is the scattering vector, h, k, I are the Miller indices.
534
C. Solans et al.
For nm spherical micelles packed in a cubic lattice, the following equations can be derived for the radius of the hydrophobic part of micelles, r, and the effective cross sectional area per surfactant molecule, as in the Ii phase:
m
r
' v T3 14W5J 3 v , (,\
a = —M -^—^
^ A «>, J
(8)
/m w
where p is the unit cell parameter, vL is the molar volume of the hydrophobic part of surfactant molecule, Vi0 is the total hydrophobic volume per micelle, A^ is Avogadro number and
Fig.20. SAXS data for I| cubic phases in water/CuECWoil systems as a function of <j>0, the volume fraction of oil(decane). The CnECWwater weight ratios are 50/50. Values correspond to a body-centered cubic structure. ( • ) Interlayer spacing, d; ( • ) Radius of lipophilic core, r; (D) Effective cross sectional area per surfactant molecule, as.
Highly Concentrated (gel) Emulsions: Formation and Properties
535
As oil is incorporated in the micellar core, micelles grow and d and r increase until an excess-oil phase separates. Hence, the solubilization of oil in the cubic phase can be estimated by the inflection point in d curve. For ndecane, as also increases as oil is dissolved in the cubic phase. The mean interaction free energy per surfactant molecule in the aggregates can be expressed by [76]: M°N= yas +Klaf
(10)
The second term {Klaf) on the left hand of Eq. (10) is the repulsive contribution to the free energy. /? can be taken as 3 for steric-type repulsion [77], as in the case of non-ionic surfactants. When the energy is minimum, df/Nl da = 0 and from Eq. (9), K = ao4y/ 3, where a0 is the optimal area. Replacing K in Eq. (10), the repulsive contribution to the free energy per amphiphile (Erep) is assumed to be proportional to (ao4/a/) [77]. For C12EO25 systems, a0 is the area per head group when the ethylene oxide chains are in an optimal conformation, namely, neither compressed nor extended [79]. The following equation has been proposed [79] for the calculation of as /a0:
ao~CdV 3 ) [ ^ ^ + O 1 / 3 - ( ^ + O " 3 } j where r0 is the length of the hydrophobic moiety corresponding to <j>0 = 0, C = (h2+k2+l2)m, h, k, I are the Miller indices, d is the measured interlayer spacing and 0S is the volume fractions of surfactant. The variation of the repulsive term Erep = (ao4/a/) calculated from Eq.(ll) as a function of >0, is shown in Fig. 21. Erep increases with oil content, and it is larger for long chain hydrocarbons such as hexadecane. The repulsion inhibits changes in curvature and favours structures with spherical aggregates, such as the I] phase. Moreover, it is suggested that the increase in repulsion between headgroups decrease the flexibility of the surfactant layer , and hence, the solubilization of oil is affected. The aggregation number in the micelles forming the cubic phase can be calculated by Nagg= 47tr3NA/3vL. The increase in surface area caused by repulsive interactions is compensated by an increase in the aggregation number, as is shown in Fig. 22. As the the hydrophobic volume in the system gets larger, more surfactants molecules should be incorporated in aggregates, which is equivalent to an increase in the attractive contribution to the free energy.
536
C. Solans et al.
Fig.21. Change of the repulsion term Erep as a function of the volume fraction of oil $,. • n-decane O n-hexadecane.
Fig. 22. Aggregation number A ^ as a function of the total lipophilic volume fraction, $>+^L, in CnEOzs/water/oil systems.The surfactant/water ratio is 50/50. Bn-dccanc O n-hexadecane
3.3. Formation and properties of highly concentrated cubic-phase-based emulsions The phase diagram of the ternary system C^EC^/water/decane at 25 °C is shown in Fig. 23. The solubilization of oil in the cubic phase is small and an excess-oil phase is separated at the solubilization limit; however, a considerable
Highly Concentrated (gel) Emulsions: Formation and Properties
537
amount of oil (up to 90 wt% n-decane) can be incorporated in the Ii+O region in transparent or translucent gel-like emulsions having an Ij cubic phase as the external phase (Fig. 24). These emulsions are quite stable, since coalescence and creaming is prevented by the extremely high viscosity of the external phase. The transparent gels, which are formed at a water/surfactant weight ratio equal to 50/50, can be diluted with water to form normal emulsions with a small droplet size (1-10 um). The formation of cubic-phase-based concentrated emulsions (although not always transparent) has also been reported in polyglycerol fatty acid esters [80], sugar esters [81] and middle-chain poly(oxyethylene) alkyl ether systems [82], and also in ionic surfactant systems [83]. As can be seen in Fig.25, even if the cubic phase does not appear at 25 °C in the binary surfactant/water mixtures, it is formed when oil is added either to a concentrated micellar solution or to a hexagonal phase. This suggests again that the solubilization of oil in the micellar cores induces a change to a more positive curvature and promotes the formation of spherical or quasi-spherical micelles arranged in a cubic structure.
Fig.23. Partial phase diagram of water/C^EChs/n-decane system at 25 °C. O is an excess-oil phase (Reproduced by permission from Ref. 72).
In the so-called D-Phase emulsification method developed by Sagitani et al, commercial nonionic surfactant is mixed with polyol, oil and a small amount of water to produce a translucent or transparent gel [69]. Then a fine emulsion is prepared by diluting the gel with water. It was established recently [82] that this gel is actually a O/Ii concentrated emulsion.
538
C. Solans et al.
Fig. 24. Schematic drawing of a highly concentrated O/Ii emulsion
Fig.25. Phase diagram of water/ C^ECVn-decane system at 25 °C. H; is a hexagonal phase (Reproduced by permission from Ref. 81).
Highly Concentrated (gel) Emulsions: Formation and Properti
539
In order to prepare a cubic-phase-based emulsion, the cubic phase should be melted to enable an adequate mixing during the emulsification process. As the mixture is agitated, the oil incorporates gradually in the surfactant phase and finally a gel is obtained by cooling the sample. A picture of a transparent gelemulsion is presented in Fig. 26. A photomicrograph of the same gel-emulsion, obtained by video enhanced microscopy (VEM), is also shown in Fig. 27. The emulsion is polydisperse and some of the droplets are polyhedral, since the volume fraction of internal phase (hydrocarbon) exceeds the maximum value for a packing of spheres, 0.74. Considering only the droplet size depicted in Fig.27, this emulsion should not be transparent, but Fig 28 gives the answer for the transparency. The values of the refractive index of the cubic phase and oil are very similar when the water/surfactant ratio is 50/50. In fact, gels became turbid for other surfactant/water ratios.
Fig.26. Photograph of a transparent 11-phase-based emulsion containing 90 wt% n-decane. The CnECWwater weight ratio is 50/50 in this sample (Reproduced by permission from Ref. 72).
540
C. Solans et al.
Fig.27. VEM photomicrograph of the 11-phase-based emulsion in Fig.26. Water-soluble dye was added to enhance contrast (Reproduced by permission from Ref. 72).
Fig. 28. Refractive index at 25 °C as a function of surfactant concentration in CnECWwater system. The dashed line indicates the refractive index of decane (Reproduced by permission from Ref. 72).
Highly Concentrated (gel) Emulsions: Formation and Properties
541
3.4. Rheology of Ii-phase-based emulsions Micellar cubic phases show viscoelastic behavior, shear thinning, a very high elastic modulus (G' > 104 Pa) and shear-induced alignment [56,84]. The results of a representative dynamic frequency sweep test for samples with different oil content in Ci2EO25/water/oil system are shown in Fig.29. As expected for cubic phases, |r|*| is very high and shows no plateau in the measured frequency range. For comparison, the complex viscosity and storage (elastic) modulus G' at fixed frequencies (low frequency region for |r|*| and high frequency region for G') are plotted as a function of the volume fraction of decane in Figure 30.
Fig. 29. Complex viscosity |T|*| as a function of frequency for different volume fractions of decane (along line B of Figure 9) in C^ECWwater/decane systems at 25°C. (strain = 1 %). The CnECWwater weight ratio is fixed at 50/50. Lines are only guides to the eyes. (Reproduced by permission from Ref. 85).
542
C. Solans et al.
Fig. 30. Complex viscosity |r|*| at co = 0.01 rad s"' and storage modulus G' at co = 10 rad s" as a function of the volume fraction of decane in C^ECWwater/decane systems at 25°C. (strain = 1 %). Lines are only guides to the eyes. (Reproduced by permission from Ref. 85).
Both |T|*| and G' first slightly increase and then decrease with increasing oil fraction. The maximum is close to the phase separation boundary. The rheological behavior of cubic-phase based emulsions (in the two-phase region) is opposite to that usually found in two-liquid concentrated emulsions, in which viscosity increases with oil content [3]. It is considered that the viscosity of the cubic-phase based emulsions is mainly determined by the cubic phase, therefore it can be expected that the viscosity will decrease with the cubic phase fraction in the system. The slight increase in viscosity with oil content at low oil fractions is probably associated with an increase in the strength of interactions as micelles swell with oil. Representative data of dynamic frequency sweep experiments at different oil contents are presented in Figure 31. The addition of very small amounts of oil changes remarkably the dynamic rheological behavior.
Highly Concentrated (gel) Emulsions: Formation and Properties
543
Fig. 31. Dynamic frequency sweep tests (strain = 1 %) for C^ECWwater/decane systems at 25°C for different volume fractions of decane. Filled symbols correspond to the storage modulus G' whereas open symbols correspond to the loss modulus G". The C^ECWwater weight ratio is fixed at 50/50. (Reproduced by permission from Ref. 85).
The systems show highly non-maxwellian behavior, as can be deduced by slope of the curves in the low frequency range and by the presence of a minimum in the loss (viscous) modulus G", which is a sign of structural arrest [86]. The variation of G' and G" with frequency is qualitatively similar to that found in solutions of star-like polymers that exhibit a micellar-like structure [87]. The crossover between G' and G" shifts to lower frequencies with increasing oil content, indicating an increase in the relaxation time, similar to what is found in concentrated two-liquid emulsions [88]. As a matter of fact, the system passes from liquid or glassy like in the absence of oil (G" prevailing over G) to gel-like (G' prevailing over G") as the oil fraction is increased. The polydispersity of the systems as well as the presence of aging phenomena [89] might play role in this behavior. 3.5. Reverse micellar cubic (I2) phases with added water and related highly concentrated emulsions Reverse cubic phases are formed in the concentrated region of poly(oxyethylene) poly (dimethylsiloxane) surfactants (Me3SiO -(Me2Si0),,,_2Me2SiCH2CH2CH2- O-(CH2CH2O)nH, abbreviated SimC3EO«), as shown in Fig. 32. Similar to the case of Ij phase (Fig. 16), the melting point of the I2 phase increases as water is solubilized in the reverse micelles until phase separation occurs, namely, the melting temperature of the I2 phase is maximum in the two-
544
C. Solans et al.
phase region. I2 can be also formed in surfactant/oil systems or even in the absence of solvent, as shown in Fig. 33.
Fig. 32. Phase diagram of binary water-Si 14C3EO7 8 system forming the I2 phase, h, O m , W and II represent reverse micellar cubic, a surfactant liquid phase containing small amounts of water, excess water phase and a two-phase region, respectively. (Reproduced bi permission from Ref. 90).
Fig. 33. Phase diagram of the water-Si2sC3EOi5 8-octamethylcyclotetrasiloxane (D4) system at 25 °C. I2, reverse micellar cubic phase; Om, reverse micellar solution or surfactant liquid phase; W, an excess water phase; S, solid present phase; II, two-phase region; III, three-phase region. SAXS measurements were performed to determine the phase boundary between I2 and I2 + W regions along Si25C3EO|5.g/D4 weight ratios of 80/20 and 70/30 indicated by lines A and B, respectively. (Reproduced by permission from Ref. 14).
Highly Concentrated (gel) Emulsions: Formation and Properties
545
The I2 phase in the absence of solvent is organized in a face-centered structure (Fd3m space group), as deduced from synchrotron SAXS results [14]. This I2 phase solubilizes a maximum of 28 wt% water (on the water-surfactant binary axis). An excess water phase separates beyond the solubilization boundary of each Om and I2 phase. The phase boundary between I2 and I2 + W regions is difficult to determine by visual observation, but it can be done using SAXS measurements. The interlayer spacing (d) of samples at constant Si25C3EOi5 8/D4 weight ratios of 80/20 and 70/30 as a function of water content is shown in Fig. 34. d first increases as micelles size increases with the solubilization of water in the reverse micellar core and then reaches a plateau when phase separation takes place. Hence, the inflection point of the J-curve clearly indicates the phase boundary between the homogeneous I2 phase and an excess water region. A large amount of water (up to 90 wt%) can be incorporated as dispersed droplets in the 12 + W region forming a turbid gel like emulsion having the I2 phase as the external continuous phase (Fig. 35). Water incorporation in the stable emulsions can be increased (up to 96wt% of water) upon addition of oil to the systems. The presence of oil molecules into the surfactant palisade layer induces the change in curvature and flexibility of the surfactant layer that allows higher amounts of water to be incorporated in the stable gel-emulsions. Excess water does not separate out even if the samples are centrifuged at 4000 rpm for 12 hours at room temperature [15]. The emulsions are also thermally stable and can be diluted with octamethylcyclotetrasiloxane (D4) to form normal emulsions with a small droplet size (1-10/m).
Fig. 34. Variation of d in the h phase in the \vater-Si25C3EO158-D4 system as a function of water content at 25 °C. Lines A and B indicate the weight ratios of Si25C3EOi5 ,8/D4 of 80/20 and 70/30, respectively. (Reproduced by permission from Ref. 14).
546
C. Solans et al.
Fig. 35. Schematic drawing of a highly concentrated W/I2 emulsion.
Fig. 36. Refractive index of water + glycerol as a function of glycerol concentration. The dashed line indicates the refractive index of the h phase of composition 50 wt% Si25C3EOi5 8/50 wt% D4. (Reproduced by permission from Ref. 15).
Highly Concentrated (gel) Emulsions: Formation and Properties
547
The W/I2 gels look turbid over the entire range of compositions in which they can be produced, due to the difference in the values of refractive indices of the cubic phase and water. However, this gap can be reduced by partially replacing water with glycerol. Fig. 36 shows the refractive index of water + glycerol as a function of glycerol concentration. The dashed line indicates the refractive index of the I2 phase of composition Si25C3EOi5.8/D4 =1. It is observed that the refractive indices of the I2 phase and dispersed phase match each other when the water/glycerol weight ratio is 1:1. A photograph of a transparent W/I2 gel emulsion of the composition: 90 wt% aqueous phase (water/glycerol = 45/45) 10 wt% I2 phase (Si25C3EO,5.8/D4 = 1) is shown in Figure 37. Inverse cubic phase emulsions are also formed in systems containing nonhydrocarbon oil, such as d-limonene [91]. In the system of Figure 38, the I2 phase is not present neither in the binary water/surfactant system nor in the binary oil /surfactant system, but it is formed when water is added to a reverse micellar solution or when oil is added to the reverse hexagonal phase (H2). A photomicrograph of a gel-emulsion in the system of Fig. 38 is shown in Fig.39. Similar to Fig. 27, the emulsion droplets are polydisperse and some of the droplets look polyhedral, since the volume fraction of the internal phase (water) is quite large (<|)w > 0.74).
Fig. 37. Photograph of a transparent I2 phase-based emulsion containing 90 wt% dispersed aqueous phase (water/glycerol = 45/45) and 10 wt% h phase (Si2sC3EOi5 8/D4 weight ratio is 5/5). (Reproduced by permission from 15).
548
C. Solans et al.
Fig. 38. Phase diagram of water/PEOS-5/d-limonene (LN) system at 25 °C. Concentrated gel emulsions can be formed in the shaded region (Reproduced by permission from 91).
Fig. 39. VEM photomicrograph of a h phase based gel emulsion containing 89% wt% water. The PEOS-5/d-limonene weight ratio is 75/25. (Reproduced by permission from Ref. 91).
Highly Concentrated (gel) Emulsions: Formation and Properties
549
4. ADHESION PROPERTIES Highly concentrated emulsions may have a rather different behaviour when placed in contact with an excess volume of the continuous phase, due to the adhesive properties between adjacent droplets, as reviewed by Babak [11]. Some emulsions (adhesive) remain stable whereas another type of emulsions (nonadhesive) relax and the volume fraction of dispersed phase decreases. The adhesive phenomena of some emulsions is also observed in its spontaneous formation by creaming or sedimentation of diluted emulsions. In some cases, concentrated emulsions, which a volume fraction of the dispersed phase (j)>0.74, may form spontaneously by creaming in diluted emulsions. For instance, O/W emulsions stabilised by SDS show this behaviour at relatively high NaCl concentrations [92]. The cream may not still contain a significant portion of continuous phase, since such cream may consist of rigid and large droplet flocks that possess a relatively open structure. However, high volume fractions, (|>>95, can easily be achieved by centrifugation of diluted emulsions [11]. These highly concentrated emulsions do not relax after centrifugation and the cream, possessing high volume fraction of dispersed phase, is stable in contact with the continuous phase. Therefore, these emulsions show adhesive properties between droplets. Nevertheless, there is another type of highly concentrated emulsions which is not formed spontaneously by sedimentation or creaming. The volume fraction of the dispersed phase can be greatly increased by centrifugation but emulsions tend to relax afterwards if placed again under normal gravity [11]. In this case, the continuous phase, in contact with the highly concentrated emulsions, is absorbed back into the emulsion, and thus the sediment spontaneously reduces its volume fraction of dispersed phase. Therefore, highly concentrated emulsions can behave in two distinct types: adhesives and non-adhesives. Adhesive emulsions can form spontaneously by sedimentation and are stable if placed in contact with the continuous phase. Non-adhesive emulsions do not form spontaneously by sedimentation at normal gravity. They can be formed by centrifugation but they relax afterwards. Examples of adhesive and non-adhesive emulsions are shown in Fig. 40, for the system H20/SDS/dodecane. Fig. 40a shows the highly concentrated emulsion, adhesive because of the addition of 0.7 M NaCl, observed after centrifugation at 2000 rpm for 5 min. This emulsion was stable two weeks after centrifugation. Fig. 40b shows that a highly concentrated emulsion is also formed by centrifugation in the system without NaCl centrifugation. However, this emulsion is non-adhesive and it relaxes in approximately two hours after centrifugation, decreasing the volume fraction of the dispersed phase, as shown in Fig. 40c.
550
C. Solans et al.
Fig. 40. Emulsions in the system H2O/SDS/dodecane (12 mM SDS, 20 wt% dodecane), observed after centrifuging 5 min at 2000 rpm a) Stable adhesive emulsion, obtained in the system containing 0.7 M NaCl, b) Unstable non-adhesive emulsion, prepared without NaCl, c) Emulsion b after approximately two hours.
The different behavior is related to the interaction forces between droplets, and more specifically, to the interaction forces in the thin liquid films between adjacent droplets. Studies of drop adhesion, investigated by optical microscopy, show that a transition between adhesion and non-adhesion can occur [93]. It has been shown that the adhesion behaviour of a pair of 50 urn emulsion droplets, stabilised by 2 mM SDS, has its adhesion transition (from zero to non-zero contact angle) at 0.4 M for room temperature [94]. The transition from adhesive to non-adhesive highly concentrated emulsions occurs at similar conditions than for a simple pair of droplets. The transition as a function of temperature and NaCl concentration, as described by Bibette et al. [94] for a pair of droplets, is shown in Fig. 41. Generally, emulsions stabilised by ionic surfactants become adhesive at high electrolyte concentrations because of the considerable screening of all electrostatic repulsions by electrolytes, as described by the DLVO theory [9596]. However, emulsions stabilised by nonionic surfactants do not possess long range repulsion forces such as electrostatic interactions and they can be adhesive even without electrolyte. Transitions from adhesive to non-adhesive emulsions, induced by increasing electrolyte concentrations, can be found in O/W systems stabilised by nonionic surfactants [97-98]. In these nonionic surfactant systems the minimum electrolyte concentrations required to prepare such adhesive emulsions is much smaller than the minimum electrolyte concentrations
Highly Concentrated (gel) Emulsions: Formation and Properties
551
generally needed to prepare adhesive emulsions in ionic surfactant systems. Therefore, without electrolyte, ionic surfactant systems are non-adhesive whereas nonionic systems can be adhesive Consequently, the sedimentation behaviour of emulsions can be studied by considering the forces of interaction between adjacent droplets. The film energy of interaction, per unit area of the film, can be calculated [93,99] from the interfacial tension, y, and the contact angle between droplets, G, as: E=2y(l-cosG)
(12)
The transition from adhesion to non-adhesion occurs when such interaction energy comes to zero. Not surprisingly, the interaction energy can not be explained in terms of DLVO theory, which only takes into consideration an electrostatic repulsion and a van der Waals attraction. Hamaker constant estimates, calculated from contact angle determinations, according to DLVO theory, are anomalously high (~100 times higher than accepted values) [11,100]. Certainly, pure DLVO theory cannot be applied to many adhesive systems, which possess additional attractive forces which have to be considered.
Fig. 41. Boundary between adhesion and non-adhesion for a pair of 50 fxm diameter dodecane-in-water emulsion drops stabilised by 2 mM SDS, as a function of temperature and NaCl concentration. (Adapted with permission from Ref. 94).
552
C. Solans et al.
One of the possible interaction forces than can explain the adhesion phenomena in highly concentrated emulsions is the attractive depletion force as reviewed by Babak [11]. The origin of these adhesive forces is related to the non-compensated osmotic pressure produced by surfactant aggregates on the emulsion droplets due to the impossibility for the aggregates to penetrate into the narrow gap between the droplets [101-104]. Some works show that the adhesion boundary of highly concentrated emulsions corresponds to the equilibria between the osmotic pressure fIm=CmRT and the capillary pressure Pc=2y/r, where Cm is the concentration of microemulsion droplets in the continuous phase, y the interfacial tension and r the droplet radius. Emulsions would be adhesive if the osmotic n m pressure exceeds the capillary pressure Pc [102]. However, depletion forces cannot account for all adhesion phenomena showed in highly concentrated emulsions. Theoretical considerations, based on availability of excess surfactant, suggest that the concentration of surfactant aggregates (microemulsion droplets) should decrease as the volume fraction of dispersed phase increases [30,32]. Electron Spin Resonance studies in W/O gel emulsions seemed to indicate that when the fraction of water in the system is very high, t()=0.99, most of the surfactant molecules are adsorbed at the oil-water interface and no surfactant aggregates are present in the continuous phase [32]. Therefore, emulsions with an extremely high volume fraction of dispersed phase have a osmotic pressure in the continuous phase close to zero, and adhesion cannot be explained by depletion. Other non-DLVO forces should be taken into consideration. These attractive interactions may arise from a variety of different forces (steric, hydrophobic forces, oscillatory forces, etc.). For instance, steric forces may be attractive if the surfactant chains are not placed in a good solvent [76]. In addition, factors arising from surface area extension and bending energy contributions may play an important role in highly concentrated emulsions, which are systems with a high degree of droplet deformation [105]. Therefore, the problem of droplet adhesion in highly concentrated emulsions is rather complex and its description is not in the scope of the present section. REFERENCES [1] [2] [3] [4] [5] [6] [7]
K.J. Lissant, J. Colloid Interface Sci., 22 (1966) 462. H.M. Princen, J. Colloid Interface Sci., 71 (1979) 55. H.M. Princen, J. Colloid Interface Sci., 91 (1983) 160. H.M. Princen and A.D. Kiss, J. Colloid Interface Sci., 112 (1986) 427. H. Kunieda, C. Solans, N. Shida and J.L. Parra, Colloids Surf., 24 (1987) 225. C. Solans, J.G. Dominguez, J.L. Parra, J. Heuser and S.E. Friberg, Colloid Polym. Sci., 266(1988)570. C. Solans, R. Pons and H. Kunieda. In B. P. Binks (ed.), Modern Aspects of Emulsion Science, The Royal Society of Chemistry, Cambridge, U.K., 1998, pp 367-394.
Highly Concentrated (gel) Emulsions: Formation and Properties
553
[8] J.C. Ravey, M.J. Stebe and S. Sauvage, Colloids Surf. A., 91 (1994) 237. [9] G. Ebert, G. Platz and H. Rehage, Berichte der Bunsengesellschaft, 92 (1988) 1158. [10] O. Sonneville-Aubrun, V. Bergeron, T. Gulik-Krzywicki, B. Jonsson, H. Wennerstrom, P. Lindner and B. Cabane, Langmuir, 16(4) (2000) 1566. [11] V.G. Babak and M.J. Stebe, J. Dispersion Sci & Tech, 23(1-3) (2002) 1. [12] H. Kunieda, N. Yano and C. Solans, Colloids Surf., 36 (1989) 313. [13] C. Solans, R. Pons, S. Zhu, H.T. Davis, D.F. Evans, K. Nakamura and H. Kunieda, Langmuir, 9(6) (1993), 1479. [14] Md. H. Uddin, C. Rodriguez, K. Watanabe, A. Lopez-Quintela, T. Kato, H. Furukawa, A. Harashima and H. Kunieda, Langmuir, 17 (2001) 5169. [15] Md. H.Uddin, H. Kunieda and C. Solans. In K. Esumi and M. Ueno (eds.), StructurePerformance relationships in Surfactants, Marcel Dekker, New York, 2003, pp 599-626. [16] E. Ruckenstein, Adv. Polym. Sci., 127 (1997) 3. [17] N.R. Cameron and D.C. Sherington, Adv. Polym. Sci. 126 (1996) 163. [18] J. Esquena, G.S.R.R. Sankar and C. Solans, Langmuir, 19(7) (2003) 2983. [19] H. Maekawa, J. Esquena, S. Bishop, C. Solans and B.F. Chmelka, Adv. Mater. 15 (2003)591. [20] P. Clapes, L. Espelt, M.A. Navarro and C. Solans, J. Chem. Soc. Perkin Trans., 2 (2001), 1394. [21] L. Espelt, P. Clapes, J. Esquena, A. Manich and C. Solans, Langmuir, 19 (2003) 1337. [22] H. Kunida and K. Shinoda, J. Disp. Sci. Technol., 3 (1982) 233. [23] H. Kunieda and K. Shinoda, J. Colloid Interface Sci., 107 (1985) 107. [24] K. Shinoda and H. Saito, J. Colloid Interface Sci., 26 (1968) 70. [25] H. Kunieda, D.F. Evans, C. Solans and M. Yoshida, Colloids Surf., 47 (1990) 35. [26] R. Pons, I. Carrera, P. Erra, H. Kunieda and C. Solans, Colloid Surf. A., 91 (1994) 259. [27] H. Kunieda, Y. Fukui, H. Uchiyama and C. Solans, Langmuir, 12 (1996) 2136. [28] K. Ozawa, C. Solans and H. Kunieda, J. Colloid Interface Sci., 188 (1997) 275. [29] U. Olson, M. Jonstromer, K. Nagai, O. Soderman, H. Wennerstrom and G. Klose, Prog. Colloid Polym. Sci., 76 (1988) 75. [30] R. Pons, J.C. Ravey, S. Sauvage, M.J. Stebe, P. Erra and C. Solans, Colloids Surf., 76 (1993) 171. [31] L. Auvray, J.P. Cotton, R. Ober, C. Taupin, J. Phys., 45 (1984) 913. [32] H. Kunieda, V. Rajagopalan, E. Kimura and C. Solans, Langmuir, 10 (1994) 2570. [33] V. Rajagopalan, C. Solans and H. Kunieda, Colloid Polym. Sci., 272 (1994) 1166. [34] R. Pons, P. Erra, C. Solans, J.C. Ravey and M.J. Stebe, J. Phys. Chem., 97 (1993) 12320. [35] T.C.Wong, N.B.Wong and P. J. Tanner, J. Colloid Interface Sci., 186 (1997) 325. [36] R.R. Balmbra, J.S.Clunie and J.F. Goodman, Nature, 222 (1969) 1159. [37] E. Blackmore and G.J.T. Tiddy, J. Chem. Soc, Faraday Trans. 2, 84 (1988) 1115. [38] J. Mitchell, G.J.T. Tiddy, L. Waring, T. Bostock and M. McDonald, J. Chem. Soc, Faraday Trans. 1, 97 (1983) 975. [39] H. Kunieda, K. Shigeta, K. Ozawa and M. Suzuki, J. Phys. Chem. B., 101 (1997) 7952. [40] B. Folmer, M. Svensson, K. Holmberg and W. Brown, J. Colloid Interface Sci., 213 (1999) 112. [41] P. Mariani, L. Amaral, L. Saturni and H. Delacroix, J. Phys. II, 4 (1994) 1393. [42] V. Luzzati, R. Vargas, A. Gulik, P. Mariani, J. Seddon and E. Rivas, Biochemistry, 31 (1992)279. [43] J.M. Seddon, N. Zeb, R.H. Templer, R.N. McElhaney and D.A. Mannock, Langmuir, 12 (1996)5250. [44] P. Alexandridis, U. Olsson and B.Lindman, Langmuir, 14 (1998) 2627. [45] P. Alexandridis, U. Olsson and B.Lindman, Langmuir, 12 (1996) 1419. [46] H. Edlund, M. Byden, B. Lindstrom and A. Khan, J. Colloid Interface Sci., 204 (1998) 312. [47] O. Soderman and L.B.A. Johansson, J. Colloid Interface Sci., 179 (1996) 570. [48] A. De Geyer, Progr. Colloid Polym. Sci., 93 (1993) 76. [49] M. Kotlarchyk, S. Chen, J. Huang and M.W. Kim, Phys. Rev. Lett., 9 (1984) 941.
554
C. Solans et al.
[50] J.A. Bouwstra, H. Jousma, M.M.Van der Meulen, C.C. Vijverberg, G.S. Gooris, F. Spies and H.E. Junginger, Colloid Polym. Sci., 267 (1989) 531. [51] H. Jousma, J.G.H. Jousten, G.S.Gooris and H.E. Junginger, Colloid Polym. Sci., 267 (1989)353. [52] J. Tabony, Nature, 319 (1986) 400. [53] E. Nurnberg, N. Pollinger, Tenside Det, 23 (1986) 26. [54] X. Auvray, M. Abiyaala, P. Duval, C. Petipas, I. Rico and A. Lattes, Langmuir, 9 (1993) 444. [55] T.A. Bleasdale, G.J.T. Tiddy and E. Wyn-Jones, J. Phys. Chem., 95 (1991) 5385. [56] M. Gradzielski, H. Hoffmann and G. Oetter, Colloid Polym. Sci., 268 (1990) 167. [57] V. Luzzati and P.A. Spegt, Nature, 215 (1967) 701. [58] P.O. Eriksson, A. Khan and G. Lindblom, J. Phys. Chem., 86 (1982) 387. [59] P.O. Eriksson, G. Lindblom and G. Arvidson, J. Phys. Chem., 91 (1987) 846. [60] P.O. Eriksson, G. Lindblom and G. Arvidson, J. Phys. Chem., 89 (1985) 1050. [61] K. Fontell, Colloid Polym. Sci., 268 (1990) 264. [62] L.B.A. Johansson and O. Soderman, J. Phys. Chem., 91 (1987) 5275. [63] J.L. Burns, Y. Cohen and Y. Talmon, J. Phys. Chem., 94 (1990) 5308. [64] H. Jousma, J.A. Bowstra, F. Spies and H.E. Junginger, Colloid Polym. Sci., 265 (1987) 830. [65] J. Charvolin and J.F. Sadoc, J. Physique, 49 (1988) 521. [66] V. Luzzati, H. Delacroix and A. Gulik, J. Phys. II, 6 (1996) 405. [67] E.M. Landau, G. Rummel, S. W. Cowan-Jacob and J.P. Rosenbusch, J. Phys. Chem. B., 11,(1997) 1935. [68] C.B. Murray, C.R. Kagan and M.G. Bawendi, Science, 270 (1995) 1335. [69] H. Sagitani, Y. Hirai, K. Nabeta and M. Nagai, Yukagaku, 2 (1986) 102. [70] H. Sagitani, J. Disp. Sci. Tech., 9 (1988) 115. [71] A. Tsugita, Y. Nishijima and T. Sasaki, Yukagaku, 4 (1980) 227. [72] C. Rodriguez, K. Shigeta and H. Kunieda, J. Colloid Interface Sci., 223 (2000) 197. [73] D. Danino, Y. Talmon, and R. Zana, J. Colloid Interface Sci., 186 (1997) 170. [74] K. Fontell and M. Jansson, Progr.Colloid Polym. Sci., 76 (1988) 169. [75] G. Lindblom, L. Johansson, G. Wikander, P. Eriksson and Arvidson, G., Biophys. J., 63 (1992)723. [76] J. Israelachvili, Intermolecular and Surface Forces, Academic Press, London, 1991. [77] C. Tanford, J. Phys. Chem. 78 (1974) 2471. [78] C. Rodriguez, PhD dissertation, 2001. [79] K. Shigeta, C. Rodriguez and H. Kunieda, J. Disp. Sci. Tech., 21 (2000) 1023. [80] H. Kunieda, K. Aramaki, T. Nishimura and M. Ishitobi, Nihon Yukagaku Kaishi, 49 (2000)617. [81] N. Kanei and H. Kunieda, Nihon Yukagaku Kaishi, 49 (2000) 957. [82] H. Kunieda, M. Tanimoto, K. Shigeta and C. Rodriguez, J. Oleo Sci., 50 (2001) 633. [83] G. Horvath-Szabo, J.H. Masliyah, and J. Czarnecki, J. Colloid Interface Sci., 257 (2003) 299. [84] V. Castelletto, I.W. Hamley and Z. Yang, Coll. Polym. Sci., 279 (2001) 1029. [85] C. Rodriguez-Abreu, M. Garcia-Roman and H. Kunieda, Langmuir, submitted. [86] F. Mallamace, S-H Chen, Y. Liu, L. Lobry and N. Micali, Physica A, 266 (1999) 123. [87] D. Vega, J.M. Sebastian, Y.H. Loo and R.A. Register, J. Polym. Sci. Part B: Polym. Phys., 39(2001)2183. [88] M. Nemer, J. Blawzdziewics and M. Loewenberg, in H. Aref and J.W., Philips (eds.), Mechanics for a new millennium, Kluwer, Dordrecht, 2001. [89] S.M. Fielding, P. Sollich and M.E. Cates, J. Rheol., 44 (2000) 323. [90] H. Kunieda, Md. H. Uddin, M. Horii, H. Furukawa and A. Harashima, J. Phys. Chem B, 105(200)5419. [91] K. Watanabe, N. Kanei and H. Kunieda. J. Oleo Sci. 51 (2002) 771. [92] H.M. Princen, M.P. Aronson and J.C. Moser, J. Colloid Interf. Sci., 75 (1980) 246. [93] M.P. Aronson and H.M. Princen, Nature, 286 (1980) 370. [94] P. Poulin, F. Nallet, B. Cabane and J. Bibette, Phys. Rev. Lett., 77 (1996) 3248.
Highly Concentrated (gel) Emulsions: Formation and Properties
555
[95] E.J.W. Verwey and J.Th.G. Overbeek, Theory of the stability of lyophobic Colloids, Elsevier Publishing, Amsterdam, 1948. [96] B.V. Derjaguin and L. Landau, Acta Physicochim., 14 (1941) 633. [97] R. Aveyard, B.P. Binks, J. Esquena and P.D.I. Fletcher, Langmuir, 14 (1998) 4699. [98] R. Aveyard, B.P. Binks, J. Esquena and P.D.I. Fletcher, Langmuir, 18 (2002) 3487. [99] M.P. Aronson and H.M. Princen, Colloids Surf., 4 (1992) 173. [100] V.G. Babak, Colloids Surf. A, 85 (1194) 279. [101] M.P. Aronson and M.F. Petko, J. Colloid Interface Sci., 159 (1993) 134. [102] M.P. Aronson, Langmuir, 5, (1989) 494. [103] J. Bibette, D. Roux and F. Nallet, Phys. Rev. Letters, 65 (1990) 2470. [104] J. Bibette, D. Roux and F. Nallet, J. Phys. II France, 2 (1992) 401. [105]D.N. Petsev. In: B.P. Binks (ed.), Modern Aspects of Emulsion Science, Cambridge, U.K., 1998, pp 56-59.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 14
Recent progress in double emulsions N. Garti and R. Lutz Casali Institute of Applied Chemistry, The Hebrew University of Jerusalem 91904, Jerusalem, Israel 1. INTRODUCTION AND PAST ACHIEVEMENTS Double emulsions are complex dispersed liquid systems known also as 'emulsion of emulsion' or 'emulsions within emulsions', in which the droplets of one dispersed liquid are further dispersed in another liquid [1]. The inner dispersed droplets in the double emulsions are separated from the outer liquid phase by a layer of another phase [2]. More than one phase can be entrapped one into the other forming multiple emulsions. However, even if such dispersions can be formed they are very non-stable. Double emulsions can be of W/O/W (water-in-oil-in-water) or O/W/O (oil-in-water-in-oil). The most common and the most studied double emulsions are of W/O/W since they have higher potential to become commercial products in food, cosmetic and pharmaceuticals systems. W/O/W double emulsions are
Fig. 1. Schematic presentation of W/OW double emulsion droplets.
558
N. Garti and R. Lutz
also very often used as intermediate phases in the preparation of microspheres and other microencapsulated particles. A schematic representation of a W/O/W double emulsion droplets is show in Fig. 1. The inner droplets (W|/O) and the droplets of the outer emulsion, sometimes called also globules, (O/W2 emulsion) are stabilized by two separate layers of surfactants. The sizes of the inner emulsion droplets (W|/O) are of lum or smaller and require hydrophobic surfactants and the size of the globules of the outer emulsion are of 10-50um and require hydrophilic surfactants [3]. The formation of double emulsions is simple and requires two consecutive emulsifications processes. Yet, the formation of stable double emulsions with a sufficient control of the release of the entrapped compounds is not easy to achieve and the number of unsolved technical related problems is very significant. Seifiriz, in 1925 [4], was the first to describe, from simple microscopic observations, the existence of emulsion-within-emulsion, but the fast film rupture along with the uncontrolled coalescence of the external droplets remained up to these days. Many of the authors brought mostly descriptive observations of the double emulsions stabilized by various blends of emulsifiers (mostly nonionic) and different oils [5-11]. Many scientific attempts were made to incorporate markers, drugs, and additives within the inner phase, and to examine the loading capacities, release patterns and stability aspects. Technologists were involved in trying to bring products to the market place but it was realized that many obstacles are standing in our way to commercialize the technology. Numerous review articles were written on double emulsions covering almost every research aspects of this technology [1, 2 and 9-15]. Advanced methods of double emulsions preparation, new analytical tools [16-19], new dynamic tests, innovative microscopic tools, and few model systems clarified some of the obscure points and cleared some of the "fog" from many of the rupture and release mechanisms. The present manuscript will review recent studies and will try to discuss the different approaches that have been recently introduced. 2. POTENTIAL APPLICATIONS Double emulsions are liquid carriers for entrapped and release of active or reactive molecules in food, pharmaceuticals and cosmetic applications. Some of the potential applications of double emulsion in the food industry include preparation of: spreads (margarine replacements), low calories mayonnaise and special sauces [3, 20, and 21]. In the flavor industry encapsulations within double emulsion were used as 'entrapment reservoirs' for masking undesired flavors and odorants. Flavor entrapment allows prolong shelf-life [22].
Recent Progress in Double Emulsions
559
Improved dissolution rates of certain active matter or solubilization of oilinsoluble materials were also suggested. The active compound is dissolved, in part, at the inner phase, and in part, at the internal interface and occasionally at the external interface [2]. Double emulsions were also mentioned as 'liquid reservoir' for sensitive molecules, protecting them from external environmental reactivity such as oxidation, light and enzymes [22-24]. O/W/O double emulsions were used [25] for entrapment of orange oil in spray-dried double emulsions. The secondary encapsulation allowed better protection and better release control of active material from the inner phase. Double emulsions, are presently used in cosmetic formulations allowing more agreeable aqueous feel then the usual oily texture of a cream and were utilized for slow release of active materials or flavors [26, 27]. Calcium thioglycolate, one of the most widely used depilatory agents both in creams and in soaps, was formulated in W/O/W double emulsions using amphiphilic copolymer and obtained good controlled release in semisolid formulation [28]. Deposition of water soluble benefit agents onto skin from wash-off systems can potentially be achieved by incorporating them within oil in multiple emulsions [29]. Microbiologists considered double emulsions as attractive technology to control microbiocidal activity. One such example is for treatment of vaginal infection [30]. When vaginal infection occurs, the pH rises. In order to reestablish the low pH it is necessary to re-generate acidic environment. The double emulsions are thus an appropriate technique for the release of acidic agents. The major potential applications of double emulsions are within human Pharmaceuticals as slow and sustained release of active matter from an internal reservoir into the continuous phase [21]. Okochai and Nakano [31] summarized some of these applications and mentioned the use of double emulsions to enhance chemotherapeutic effect in cancer [32-34], lymphatic absorption of drug [35], eternal absorption of drug [36], drug immobilization [36, 37], treatment of drug overdoses [38], and substitution of red blood cells [39]. In order to protect insulin from enzymatic degradation double emulsions were prepared [40]. Salmon calcitonin (sCT) (polypeptide hormone consisting of 32 amino acid residues) which can be used successfully for the treatment of osteoporosis, in W/O/W in vitro and in vivo by using a rat model was widely studied [41]. In the simulated gastric medium, the amount of sCT diffused from the solution form was detected as 20% at the end of first hour of the study. In the presence of gastric peptide (pepsine) its activity is further inhibited by the inclusion of aprotinin. The W/O/W formulation containing entrapped aprotinin apparently produced the best profile, with almost 100% of the peptide present in the diffusion medium during the experiment.
560
N. Garti and R. Lutz
Double emulsions of W/O/W are potential drug carriers, but the penetration of drug through the skin is mostly based on O/W/O double emulsions. Laugel et al. investigated triterpene compounds formulated with polymeric surfactants as emulsified in O/W/O emulsion for drug delivery [42]. Additional application of O/W/O includes camphor encapsulation [43]. There are also some potential applications in agriculture and housekeeping products [44]. New industrial potential applications were lately investigated. These include the use in double emulsions of O/W/O as fuels [45]. Double emulsions are intermediate precursors in making solid or semi-solid nanospheres, or microcapsules [1, 46, and 47]. 3. IMPROVED PREPARATION METHODS The preparation of double emulsions was always considered to be simple and unsophisticated. However, it seems that the classical two-step emulsification does not form sufficiently small enough, and monodispersed globules. Recently investigators used some advanced techniques to improve double emulsion preparations. We will review some of the most interesting techniques. After years of studies, two types of preparations were considered as the major routes to the formation of the double emulsions: (1) one-step emulsification that is accomplished either by phase inversion or by mechanical treatment of the mixture of the components, and (2) two-step emulsification. Detailed description of the two methods can be found in review articles [2, 13, 20, 48 and 49]. In early reports [9, 48-51], double emulsions were prepared by a phase inversion process using only one set of emulsifiers. The process is difficult to control since the emulsifiers' distribution within the two interfaces is difficult to control, and the stability in most cases was very limited. Fast exchange of the emulsifiers between the two interfaces occurred. Most researchers have abandoned this method of making emulsions. In some specific cases, under special conditions, one can still make stable double emulsions by one step emulsification. Those special cases will be further discussed. Most double emulsions are prepared in two steps (Fig. 2). At first, a high shear homogenization is applied on the system containing water, oil phase and hydrophobic emulsifiers. High shear and prolonged homogenization is needed in order to obtain small submicronal water droplets and stable W/O emulsion. In the second step, the pre-prepared W/O emulsion is gently and slowly added to an aqueous phase containing the hydrophilic emulsifiers. Slow and stepwise addition under moderate stirring and without any high-shear homogenization should be applied. The droplets size distribution of a typical double emulsions made by this technique ranges from 10 to 50 urn.
Recent Progress in Double Emulsions
561
Fig. 2. Schematic illustration of the two-step emulsification of W/O/W double emulsions.
3.1. Droplets fragmentation The first concept of preparation double emulsions is by breakup the droplets into two or more segments. For a droplet to break into fragments, it must elongate sufficiently. This implies that the hydrodynamic shear must exceed the cohesion constraint due to capillary forces [52]. Rumscheidt and Masom [53] observed that droplets break-up into two or more segments happen when the capillary number exceeds a critical rupture value. The critical value depends on the type of flow and on the viscosity ratio of the two liquids. The critical value of the capillary number must be approximately 1 [12]. Double emulsions consist of two different interfaces that require two sets of different
562
N. Garti and R. Lutz
types of emulsifiers and the critical rupture values are difficult to be estimated. Within double emulsions two types of droplets fragmentation are possible: (1) fragmentation of the W,/O droplets into W2; and (2) fragmentation of the W, droplets into the oil phase. The critical value of the capillary number for fragmentation is more complex to analyze and achieve. For W/O/W double emulsions the inner emulsifiers are hydrophobic while the outers are hydrophilic. In many cases a blend of two or more emulsifiers in each set are recommended for better stabilization results but their internal exchange is complicating our estimations on the main factors that are essential for controlling the stability. 3.2. Selection of emulsifiers Low-molecular-weight hydrophobic emulsifiers migrate from the W/O interface to the oil phase and the hydrophilic surfactants can migrate to the inner interface and alter the required hydrophilic/hydrophobic balance of each of the phases. Most of the studies, in the years 1970 to 1985, searched for a proper monomeric emulsified' blend or combination of hydrophobic and hydrophilic surfactants which suitable proper ratios between the emulsifiers. Matsumoto et al. [54-59] established an empirical weight ratio of 10 of the internal hydrophobic to the external hydrophilic emulsifiers. Garti et al. [60-65] proved that the free exchange between the internal and the external emulsifiers' required a calculation of an 'effective HLB value' of emulsifiers to optimize the stabilization of the emulsion. Parameterization work was done on almost every possible variation of the ingredients and compositions [5-11]. In most cases the internal emulsifiers are used in great excess to the external emulsifiers. The nature of the emulsifiers also dictated the number of compartments and the internal volume that the inner phase occupies. Many of the more recent studies explore various more sophisticated emulsifiers such as sphingomyelins [66], modified or purified phospholipids [67], etc. The principals for selecting the proper emulsifiers are similar to those known for classical emulsions. Some of the emulsions might have better stability than others but the general trend remained unchanged. It should be also stressed that one must adjust the emulsifiers to the in-mind application and must substitute one emulsifier by the other depending on the total composition of the system. Vasudevan and Naser, [29] used a blend of hydrophilic/hydrophobic surfactants that reduced the interfacial tension between the oil-water interface to 10"2 mN/m. The effect was not seen for oils or surfactants with unsaturated lipophilic chains due to their incompatibility to form closed packed, condensed interfacial films. Close packing is essential to prevent solubilization of the adsorbed low HLB emulsifier by the high HLB surfactant. Very long-term stability of double emulsions was found when the double emulsions with a
Recent Progress in Double Emulsions
563
silicone-based polymeric emulsifiers and light mineral oil were prepared. The double emulsions were stable for ca. few months. It must be also recognized that 'empty' double emulsions will behave differently from those containing active matter (electrolytes, biologically active materials, proteins, sugars, drugs, etc.) due to osmotic pressure gradients (caused by the additives) between the outer and the inner phases. In addition, many of the active ingredients have some hydrophobicity and surface properties. Such molecules (peptides, drugs, pesticides) will migrate from the inner bulk and will adsorb onto the interface changing the delicate emulsifiers HLB. The emulsifiers around the water or the oil droplets will not fully cover the droplets, and the stability will be reduced. 3.3. Emulsified microemulsions It is well established that the best possibility to enhance the double emulsions stability is to reduce the droplets' sizes of the inner phase. Reduction of the inner droplets of the W/O emulsion to submicronal sizes may increase the emulsion stability, and as a consequence may reduce also the sizes of the droplets of the external phase. The ideal preparation is to form swollen micelles or microemulsions in the inner phase. The concept of 'emulsified microemulsion' was coined many years ago by Pilman et al. [68] and was also patented [69] but never fully explored or utilized. The concept is very logic, innovative and promising but difficult to implement and even more difficult to prove that microemulsion stays intact after the second emulsification. Larsson in is search for new 'liquid crystalline lyotropic mesophases' based on polar lipids such as mono fatty acid esters of glycerol (monoglycerides of fatty acids). He and other [70-74] characterized the structures of L2, and L3 mesophases, and concluded that the mesophases are 'microemulsion-like' with symmetry patterns of cubic or lamellar phases. The L3 systems are fluid isotropic phases with nanosized domains. Upon further water dilution and with the addition of polymeric emulsifiers, such as block copolymer of polyethylene and polypropylene (Pluronics), and applying high shear, new dispersed 'emulsionlike' structures were formed. Larsson described those structures as cubosomes derived from 'microemulsion-like mesophases', that according to him, are "actually double emulsions of W/O/W, where the inner phase is the L2 or L3 microemulsion-like system" (Fig. 3). The analogy of the emulsified mesophases (cubosomes) to double emulsions was done because the dispersed structures were composed of compartments of microemulsion within an emulsion. Larsson did not further elaborate on the concept and it remained in the literature as a descriptive analogy. The very fast exchange of the monomeric surfactants, regularly in use for making micellar structures and emulsions, results in most cases, in the formation of empty O/W emulsions with no internal core but the formation of semisolid mesophases is retarding such exchange and gives us new
564
N. Garti and R. Lutz
option to form emulsified microdiomains. The concept was challenged, recently, by several authors showing some promising results. We believe that the concept for making "emulsified emulsions" should be further challenged in future studies. Larsson concept was expressed in his book back in 1994 [75] mainly in order to better explain the formation of the cubosomes but it was revolutionary for those dealing with double emulsions. His work attracted the attention of some scientists working in the area of double emulsions. Recently, and possibly, independently of Larsson ideas, work was done on that concept. In our most recent work, in a search for new non-viscous mesophases we found, in the three component phase diagram, 'ill-defined cubic phases' made out of alcohol, water and monoglycerides of fatty acids. At certain surfactants compositions a very small isotropic region was detected and characterized as being 'ill-defined structure' that have some characteristics of cubic phases, some of lamellar phase and some of liquid-like microemulsions. The domains are very small (few nanometers), non-viscous and transparent. We termed them QL mesophases. Dispersion of the QL systems in water, in the present of Pluronic (or other amphiphilic compounds), led to the formation of "cubosomes-like" structures which can be termed also 'emulsified microemulsion-like mesophases' or 'double emulsions with nanosized inner droplets' (results to be published soon) (see Fig. 4).
Fig. 3. Dispersion of a GMO/P/W cubic phase as seen in cryo-TEM, reproduced by permission. The cubic phase is a Cp phase with edges of the particles along the cubic axes, and the inner periodicity has a repetition distance of about 10 nm. The largest particle of the hexagonal phase is seen along the c-axis, and the bar in this micrograph is 100 nm (Reproduced with permission from Ref. 70).
Recent Progress in Double Emulsions
565
Fig. 4. Typical structure of cubosomes, preparation from Q\ microemulsion like cubic mesophases dispersed in water and stabilized by Pluronic. Cubosomes are 'emulsified
mesophases' (Reproduced with permission from Ref. 76). Other scientists adapted the general concept of emulsified mesophases but explored somewhat different directions. Three major approaches are very interesting and should be discussed in more details. The first approach is based on the so-called 'lamellar phase dispersion process' reported by Vigie [77]. The procedure (Fig. 5) is based on preparation of 'liposome-like vesicles' with nonionic emulsifiers. The process can be used when the constituents form a lamellar phase, by mixing certain surfactant(s) with water in certain proportions. This procedure offers some advantage since it requires only a simple emulsification step. The mesophase formed by specific ratio of lipophilic emulsifiers in water is thermodynamically stable and can be obtained rapidly.
566
N. Garti and R. Lutz
The method main limitations are derived from the fact that double emulsifiers form lamellar phases (phospholipids or monoglycerides) at very limited levels of water. In addition, the quantity of oil that is incorporated into the lamellar phase is always low, rarely higher than 20 wt%. Another drawback of this process is the weak control on the rate of encapsulation of the active substances. The criteria for making such double emulsions are similar to those established by Larsson. Vesicles are, therefore, an interesting option for the inner phase in double emulsions. Readers of this chapter are urged to explore new ways of emulsifying mesophases into multi-compartment subunits that will be capable of entrapping hydrophilic matter and release it in a more controlled way. The second interesting emulsification process was described by Grossiord and Seiller [78]. They introduced a method termed 'the oily isotropic dispersion processes'. The author idea was to disperse an oil phase within water by excess of suitable surfactant and to form L2 phase (or W/O microemulsion). This isotropic phase is further emulsified with water to form double emulsions. However, there is no evidence in the paper that the L2 phase remains intact after the second emulsification and that the process, indeed, leads to the formation of
Fig. 5. Preparation of W/O/W double emulsions by lamellar-phase dispersion (Reproduced with permission from Ref. 77).
Recent Progress in Double Emulsions
567
Fig. 6. Preparation of W/O/W double emulsions by "oily isotropic dispersion" (emulsified microemulsion) (Reproduced with permission from Ref. 78).
multiple-compartment emulsions with submicronal or nanosized droplets. This process is worth further investigation and should be more carefully evaluated. If one can prove that the internal compartmentalization remains in the form of microemulsion droplets, it might be a real breakthrough, since the external droplets could be reduced to submicronal sizes. Such formulations will allow formation of injectable double emulsions for controlled drug release. Fig. 6 illustrate the preparation of double emulsion by this method. The third approach [79] is based on using the "middle" phase microemulsion, (M-phase) along with an oil phase (O-phase) as the ingredients for their further emulsification. The authors called the so-formed double emulsion M/O/M. When a two phase emulsions (2-butoxyethanol C4E|/ndecane/water), one of middle phase microemulsion (M-phase) and other top phase of oleic (O-phase) phase are agitated at constant pressure, and temperature, double emulsions of M/O/M is formed (Fig. 7). The existence of
568
N. Garti and R. Lutz
Fig. 7. Optical micrographs of stagnant, two-phase multiple emulsions formed C4E,/«-decane for §j 0.133 =(a) and for <)| j 0.298 = (b), (magnification x600) (Reproduced with permission fromRef. 79).
such microemulsion in the core of the inner phase was confirmed by conductivity measurements. Inversion of O/M emulsion to the M/O emulsion occurred when the volume fraction of the microemulsion was decreased and was reduced to zero. In summary-we believe that the concepts of "emulsified microemulsions" or "emulsified mesophases" are promising preparations of W/O/W systems. 3.4. Monodispersed double emulsions Authors searched for years for new engineering methods to make monodispersed emulsions droplets. Once such methods were well established they were also used for making monodispersed double emulsions. Two major methods will be reviewed: 'transmembrane emulsification' and ' controlled twostep preparation'. The engineering modifications to the one step and the twostep emulsification were reported by various researchers. 'Membrane emulsification technique' [80] was described as a new method of producing W/O/W multiple emulsions. The method permits the formation of 'monodispersed liquid microdroplets containing aqueous microdroplets'. The idea was to permeate the pure dispersed phase through a porous membrane into moving continuous phase or the passage of previously prepared pre-emulsion through the membrane. The aqueous internal phase is based on oil phase containing lipophilic emulsifier that is mechanically mixed. The mixture is then sonicated to form W/O emulsion that is filled into the upper chamber of a special apparatus. The external aqueous phase, containing the hydrophilic emulsifier, is continuously injected into the lower chamber to create a continuous flow. The emulsion is progressively removed from the apparatus. More details on the experimental procedure can be found in Joscelyne and Tragardh [81] that reviewed the preparation of monodispersed W/O/W by membrane.
Recent Progress in Double Emulsions
569
Similar concept with different membranes was used by Vladisavljevic and Schubert [82]. The technology is based on previously reported technique developed by Nakashima et al. [83]. W/O/W double emulsions were prepared over a wide range of membrane'wall shear stresses' (0.37-40 Pa) with dispersed phases of up to 20 vol% and transmembrane pressure of 1.2-5.7 times larger than the capillary pressure. Small droplets are directly formed at the pore opening and detached from the membrane surface by the shear stress in continuous phase. The authors prepared W/O emulsions by using classical microfluidizer technique with small droplets of less than 0.2 \x.m and then dispersed them into water continuous phase passing through the SPG membrane (5 \\m) (Fig. 8). O/W emulsions were prepared using rapeseed oil with 2 wt% of Tween 80 dissolved in the aqueous phase. The oil phase in the W/O/W double emulsions contained 10% polyglycerol polyricinoleate (PGPR). The membrane had pore sizes of 1.4-6.6 um. Shear stresses were of 0.33-40 Pa. Fig. 9 is showing photomicrographs of emulsion droplets prepared utilizing SPG membranes of different pore sizes. The influence of the transmenbrane pressure on droplet size distribution was carefully studied. In Fig. 10 we can see the droplets sizes distribution as a function of transmembrane pressure in a 4.8 um pore size membrane.
Fig. 8. Schematic diagram of the preparation of multiple W/O/W emulsion using SPG membrane (Reproduced with permission from Ref. 82).
570
N. Garti and R. Lutz
Fig. 9. Photomicrographs of the emulsion droplets prepared by utilizing SPG membranes of different mean pore sizes: (a) J p =6.6 |j.m, ^3 2=24 |im, 1-day-old emulsion; (b) ^=4.8 |j.m, c?3,2= 17.5 (j.m, fresh emulsion; (c) dv=\A\x.m,
Fig. 10. Influence of transmembrane pressure (1.2, 2.9, 3.6, 5.7 Pcap) on the droplet size distribution for the 4.8 nm-SPG membrane (Reproduced with permission from Ref. 82).
3.5. New fragmentation device for monodispersed double emulsions A team of French scientists [84-86] have considered W/O/W double emulsions as having a thin films of metastable bilayer of surfactants with limited thermodynamic stability. The film life-time governs the existence of many dispersed structures. The thin film growth determines the life-time of the film and the stability of the double emulsions: co = co0 exp(-Ea/kT),
Recent Progress in Double Emulsions
571
where, co0 is the natural frequency, and Ea is the activation energy, (the energy cost for reaching the critical hole radius) and co is the so called by the authors the nucleation frequency (co) of the hole that is formed when reaching a critical size above which it becomes unstable before rupturing. In other words it is the unique rupturing frequency per unit film area. The researchers developed unique unambiguous method for the measurement of the microscopic parameters coo and Ea based on the use of monodispersed W/O/W double emulsion. Based on those values they could evaluate the nucleation frequency and as a result, monodispersed emulsions were made in a two-step process using couette-type fragmentation device [85-86]. Span 80, at fixed concentration of 2% with respect to the continuous phase, was used for the inner interface, in the presence of 0.4 mol/liter of NaCl to avoid coursing of the emulsion. W/O droplets of 0.36 um with polydispersity of 15%, were formed. The inverted emulsion was added to water containing a very low concentration (10°-10"4 mol/1) of SDS using high-pressure jet homogenizer. The double emulsion globules that were obtained have polydispersity of >25%. The ready made double emulsions were further stabilized by additional amount of SDS and glucose (to adjust the osmotic pressure). The thin film, formed between the internal droplets and the globules surface is composed of mixed monolayers covered by Span 80 and SDS molecules and separated by oil. SDS migrated to the inner phase within very short period of time (1 min) which caused the film to be regarded as in thermodynamic equilibrium. 3.6. Bidimensional formulation-composition map The use of polymeric amphiphilic and surfactants mixture allows to strongly inhibiting the mass transfer and to considerably lengthen the equilibrium between interfaces. The concept of using polymeric amphiphiles
Fig. 11. Formulation composition bidimensional map (Reproduced with permission from Ref. 87).
572
N. Garti and R. Lutz
will be further discussed in the chapter related to stability of the double emulsions. We will bring here an interesting preparation approach that is not engineering based but rather conceptual. Allouch et al. [87] described the "region corresponding to different emulsion morphology occurrences in a form of a map of formulation vs composition (Fig. 11). An increase in the oil viscosity was found to alter the map and to modify the multiple emulsion characteristics. The central horizontal line corresponds to HLD = 0, i.e., (HLDhydrophilic-lipophilic deviation, expressing the surfactant affinity difference (SAD)), i.e., the variation of the chemical potential of a surfactant molecule when it passes from the aqueous and oil phases, which can be calculated from the measurement of the partition coefficient of the surfactant between these phases optimum formulation for three-phase behavior, where the surfactant exhibits an equal affinity for the oil and water phases. Above it at HLD > 0 and below it at HLD < 0, the affinity of the surfactant for the oil (respectively for water) dominates. The bold line separates the regions where oil external (W/O) and water external (O/W) emulsions dominate, when an equilibrated surfactantoil-water system, having the corresponding formulation and composition are formed. The central region is labeled A + (respectively -) indicating that the region is located above (respectively below) the optimum or neutral tendency formulation line (HLD = 0). All of these A" C" and A+ B+ regions are called normal regions because the emulsion type corresponds to the curvature that is
Fig. 12. Two-step emulsification procedure to make a multiple emulsion in C+/CT. Sample with small amount of oil phase containing the Span 20 (lipophilic surfactant) and of an aqueous phase containing 1 wt % NaCl. The sample is left to equilibrate for 24 h, and the emulsification is carried out with the Ultra-Turrax mixer. A first emulsification is carried out at 15 000 rpm for 1 min to form W/O/W emulsions. Immediately after the first emulsification is carried out, a certain amount (see below) of hydrophilic polymer is added as an aqueous aliquot, and the stirring is resumed for an additional 1 min at 8500 rpm. The outer O/W emulsion is reformed with an external aqueous phase containing a hydrophilic amphiphile to ensure the proper formulation for its representative point to be located in the C region, where O/W emulsions are known to be stable. (Reproduced with permission from Ref. 87).
Recent Progress in Double Emulsions
573
Fig. 13. Schematic representation of phase inversion of W/O/W emulsions by extrusion through polycarbonate membrane. W/O/W emulsions are extruded through the polycarbonate membrane. The concentration balance of surfactant between the hydrated layer at the surface of multiple droplets and the bulk outer aqueous phase might be destroyed due to velocity gradient. In the locally surfactant-rich region, surfactant molecules would involve water molecules and are carried into the oil membrane. Consequently, the inner aqueous phase might increase. In the locally surfactant-poor region, the aqueous film at the outer surface of multiple droplets would be ruptured and coalescence of multiple droplets would occur. Migration of water from the outer to the inner aqueous phase and coalescence of droplets would occur simultaneously and cause intermixing of oil, water, and hydrophilic and hydrophobic surfactant molecules in the pores. Consequently, the emulsion sheared by the velocity gradient in pores might form D-phase-like structures, where the surfactant molecules are oriented and imply water molecules between hydrophilic surfaces and oil molecules between hydrophobic surfaces. In pores of the membrane, surfactant molecules might be oriented with their hydrophobic groups toward the pore wall surface and with hydrophilic groups toward the solubilized water molecules. Therefore, a lamellar structure would be formed in the pores and it would be ruptured at the outlet of pores by abrupt changes in velocity due to an increase in area and a reversed micelle would be formed. The resultant reversed micelle would coalesce and a w/o emulsion would be formed, (Reproduced with permission from Ref. 88).
favored by the physicochemical formulation effect. On the contrary, the C+ and B" regions are so-called abnormal, it is the phase existing in a higher volume in the system. In these abnormal regions, double emulsions are often produced by simple stirring. They are of the O/W/O type in B" and the W/O/W type in C+, where the lower case letter indicates the most internal phase, i.e., the droplets inside the drops. It can be said that the spontaneous occurrence of a multiple
574
N. Garti and R. Lutz
emulsion is a way for the system to satisfy conflicting inclinations, as far as interface curvature is concerned. The most interior or "inner" emulsion is of the type imposed by the formulation, e.g., W/O in the C+ region, whereas the external or "outer" one is the type demanded by composition, e.g., O/W in the C+ region. Fig. 12 shows the schematic process used to make the double emulsions. 3.7. Use of viscosity agents Hino et al. [88, 89] applied three different protocols for achieving best stability, good entrapment and release results: (1) keeping the inner aqueous phase at hypertonic conditions to prolong the release of the drug, (2) using chitosan as viscosity agent to improve stability and, (3) applying phase inversion with porous membrane. The emulsions were extruded through a polycarbonate membrane with 3 um diameter pores (Fig. 13). The double emulsions could be spray-dried to form powdered preparation. The dry powder emulsion could be easily redispersed to form reconstituted double emulsions. The process was used in TAE (Transcatheter Arterial Embolization) therapy. 4. STABILITY CONSIDERATIONS Double emulsions, even more than regular emulsions, because of their droplets sizes, have inherent thermodynamic instability [17]. In addition, the release of active material from the inner phase to the outer phase is mostly uncontrolled. Many studies dealt with those problems [8, 10, 13, 17, 48, 49, 90 and 91]. Every scientist is aware of the instability characteristics of any double emulsions. Many of the studies do not cover the full instability process since many of the pathways involved in the breakdown of the double emulsion are difficult to detect. Some of the instability pathways are occurring in parallel and some are rate-determining steps (Fig. 14). In many of the cases the pathways are depending on the internal composition, nature of ingredients, methods of formation, etc. There is no one single study that covers all the aspects related to stability. There are always problems related to the analytical methods. Despite all the difficulties, researchers have improved the stability of double emulsions, and in many of the cases, the stability is sufficient for specific application. The today bottle neck is the release more than the stability problems. In thermodynamic terms the stabilization is based on three major mechanisms: (1) electrostatic stabilization mechanism, (2) steric stabilization mechanism, (3) mechanical stabilization mechanism [92]. Monomeric emulsifiers can impact electrostatic stabilization mechanisms, as well as, steric stabilization. The polymeric amphiphiles are multi-anchoring surfactants with strong steric stabilization, while solid microparticles stabilize emulsion mostly via mechanical stabilization mechanism.
Recent Progress in Double Emulsions
575
Fig. 14. Schematic representation of the possible pathways for breakdown in multiple emulsions.
4.1. Electrostatic stabilization mechanism The electrostatic stabilization mechanism is well documented for simple O/W emulsions. Yet, in terms of electrostatic force, there is no significant difference between single dispersed droplets and the double emulsions droplets except for the droplets sizes. Double emulsions droplets are much larger in size and therefore, the repulsive electrostatic forces are less pronounced. The reader can find more on electrostatic repulsion in any text book dealing with emulsions stability. We, therefore, will not elaborate on the repulsive forces. 4.2. Steric stabilization mechanism Steric stabilization was also widely studied in O/W simple emulsions. The significance of steric stabilization becomes dominate factor when dealing with polymeric amphiphiles used as macromolecules surfactants that cover mainly the external interface. Polymeric amphiphiles can improve the stability of double emulsions but can also retard the transport through interfaces since they form thick films and do not aggregate in reverse micelles in the oil phase.
576
N. Garti and R. Lutz
Fig. 15. Two limiting models for steric stabilization with amphiphilic macromolecular surfactants: (a) interpenetration of adsorbed layers without compression; (b) compression without interpenetration. The
literature
distinguishes
between
three
main
mechanisms
of
stabilization with polymeric amphiphiles: (1) depletion stabilization by nonadsorbing macromolecules that prevent collision between particles (droplets) and provide elasticity to the system; (2) electrostatic repulsions between two droplets carrying the same charge: and (3) steric stabilization resulting from hydrophobic interactions between adsorbed polymers [92]. The main contribution to the stability of maeromolecular-stabilized emulsions is related to droplets approaching a distance where compression due to polymer-polymer interactions occurs (without inter- penetration) (Fig. 15). The total free energy, Gh of systems stabilized with macromolecules includes three main terms, 'volume restriction', termed Gvr, 'mixing' term, Gmix and free energy of attraction, Ga. If an interaction exists between the polymeric chains adsorbed onto two neighbor particles (droplets), Gvrf describes the reduction in the entropy of polymeric chains (configurationally entropy) due to restriction of the total volume available to each chain. Gmix describes the buildup in polymeric amphiphile segment concentration in the interaction zone
Recent Progress in Double Emulsions
577
Fig. 16. Variation of the different free energies involved in the steric stabilization mechanism.
between the particles (droplets) that leads to an increase in the local osmotic pressure and in the steric free energy AGS =AGmix+ AGvr. When decreasing the bulk solvency toward the dispersed phase, a significant minimal value of the total free energy of interaction {GJ could be reached. Decrease solvency leads to a contraction in the thickness of the adsorbed layer, an increase in segment concentration and consequently a minimum in the value of Gt. The segment concentration in the adsorbed layer is significantly increased and results in enhanced free energy of attraction {GQ>. In the case of low-molecular-weightpolymeric-anchored-chains the effect will be more gradual than for high molecular weight polymers due to lower segment concentration in the interaction zone between the droplets. The different variations of energy are presented in Fig. 16. Total contribution to the free energy of the system AG, in the total free energy of the system compared of AGa (attraction) and AGr (repulsion) or AGS (steric), AGmix in the steric stabilization attributed to the mixing term and AGvr is attributed to the volume restriction when the polymer is of high molecular weight and has good solvency will impact high stabilize to the system. 4.2.1. Polymeric amphiphiles Steric stabilization was widely studied in O/W simple emulsions. The significance of steric stabilization becomes dominate factor when dealing with polymeric amphiphiles used as surfactants that cover and stabilize mainly the globules by adsorbing onto the external interface. However it must be stressed that certain hydrophobic polymeric amphiphiles have also been mentioned as
578
N. Garti and R. Lutz
steric stabilizers for the internal phase of W/O [62-63, 90, 93-97]. Polymeric amphiphiles improve the stability of double emulsions and retard the transport through the external interfaces by forming semi-solid thick films. The polymer is anchored by several segments at the interface, and its free energy of adsorption is much higher than that of a monomeric surfactant. For industrial purposes, this approach seems to be the most promising, since it answers most of the stability problems and solves in part the release problems and requires lower amounts of the surfactants. Polymeric amphiphiles also block the reverse micellar transport because they do not form reverse micelles in the oil phase. Improved stability of double emulsions was demonstrated for synthetic amphiphilic copolymers, [98], and natural occurring macromolecules [99-104].
Fig. 17. Structure of the surfactants that allow preparing a stable W1/O/W2 emulsions (Reproduced with permission from Ref. 48).
Recent Progress in Double Emulsions
579
Fig. 18. Photomicrograph of double emulsions of W/O/W stabilized with Abil EM90 as emulsifier I and PHMS-PDMS-52% UPEG-45 EO as emulsifier II, (X 200) (Reproduced with permission from Ref. 93). Garti et al. were the first to demonstrate that hydrophobic polymeric
surfactant such as polysiloxane, cetyldimethicone coplyol (Abil EM 90) and poly-glycerol-poly-ricinolate (PGPR) [19, 90, 93-97 and 101] improve double emulsions stability in several independent cases using different oils and various hydrophilic surfactants (Fig. 17). The hydrophobic polymers stabilized the inner droplets and formed submicronal droplets while the hydrophilic surfactants stabilized the external interfaces. Fig. 18 shows the droplets sizes that were obtained with the combination of the hydrophilic and hydrophobic polymeric surfactants. It should be stresses that the use of polymeric hydrophobic emulsifiers (Abil 90 EM or PGPR) and polymeric hydrophilic emulsifiers (hydrophilic siliconic emulsifiers or WPI/xanthan) are showing extremely good results. Excellent double emulsion stabilities with globules smaller than any other prepared double emulsions by any other technique were obtained both with the synthetic and with the natural occurring emulsifiers. De Luca et al. [98] examined double emulsions based on mixture of a polyester non ionic surfactant (Hypermer A60®) and an ethoxylated polypropylene oxide (Synperonic PF/127®). The double emulsions were viscosified to minimize mobility between to phases. They claimed that even after ten years of aging the double emulsions remained stable. In another work [105, 106], W/O/W double emulsions were loaded, fore example, with high-molecular-weight-graft-copolymers of poly(acrylic acid) and Pluronics (block copolymer of polyethylene oxide and polypropylene
580
N. Garti and R. Lutz
oxides) as stabilizing agents. The double emulsions had high viscosity, and were claimed to be thermo-reversible systems, and extremely stable for at least 3 months at room temperature. The double emulsions were loaded with high fractions of drug. The drug was released under moderate shear, around 1000 s"' (30% at 20°C and 100% at 35°C). Hydrophobically modified poly (sodium aery late) (10 mol% of carbon chain length of 8-16 on the backbone, MW of 50000 g/mol) was used as emulsifier at the outer interface [107]. The inner phase emulsifier was classical hydrophobic nonionic Span 80 Stable. Highly concentrated double emulsions were prepared that exhibited long- shelf life. The authors used diffusing wave spectroscopy to study the interactions between the droplets and the globule surface which are important for understanding the steric stabilization. The emulsions were very stable and the release of a marker (NaCl) initially encapsulated in the aqueous droplets showed very slow release. Synthetic amphiphilic polymers are not food permitted and therefore, they must be replaced by biopolymers. Natural-occurring macromolecular surfactants (proteins, polysaccharides) are food-grade surfactants. Some of them are known to interact with the oil phase or with other emulsifiers and can form thick interfacial layer [19]. The emulsions stabilized by natural biopolymers are nonviscous liquids. In our most recent studies we have demonstrated that proteins
Fig. 19. Phase diagrams of aqueous phase of WPI and aqueous phases of: (1) LBG, and (2) xanthan gum.
581
Recent Progress in Double Emulsions
(BSA, WPI, caseins) are capable of forming soluble complexes (hybrids) with certain hydrocolloids (xanthan, guar, LBG). Phase diagrams of WPI and xanthan gum or locust bean gum (LBG) have been constructed in order to reveal the compositions that are forming the complexes and those that are causing phase separation and conservation or precipitation. Fig. 19 is showing the differences in the complexation of xanthan vs LBG with WPI. The complex is amphiphilic macromolecule that strongly adsorbs onto the oil/ware outer interface and forms smaller globules with improved stability over protein stabilized double emulsions more than the protein alone (Table 1). The emulsions behave somewhat like of microcapsules, microspheres or mesophasic lyotropic liquid crystals, since the interfacial film is rather thick and in most cases, multilayered. Fig. 20 demonstrates droplets sizes formed with BSA as external emulsifier and Fig. 21 shows the double emulsion droplets of preparation based on xanthan gum/WPI as external emulsifier. Some of the major biopolymers that have been studied include: gelatin [100, 103], bovine serum albumin (BSA) [101], protein [102, 104], WPI and caseins [19], etc. Table 1 Globule size distribution of double emulsions made with WPI, xanthan gum, LBG and the blends (hybrids) of the gums with WPI (at their optimum ratios).dn: average droplet diameter of 90% of the droplets; d4!: average droplet diameter given by
rf(4,3)=^v,4/E^3 Xanthan LBG WPI/xanthan gum 4/0.5 0.5wt% 0.5wt% 23.4 12.0 28.4 3.3
4.8
10.8
4.0
WPI 4wt%
WPI/LBG 4/0.5
(urn) (jim)
19.8
23.6
3.1
582
N. Garti and R. Lutz
Fig. 20. Photomicrographs of W/O/W double emulsions prepared with 10 wt% Span 80/Tween 80 (9:1) in the inner phase and 0.1 wt% BSA in the outer phase (Reproduced with permission from Ref. 101).
Fig. 21. Optical microscopy images of double emulsions containing 30 wt% Wi/O (Miglyol 812N) contain 1% of vitamin Bi stabilized by 8% PGPR and 2% glycerol monooleate (GMO), the external droplets stabilized by 5% of WPI/xanthan gum 4/0.5.the PSD of (W/O) is 130 nm (left) and PSD of (W/O/W) is 10 mm (right).
Recent Progress in Double Emulsions
583
4.2.2. Depletion stabilization, viscosity and gelation effects Three options to reduce the mobility within W/O/W emulsions have been suggested [108]: (1) use high viscous oil in order to prevent diffusion of water/surfactant/active material from the inner phase through the oil phase, (2) gelation of the oil or the aqueous phase, and (3) polymerization of the interfacial adsorbed surfactant molecules. The two first options are the essence of the socalled "depletion stabilization" while interfacial polymerization is modified the steric stabilization. All three options were tested by various groups and found to be efficient way to improve stability and to reduce uncontrolled release. The depletion mechanism of stabilization is based on placing viscosity or gelling agents in one of the phase to reduce the mobility of the entrapped ingredients (Fig. 22). The viscosifiers do not absorb into the interface and are therefore, not emulsifiers. The emulsions are 'semi-solid' or 'gel-like' formulations (creams) that are highly recommended for cosmetic application but are non applicable if non-viscous, diluted emulsions give required applications such as, beverages. Variety of additives [109] such as hydroxypropyl cellulose (HPC), hydroxypropylmethyl cellulose (HPMC), hydroxyethyl cellulose (HEC), xanthan, guar gum, carrageenan, carbopol (lightly crosslinked PAA), synthetic
Fig. 22. Schematic presentation of depletion stabilization by viscosity or gelling agents.
584
N. Garti and R. Lutz
acrylic polymers, were tested. The viscosity or jelling agents affect many of double emulsions properties such as: entrapping yield, stability, droplets sizes, consistency, skin feel, etc. Some of the more recent suggesting includes adding new viscosity agent along with improving the formula preparation methods. In another study [110], the double emulsions were stabilized viscosifying or gelling such as, hydroxypropyl cellulose in the outer aqueous phase and showed that their double can break emulsions under shear but the release process is moderated by those agents. The emulsifiers ware fit Taylor's theoretical framework indicating that the bursting mechanisms of the globules under shear are the same for whatever the composition of the multiple emulsions is. As a result stability and release were controlled by shear. The thick interfacial layer and the viscous or gelled aqueous phase might have some advantages when protection of sensitive addenda was required. Kim, et al. [ I l l ] treated the internal aqueous phase with xanthan gum and introduced Kojic acid (5-hydroxy-2-(hydroxymethyl)-4-pyrone) as a model antioxidant. They found improved stability to coalescence and improved chemical stability in the presence of the polymeric hydrocolloids. In anther study, the authors [48] used in the inner water phase contains Abil EM90, while the outer phase is based on the use of betaine and SDS on hydrophilic surfactants and xanthan gum as stabilizer (viscosity agent). From microscopic observations along with surface isotherms and linked to rheology measurements the authors concluded that for a double emulsion to be stable it is essential to use low-HLB surfactant that can provide reversibly expandable and compressible and irreversibly adsorbed. The long-term stability required, in addition, a balance between Laplace and osmotic pressures between W| droplets ether sulfate) as ideal blend of emulsifiers for the external interface (Table 2) and the oil phase. Small quantities of salt in the inner phase were able to balance the excess pressure. As for the outer phase, it was concluded that the presence of thickener in the W2 phase is necessary in order to read the viscosity ratio (permeability «1) between Wi/O and W2 that will allow dispersion of the viscous primary emulsion into the W2 aqueous phase. The interactions between the low and high HLB emulsifiers at the O/W2 interface should not destabilize the films. They have, therefore, suggested using high HLB surfactant (betaine derivative) in combination with an anionic surfactant (SLES- sodium lauryl ether sulfate) as ideal blend of emulsifiers for the external interface (Table 2). 4.3. Mechanical stabilization mechanism (paragraph extracted from Ref. 112) Some attempts have been made at improving the shelf-life of multiple emulsions by incorporating small solid particles into the surfactant formulations [113115]. The idea, as before, is that particles act as a mechanical barrier to
Recent Progress in Double Emulsions
585
Table 2 Double emulsion prepared with blends of emulsifiers and viscosity agents (Reproduced with permission from Ref. 48).
Wi
H2O NaCl
18.78 0.72
Composition% (w/w) 18.78 0.72
Oil
Light mineral oil Abil EM90
10.02 0.48
10.02 0.48
w2
H2O Betaine SDS Xanthan gum
54.8 8.6 6.2 0.4
54.9 10.9 3.9 0.4
coalescence if adsorbed at interfaces. Oza and Frank [113] were the first to develop the concept by using colloidal microcrystalline cellulose (MCC) as the water-soluble surfactant in W/O/W emulsions containing oil-soluble hydrophobie surfactants (Spans). Emulsions, stable for up to 1 month, contained a network of MCC particles adsorbed at the outer oil-water interface. Similar improvement in stability but for O/W/O emulsions was reported by Sekine et al. [114] in which the external W/O interface was partially coated with a layer of hydrophobically modified clay particles causing it to become rigid. The organoclay particles also caused gelling of the outer oil phase, preventing sedimentation of the water globules. Unlike the previous studies, in which both surfactants and particles have been employed in combination, Binks et al. [116 and 117] have successfully prepared multiple emulsions using particles alone as emulsifier. It is predicted that multiple emulsions should form in oil phase and water phase types of particles differing only in their hydrophobicity. Extremely stable multiple emulsions were made using a mixture of silica particles differing by only 25% in their SiOH content. This novel class of multiple emulsions should have great benefit in the pharmaceutical field where controlled release of substances is required in certain applications. Double emulsions droplets, W/O/W, based on hydrophobie silica particles are showing in Fig. 23 (upper). The oil globules are approximately 40 \xm in diameter and the water inner drops are of 2 (j.m. Increasing the concentration of the inner hydrophobie silica leads to an increase in the average globule diameter along with an increase in the number of inner water drops per globule. By contrast, an increase in the outer hydrophilic silica concentration causes the globule diameter to decrease and the viscosity of the outer water phase to increase. Fig. 23 (lower), shows typical O/W/O double emulsions formed by the same technique.
586
N. Garti and R. Lutz
Fig. 23. Optical microscopy images of multiple emulsions stabilized entirely by two types of silica particles of different hydrophobicity: (upper) W/O/W with triglyceride oil and (bar=50 |j.m) (lower) O/W/O with toluene (bar=20 (j.m) (Reproduced with permission from Refs. 116 and 117).
Fig. 24. Double emulsions stabilized with 1% solid nanoparticles of tristearin and GMO in the inner interface of the W/O emulsions. The outer interface contained WPI.
Garti et al. [115] demonstrated the enhancement in stability to coalescence upon addition of submicron crystalline fat particles as co-stabilizers
Recent Progress in Double Emulsions
587
of the inner interface of water-in-soybean oil-in-water emulsions (Fig. 24). They have, therefore, suggested using high HLB surfactant (betaine derivative) in combination with an anionic surfactant (SLES- sodium lauryl ether sulfate) as ideal blend of emulsifiers for the external interface (Table 2). 5. DESTABILIZATION PATHWAYS The instability and release of double emulsion droplets can be derived from several possible pathways: (1) coalescence of the double emulsion globules. This process should be the fastest because the droplets are very large, (2) coalescence of the inner water droplets. It should be slower process because the droplets are much smaller, (3) release of the entire inner (W/O) droplets when approaching the outer interfacial lamella. This mechanism is known as 'lamellar transport from the inner phase to the outer aqueous phase', (4) transport of water, due to osmotic pressure gradient, into the inner phase causing swelling of the double emulsion droplets or removal of water from the inner to the outer phase causing shrinkage of the double emulsion droplets, (5) release of addenda in a molecular form via a mechanism known as 'reverse micellar transport'. The present review will not discuss in details the different mechanisms but will bring some new interesting contributions that elucidated, some new points in the stability mechanisms using new approaches or new methods. Some of the papers deal with specific case studies, such as preparation of single double droplet [16, 118, 119] etc. but, nevertheless, they contribute to the understanding of the stability phenomena. 5.1. Coalescence of the inner phase droplets vs the coalescence of the outer phase globules Coalescence of the inner O/W interface vs the coalescence of the outer O/W emulsion was examined by Villa, Papadopoulos et al. [16, 118, 119] using capillary microscope. The work was done on a single W]/O/W2 double emulsion globule using a classical monomeric hydrophobic (Span 80) and hydrophilic (Tween 80 or SDS) surfactants. Two extreme and one between situations were discussed: (1) coalescence starting at the outer interface (dominated by external coalescence) that takes place when Wi and W2 phases were not too similar in their overall surfactant concentration (Fig. 25), and (2) Coalescence starting in the inner phase (inner phase dominate coalescence) (Fig. 26). The authors found that the main responsibility for the coalescence events is of the hydrophilic emulsifier. Based on the direct microscopic observations the authors managed to calculate the concentration of the external emulsifiers that will lead to coalescence of the inner droplets and similarly the concentration of the inner surfactant that leads to the coalescence starting in the outer interface. When the W| droplets were deformed the coalescence starts in the inner phase. The
588
N. Garti and R. Lutz
Fig. 25. External coalescence with no internal coalescence in a double-emulsion globule with surfactant in the external aqueous phase. Wi = 25 CMC Tween 80 solution, O = «-hexadecane with 0.05 M Span 80 (2% w/w), and W2 = 25 CMC Tween 80 solution (Reproduced with permission from Ref. 119).
Fig. 26. Internal coalescence followed by external coalescence in a W1/O/W2 double-emulsion globule. Wi = 50 CMC Tween 80 solution, O = n-hexadecane with 0.05 M Span 80 (2% w/w), and W2 = pure water (Reproduced with permission from Refs. 119).
emulsifiers concentration needed to avoid such dominant inner phase coalescence was termed 'critical internal emulsifier concentrations' and calculated to be of 1-3 times the CMC values for SDS and 50-100 times the CMC of Tween 80. On the other hand, when the outer phase deforms the coalescence starts in the outer interface and the coalescence is outer phase dominated. Such pathway occurs when the hydrophilic surfactant concentrations is very high (>100 CMC for Tween 80). With certain levels of Tween 80 or SDS (<25 CMC and <0.5 CMC respectively) spontaneous emulsification occurs. The main conclusion from this study is reconfirmation and illustration of what was known to many double emulsions formulators, that the hydrophobic surfactant (Span 80 in this case) have a stabilizing effect on the double emulsion while the hydrophilic surfactants (Tween 80, or SDS in this case) have a destabilizing effect. In addition, it was demonstrated that by varying the surfactant concentrations and controlling the hydrophobic/lipophilic ratio in both phases, one can minimize the internal coalescence. The work is mostly observation-based and empirical without strong scientific support. However, it is
Recent Progress in Double Emulsions
589
an excellent contribution to those that are searching for the optimal surfactants ratio to get best stabilities. The work allows also future manufactories to be able to minimize internal coalescence in their preparations. 5.2. Coalescence derived from droplets closer to the external interface and those closer to the center of the droplet Coalescence can be triggered by interface instabilities occurring when excess of hydrophilic surfactant is migrating to the inner interface and changing the bending energy and the interface curvature of the droplet in contact with the oil continuous interface producing an opening and a sudden release of the droplet content to the water continuous phase [21, 120]. The two-stage coalescence of the inner droplets [121] in double emulsions was recently followed step by step by fast digital video microscopy. At the first stage, a transient water-in-water emulsion is formed i.e. the oil enveloping the water droplets is removed (oil retraction), leaving the droplet immersed in the continuous water phase supported by a film of oil and surfactant. The retraction of the oil occurs in a time span of 1 milliseconds. During the second stage, the film covering the water droplets wears of and the drop breaks, releasing its contents to the continuous water phase. The second stage occurs in a time span of few tens of milliseconds. Those studies contribute very significant information to the understanding of sub processes occurring during the coalescence. The above mentioned mechanism is typical to any coalescence processes caused by drainage of the continuous phase from interfaces droplets films, film thinning and film rapture. The major difference between the single phase emulsion and double emulsion is derived from existence of the second interface. The authors distinguish between two types of droplets within the double emulsions, those close to the external interface and those closer to the center of the droplet (far from the external interface). Droplets close to the external interface are more likely to exit (be released) earlier than those closer to the center of the oil globule, which then have the time to grow larger by internal coalescence. This time scale difference depends, to some extent, on the time required to drag out the fluid between the two interfaces. Once in droplets are contact, temperature driven fluctuations break the external interface and the oil retracts to its shape of minimal energy. Thus, the time of the first stage is determined by the properties of the oil phase such as viscosity, surface tension, initial and final curvature, and so forth. The time span of the second stage depends on the formation of the film of oil and surfactant around the water droplets and on its properties such as thickness structure and molar composition. It is again clear that the first stage of oil drainage can be minimized by controlling the oil viscosity, its interfacial tension and use surfactant covering properly the water droplets. Similarly, the second stage is also controlled using the same tool. The only problem with the study is that in reality we can not see
590
N. Garti and R. Lutz
(or foresee) when type 1 (close to interface) or type 2 (close to the center) droplets are formed and the controls are difficult to quantify and to be predicted. 5.3. Osmotic pressure driven coalescence and volume fraction Many of the stability/instability phenomena are derived from osmotic pressure gradient that might exist in double emulsions, i.e the osmotic pressure in the inner phase does not match the osmotic pressure of the outer phase. The osmotic pressure gradient is a major driving force of water flow from the inner to the outer phase, and vice versa for the swelling phenomena of double emulsions globules. The osmotic pressure gradient can be driven by the existence of different molecules in each of the phases or concentration differences between the two phases. Scientifics have recognized that osmotic pressure gradient is a key factor in the stabilization/destabilization of double emulsions. Many studies were conducted in order to control the pressure gradient by adding electrolytes, sugar, etc. and the parameters affecting it are well quantified. Diagram depicting swelling/breakdown kinetics derived from a situation when the internal phase consists of matter of higher concentration than the external phase was the basis of an interesting work done by Grossiord and Seiller [12]. Under osmotic pressure gradients condition, water flow is observed from the external to the internal phase in order to reduce the concentration gradient. This water flow makes the droplet swell, and beyond a critical size the oil membrane breaks, a coalescence occurrs [122, 123]. The authors figured that a good solution is overcome such pressure gradient is to minimize the water activity. One way of restricting water mobility is by increasing the volume fraction of the inner phase. When highly concentrated emulsions are produced, with volume fraction of the oil phase close to 80%, the dispersed droplets are closely packed, so that there is not sufficient room and not much water to swell the droplets and the emulsion remains stable with respect to the swellingbreakdown mechanism, but upon dilution the breakdown mechanism will become dominant. In conclusion - the major instability pathways are coalescence driven of either the inner droplets or the double emulsions globules. Osmotic pressure gradient, viscosity, film coverage (proper balance between hydrophilic and hydrophobic surfactants) will dictated the coalesce nature that eventually can be reduced or stopped by various preparations tricks and techniques.
6. RELEASE AND TRANSPORT PHENOMENA The uncontrolled and fast released of entrapped material are the main obstacle that prevents commercializing the double emulsions technology. The only applications that are in use today are heavily thickened formulations (in their inner and outer phases) by hydrocolloids, regarding the transport to the outer
Recent Progress in Double Emulsions
591
phase. However, many of the pharmaceutical and food applications and most of the industrial applications consist of pourable liquids with dilution capabilities. In addition, many of the applications require long storage stability (shelf-life of months or years) without any release of addenda. In addition, triggered release upon addition or change of pH should be possible. Such requests are difficult to accomplish. The release from double emulsions was discussed in great details by many researchers [2, 10, 12, 65, 108 and 128] but up to this day there is no full consensus on the release mechanisms and how to control or to prefer one over the other, and whether one can separate the release mechanism from the instability processes. It should also be noted that researchers are using different terminologies for similar, events instability which cusses additional confusion. The biggest confusion is derived from the fact that in order to prove certain pathway the researchers are working on ideal or simple case studies (single bubble [118, 119], Brownian flow, equalizations of osmotic pressure [12], gelation of one of the phases [109, 110], hypertonic conditions [88], etc. Simplistic cases studies conclusions are extracted, that are not sufficiently general and many of them are even not relevant, or do not reflect the reality of double emulsion release patterns. Nevertheless, the studies have importance since they help to stress certain mechanisms or pathways that otherwise are masked and impossible to detect. Generally, common double emulsions (W/O/W) are composed of two aqueous phases and each contains soluble ingredients. Consequently, as previously explained, osmotic pressure builds up in the double emulsion system. The active matter tends to diffuse and migrate from the internal phase to the external phase. In the early work by Mansumoto et al. [54, 58] two possible mechanisms for permeation of water and water soluble materials through the oil phase were suggested; the first being via 'reverse micellar transport' and the second by 'diffusion across a very thin lamella' of surfactant, formed where the oil layer is vary thin. These two transport mechanisms were further studied by various groups and many evidences were providing to straighten their existence. Colinart et al. [124] described two other major pathways. It seems therefore, that by consolidating the transfer mechanisms described by the two groups. One can distinguish between three major transport mechanisms: (1) reverse micellar, (2) thin lamella transport and (3) hydrated surfactants transport. Three mechanisms are schematically shown in Fig. 27. 6.1. Reverse micellar transport Garti et al. [60-65] studied the release of electrolytes from double emulsions stabilized by of monomeric and polymeric emulsifiers. They have indicated that for monomeric emulsifiers the electrolyte transport takes place
592
N. Garti and R. Lutz
Fig. 27. (a) water transport through thin lamella of surfactant; (b) water transport via reverse micelles; (c) water transport via hydrated surfactants (Reproduced with permission from Ref. 118).
Recent Progress in Double Emulsions
593
mostly by reverse micellar mechanism, because even if the droplets are made to be very stable to coalescence and even when the osmotic pressure of the two phases has been equalized, the electrolytes were found to be transported. The Higuchi model for the release of matter from polymeric matrices was adopted [63-65 and 125]. A modified model was worked out and a modified release equation was offered. Excellent correlation coefficients confirming the release model were obtained in these experiments. It has been also demonstrated that if the osmotic pressure is not equalized the external water can flow into the inner W/O droplets and swell them or can migrate from the inner droplets to the outer phase. The migration direction and quantity is osmotic pressure gradient driven and the transport is via reverse micelles. It was demonstrated that surfactant molecules, water molecules and water soluble addenda are transported in a significant rates, even in the absence of osmotic pressure gradient, if sufficient number of reverse micelle exist in the oil phase. The mechanism was termed "diffusion controlled release mechanism of reverse micellar transport". According to many observations (supporting these pathways) hydrophobic surfactants, that are always in great excess, aggregate in the oil phase to form reverse micelles that can solubilize water molecules and watersoluble molecules (including electrolytes) in their core. The swollen micellar transport with the oil phase and upon approaching the outer interface they transport the addenda through the "surfactant layer" to the outer watercontinuous phase. It were demonstrated, that by diminishing the amount of the hydrophobic surfactant (Span 60, sorbitan monostearate), the numbers of the reverse micelles is reduce. If the hydrophobic surfactants are not in great excess the formation of the reveres micelle is reduced, and the release rates of the addenda transported to the outer phase is minimized. In addition, we have demonstrated that adding 'polymeric amphiphilic silicon-based surfactants' in the inner and outer phase reduced the release rate dramatically since the polymers can not form reverse micelles. 6.2. Thin film lamella transport Transport via a thin-lamella was also proved to exist, mainly when the inner droplets are free to move in the oil phase (non viscous system) and can approach the outer interface. Double layered film is formed that drains easily, and ruptures the film and releases the entire water droplets to the continuous phase [54, 59]. Such pathway will empty the inner phase without causing coalescence. The pathway was demonstrated also microscopically and by other techniques including non-invasive measurements that monitored the electrolytes content at the outer phase with time. Papadopoulos et al. [16, 118] proposed two competing release mechanisms of water that entrapped in the inner phase depending on the position of the water droplet within the oil phase. They distinguished two cases
594
N. Garti and R. Lutz
of the water droplets: those in visual contact with the Wt/O in O/W2 interface or those with no visual contact. In Fig. 28, one can see the under microscope (picture and illustration), the two positions of the water droplets. They assumed, based on the results of transport rates (droplets in visual contact transport much greater than those that are in the center, that in case of visual contact 'an hydrate surfactant mechanism' is the main controlling transport, whereas at no contact spontaneously emulsified droplets and reverse micelles are mainly responsible to the transport of water. The authors [16] investigated the influence of two surfactants on the rate of transport in double emulsion system with monomeric surfactant (Span 80 the hydrophobic surfactant and Tween 80 the hydrophilic surfactant). It was found that, although, in general, water transport rates increase with increasing water-soluble surfactant concentration in W|, minute amounts of Tween 80 in Wi retard water transport from Wi to W2 irrespective of the transport mechanism as compared to when there is no surfactant in Wi. Unlike
Fig. 28. Wi/O and O/W2 interfaces are in (a) visual contact and (b) no visual contact (Reproduced with permission from Refs. 16 and 118).
Recent Progress in Double Emulsions
595
Fig. 29. Effects of Span 80 concentration in oil on water transport rates in W1/O/W2, when the W|/O and O/W2 interfaces are at (a) visual contact and (b) no visual contact. W, = pure water; O = rc-hexadecane + Span 80; W2 = 5 M NaCI solution (Reproduced with permission fromRef. 16).
when present in W|, water-soluble surfactant in W2 always increases the water transport rates, even at very low concentrations (Fig. 29).
596
N. Garti and R. Lutz
6.3. Droplets and globules breakdown and release The two transport mechanisms are well established, and tend to well correlate to the release of water, electrolyte or active material from the inner phase to the outer phase in stable double emulsions stabilized by two sets of monomeric emulsifiers. However, in many cases the release might be due to other two major events which result from film rupturing and not lamellar or micellar transport: (1) coalescence of the thin liquid film separating the internal droplet, and (2) rupture of the outer interface at the double emulsion globule surface. The two pathways were discussed in the previous paragraph, dealing with stability mechanisms. The second pathway is very relevant to the release pathways and will be further discussed. Authors termed it 'compositional ripening' witch occurs as a result of diffusion and/or permeation of the active material/water across the oil phase [126]. The authors found suitable experimental conditions where the two mechanisms occur over time scale and could separate them. The authors claimed that there is a possibility to shift from one mechanism to the other by varying the proportions and/or the chemical nature of the surface active agents. The work shows that using polymeric amphiphiles and manipulating type of oil and type of surfactant linked to temperature variations allows a controlled release over time dictated by mechanism 1. Fig. 30 shows release patterns dictated by mechanism 1 while Fig. 31 shows pathway dictated by mechanism 2 once using monomeric surfactants or dictated by mechanism 2 using polymeric surfactants (Fig. 32).
Fig. 30. Kinetics of leakage of an emulsion stabilized by Arlacel P 135 and Synperonic PE/F68 as a function of temperature. [NaCl]°=0.4 mol/l, (|>,0=20%, ^=18%, d,=0.36 |j.m, dg=4 urn, G = 5 % , C7=2% (w/w); oil:dodecane [126].
Recent Progress in Double Emulsions
597
Fig. 31. Kinetics of leakage of an emulsion stabilized by Span 80 and SDS as a function of temperature: [NaCl]°=0.4 mol/l, <|>/0=20%,<|)g=l2%, d,=036 urn, 4=3.6 urn, C/=2% w, C>0.3 CMC, oil:dodecane[l26].
Fig. 32. Kinetics of release in presence of various natural polyelectrolytes: [NaCI]°=0.4 mol/l, c|>,0=40%, (|)g=60%, d,= \ nm, <4=17 ^m, C/=5% w; C/,=3% E; oil:sunflower oil; T=25 °C. The dashed lines are only guides to the eyes (Reproduced with permission from Ref. 126).
A relative percentage of salt released, during the first 5 days following the preparation, of the double emulsion, stabilized by hydrophobic polyethylene-30 dipolyhydroxystearate (Arlacel PI35) and hydrophilic block copolymer of polyethylene glycol (PGE) and polypropylene glycol (PGE) (Synperonic, PE/F68) at three different temperatures is shown in Fig. 31. Since the water droplet concentration (Pi) in the globules did not change, it was concluded that the release was controlled by diffusion and/or permeation of the salt across the oil globule without film rupturing.
598
N.GartiandR.Lutz
On the other hand, one can find the evolution of the salt from two emulsions formulated with identical stabilizing agent, with globules and droplets of the same colloidal size, but with oils of different chemical structure: in one case the oil used is dodecone while in the other one, it is sunflower oil. The nature of kinetic evolution is quite comparable. Under these conditions it seems that the oil is not a dominant factor in the release patterns. From Fig. 32 it was observed that biopolymers such as carboxymethyl cellulose (CMC); HEC (hydroxyethyl cellulose), and alginate that were introduced both in the inner and outer phases (same yield of entrapped, same droplet sizes stabilized by Arlacel PI35 are affecting the release of the salt. It suggests that the release is compositional ripening department (without film rupturing). By varying the proportions and/or the chemical nature of the surface active species it is possible to shift from one mechanism to the other. Those observations are confirming Matsumoto, and our, early findings on the dilemma of using the combination of excess hydrophobic surfactant for additional stability but enhanced release vs. the use of excess hydrophilic surfactant for less stability and less transport. 6.4. Ion-pair controlled transport Most of the transport mechanisms that were previously discussed deal with nonionic surfactants and addenda that do not affect the pH of the double emulsions. However, in many of the potential applications ionic addenda is encapsulated, and the pH varies accordingly. The transport of acidic species in W/O/W double emulsions was examined [127]. Diffusion mechanism (with breakdown of oil globules) was the dominant mechanism, similarly to other system [128]. However, the pH is the driving force for the uncontrolled releases. If a more controlled is required one should add an "ion-pair" additive (octadecylamine) in the oil phase that limits the acidification of the external aqueous phase and slowed down the transport. Fig. 33 is showing the release patterns of double emulsions stabilized by two acidulants. EMI is double emulsions which contain lactic acid in the inner phase and EM2 is a double emulsion containing hydrochloric acid. For hydrochloric acid (EM2), the pH of the external phase decreased just after preparation, meaning very fast and uncontrolled released. However, with lactic acid the pH was almost constant over period of time. The results pointed out the rapid diffusion of lactic acid from the inner to the outer phase just after preparation, whatever was the mechanism involved. At pH 2 in the internal phase, lactic acid would exist
Recent Progress in Double Emulsions
599
Fig. 33. pH variation versus time for multiple emulsions containing lactic acid (EMI) or hydrochloric acid (EM2) (Reproduced with permission from Ref. 127).
almost exclusively as the unionized form. So it could easily pass through the oil layer to the external aqueous phase along a concentration gradient. This diffusion is faster since the lactic acid in the oil phase forms (probably) dimmers with oil. The slow diffusion observed for hydrochloric acid was more surprising: hydrochloric acid is a strong acid and was present in the internal aqueous phase as H3O+ species. Theoretically, a simple diffusion of these ionized species through the oil layer appears impossible even for its hydrated form H3O+(H2O)3. Nevertheless, the hypothesis of Fick's diffusion is strengthened by the study of the influence of several oils with different polarities. 6.5. Controlling the release - possible strategy Many possible strategies can be examined in order to control the release. Every strategy is suitable for the corresponding release mechanism. Diminishing the repute of the inner droplets can be achieved by reducing the droplets size, (by homogenization, new surfactants, new membrane separation technique formation of microemulsions), increasing the viscosity, sealing better the interface, use surfactants that do not form, reverse micelle, etc. will reduce reverse micelle transport and will to minimize excess of reverse micelle at the oil phase (less surfactants, polymeric amphiphilic), will reduce hydrate transport or ion-pair release. Adequate additives must be incorporated. The osmotic pressure gradient can be minimized by adjusting the osmotic pressure of the two
600
N. Garti and R. Lutz
phases. From recent studied it seem that polymeric amphiphilic are adequate solution for many of the problems related both to stability and release. We will bring only some new example to the importance of the polymeric amphiphilic in stabilization the double emulsions. 7. CONCLUSIONS AND FUTURE PROSPECTIVE Double emulsions are considered extremely promising formulations for slow and controlled release of entrapped active matter from the inner phase to the outer continuous phase. However, intrinsic stability and release problems discouraged scientists and technologists and the promising formulation was almost abandoned. However, recent progress in utilization of advanced analytical tools to study stability, release mechanisms and pathways along with new and innovative ideas were the driving force for the revival of the double emulsion technology. Refreshing flow on new studies suggesting interesting ideas to overcome the stability and release difficulties are promising that soon we can see food, cosmetic or pharmaceutical applications based on double emulsions. Advanced engineering methods of preparation of double emulsions were used to form small and monodispersed globules. Upon the most impressing technologies are the use of specific porous membranes, the use of polymeric amphiphilic, viscosity enhancing agents and knowledgeable selection of the emulsifiers. One of the most promising techniques for making double emulsions is using the concept of 'emulsified microemulsion' or 'emulsified mesophases'. Inner W/O droplets with submicronal sizes are probably the key to more stable double emulsion droplets with narrower droplet distribution and smaller in size (1-5 m). Stability considerations are an important factor to establish new ways for making double emulsions. Coalescence of the inner phase is now minimized by reducing the right droplet sizes, achieving better (full) coverage of the interface by using polymeric amphiphiles of synthetic origin or natural occurring. Gelation or viscosifying of the inner phase are yet additional option. Steric stabilization, mechanical stabilization and depletion stabilization can improve stability considerably. Release pathways are numerous and difficult to control. The major release pathway is based on film rapture of the inner or the outer interface. Scientists learned how to minimize the excess of the outer surfactant and to maximize the content of the inner (hydrophobic) one. However the release is dictated also by at least four other pathways. Reverse micellar transport, diffusion and thinning the lamellar mechanism,
Recent Progress in Double Emulsions
601
surfactant hydration transport and ion-pairing release. All four pathways were discussed in great details and scientists learned to deal with them one at a time. However, we are far from being able to control all four of them in parallel. In short, double emulsion of today's formulations are much more stable, smaller in globule sizes, smaller in inner droplets sizes, more monodispersed and more viscous. Double emulsions of today's formulation can retain the addenda for longer periods of time of the shelf and release them in a more controlled pattern. Yet, full control is far from being achieved. Double emulsions are a mature technology for some applications and technology in its infancy for other applications. A very stable double emulsion (for over 40 years) can be made by cosmetic creams but a non-stable formulation (less than few days) can be manufactured for beverages. We need much more work in order to form nonviscous, dilute emulsion with prolonged stability and good controlled release. This technology is waiting for new approaches and new ideas for fast future progress. REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [ 18] [19] [20]
N. Garti, Colloids Surf. A, 123 (1997) 233. J. Sjoblom (ed.), Encyclopedic Handbook of Emulsion Technology, Marcel Dekker, New York, 2001,377-407. Y.F. Maaand C. Hsu, J. Controlled Release, 38 (1996) 218. W. Seifriz, J. Phys. Chem., 29 (1925) 738. D.H. Lee, Y.M. Goh, J.S. Kim, HK Kim, H.H. Kang, K.D. Suh and J.W. Kim, J. Dispersion Sci. Technol., 23 (2002) 491. A.K. Chattopadhyay and K.L. Mittal (eds.), Surfactants in Solution, Marcel Dekker, New York, 1996, 297-332. Y. Barenholtz and D.D. Lasic (eds.), CRC, New York, 1996, 143-199. M. Seiller and J.L. Grossiord (eds.), Influence of the Formulation on the Characteristics and Stability of Multiple Emulsions, Editions de Sante', Paris, 1998, 81-116. S. Matsumuto and W.W. Kang, J. Dispersion Sci. Technol., 10 (1989) 455. N. Garti, Progress in stabilization and transport phenomena of double emulsions in food applications. Lebensm. Wiss. Technol., 30 (1997) 222. N. Garti and C. Bisperink, Curr. Opin. Colloid Interface Sci., 3 (1998) 657. J.L. Grossiord and M. Seiller, S.T.PPharma Sci., 11 (2001)331. A.T. Florence and D. Whitehill, Int. J. Pharm., 11 (1982) 277. M. Nakano, Adv. Drug Delivery Rev., 45 (2000) 1. J. Bibette, F. Leal-Calderon, V. Schmitt and P. Poulin, Springer Tracts Mod. Phys., 181 (2002) 117. L. Wen, and K.D Papadopoulos, Langmuir, 16 (2000) 7612. J. Jiao, D.G. Rhodes and D.J. Burgess, J. Colloid Interface Sci., 250 (2002) 444. R.H. Tromp, F. van de Velde, J. van Riel and M. Paques, Food Res. Int., 34 (2001)931. A. Benichou, A. Aserin andN. Garti, Polym. Adv. Technol., 13 (2002) 1019. S. Mastumoto, J. Texture Stud., 17 (1986) 141.
602
N. Garti and R. Lutz
[21] K. Pays, J. Giermanska-Kahn, B. Pouligny, J. Bibette and F. Leal-Calderon, Langmuir, 17(2001)7758. [22] Y.H. Cho and J. Park, J. Food Sci., 68 (2003) 534. [23] K. Yoshida, T. Sekine, F. Matsuzaki, T. Yanaki and M. Yamaguchi, J. Am. Oil Chem. Soc, 76(1999) 195. [24] S.Y. Kim and Y.M. Lee, Ind. Eng. Chem. Res., 5 (1999) 306. [25] A. Edris and B. Bergnstahl, Nahrung, 45 (2001) 133. [26] H. Fukuda and M. Tanaka, Japan Kohai, 77 134029, 1977, through, Chem. Abst. 88, pp. 78962b, 1978. [27] M. Gallarate, M.E. Carlotti, M. Trotta and S. Bovo, Int. J. Pharm., 188 (1999) 233. [28] M. Gallarate, M.E. Carlotti, M. Trotta and M. Aimaretti, J. Dispersion Sci. Technol., 22 (2001) 13. [29] T.V. Vasudevan and M.S. Nase, J. Colloid Interface Sci., 256 (2002) 208. [30] G.M. Tedajo, S. Bouttier, J.L. Grossiord, J.P. Marty, M. Seiller and J. Fourniat, Int. J. Antimicrob. Agents, 20 (2000) 50. [31] H. Okochi and M. Nakano, Adv. Drug Delivery Rev., 45 (2000) 5. [32] Y. Nakamoto, M. Hashida, S. Muranishi and H. Sezaki, Chem. Pharm. Bull, 23 (1975) 3125. [33] H. Higashi, N. Tabata, K.H. Kondo, Y. Maeda, M. Shimizu, T. Nakashima and T. Setoguchi, J. Pharmacol. Exp. Ther, 289 (1999) 816. [34] C.J. Benoy, R. Schneider, L.A. Elson and M. Jones, Eur. J. Cancer, 10 (1974) 27. [35] J.A. Omotosho, A.T. Florence and T.L. Whateley, Int. J. Pharm., 61 (1990) 51. [36] M. Shichiri, Y. Shimizu, R. Kawamori, M. Fukuchi, Y. Shigeta and H. Abe, Diabetologia, 10(1974)317. [37] S.W. May andN.N. Li, Biochem. Biophys. Res. Commun., 47 (1986) 1179. [38] Y. Morimoto, K. Sugibayasi, Y. Yamaguchi and Y. Kato, Chem. Pharm. Bull., 27 (1974)3188. [39] C M . Borwanker, S.B. Pfeffer, S. Zheng, R.L. Beissinger, D.T. Wasan and L.R.S. Rosen, Biotechnol. Prog., 4 (1988) 210. [40] Y. Onuki, M. Morishita, H. Watanabe, Y. Chiba, S. Tokiwa, K. Takayama and T. Nagai, S.T.P Pharma Sci., 13(2003)231. [41] S.T. Dogru, S. Calis and F. Oner, J. Clin. Pharm. Ther., 25 (2000) 435. [42] C. Laugel, P. Rafidison, G. Potard, L. Aguadisch and A. Baillet, J. Controlled Release, 63 (2000) 7. [43] S.C. Yu, A. Bochot, G. Le Bas, M. Cheron, J. Mahuteau, J.L. Grossiord, M. Seiller and D. Duchene, Int. J. Pharm., 261 (2003) 1. [44] J. Verstee, Stabilized static liquid membrane compositions, US Patent No. 4 083 798 (1978). [45] C.Y. Lin and K.H. Wang KH, Fuel, 82 (2003) 1367. [46] F.T. Meng, G.H. Ma, Y.D. Liu, W. Qiu and Z.G. Su, Colloids Surf. B, 33 (2004) 177. [47] T. Uchida, K. Yoshida and S. Goto, J. Microencapsulation, 13 (1996) 219. [48] Y. Kita, S. Matsumoto and D. Yonezawa, J. Colloid Interface Sci., 62 (1977) 87. [49] C. Fox, Cosmet, Toiletries, 101 (1986) 101. [50] P. Becher (ed.), Encyclopedia of Emulsion Technology, Marcel Dekker, New York, 1983,337-368. [51] T.J. Lin, H. Kurihara and H. Ohta, J. Soc. Cosmet. Chem., 26 (1975) 121. [52] M. Kanouni, H.L. Rosano and N. Naouli, Adv. Colloids Interface Sci., 99 (2002) 229. [53] F.D. Rumscheidt and S.G. Masom, J. Colloids Sci., 16 (1961) 238. [54] S. Matsumoto, Y. Kida and D. Yonezawa, J. Colloid Interface Sci., 57 (1976) 353.
Recent Progress in Double Emulsions
[55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90]
603
S. Matsumoto, M. Koda and S. Murata, J. Colloid Interface Sci., 62 (1977) 149. S. Matsumoto, Y. Ueda, Y. Kita and D. Yonezaw a, Agric. Biol. Chem, 42 (1978) 739. S. Matsumoto, J. Colloid Interface Sci., 94 (1983) 362. S. Matsumoto, T. Inoue, M. Koda and T. Ota, J. Colloid Interface Sci., 77 (1980) 564. S. Matsumoto, Y. Koh and A. Michiura, J. Dispersion Sci. Technol, 6 (1985) 507. M. Frenkel, R. Shwartz and N. Garti, J. Colloid Interface Sci., 94 (1983) 174. N. Garti, M. Frenkel and R. Schwartz, J. Dispersion Sci. Technol., 4 (1983) 237. S. Magdassi, M. Frenkel and N. Garti, J. Dispersion Sci. Technol, 5 (1984) 49. S. Magdassi, M. Frenkel, N. Garti and R. Kasan, J. Colloid Interface Sci., 97 (1984) 374. S. Magdassi, M. Frenkel and N. Garti, Drug Dev. Ind. Pharm., 11 (1985) 791. S. Magdassi andN. Garti, J. Controlled Release, 3 (1986) 273. A.J. Khopade and N.K. Jain, Drug Delivery, 6 (1999) 107. M.P.Y. Piemi, M. De Luca, J.L. Grossiord, M. Seiller and J.P. Marty, Int. J. Pharm.. 171 (1998)207. E. Pilman, K. Larsson and E. Torenberg, J. Dispersion Sci. Technol., 1 (1980) 267. A.G. Gaonkar, Method for preparing a multiple emulsion, US Patent 5 322 704 (1994). M. Monduzzi, H. Ljusberg-Wahren and K. Larsson, Langmuir, 16 (2000) 7355. J. Gustafsson, H. LjusbergWahren, M. Almgren K. Larsson, Langmuir, 12 (1996) 4611. D. Anderson, H. Wennerstrom and U. Olsson, J. Phys. Chem., 93 (1989) 4243. D. Anderson, H. Wennerstrom and U. Olsson, Abstr. Pap. Am. Chem. Soc, 201 (1991) 168-COLLPart 1. T. Landh, J. Phys. Chem, 98 (1994) 8453. K. Larrson, Dispersions of lipid-water phases, in Lipids-molecular organization physical functions and technical applications, The Oily Press Ltd, Scotland, 1994. N. Garti, A. Aserin and R. Efrat, L4 Cubosomes from Ternary, US Patent 283400 (2003). L. Vigie, Emulsions multiples h/l/h: nouveaux proce'de's de fabrication et decaracte'risation, Graduate thesis, University Paris XI, 1992. M. Seiller and J.L. Grossiord (eds.), Multiple Emulsions: Structure, Properties and Applications, Editions de Sante', Paris, 1998, 57-80. J.M. Lee, K.H. Lim and D.H. Smith, Langmuir, 18 (2002) 7334. S. Higashi, M. Shimizu, T. Nakashima, K. Iwata, F. Uchiyama, S. Tateno, S. Tamura and T. Setoguchi, Cancer, 7 (1995) 1245. S.M. Joscelyne and G. Tragardh, J. Membr. Sci, 169 (2000) 107. G.T. Vladisavljevic and H. Schubert, J. Membr. Sci, 225 (2003) 15. T. Nakashima, M. Shimizu and M. Kukizadi, Proceeding of the 2nd International Conference on Inorganic Membranes Montpellier, France, 1991. K. Pays, J. Giermanska-Kahn, B. Pouligny, J. Bibette and F. Leal-Calderon, Phys. Rev. Lett, 87(2001) 178304-1. T.G. Mason and J. Bibette, Langmuir, 13 (1997) 4600. C. Mabille, V. Schmitt, P. Gorria, F.L. Calderon, V. Faye, B. Deminiere and J. Bibette, Lamgmuir, 16 (2000) 422. J. Allouche, E. Tyrode, V. Sadtler, L. Choplin and J.L. Salager, Ind. Eng. Chem. Res, 42 (2003) 3982. T. Hino, Y. Kawashima and S. Shimabayashi, Adv. Drug Delivery Rev, 45 (2000) 27. T. Hino, A. Yamamoto, S. Shimabayashi, M. Tanaka and D. Tsujii, J. Controlled Release, 69(2000)413. H.L. Rosano, F.G. Gandolfo and J.D.P. Hidrot, Colloids Surf. A, 138 (1998) 109.
604
N. Garti and R. Lutz
[91] M.F. Ficheux, L. Bonakadar, L. Leal-Calderon and J. Bibette, Langmuir, 14 (1998) 2702. [92] D. Myers, Emulsions in Surfactant Science and Technology, VCH Publishers, Inc., New-York, 1998. [93] Y. Sela, S. Magdassi and N. Garti, Colloids Surf. A, 83 (1994) 143. [94] Y. Sela, S. Magdassi and N. Garti, Colloid Polytn. Sci., 272 (1994) 684. [95] Y. Sela, S. Magdassi and N. Garti, Proceedings of the l sl World Congress on Emulsions, Paris, 1993, 1. [96] N. Garti and Y. Sela, Acta Polymer., 49 (1998) 606. [97] Y. Sela, S. Magdassi andN. Garti, J. Controlled Release, 33 (1995) 1. [98] M. De Luca, J.L. Grassiorod and M. Seiller, Int. J. Cosmet. Sci., 13 (2000) 1019. [99] C M . Bryant and D.J. McClements, J. Food Sci., 65 (2000) 259. [100]A. Vaziri and B J . Warburton, J. Microencapsulation, 11 (1994) 649. [101]N. Garti, A. Aserin and Y. Cohen, J. Controlled Release, 29 (1994) 41. [102] A. Koberstein-Hajda and E. Dickinson, Food Hydrocolloid, 10 (1996)251. [103] W. Zhang, T. Miyakawa, T. Uchida and S. Goto, J. Pharm. Soc. Jpn., 112 (1992) 73. [104]G.O. Phillips, P.A. Williams and D.J. Wedlock (eds.), Gums and Stabilizers for the Food Industry, IRL Oxford University Press, Oxford, 1994, 91-101. [105]L. Olivieri, M. Seiller, L. Bromberg, M. Besnard, T.N.L. Duong and J.L. Grossiord, J. Controlled Release, 88 (2003) 401. [106]L. Olivieri, M. Seiller, L. Bromberg, E. Ron, P. Couvreur and J.L. Grossiord, Pharm. Res., 18(2001)689. [107]F. Michaut, P. Hebraud and P. Perrin, Polym. Int., 52 (2003) 594. [108]D.O. Shah (ed.) Macro- and microemulsions: theory and application, ACS Symposium Series, Washington, 1985, 359. [109]O. Ozer, V. Muguet, E. Roy, J.L. Grossiord and M. Seiller, Drug Dev. Ind. Pharm., 26 (2000) 1185. [110]V. Muguet, M. Seiller, G. Barratt, O. Ozer, J.P. Marty and J.L. Grossiord, J. Controlled Release, 70(2001)37. [111JJ.W. Kim, H.H. Kang, K.D. Suh and S.G. Oh, J. Dispersion Sci. Technol, 24 (2003) 833. [112]R. Aveyard, B.P. Binks and J.H. Clint, Adv. Colloid Interface Sci., 100-102 (2003) 503-546. [113]K.P. Oza and S.G. Frank, J. Dispersion Sci. Technol. 10 (1989) 163. [114]T. Sekine, K. Yoshida, F. Matsuzaki, T. Yanaki and M. Yamaguchi, J. Surfactants Deterg, 2(1999)309. [115]N. Garti, A. Aserin, I. Tiunova and H. Binyamin, J. Am. Oil Chem. Soc, 76 (1999) 383. [116]B.P. Binks, A.K.F. Dyab, P.D.I. Fletcher and H. Barthel, Multiple emulsions, German Patent DE 10 211 313. [117]B.P. Binks, A.K.F. Dyab and P.D.I. Fletcher, Proceedings of 3 rd World Congress on Emulsions, Lyon, CME, Boulogne-Billancourt, 2002. [118]L. Wen and K.D. Papadopoulos, Colloids Surf. A, 174 (2000) 156. [119]C.H. Villa, L.B. Lawson, Y.M. Li, Y. Li and K.D. Papadopoulos, Langmuir, 19 (2003) 244. [120] A. Kabalnov and H. Wennerstrom, Langmuir, 12 (1996) 276. [121]H. Gonzalez-Ochoa, L. Ibarra-Bracamontes and J.L. Arauz-Lara, Langmuir, 19 (2003) 7837. [122]S. Matsumoto, T. Inoue, M. Kohda and K. Ikura, J. Colloid Interface. Sci., 77 (1980) 555.
Recent Progress in Double Emulsions
605
[123]M. Tomita, Y. Abe and T. Kondo, J. Phartn. Sci., 71 (1982) 332. [124]P. Colinart, S. Delepine, G. Trouve and H. Renon, J. Membr. Sci., 20 (1984) 167. [125JR.G. Stehle and W.I. Higuchi, J. Pharm. Sci., 61 (1972) 1922. [126]K. Pays, J. Giermanska-Kahn, B. Pouligny, J. Bibbette and F. Leal-Calderon, J. Controlled Release, 79 (2002) 193. [127]G.M. Tedajo, M. Seiller, P. Prognon and J.L. Grassiord, J. Controlled Release, 75 (2001)45. [128]S. Magdassi andN. Garti, Colloids Surf, 12 (1984) 367.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 15
Stability of concentrated emulsions V. Schmitta, S. Arditty a, F. Leal-Calderon b a
Centre de Recherche Paul Pascal (CRPP) / CNRS, Avenue du Dr Albert SCHWEITZER 33600 PESSAC, FRANCE
b
Laboratoire des Milieux Disperses Alimentaires, Institut des Sciences et Techniques des Aliments de Bordeaux (ISTAB), Avenue des Facultes 33405 TALENCE, FRANCE
1. INTRODUCTION The lifetime of emulsions may considerably vary from one system to another; it can change from minutes to many years, depending on the nature of the surfactants, the nature of both phases and their volume ratio. Despite the large amount of work devoted to this issue, predicting the destruction scenario and the emulsion lifetime still raise challenging questions, especially in concentrated emulsions. Irreversible coarsening of emulsions may proceed through two distinct mechanisms. The first mechanism, known as Ostwald ripening [1], is driven by the difference in Laplace pressure between droplets having different radii: the dispersed phase is transferred from the smaller to the larger droplets. The rate of droplet growth may be determined by the molecular diffusion across the continuous phase and/or by the permeation across the surfactant films. The second mechanism known as coalescence consists of the rupture of the thin film that forms between droplets leading them to fuse into a single one. At a microscopic scale, a coalescence event proceeds through the nucleation of a thermally activated hole that reaches a critical size above which it becomes unstable and grows. In principle, understanding the metastability of emulsions requires two types of information that account for two distinct phenomena. The first one concerns the microscopic mechanism of the instabilities. The second one concerns the scenario of destruction, i.e. the time and space distributions of the
608
V. Schmitt, S. Arditty and F. Leal-Calderon
coarsening events. Generally, the destruction scenario of emulsions results from the interplay between coalescence and Ostwald ripening. For the sake of simplicity, most of the studies have been performed in conditions such that one type of instability is dominating the other one, enabling one to monitor its progress quite precisely. In this limit, theoretical models as well as experiments have revealed that Ostwald ripening generates emulsions with narrow size distributions in the asymptotic regime [2-7]. On the contrary, coalescence favors the diverging growth of large droplets at the expense of the smaller ones [8, 9]. Due to their different consequences on the droplet evolution and size distribution, coalescence and Ostwald ripening can be easily identified: a system evolving under the effect of Ostwald ripening alone exhibits a narrow size distribution with a droplet growth rate that progressively vanishes; instead, coalescence events have the effect to increase the polydispersity and to accelerate the rate of coarsening due to the divergent growth of large nuclei. In this chapter, we propose a review of some recent advances in the field of emulsion metastability, emphasizing the destruction of concentrated emulsions (droplet volume fraction (j)>70%) through coalescence. The review concerning Oswald ripening (section 2) will be more concise since this mechanism is fairly well understood and has been extensively documented in the literature. So far, the destruction of concentrated emulsions through coalescence is much less understood and has motivated many recent studies and developments that we propose to summarize (section 3). 2. OSTWALD RIPENING Ostwald ripening consists of a diffusive transfer of the dispersed phase from the smaller to the larger droplets. Ostwald ripening is characterized by either a constant volume rate Q,3 [2, 3] (diffusion-controlled ripening), or a constant surface rate Q2 [6] (surface-controlled ripening) depending on the origin of the transfer mechanism:
Diffusion-controlled ripening (a=3) has been recognized in sub-micron diluted emulsions stabilized by ionic or nonionic surfactants; so far, permeationcontrolled ripening (a=2), has been proposed to account for the coarsening of concentrated air foams [6].
Stability of Concentrated Emulsions
609
If the ripening is controlled by diffusion across the continuous phase, then the cube of the diameter increases linearly with time (a=3) and the ripening rate £23 can be derived using the Lifshitz and Slyozov theory [2, 3]. ^
6
4
1
^
(2)
K.1
where S is the molecular solubility, Vm is the molar volume, Diff is the molecular diffusion coefficient of the dispersed phase in the continuous one, y is the surface tension of the oil-water interface and R is the molar gas constant. The theory also predicts that the size distribution becomes quite narrow and asymptotically self-similar, in agreement with experiments. In principle, Eq. (2) is valid in the limit of very dilute emulsions. In general, it is expected that emulsions with higher volume fractions of disperse phase will have faster absolute growth rates [10-14] than those predicted by Lifshitz and Slyozov model. The theoretical rate of ripening must therefore be corrected by a factor f(t))) that reflects the dependence of the coarsening rate on the dispersed phase volume fraction (J) [10, 14]. Kabalnov [15] and Taylor [16] found that the presence of ionic micelles in the continuous phase had a surprisingly small effect on the rate of ripening, despite the fact that the solubility of the dispersed phase is largely enhanced. It is argued that due to electrostatic repulsion, ionic micelles cannot absorb oil directly from emulsion drops. In the presence of non-ionic surfactants, much larger increases in the rate of ripening might be expected due to larger solubilization capacities and to the absence of electrostatic repulsion between droplets and micelles. Weiss et al. [17] showed that a significant increase in average diameter can be achieved for tetradecane-in-water emulsions diluted with a fresh solution of non-ionic surfactant. In this experiment, the variation of the droplet diameter results from a complex interplay between oil solubilization by the micelles which tend to reduce droplet diameter and Ostwald ripening which tends to increase it. When the dispersed phase is composed of a binary mixture, the growth may be arrested if one component is almost insoluble in the continuous phase, therefore retaining the soluble one, due to the gradual loss of mixing entropy [18]. The osmotic pressure of the trapped species within the droplets can overcome the Laplace pressure differences that drive the coarsening and 'osmotically stabilize' the emulsions. Webster and Cates [19] gave rigorous criteria for osmotic stabilization of monodisperse and polydisperse emulsions in the dilute regime. The same authors [20] have also examined the concentrated
610
V. Schmitt, S. Arditty and F. Leal-Calderon
regime in which the droplets are strongly deformed and therefore possess a high Laplace pressure. These authors conclude that osmotic stabilization of dense emulsions also requires a pressure of trapped molecules in each droplet that is comparable to the Laplace pressure that the droplets would have if they were spherical, as opposed to the much larger pressures actually present in the system. Mass transfers in emulsions may be driven not only by differences in droplet curvatures but also by differences in their compositions. This is observed when, e.g. two chemically different oils are emulsified separately and the resulting emulsions are mixed. This phenomenon is called composition ripening [21, 22]. Mass transfer from one emulsion to the other is controlled by the entropy of mixing and proceeds until the compositions of the droplets become identical. The most spectacular evidence of composition ripening comes from the so-called 'reverse recondensation' which occurs when the two emulsions differ significantly both in their initial size and in their rate of molecular diffusion. If the larger sized emulsion is composed of the faster diffusing oil, then molecular diffusion occurs in the 'reverse' direction, i.e. from large to small droplets [23]. Most of the experimental studies concerning Ostwald ripening have been performed in the limit of highly dilute emulsions ((|)<1%) [4, 5, 17, 24, 25], i.e. in conditions such that the droplets are not in permanent contact. Ostwald ripening is then preferentially controlled by the molecular diffusion of the dispersed molecules across the continuous phase, as deduced from the experimental measurements. More recently, Taisne and Cabane [26] have examined the coarsening of concentrated alkane-in-water droplets ((f>=20-40%) stabilized by a non-ionic poly(ethoxylated) surfactant, following a temperature quench. Since the rate of ripening does not depend on the alkane chain length, they conclude that the transfer of oil from the smaller drops to the larger ones does not occur by diffusion across the continuous phase but rather through the direct contact of the droplets (permeation). A similar conclusion is drawn by Schmitt et al. [27]. These authors produce different alkane-in-water concentrated emulsions (4>~8O%) stabilized by the same non-ionic surfactant and follow the kinetic evolution by granulometry. The size distribution becomes remarkably narrow during the first stages of coarsening and progressively turns to a wide function as time passes. They get evidence that the size evolution is first determined by Ostwald ripening and then by coalescence. When a second hydrophobic species of large molecular size is dissolved in the dispersed phase, it is expected to inhibit the permeation mechanism and coalescence should act alone. Surprisingly, coalescence is also suppressed, even at very low concentration of the second component (-1% w/w). The fact that the two types of instability disappear simultaneously strongly suggests that they possess the
Stability of Concentrated Emulsions
611
same microscopic origin: hole nucleation in the thin liquid films. Due to the activated nature of the process, only the largest holes grow spontaneously and produce coalescence events. Although the smaller holes are evanescent, they allow the diffusion of matter between droplets and therefore they contribute to the permeation, increasing and even dominating the overall mass flux and hence ripening rate. Since the ripening is surface-controlled, the square of the diameter should increase linearly with time. Schmitt et al show that the quadratic scaling (a=2) correctly accounts for the evolution of the diameter as a function of time for emulsions stabilized by nonionic poly(ethoxylated) surfactants. 3. COALESCENCE 3.1. General phenomenology and microscopic description. An emulsion which is for instance stable over many years at low droplet volume fraction may become unstable and coalesce when highly compressed. As an example, when an oil-in-water emulsion stabilized by Sodium Dodecyl Sulfate (SDS) is introduced in a dialysis bag and is stressed by the osmotic pressure imposed by an external polymer solution, coarsening occurs through the growth of a few randomly distributed large droplets [8]. A microscopic picture of such growth is shown in Fig. 1. By using centrifugation, the same type of compression may be applied. In this case, the osmotic pressure results from the buoyant stress of the droplet layers below the top of the cream. The destabilization proceeds through the growth of a macroscopic domain that nucleates at one side of the sample and progress as a front. Careful observation [28] reveals that the front is a thin zone
Fig. 1. Microscopic picture of an emulsion, initially composed of monodisperse droplets having a diameter of about 1.5um, which has been submitted to an osmotic stress of 0.6 atm for 15 days at room temperature (from Ref. [8]).
612
V. Schmitt, S. Arditty and F. Leal-Calderon
where the same type of domains grow before merging into a single macroscopic phase. The progress of the coalescence front during centrifugation, h(t), has been very instructive and reveals interesting issues. The function h(t) is found linear at the beginning with a slope p and saturates at longer times at a value h*, that depends upon the applied centrifugation acceleration. It can be deduced that below a certain height h*, the rate of coalescence tends to zero, allowing to define a threshold osmotic pressure IT* to induce coalescence. (3 is found to depend on the thermodynamic properties of surfactant layer. For instance, adding electrolytes in presence of ionic surfactants or increasing the length of the hydrophobic tail of non-ionic surfactants reduces the rate of coalescence |3. At a microscopic scale, a single coalescence event proceeds through the nucleation of a thermally activated hole that reaches a critical size, above which it becomes unstable and grows [29]. We shall term as E(r) the energy cost for reaching a hole of size r. A maximum of E occurs at a critical distance r*, E(r*)=Ea being the hole nucleation activation energy (Fig. 2). The origin of the activation energy is still a matter of debate but it is now well admitted that the so-called spontaneous curvature of the surfactant monolayer, covering the emulsion drops is one of the most determining parameters [30]. Convincing evidence is given by a general experimental observation: it is well known that the micellar phase always tends to be the continuous phase of an emulsion. Shaking an O/W microemulsion coexisting with excess oil preferentially leads to the formation of an O/W emulsion, the continuous phase being the O/W microemulsion. Conversely, shaking a W/O microemulsion coexisting with excess water leads to the formation of a W/O emulsion with W/O swollen micelles in the continuous phase [31]. Recently, an interpretation for this correlation was proposed by Kabalnov and Wennerstrom
Fig. 2. Scheme of the hole nucleation process and variation of the energy cost with the hole radius r.
Stability of Concentrated Emulsions
613
[30]. Let us consider an oil film separating two water droplets (Fig. 3). If the surfactant covering the droplets is essentially oil-soluble, it possess a large tail and small polar head. The spontaneous curvature Co of the monolayer is negative, meaning that it tends to curve towards oil. If a tiny hole is formed in this film, the monolayer at the edge of the hole is curved against the direction favored by the spontaneous curvature. Because of this, for the film rupture to occur, the system must pass through an energy barrier, after which the growth becomes spontaneous. This state can be reached only by a thermal fluctuation and has a low probability because of the unfavorable spontaneous curvature Co. However, if the surfactant covering the droplets is essentially water-soluble, the spontaneous curvature Co of the monolayer is positive. The hole propagation occurs with a very moderate energy barrier because the local curvature C of the edge fits the surfactant monolayer spontaneous curvature. The consequence is that W/O emulsions persist in presence of oil-soluble surfactants and are rapidly destroyed in presence of water-soluble surfactants. Conversely, O/W emulsions persist in presence of water soluble surfactants. This model is in perfect agreement with the phenomenology previously described and also with the well known Bancroft rule [32].
Fig. 3. Scheme explaining the influence of the spontaneous curvature on the activation energy for coalescence in a W/O emulsion.
614
V. Schmitt, S. Arditty and F. Leal-Calderon
The spontaneous curvature is not the only parameter controlling the activation energy. The experiments of Sonneville-Aubrun et al [33] suggest that Ea is influenced by the short range surface forces and by the surfactant packing at the oil-water interface. These authors carefully examine the influence of centrifugation on O/W emulsions stabilized by ionic surfactants. The cream that forms after most of the water has been removed has the structure of a biliquid foam. Examination of this cream through electron microscopy shows polyhedral oil cells separated by thin films. The thickness of these films was measured through small-angle neutron scattering. The results yield a disjoining pressure isotherm, where the film thickness is solely determined by the pressure applied to extract water during centrifugation. For hexadecane-in-water biliquid foams stabilized with SDS, this isotherm has two states, the common black film (CBF; water thickness beyond 25 A) and the Newton black film (NBF; water thickness of 13 A). The thickness of the NBF is stabilized by hydration forces, which resist the dehydration of counter-ions and head-groups. The surface density of SDS molecules in these films has also been measured. As water is extracted, the concentration of counter-ions increases, and the head-groups are more efficiently screened; as a result, the surface density of SDS in the monolayers rises. In the NBF state, the monolayers are tightly packed, with an orientational order that exceeds that of the lamellar phase. Consequently, some fluctuations that could cause the rupture of the films (vacancies in the monolayers and defects where opposing monolayers recombine) will be inhibited. This suggests a general route for improving the metastability of biliquid foams, which is simply to improve the packing of the surfactant molecules in the monolayers. This tighter packing of surfactant molecules may explain the surprisingly high metastability of biliquid foams when the films are in the NBF state. Other experiments performed by Bergeron [34] on air foams stabilized with ionic surfactants reveal that the so-called Gibbs or dilatational elasticity e may play an important role in the coalescence process. The Gibbs elasticity measures the variation of surface tension y associated to the variation of the surfactant surface concentration T: e= - Tdy/dr
(3)
Note that e is also linked to the packing of the surfactant molecules in the monolayers. Foam and emulsion films undergo both spatial and surfactant density (i.e. charge) fluctuations occurring at the interface (Fig. 4). The hydrodynamic influence associated with this surface moduli effect is often qualitatively expressed as a Gibbs-Marangoni stabilization mechanism. Bergeron argues that the surface elasticity plays a key role in dampening both
Stability of Concentrated Emulsions
615
Fig. 4. Scheme of (a) a spatial fluctuation and of (b) a local depletion zone due to monolayer density fluctuations in thin liquid films.
spatial and density fluctuations in foam and emulsion films. When these fluctuations are dampened, the probability of overcoming the activation barrier which holds a film in a metastable state is lower and the film will be more stable. Whether or not disturbances are thermally or mechanically induced, a cohesive surfactant monolayer with a high surface elasticity will promote film stability. The influence of 8 was revealed by comparing the relative coalescence rate of foams stabilized by ionic surfactants (alkyltrimethylammonium bromides) with the same polar head but different alkyl chain lengths: the resistance to coalescence was clearly correlated to e, which was an increasing function of the alkyl chain length. 3.2. Measurements of the coalescence frequency Following the mean field description of Arrhenius, we can define co, the frequency of coalescence per unit surface area of the droplets: co = co 0 exp(-E a /kT)
(4)
616
V. Schmitt, S. Arditty andF. Leal-Calderon
In this expression, kT is the thermal energy and coo is the natural frequency of the hole nucleation process. Assuming the existence of a unique rupturing frequency co per unit film area, numerical simulations performed on a 2-D cellular material [9], reveal that the scenario of destruction through coalescence is intrinsically inhomogeneous with a few giant cells growing much more rapidly than the average, in agreement with the experimental observations previously described. Due to the intrinsic complexity of the destruction scenario, the measurements of coo and Ea are scarce. However, any progress in understanding the metastability of emulsions is only possible trough a quantitative determination of co. In the following, we describe some recent attempts to measure co in two limiting situations (i) activated coalescence, i.e film rupturing in presence of a significant activation energy and (ii) nonactivated coalescence, /. e coalescence with a negligible energy barrier. 3.2.1. Simple emulsions stabilized by surfactants (activated coalescence) The first measurements of co performed in simple emulsions were reported by Kabalnov and Weers [35]. The emulsion is modeled as a stack of monodisperse cells with characteristic size D. The total number of drops per unit volume, n, is related to the volume fraction § of the dispersed phase:
Since coalescence is a completely random process, the total number of coalescence events per unit time is assumed to be proportional to the total surface area A of the droplets: _ dn _ 2 — coA — con7iD (o) dt From the two previous equations, we conclude that the mean size in the emulsion increases with time according the following law:
D02
D2
3
where Do is the initial diameter. This equation predicts a divergence of the diameter after a finite time x given by:
Stability of Concentrated Emulsions
T=
617
<8)
id^[
The coalescence frequency is obtained by monitoring the characteristic time at which the layer of free oil formed at the top of the sample corresponds approximately to half of the volume of the dispersed phase. This time is assumed to be equal to T. By measuring T at different temperatures, the activation energy and the natural frequency are deduced from an Arrhenius plot. The authors were able to measure the activation energy for a water-in-octane emulsion at $=50%, stabilized by the nonionic surfactant C12E5 (pentaethylene glycol mono n-dodecyl ether), above the phase inversion temperature (PIT) and found a value of 47 kTr, Tr being the room temperature. This method is based on a mean field description that assumes a monodisperse growth and predicts a finite lifetime. Clearly, this is not consistent with the experiments which reveal that the size distribution becomes extremely wide during the destruction process. At $-50%, the emulsion viscosity is quite low and local rearrangements take place: the larger drops can migrate and concentrate at the top of the samples under the effect of buoyancy. Since the biggest drops become progressively predominant, coalescence is accelerated. Indeed, neighboring drops are in permanent contact and larger drops will grow faster because they exhibit a larger surface contact area with their neighbors. In other words, the appearance of a macroscopic oil phase after a finite time may be due to creaming effects. In highly concentrated emulsions ((|)>75%), rearrangements due to creaming are no longer possible, the size distribution remains spatially homogeneous and the coarsening generally does not exhibit the diverging growth predicted by Eq. (7). A different approach has been proposed quite recently by Schmitt and Leal-Calderon [36]. They produce O/W emulsions at $=78%, stabilized by ionic and nonionic surfactants. The emulsions were stored in that concentrated state at 20°C. The droplet size distribution was measured at regular time intervals by static light scattering using Mie theory. Starting from the volume distribution, a little bit of algebra allows a straightforward calculation of the average diameter D[p,q] defined as:
'ZNiDfV^ V i
)
618
V. Schmitt, S. Arditty and F. Leal-Calderon
where Nj is the total number of droplets with diameter Dj. The emulsions were characterized in terms of their surface-averaged diameter DS=D[3,2] and polydispersity:
YN.DfD-D: i
Z_J
1
1
I
i
where D is median diameter, i.e. the diameter for which the cumulative undersized volume fraction is equal to 50%. For all the emulsions under study, the size distribution followed the same qualitative evolution as the one reported in Fig. 5 for octane-in-water droplets stabilized by SDS. In figure 6 are reported the evolutions of Ds, D[3,0] and U as a function of time for the same emulsion. The time interval where the polydispersity is lower than 20% has been shaded: in terms of droplet diameter, the interval lies between lower and upper limits, denoted Ds] and Ds2, respectively. Initially, the polydispersity is of the order of 30-40%. As time passes, Ds increases until reaching Dsl and U decreases down to around 13-20%, a low value indicating the formation of an emulsion with high degree of monodispersity. For Ds
Fig. 5. Size distribution of an octane-in-water emulsion stabilized by SDS (adapted from Ref. [36]).
Stability of Concentrated Emulsions
619
Fig. 6. Time evolution of the diameters Ds ( • ) and D[3,0] ( A ) and the polydispersity U ( • right scale) for an octane-in-water emulsion stabilized by SDS (adapted from Ref. [36]).
controlled ripening (the experimental precision is insufficient to differentiate the two possible origins on the basis of U measurements). The same mechanism is operative when the average diameter is smaller than Ds, but the asymptotic regime is not yet achieved because of the large polydispersity of the initial emulsions. At longer times, for Ds>Ds2, the distribution becomes broad again (U>20% and continuously increases) with the appearance of drops much larger than the average (Fig. 5). A general feature of this regime is the divergent diameter growth (d2Ds/dt2 > 0) which allows concluding that this evolution is consistent with a coalescence-driven mechanism. Hence, for all the systems under study, the coarsening is determined first by Ostwald ripening followed by coalescence. It is only the time scale of these two mechanisms that varies epending on the surfactant and oil chemical natures. The homogeneous growth observed in the shaded part of Fig. 6 (Dsl
620
V. Schmitt, S. Arditty andF. Leal-Calderon
coalescence becomes the dominating instability. At DS=D*, the two coarsening mechanisms occur at comparable rates
—V
UL
~ —/Coalescence
,
V
/Ostwaldripening
thus:
The right hand side was derived from the variation of the drops number considering the volume conservation principle. From Eq. (11), the authors deduce an estimation of the frequency co* valid for Ds~ D* and some values are reported in table I.
n3 Surfactant
'
(K^VS1)
n2 (mV)
D* (um)
co* (m'V) DS=D* Equation
co (m"V) Ds=D*+lum Equation
(11) 16
3.1±0.2
(13) 5
(5.2±1.5) 10
(3.4±0.6) \0b
Ifralan
Heptane
-
(1.0±0.1) 10"
Ifralan
Octane
-
(1.9±0.1) 10"'7 3.0±0.2
(l.l±0.4)10 5 (0.8±0.2) 105
Ifralan
Nonane
-
(5.9±0.3) 10"18 2.5±0.1
(7.3±2.1) 104 (6.5±1.3)10 4
Ifralan
Dodecane
-
(2.5±0.1) 10"19 2.4±0.1
(3.6=1=1.2) 103 (4.2±0.8)10 3
SDS
heptane
(6.3±0.3)
-
6.6±0.3 (1.6±0.5) 104 (2.0±0.4)10 4
SDS
Octane
(2.5±0.1)
-
6.8±0.4 (5.4±1.6)10 3 (7.2±1.4)10 3
Table I: Characteristic coalescence frequency for different alkane-in-water emulsions. Ifralan is a commercial nonionic surfactant essentially comprised of a mixture of C12E5 and C10E5. In Eq. (11), a=2 was adopted for Ifralan surfactant and cx=3 was adopted for SDS. (Adapted from Ref. [36]).
It is interesting to examine the impact of the alkane molecular weight on the coalescence rate, for a given surfactant: smaller alkanes coalesce more rapidly. The authors [27] argue that this dependence is related to the spontaneous curvature of the surfactant monolayers. The longer alkane chains, like hexadecane, can hardly penetrate the hydrophobic surfactant brush covering the surfaces and therefore the natural spontaneous curvature is quite elevated thus stabilizing the direct films against hole nucleation. Instead, shorter oil chains like octane can more easily penetrate and swell the surfactant brush
Stability of Concentrated Emulsions
621
providing a less positive average curvature, which allows rapid formation of holes in the O/W/O films. In order to test the reliability of the previous method, the authors compare it to an independent measurement of co. They thus propose an extended version of the previous mean field model, valid at any stage of the coalescence regime, even in presence of broad droplet size distributions. It is obtained by considering that the variation of the total number of coalescence events is proportional to the total surface area per unit volume developed by the droplets of different sizes. The total number of drops and total surface are replaced by summations over all the granulometric size intervals:
Using the definition of the averaged diameters D[3,0] and DS=D[3,2] (relation (9)), the previous equation can be rewritten as:
»=-I!(—LJD.
(13)
5
ndl(D[3,0] J This equation reflects the possibility to measure co from the experimental evolution of D[3,0] and Ds, both diameters being directly deduced from the experimental droplet size distributions (Fig. 6). Of course, this procedure is to be applied at long times, i.e. in the regime governed by coalescence (DS>D*). In Fig. 7, it appears that co exhibits a regular decrease with time. In a concentrated emulsion, the droplet surfaces contain facets, along which droplets press against each others across thin films of water. The inward pressure exerted on the films, called the disjoining pressure Jid, is opposed by repulsive interactions between the oil-water interfaces, and determines the film thickness. This pressure also determines the mean radius of curvature rc on the remaining curved sections of the droplets surfaces by Laplace equation: nd=2y/rc. For (j)=78% and polydisperse systems, the droplets are weakly deformed spheres, and press against each other across small facets. Hence, 7id approaches the average Laplace pressure 7tL=4y/Ds of the undeformed droplets: ftd=ftL. The variation of co essentially reflects the dependence of the coalescence frequency on the disjoining pressure in the films. With an increase on the droplets diameter, the average Laplace pressure decreases and each plane parallel film is moving along the curve of the repulsive disjoining pressure isotherm so that to increase the average film thickness. In other words, the
622
V. Schmitt, S. Arditty and F. Leal-Calderon
Fig. 7. Evolution of the coalescence frequency with time deduced from Eq.(13) for an octanein-water emulsion stabilized by SDS (adapted from Ref. [36])
evolution of co reveals the fact that %A is not a pure "hard-sphere" profile but rather a continuous decreasing function of the film thickness. For the sake of comparison, table I contains the numerical values of (0 obtained following the two previous independent methods, at the same (|)=78%. For the second one (Eq. (13)), the reported data are those obtained in the vicinity of D* (at Ds ~ D*+lum). The comparison was made at a Ds value larger than D* to be sure that the measured frequency is negligibly perturbed by Ostwald ripening. The agreement between the two methods is rather satisfactory considering that the calculation procedures are completely different. Thus, using Eq. (11), it is possible to describe the metastability of concentrated emulsions in terms of only two independent parameters: the rate of diffusive ripening D.a and the critical diameter D* that defines the cross-over between Ostwald ripening and coalescence. The characteristic coalescence frequency nearby D* can then be deduced from Eq. (11). The phenomenology previously described is quite general since it was reproduced in presence of different alkanes stabilized by both ionic and non-ionic surfactants. Ionic surfactants generally provide a better stabilization against coalescence, revealed by larger D* values. 3.2.2. Double emulsions stabilized by surfactants (activated coalescence) A rigorous method for the measurement of the coalescence frequency has been proposed by Pays et al. [37-39]. Their method is based on the use of monodisperse (W/O/W) double emulsions. The so-called water-in-oil-in-water (W/O/W) double emulsions are comprised of oil globules dispersed in water, each globules containing smaller water droplets. Salt, sodium chloride, was initially dissolved in the internal droplets as a tracer to probe the coalescence
Stability of Concentrated Emulsions
623
kinetics. The amount of salt released through coalescence from the internal droplets to the external water phase was measured by an electrode selective to chloride ions immersed in the continuous phase of the double emulsion. The method of Pays et al exploits the fact that the total number of internal droplets adsorbed on a globule surface governs the rate of release. An attractive interaction exists between the small internal droplets and between the droplets and the globule surface. However, since the globules are at least 10 times larger than the entrapped droplets, the attraction between the almost flat globule and a small droplet is nearly twice as large as that between inner droplets. This attraction is small enough for the small droplets to behave like a gas which adsorbs reversibly onto the globule surface. By varying the concentration of the hydrophilic surfactant within the external water phase, the authors found a regime where the leakage is controlled by the droplet/globule coalescence only. Under such conditions, measuring the rate of release allows a direct determination of the average lifetime of the thin film that forms between a small internal droplet and the globule surface. A determination of both the activation energy and the natural frequency of the hole nucleation process is then possible, by exploring the temperature dependence of the rate of release. The thin liquid film that forms between the internal droplets and the globule surface is comprised of two mixed monolayers covered by both oil and water-soluble surfactant molecules. Since the water-soluble molecules rapidly migrate from the external to the internal water phase, the film can be regarded as close to thermodynamic equilibrium with respect to surfactant adsorption. As was previously explained, such inverted films possess a long-range stability when essentially covered by hydrophobic surfactant (C0<0) but become very unstable when a large proportion of hydrophilic surfactant (C0>0) is adsorbed. The transition from long range to short range stability may be achieved by varying the concentration of the hydrophilic surfactant in the external water phase. In the experiments described by Pays et al [38], the double globules are comprised of dodecane and the surfactant used are sorbitan monooleate Span 80, which is oil soluble and the water-soluble SDS. The concentrations of both surfactants are fixed and the initial internal droplet volume fraction is varied between 5% and 35 %. The data are represented in Fig. 8. The ordinate corresponds to the amount of salt released trough coalescence expressed in relative %. A significant decay of the characteristic time of release as a function of the internal droplet volume fraction is observed. A careful observation under microscope reveals that a fraction of the internal droplets is adsorbed on the globule surface. This is a natural consequence of the Van der Waals attraction that exists between the internal droplets and the
624
V. Schmitt, S. Arditty andF. Leal-Calderon
external water phase, nj is defined as the total internal droplet concentration within the globules, and na is the concentration of adsorbed droplets. It can be assumed that the number of coalescence events per unit time is simply proportional to the concentration of adsorbed droplets:
JlL = -A n a =-A f(ni)
(14)
where A is the characteristic frequency of coalescence between an adsorbed droplet and the globule surface. In the model described by Pays et al, the coalescence rate A is defined as a number of coalescence events per unit time and is not normalized by the droplet surface. At any time t, n; is calculated from the ordinate of the curves and the number of coalescence events dn/dt is deduced from the derivative. All the experimental points in Fig. 8 were transformed and plotted again in (n,, dnj/dt) coordinates on Fig. 9. The data lie within a single curve, meaning that the rate of coalescence dn,/dt depends only on n;. Following the previous equation, this function is proportional to A and corresponds to the adsorption isotherm of the water droplets on the globule surface. The adsorption isotherm was modeled in order to deduce a numerical value for A. na was calculated following the model of Frumkin and Fowler [40] : the adsorbed concentration depends upon the total number density of available
Fig. 8. Influence of the initial volume fraction of internal droplets on the kinetics of release: globule diameter = 3.6ujn, globule volume fraction = 10%, droplet diameter = 0.36|a.m, SDS concentration in the continuous phase = 2.4 10~2 Mol.l"1, Span 80 concentration in the oil phase = 2wt% (adapted from Ref [38]).
Stability of Concentrated Emulsions
625
sites for adsorption, n0, the adsorption energy, ua, and the lateral energy of interaction between the droplets, U|. n0 was deduced from simple geometrical considerations. From the average length of the surfactant tails, an estimation of the Van der Waals interactions can be obtained: ua and u(. The coalescence frequency is therefore the unique free parameter in the model and was determined from the best fit to the experimental curves. In Fig. 10 is plotted the kinetic evolution of n, at constant surfactant concentrations for two different globules diameters: 3.6 um and 14.5um. Using one and the same value of A, the theoretical points (continuous line) correctly fit the experimental data: A=6.10"3 min"1. From the obtained numerical value of A, it can be estimated that a droplet of 0.36 |im spends on average 3 hours on the globule surface before a coalescence event to occur. Measurements of the same type were performed under the same conditions but with oils of different molecular weight: for octane globules A was very small and could not be measured, while for hexadecane A=2.5 10"2 min"' [39]. Again, the influence of the spontaneous curvature may be invoked to explain the impact of the molecular chain length. Short alkanes like octane easily penetrate the surfactant brush and provide a negative spontaneous curvature to the surfactant monolayer, which tends to stabilize W/O/W films. Conversely, large alkanes like hexadecane are essentially excluded from the short surfactant brush and the spontaneous curvature becomes less negative.
Fig. 9. Rate of coalescence as a function of the number density of internal droplets in the globules. Globule diameter = 3,6 p,m, droplet diameter =0.36 |im, globule volume fraction = 10%, SDS concentration = 2.4 10"2 Mol.l"1, Span 80 concentration in the oil phase = 2wt%. The dashed line is a guide to the eyes (adapted from Ref [38])
626
V. Schmitt, S. Arditty andF. Leal-Calderon
Fig. 10. Number density of internal droplets for two different globule diameters. Droplet diameter =0.36 |j.m, globule volume fraction = 10%, SDS concentration = 2.4 10" Mol.l" , Span 80 concentration in the oil phase = 2wt%. The continuous lines are theoretical predictions (adapted from Ref [38]).
The previous experiments were performed at room temperature. Within the same conditions the temperature was varied between 20 and 60°C for double emulsions comprised of dodecane globules. In Fig. 11 is plotted the evolution of
Fig. 11. Frequency of coalescence as a function of 1/kT. Globule diameter = 3,6 urn, droplet diameter =0.36 urn, globule volume fraction = 10%, SDS concentration = 2.4 10"2 Mol.l"', Span 80 concentration in the oil phase = 2wt%. Initial internal droplet volume fraction = 20% (adapted from Ref [38]).
Stability of Concentrated Emulsions
627
A as a function of 1/kT in a semi-log plot for the same system. The experimental points roughly follow an Arrhenius law and from the best fit to the data, the activation energy is obtained: Ea=30 kTr. From the intercept, it is possible to obtain the natural frequency which corresponds to the total number of holes generated per unit time in the film between the internal droplets and the globule surface: Ao =4.1010 min"'. Only a fraction of them will grow and ultimately produce a coalescence event while the other ones simply vanish. From the ratio between the measured values of A and Ao, it can be concluded that only one over 1013 holes generated in the film leads to a coalescence event. 3.2.3. Simple emulsions stabilized by solid particles (non-activated coalescence Arditty et al. describe a completely different scenario of coalescence in solid-stabilized emulsions and propose a precise determination of oo [41]. In the experiments described by Arditty et al., the coalescence rate is not an activated process, but is more probably controlled by the drainage of the continuous phase in between the emulsion droplets. It is well established that solid particles of colloidal size may be employed to kinetically stabilize emulsions. Solidstabilized emulsions can be obtained with a wide variety of solid organic or mineral powders. The surface of such particles is made partially hydrophobic in order to promote adsorption at the oil-water interface and generally the adsorption is totally irreversible. Emulsions can be prepared by manually shaking a mixture of oil and an aqueous dispersion of silica. When shaking is stopped, the oil drops coalesce to form drops of macroscopic size. The sequence in Fig. 12 reveals the typical evolution of a concentrated emulsion. The average droplet size increases at short times and rapidly saturates at some asymptotic value. Because in the following experiments the solid content is initially insufficient to fully cover the oil-water interfaces, emulsion droplets coalesce to a limited extent. Under the effect of coalescence, the total interfacial area between oil and water is progressively reduced thus increasing the degree of coverage by particles because they are irreversibly adsorbed. This results in the formation of dense solid interfacial films that ultimately inhibit coalescence and kinetically stabilize the emulsions. Fig. 13 shows the images of O/W emulsions obtained for different solid contents, mp, expressed as a mass of particles in the emulsion. It can be clearly seen that the final droplet size decreases upon increasing the total amount of solid particles. The volume fraction of the dispersed phase is quite elevated, being larger than 80%. Such emulsions can therefore be considered as biliquid
628
V. Schmitt, S. Arditty andF. Leal-Calderon
Fig. 12. Images of an o/w emulsion containing 90 wt% of oil and 24.4 mg of hydrophobically modified silica particles taken at different times since formation. The times are (a) 9s, (b) 21s, (c) 54s, (d) 141s and the scale bar = 7.5 mm (from Ref [41]).
foams. Surprisingly, these foams can be stored at rest for months without any structural evolution. The surface-averaged final diameter Dst can be perfectly controlled by adjusting the amount of particles. Because the solid particles are irreversibly adsorbed at the oil/water interface, the inverse average droplet diameter varies linearly with the amount of particles according the simple equation: 1
s m
f
— = ^ ^D DSf 6V d
(15)
where Vd is the volume of dispersed phase and sf the specific surface area, i.e. the droplet surface covered per unit gram of silica. Fig. 14 represents the evolution in time of the surface-averaged droplet diameter for different amounts of solid particles. The kinetic curves confirm the qualitative evolution previously described. The droplet growth is initially rapid but the coalescence rate progressively decreases until the average diameter reaches an asymptotic value. Fig. 15 shows the change in the droplet size distribution with time.
Stability of Concentrated Emulsions
629
Fig. 13. Images of o/w emulsions at long times containing 80 wt.% oil obtained for different masses hydrophobically modified silica particles, mp (given). Scale bar = 1.2 cm (from Ref [41]).
Fig. 14. Evolution in time of the surface-averaged drop diameter in o/w emulsions containing 90 wt% oil and different masses of silica particles (adapted from Ref [41]).
630
V. Schmitt, S. Arditty and F. Leal-Calderon
For all the plots reported in this figure, the uniformity U (Eq. (10)) is lower than 30%, a surprisingly low value considering that the distribution is achieved by coalescence. The change in the droplet concentration n as a function of time is deduced from Fig. 14 considering Eq. (5). The curves obtained were numerically differentiated with respect to time in order to obtain the characteristic frequency of coalescence through Eq. (6). In Fig. 16 is reported the evolution of co as a function of the degree of surface coverage defined as T=DS/Dst. As could be expected, the curves reveal a dramatic decrease of the coalescence frequency with x. As T approaches unity, the degree of surface coverage becomes sufficient to kinetically stabilize the droplets such that no further structural evolution is observed. In this limit, one expects that CO reaches very small, but non zero values, the stability being only of kinetic order. In Fig. 17 is reported the evolution of co as a function of the average droplet diameter Ds, at constant degree of surface coverage x. It is interesting to note that co is a decreasing function of droplet diameter, at constant concentration of solid particles at the oil/water interface. The qualitative evolution can therefore be considered as a general feature of non-activatedcoalescence in biliquid foams. It is argued that the kinetics of the limited coalescence process is determined by the uncovered surface fraction 1-xand by the rate of thinning (drainage) of the films formed between the deformable droplets.
Fig. 15. Evolution of the droplet size distribution with time for an o/w emulsion containing 90 wt% oil and 24.4 mg of silica particles. The values of the uniformity index U and the average drop diameter Ds are given for different times since formation (Adapted from Ref.[41])
Stability of Concentrated Emulsions
631
Fig. 16. Variation of the coalescence frequency CO with i for O/W emulsions. The composition is the same as in Fig. 12. (adapted from Ref [41])
The homogeneous and monodisperse growth generated by limited coalescence is intrinsically different from the polydisperse evolution observed for surfactant-stabilized emulsions. As noticed by Whitesides et al [43], the mere fact that coalescence halts due to surface saturation does not provide an
Fig. 17. Variation of the coalescence frequency co with Ds at constant T for o/w emulsions stabilized by silica particles. The lines are only guides to the eyes (adapted from Ref [42]).
632
V. Schmitt, S. Arditty and F. Leal-Calderon
obvious explanation of the very narrow droplet size distributions that are frequently obtained in solid stabilized emulsions. The same authors propose a theoretical analysis for either diffusional or turbulence-driven droplet collisions. Assuming that the coalescence probability between two drops is simply proportional to their individual uncovered surface fraction 1-x, Monte Carlo simulations predict droplet size distributions that are in close accord with experimental results, i.e. and that are much narrower than those resulting from unlimited coalescence. One interesting conclusion from the numerical calculations is that the final narrow size distribution is insensitive to the agitation conditions and to the details of the initial droplet size distribution (uniformity, surface coverage) within fairly wide limits. 3.3. Gelation and homothetic contraction The following section is devoted to an again distinct scenario occurring in emulsions comprised of highly viscous drops. Coalescence involves two different steps. The first step consists of the nucleation of a small channel between two neighboring droplets. In the following, we shall term as xn the characteristic time separating two nucleation events. This first nucleation step is followed by a shape relaxation driven by surface tension, which causes two droplets to fuse into a unique one. The characteristic time for shape relaxation is governed by the competition between surface tension and viscous dissipation and is given by: T=r|D/y
(16)
where r| is the viscosity of the droplets, D is their characteristic diameter and y is their surface tension. When there is no energy barrier for coalescence, the droplets coalesce as soon as they collide under the effect of Brownian motion. In the limit where the characteristic shape relaxation time xr is shorter compared to the time xn separating two droplet collisions, it was found both theoretically and experimentally that the average droplet size scales with time t as t1/3. A very different scenario is expected in the limit where xr is much larger than xn. The coarsening is now limited by shape relaxation leading to very different structures and kinetics than in the previous case. This limit is frequently encountered in systems like emulsions of highly viscous substances (bitumen) or phase separations in binary mixtures of polymers. We now summarize the work of J. Philip et al. [44] who described the limit where x,»x n . They used model emulsions of highly viscous bitumen droplets which can be made suddenly unstable towards coalescence upon
Stability of Concentrated Emulsions
633
Fig. 18. Sequence showing the homothetic contraction of a bitumen-in-water emulsion (Adapted from Ref [44])
addition of a suitable chemical. Once the emulsion is made unstable, the droplets form a macroscopic gel made of an array of fused droplets. Then the gel continuously contracts with time in order to reduce its surface area. In order to study the gelation and the contraction phenomena, the emulsion were introduced in a rectangular cuvette and a destabilizing agent was added in the continuous phase. Initially, the system remains liquid-like, but after some time, the emulsion does not flow any more. At this stage, observation under microscope reveals that the droplets stick together and form a three-dimensional gel
Fig. 19. Evolution of the volume fraction in the bitumen contracting gel as a function of time for emulsions with different initial volume fractions (adapted from Ref [44])
634
V. Schmitt, S. Arditty and F. Leal-Calderon
network. Once this network is formed, the gel starts to contract by reducing its surface area. In this process water is expelled from the space filling network. The contraction remains remarkably homothetic meaning that it preserves the geometry of the container (Fig 19). Fig. 19 shows the evolution of the bitumen volume fraction in the contracting gel as a function of time for emulsions with different initial volume fractions, at room temperature (25°C). When the destabilizing agent is introduced in the emulsion, a delay Tg has to be expected before the gel starts to shrink. Tg corresponds to the time required for the droplets to form a continuously interconnected network that fills the whole volume. Once the contraction starts, two different regimes can clearly be distinguished. The rate of contraction at the initial stage is quite rapid and becomes much slower towards the final stages of contraction. The first regime is roughly linear, the slope being an increasing function of the initial droplet volume fraction. The contraction kinetics can be described using the so-called 'cylindrical model' for sintering [45-47]. The model considers a cubic array formed by intersecting cylinders that are made up of strings of particles. The initial cylinder radius corresponds to the average radius of the particles in the material. Although the choice of cubic array is an idealized approximation compared to the complicated microstructures formed in real situations, studies have shown that the geometry of the unit cell chosen (eg: octahedral, tetrahedron, inverse tetrahedron) have very little influence on the kinetics. To reduce their surface area, the cylinders tend to become shorter and thicker. In these calculations, it is assumed that the geometry of the cell is preserved. A brief description of the calculations is as follows. If '1' and 'a' are the length and radius of the cylinders respectively(Fig. 20), the energy dissipated in viscous flow (Ef) varies as:
BnnaVdn 2
•
'"
1 Idtj
(W)
The superscript dot indicates a derivative with respect to time. The energy change due to the reduction of surface area (E s ) is given by: Es=Y^dt
(18)
Stability of Concentrated Emulsions
635
Fig. 20. Cylindrical model for sintering. The gel is modeled as a cubic array made of intersecting cylinders with length I and radius a (adapted from Ref. [47]
where Sc is the surface area of a full cylinder. The energy balance requires the following condition: Ef + E s = 0
(19)
From equations (18) and (19), the rate of densification is deduced: (Y/Tllo) (1 / «M'/3 (t-to) = Jo" 2 dx/ (3 JI - 8 V2 x)1/3 x2/3 (20) Where x= a/1. For a cubic cell, x is related to the cylinder volume fraction as: <|>= 3 7T (a/1)2 - 8 ^ 2 (a/1)3
(21)
(|) corresponds to the measured volume fraction of bitumen inside the gel. t0 is the fictitious time at which x=0. In Eq. (20), (y/r)lo)(l / <\>0)in = K is a constant for a given initial volume fraction §0. Indeed §0 sets the initial cylinder height l0. When the ratio of cylinder radius to its height is equal to 1/2, the neighboring cylinders touch and the cell contains only closed pores. The corresponding theoretical density (volume fraction) of the sample would be 0.942. Therefore, the cylindrical model is not valid anymore for <)| values larger than 94.2 %. Fig. 21 shows the evolution of the gel volume fraction § as a function of reduced time K(t-t0). The solid line represents the theoretical curve obtained using Eq. (20). For (j) values between 0.2 and 0.8, the theoretical curve is roughly linear with a slope of 1 (see dashed line). Equivalently, within the same § range, the volume fraction should vary linearly with time with a slope equal to K. The previous experimental data where recalculated in order to be plotted in reduced
636
V. Schmitt, S. Arditty and F. Leal-Calderon
coordinates. For each initial volume fraction, K is deduced from the initial slope of the curve <))=f(t) (for (j) between roughly 0.2 and 0.6). All the data lie within a unique curve that is in reasonable agreement with the theoretical one. Philip et al explored the influence of viscosity on the rate of contraction by changing the temperature. Two different asphalts were used for that purpose. According to viscous sintering theory, the rate of densification K curve should vary as the inverse viscosity. Fig. 22 shows the evolution of K as a function of viscosity. It can be noticed that lowering the viscosity has the effect to increase the rate of contraction K. As expected, the observed contraction is controlled by the viscous flow of asphalt through the gel network. The experimentally observed slopes are in reasonable agreement with the expected value. In order to see whether this contraction phenomenon is a general one, various O/W emulsions using highly viscous oils were prepared [48]. As in the case of bitumen, similar contraction mechanism were observed when the emulsion was allowed to break by adding a suitable rupturing agent. From all the experiments performed with different types of oils and rupturing agents, it can be stated that the gel contract in a homothetic way when drop viscosity exceeds around 100 Pa.s. The sintering process may be of great technological importance since it allows transforming an initially liquid emulsion, into a dense and highly viscous material within a short period of time and at room temperature.
Fig. 21. Evolution of the gel volume fraction § as a function of the reduced time K(t-to). The solid line represents the theoretical curve obtained using Eq. (20) (adapted from Ref [44])
Stability of Concentrated Emulsions
637
Fig. 22. Evolution of K as a function of viscosity (adapted from Ref [44])
4. CONCLUSION AND PERSPECTIVES The recent experimental studies have certainly led to a better understanding of the coalescence phenomena in concentrated emulsions. Despite the complexity and variety of the destruction scenarios, different methods for measuring the coalescence frequency, co, have been proposed. It should be within the reach of future work to measure co for a large variety of systems in order to establish a comparative stability scale. This is a necessary step to determine the microscopic parameters that control the activation energy Ea and the natural hole nucleation frequency C0o- It is probable that numerous interfacial parameters are involved (surface tension, spontaneous curvature, Gibbs elasticity, surface forces) and differ from one system to the other, according the nature of the surfactants and of the dispersed phase. Only systematic measurements of co will allow going beyond empirics. Besides the numerous fundamental questions, it is also necessary to measure co for an evident practical reason: predicting the emulsion lifetime. This remains a serious challenge for anybody working in the field of emulsions because of the polydisperse and complex evolution of the droplet size distribution. The mean field approaches adopted to measure co are acceptable as long as the droplet polydispersity remains quite low (U<50%) but need to be revisited for very polydisperse systems.
638
V. Schmitt, S. Arditty andF. Leal-Calderon
REFERENCES [I] W. Ostwald, Z. Phys. Chem., 37 (1901) 385 [2] I.M. Lifshitz and V.V. Slyozov, Soviet Physics JETP, 35 (1959) 331 [3] I.M. Lifshitz and V.V. Slyozov, J. Phys. Chem. Solids, 19 (1961) 35 [4] A.S. Kabalnov, A.V. Pertzov and E.D. Shchukin, J. Colloid Interface Sci., 118 (1987) 590 [5] A.S. Kabalnov, K.N. Makarov, A.V. Pertzov and E.D. Shchukin, J. Colloid Interface Sci., 138(1990)98 [6] D.J. Durian, D.A. Weitz and D.J. Pine, Phys. Rev. A, 44 (1991) R7902 [7] C. Wagner, Z. Electrochem., 65 (1961) 581 [8] J. Bibette, D.C. Morse, T.A. Witten and D.A. Weitz, Phys. Rev. Lett., 69 (1992) 2439 [9] A. Hasmy, R. Paredes, O. Sonnneville-Aubrun, B. Cabane and R. Botet, Phys. Rev. Lett., 82(1999)3368 [10] R. Lemlich, Ind. Eng. Chem. Fundam, 17 (1978) 89 [II] M. Tokuyama and K. Kawasaki, Physica A, 123 (1984) 386 [12] Y. Enamoto and K. Kawasaki, Acta Metall, 35 (1987) 907 [13] T.W. Patzek, AIChE J., 39 (1993) 1697 [14] S.P. Marsh and M.E. Glicksman, Acta Mater., 44 (1996) 3761 [15] A.S. Kabalnov, Langmuir, 10 (1994) 680 [16] P. Taylor, Colloids Surfaces, 99 (1995) 175 [17] J. Weiss, N. Herrmann and D.J. McClements, Langmuir, 15 (1999) 6652 [18] W.I. Higuchi and J. Misra, J. Pharm. Sci, 51 (1962) 459 [19] A.J. Webster and M.E. Cates, Langmuir, 14 (1998) 2068 [20] A.J. Webster and M.E. Cates, Langmuir, 17 (2001) 595 [21] A.S. Kabalnov, P. A.V. and E.D. Shchukin, Colloid Surfaces, 24 (1987) 19 [22] L. Taisne, P. Walstra and B. Cabane, J. Colloid Interface Sci., 184 (1996) 378 [23] R.A. Arlauskas and J.G. Weers, Langmuir, 12 (1996) 1923 [24] J. Weiss, J.N. Coupland, D. Brathwaite and D.J. McClements, Colloids Surfaces A, 121 (1997) 53 [25] N. Hedin and I. Furo, Langmuir, 17 (2001) 4746 [26] L. Taisne and B. Cabane, Langmuir, 14 (1998) 4744 [27] V. Schmitt, C. Cattelet and F. Leal-Calderon, Langmuir, 20 (2004) 46 [28] O. Sonneville thesis "Biliquid Foams" Universite Paris VI, 1997 [29] A.J. De Vries, Reel. Trav. Chim. Pays-Bas, 77 (1958) 383 [30] A.S. Kabalnov and H. Wennerstrom, Langmuir, 12 (1996) 276 [31] K. Shinoda and H. Saito, J. Colloid Interface Sci., 26 (1968) 70 [32] W.D. Bancroft, J. Phys. Chem., 17 (1913) 501 [33] O. Sonneville-Aubrun, V. Bergeron, T. Gulik-Krzywicki, B. Jonsson, H. Wennerstrom, P. Lindner and B. Cabane, Langmuir, 16 (2000) 1566 [34] V. Bergeron, Langmuir, 13 (1997) 3474 [35] A.S. Kabalnov and J. Weers, Langmuir, 12 (1996) 1931 [36] V. Schmitt and F. Leal-Calderon, to be published in Europhys. Lett. (2004) [37] K. Pays thesis "Double emulsions: coalescence and compositional ripening" Bordeaux 1, 2000
Stability of Concentrated Emulsions
639
[38] K. Pays, J. Kahn, P. Pouligny, J. Bibette and F. Leal-Calderon, Phys. Rev. Lett., 87 (2001) 178304 [39] K. Pays, J. Kahn, B. Pouligny, J. Bibette and F. Leal-Calderon, Langmuir, 17 (2001) 7758 [40] R.H. Fowler and A. Guggenheim (eds),Statistical thermodynamics, University Press: Cambridge, 1939 [41] S. Arditty, C. Whitby, B.P. Binks, V. Schmitt and F. Leal-Calderon, Eur. Phys. J. E, 11 (2003) 273 [42] S. Arditty, C. Whitby, B.P. Binks, V. Schmitt and F. Leal-Calderon, Eur. Phys. J. E, 12 (2003) 355 [43] T.H. Whitesides and D.S. Ross, J. Colloid Interface Sci., 169 (1995) 48 [44] J. Philip, L. Bonakdar, P. Poulin, J. Bibette and F. Leal-Calderon, Phys. Rev. Lett., 84 (2000)2018 [45] G.W. Scherer, J. Am. Ceram. Soc, 60 (1977) 236 [46] G.W. Scherer, J. Am. Ceram. Soc, 60 (1977) 243 [47] G.W. Scherer (eds),Surface and Colloid Science, Plenum Press: New Jersey, 1987 [48] J. Philip, J.E. Poirier, J. Bibette and F. Leal-Calderon, Langmuir, 17 (2001) 3545
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 16
Emulsions stabilised by solid particles: the role of capillary pressure in the emulsion films P. Kruglyakov and A. Nushtayeva Chemical Department, Penza State University of Architecture and Building, 28, Titov str., Penza, 440028, Russia 1. INTRODUCTION It has been known for a long time that small solid particles can be effective emulsifiers like surfactant molecules. Emulsions stabilised by solid particles alone are often referred to as Pickering emulsions [1-4]. Like surfactant molecules, small solid particles can attach to the water-oil interface even though they are not usually amphiphilic. Solid particles (stabilisers of emulsions or as modifiers of different properties) are present in crude-oil and bitumen emulsions, paints, agricultural and pharmaceutical preparations, drilling fluids, food emulsions (water or fat crystals), in waste water, etc. A short historical essay of emulsions stabilised by solid particles and a list of substances used as solid stabilisers of emulsions could be found in the book by Clayton [5] and in the work of Yan and Masliyah also [6]. During the last years, nanoparticles (diameters between about 5 and 30 nm) are used as solid stabilisers. For example, it is possible to prepare very stable multiple emulsions of the w/o/w (water in oil in water) and o/w/o (oil in water in oil) type using two kinds of silica particles with a different degree of hydrophobicity without surfactants [7]. Basic factors affecting the stability of emulsions stabilised by solid particles are summarised in a number of reviews and monographs [6,8-12]. Nevertheless, the stabilisation mechanism is not fully clear. To obtain stable emulsions, the following conditions should be fulfilled: I. The particles should be selectively wetted by one of the phases (the contact angles should be 0° < 6 < 180°) and should accumulate at the surface as a result of adhesion interaction (at the present time, many researches use the term adsorption primarily for surfactant molecules [10,13,14]).
642
P. Kruglyakov and A. Nushtayeva
2. The particle concentration should be sufficient to form a closely packed layer on the surface of the emulsion droplets. 3. The size of the solid particles should be much smaller than that of the droplets. 4. The particles should coagulate to some extent but at the same time they should be characterised by a high initial monodispersity. The most stable oil in water (o/w) emulsions are formed for contact angles in the range of 60°-85° and the most stable water in oil (w/o) emulsions - for angles ranging between 95° and 120° [12,15-18]. Most authors suppose [12] that the main mechanism of the emulsion films rupture at the droplet coalescence is detachment (ejection) of the particles from the oil/water interface into the continuous or dispersed phase. Therefore, an estimate of the particle attachment energy (or the wetting energy of the particles by water or oil phase at its transfer from the interface into bulk phase) is an important task. Kruglyakov and Koretski [17,18] were the first to investigate the relationship between the stability of emulsions stabilised by dispersed glass and graphite and the work of wetting of the particles. They showed that the emulsion type and stability depends on the work of particle removing from the interface into both phases. The minimal works, necessary for stabilisation of o/w and w/o type of emulsions, were determined [17,18]. The calculations of the energy, required to remove the particle from the interface into one of the bulk phases, were later reproduced by many researchers [6,10,11,14,19] and used for discussion of different aspects of the emulsion stabilisation. 2. ADSORPTION OF SOLID PARTICLES AND ENERGY OF PARTICLE TRANSFER FROM THE OIL/WATER INTERFACE INTO A VOLUME PHASE The position of a small spherical particle of radius R at the interface between two liquids, for negligible gravity force in comparison with the surface forces, is determined by the equilibrium contact angle (Fig. 1). For the transfer of the particle into one of the bulk phases it is assumed that the particle consists of a single component, and the transfer of medium components can be neglected*. Let us consider a surface element with the area of nR1 (R is radius of the particle) at the liquid/liquid interface. For a system comprised of the particle and liquid/liquid interface with area nR2, the equilibrium surface free energy Gs of all the interfaces involved consists of three terms Gl + Gs + Gssw (here the sub " Expressions for the work of particle transfer including adsorption and transfer of the medium components were considered in [20].
Emulsions Stabilised by Solid Particles
643
Fig. 1. Position of a spherical solid particle at the water/oil curved interface. 9 is the contact angle, R is the particle radius, Rj is the droplet radius, x is height of the particle segment, / is the radius of the segment foundation (i.e. of three-phase contact line), ^and r/ are auxiliary angles, crvw, aso, crow are solid/water, solid/oil and oil/water interfacial tensions respectively.
scripts ow, so, sw refer to the oil/water, solid/oil and solid/water interfaces respectively): Gl = 2nRxa0S + {AnR2 - 2*Rx)crw + n(R2 - I2)aow.
(1)
Taking into account that R2 -I2 = (R-x)2 = (R cos df, one obtains
Gswo=7rR2[^aos
+4aws-(2x/R)aws+<Jowcos2
= nR2{Aows + 2CTOWCOS 0 - crowcos2 0),
0^ = (2)
where 0 is the contact angle that the oil/water interface acquires at the particle surface. The angle 0 is usually measured in the water phase. When the particle is completely immersed in the oil phase, then Gs0 = AnR2a0S + nR2cjm = nR2{4a0S + aow),
(3)
and when it is in the water phase G'w = 4nR2<7m + 7tR2aow = nR1 (4cr0,,, + am).
(4)
To move the particle from its equilibrium position at water/oil interface into the oil bulk, the wetting work (or Gibbs energy of adsorption) is
644
K =G°V -Gl
P. Kruglyakov and A. Nushtayeva
= /rf?Vwo(l + cos£) 2 ,
(5)
while for the transfer of the particle into the aqueous phase, the wetting work required is W: = G* - Gswo = 7rR2awo (1 - cos
ff)2.
(6)
The functions W£(Q) and Wos (B) are shown in Fig. 2 for R = 100 nm and o o w = 10 mNnT 1 . Eqs. (5) and (6) show [10,11] that for small particles (for example, for R = 100 nm, crow = 25 mN m~' and 9= 90°) the energy needed to transfer the particle from the interface into one of the volume phases is equal to approximately 5xlO4 kT. For particles with radius R = 0.3 urn the energy is equal to 2xlO6 kT [10]. For relatively large particles at the same conditions and at the oil-water density difference Ap = 1 kg dm"3, the gravity force becomes equal to the restoring ("flotation") force at the particle radius R=\ mm, and it is less by a factor of 103 than the attachment force at R = 0.1 mm. This means that the isolated solid particles are in a very deep energy minimum at the interface and therefore, at a large enough concentration they may form a close packed monolayer at the droplet interface [10].
Fig. 2. The dependence of the wetting work by water, W*v (1), and by oil phase, W* (2), on the contact angle 6*for the particle radius R= 100 nm and the interfacial tension aow = lOmN m1.
Emulsions Stabilised by Solid Particles
645
The ratio of the quantities given by Eqs. (5) and (6) expresses the hydrophilic-lipophilic balance of spherical solid particle (HLRQ) [12]:
^ 1 - cos 0 ) For the compact gel-like bilateral layers a different expression for HLRS is obtained [21] (see also [12]). Eqs. (5) and (6) are derived assuming equilibrium contact angles. Under practical conditions, the particles of solid emulsifiers are characterised by a contact angle hysteresis. For a meta-stable equilibrium, the advancing and/or receding angles depend on the direction in which the wetting perimeter moves. In this case, the derivation presented above is, strictly speaking, invalid because the surface energy cannot depend on the direction of the motion. However, it was shown in [22] that when the contact angle hysteresis exists, the wetting work can be calculated from the integration of an elementary work performed against the holding (flotation) force which acts along the perimeter of the particle as the particle is moved from the interface into the bulk phase. In the case of spherical particles, the expression for the wetting energy is derived in [22]: W' = nR2a{\- cos 9)2CgR,
(8)
where CgR is a constant, which weakly depends on the values of 9 and R. If the hysteresis contact angles (0m) are used to calculate the wetting work (Wg and W^\ then the metastable equilibrium is determined by the following correlation [23]: cos em = a°° ~ Usw ± ~^~ = cos ee ± -¥-, WO
OW
(9)
OW
where 0e is the equilibrium contact angle; cp is the resistance (friction) force related to the wetting (free-phase) line length, creating an additional barrier for the particle transfer. Taking this barrier into account, the particle work of wetting by the oil phase can be determined by Eq. (5) using the minimum receding angle of water (#,„) instead of the equilibrium angle 6e. The wetting work by the water phase is determined by Eq. (6). The calculations by Eqs. (5) and (6) show that the contact angle hysteresis increases both wetting works when compared to the equilibrium contact angle.
646
P. Kruglyakov and A. Nushtayeva
Eqs. (5) and (6) give the wetting works of desorption of the particles from the plane interface. The adsorption energy, taking into consideration the line tension (as), was derived [19]: AG
= TTR2CT0W ( 1 ± C O S Of
——(1±
cos
6).
i?usin 6
(10)
\
In the recent paper of Aveyard et al. [13] the effect of the liquid drop curvature and line tension on the wetting work was examined. The following expression was obtained: t,G = 2na0W
L
R2(l ± cos /)• cos 6 • 1 -
I
S
' °°S 7
- R2d{l - cos rj) +2nl<e,
°J )
(11)
where / is the radius of the three-phase contact line; R is the radius of the solid spherical particle; Rd is the radius of the emulsion drop; y = arcsin (l/R) and q = arcsin (l/Rd) are auxiliary angles (see Fig.l); 33 is the line tension. Detailed analysis in [13] has shown that the influence of the drop curvature on the adsorption energy is very small. For particles with a radius about 10% of the drop radius, the Gibbs energy of adsorption is rendered less negative by about 1%. Analysis of the line tension influence in the same paper has shown that if the expected value is as = +5-10~uN [13] then it becomes essential only for the particles with radii R < 15 nm at the contact angles 6< 30° and 6> 150°. In other papers [11, 24, 25], higher values of ae have been obtained. Investigation of the emulsion stability in the region of HLR ~ 1 [12], when both wetting works are large enough, shows that the mechanism of emulsion breakdown associated with pushing out of the particles can not be used for the explanation of the emulsion films instability. Hence, there are some other factors regulating the emulsion stability. One of them is capillary pressure. 3. CAPILLARY PRESSURE IN THE EMULSION FILMS 3.1. Calculation of the capillary pressure The first quantitative treatment of the effect of the particle wettability on stability of solid-stabilised films was performed on asymmetrical films containing solid particles that bridge the foam film [26]. In this paper, the heterogeneous defoaming mechanism is explained by the wettability effect (at 9> 90°). Analysis of the stability of an emulsion film containing one layer of solid spherical particles that bridge the film has been reported by Denkov and coworkers [27]. In stable emulsions with a high concentration of the particles, each droplet is covered by close-parked layer of the particles (often by
Emulsions Stabilised by Solid Particles
647
Fig. 3. The bilayer model of the emulsion film stabilised by two layers of the solid spherical particles: a) the scheme of formation of the film between the solid-stabilised o/w emulsion droplets; b) the equilibrium bilayer film (when Pa= 0); c) the thinning bilayer film (Pa> 0). Here h is the film thickness, R is the particle radius, 6 is the contact angle, a is the angle between the equatorial plane EE and the particle radius at the contact line, r is the radius of the interface curvature in the porous space, P{, Pi, Pb are the pressure in the film, in the drop and in the Plateau body, EE is equatorial plane of the particles, O\, O\, O\, O\ are the particles centres.
monolayer), that is confirmed by microphotographs and electron micrographs published in some papers [10,13,16]. At collision and deformation of such droplets (Fig. 3(a)) a bilayer emulsion film is formed. The simple realistic model of such a film (between droplets covered by the monolayers) is shown in Fig. 3. At equilibrium, it consists of two touching interfacial monolayers of the solid particles and continuous liquid phase filling in the porous space between the particles. We imply here that the film thickness h is the distance between the levels of local oil/water interfaces in the porous space of the opposite interfacial layers. As it is seen in Fig. 3(b), if the opposite interfacial layers touch, then the film thickness (denoted as the equilibrium thickness, heq) depends on the particle radius R, packing type of the interfacial layers and the contact angle #at selective wetting. Further thinning of the liquid part of the film creates oil/water menisci in the porous space between the particles of the interfacial layer (Fig. 3(c)). Hence, the capillary pressure P^in the film rises and it is equal to
648
P. Kruglyakov and A. Nushtayeva
Ptr={Pd-Pf)=C-
(12)
where / / a n d Pd are the pressures in the film and in the drop phase respectively; C is the curvature of the oil/water meniscus. In the presence of Pa the film thickness h equals the nearest distance between the menisci arising on the opposite sides of the film. In the absence of additional external pressure in gravitational field, the pressure in the continuous phase (i.e. in the emulsion film) at the emulsion layer height z is equal to Pf = P0-Apgz,
(13)
where Po is the pressure in the lowest layer of the continuous phase of the emulsion; A/? is the pressure difference of oil and water phases; g is the gravitational constant. The pressure in the droplet is Pd=PQ+^T-
(14)
Then the capillary pressure in the film is equal to P,=^-
+ ^Pgh.
(15)
If reduced external pressure is applied to the continuous phase (for example, in the case of emulsion sample placed on a porous plate) and the pressure difference is AP = Palm - P' (here Patm is atmospheric pressure, P' is pressure under a porous plate), then Pf = P0-Apgz-AP,
(16)
and pa =7^™L + Apgz + A P
(17)
R
d
In polyhedral emulsions the term 2aow/Rd becomes equal to aowl Rd. In the thin layer of the emulsion, the hydrostatic part Apgz is negligible. We can calculate the change of Pa during the bilayer solid-stabilised film thinning and the maximum (threshold) capillary pressure P^max, which leads to the film rupture [28,29]. A compact (hexagonal and/or cubic) packing type of
Emulsions Stabilised by Solid Particles
649
the monodisperse, spherical, selectively wetted solid particles in the interfacial monolayer is assumed. The deformed oil-water interface in the void space between the contacting spheres (Fig. 4) has a complicated shape and consists of a central meniscus and three wedge menisci [30] (in cubic packing there are four wedge menisci). The central meniscus at its lowest point is spherical and the meniscus curvature Cc is characterised with two equal radii (Cc - 2/rc). The wedge meniscus is characterised by a single radius of curvature (Cw = l/rw) because the second radius of the wedge meniscus is relatively large and may be neglected.
Fig. 4. Liquid/liquid meniscus in the porous space between three contacting spheres («=£ 0°): a) scheme of the meniscus surface; b) the central meniscus with radius rc. 1 is the three-phase contact line; 2 is the central meniscus and 2' is its projection to the cross section in plate (k); 3 is the wedge meniscus and 3' is its projection to plate (/); AA is the central axis of the porous space; bmm is the radius of a sphere inscribed into the porous space in the equatorial plate EE\ R is the particle radius; 6 is the contact angle; a is the angle of the slope of the particle radius at the contact line; p is the angle between the particle radius at the contact of the wedge meniscus and the axis BB.
650
P. Kruglyakov and A. Nushtayeva
It must be stressed that the wedge meniscus curvature is always equal to the curvature of the central meniscus (i.e. rc = 2rw). Hence the meniscus will adopt a configuration with the lowest total curvature and therefore, the lowest Pa ("minimum surface area" principle [30]). The total curvature magnitude is regulated by one of the menisci, which provides the lowest value. The central meniscus radius rc can be calculated using sphericalmeniscus-method (SM-method) [28,29]. Fig. 4(b) shows the geometry of the central meniscus. Here 6 is the contact angle measured in the internal phase of the film (an water-in-oil film is considered), a is an auxiliary angle associated with the film thickness; it is the angle between the equatorial plate of the particle and the particle radius at the contact line. The angle a may range from 0° to (90° - Of. Note that a may differ along the three-phase contact line, which is due to perturbations from neighbouring particles. In our calculation we have used the angle a, which is situated on a plate passing through the particle centre and the central axis on the porous space (see Fig. 4b). It follows, from Fig. 4(b), that the length of the segment K\N is equal to rc cos(« + 6). On the other hand K\N = B]A^ = (R + bmm) - R cos a, where bmm is the radius of a virtual sphere placed into the space between the particles in the equatorial plate (at a = 0°). For hexagonal packing of the particles bmm = R [l/sin60° - 1] = 0.155i?; for cubic packing: bmm = R [1/sin 45° - 1] = 0.4147?. Thus the central meniscus radius is
cos(6> + a) and the capillary pressure in the porous space and respectively in the bilayer (solid particle-stabilised film) determined by the SM-method is equal to p
1(J =
OW ^ rc
2g
Ow
COS 6>
R(l-cosa)
( + °0
,,m
+ bmm'
The film thickness h can be calculated by the formula (using the SM-method): h = RA
+
2sin« - 2(l + 6 m i n - c o s a ) ( l - s i n ( g + a))l cos(# + a)
(2Q)
where AR is the distance between the equatorial plates of the two opposing interfacial particle layers of the film. Here we also consider hexagonal and cubic packing type of the opposite interfacial layers. For the most compact, i.e. hexagonal packing, AR is the height of the pyramid O1O2O3O4 (see Fig. 3(b)) with
Emulsions Stabilised by Solid Particles
651
edge length 2R; it follows that A - 2^2/3 «1.633. For cubic packing of the interfacial layers A = 2. From Eq. (20) the equilibrium thickness of the film for an angle a = 90° <9 can be obtained: heq = R (A+ 2cos&).
(21)
Using Eq. (21), the radius of the spherical solid particles (or quasispherical aggregates) contained in the film can be determined by the equilibrium thickness of the film. The radius rw of the wedge meniscus can be derived from the theory of Mayer, Stowe, Princen [31,32] (MSP-method [30]) using rw=—, " L(rw)
(22)
where S is the area of central meniscus projection onto the cross-section of the capillary space, i.e. the area of figure EFGHDC in Fig. 4(a); L = L\ + Ls • cos 6, where Lt is the "liquid" part of the perimeter of the projection (EF, GH, DC) and Ls is the "solid" perimeter part (FG, HD, CE in Fig. 4(a)). Eq. (22) can be obtained from the work balance of an infinitesimal meniscus displacement dx [30] in a capillary of a uniform circular cross-section: Pa Sdx = amv(Ls cos <9 + Li)dx.
(23)
Using the MSP-method, we may calculate the dependence Pa{6, a, R). For this purpose the following formulae for rw, S and L are obtained (e.g., for hexagonal packing of the particles in the interfacial layer): rw = /?/sin(60°+ 0-G) - (\.\55R -rff)sin(60°+ ^)/sin(60° + jB-ff)
(24)
S = r,J(60° + p- (9)/z7360°- 0.5 rw2 sin<9 sin(60° + p~ <9)/sin(60° + P) + + 0.577i?2sin y9sin(60° + /J) - (1.155-7? - raf fi ^/360°
(25)
L=rw(6Oo + p-0)xnmo
(26)
+ (\.\55R-ra)cos0
/?;r/180o
where (3 is the angle between the particle radius at the contact line to the point of contact of the wedge meniscus with the particle (point F) and axis BB; ra=i?(1.155-cosar)/cosar is the radius of the sphere inscribed in the capillary space. In the MSP-method, the capillary pressure is equal to
652
P
P. Kruglyakov and A. Nushtayeva
(27)
Detailed analysis of Eqs. (24-26) has been reported in [29]. The difference in the values obtained by Eq. (19) and by the system of Eqs. (22), (24-27) is about 10-15%, which is considerable only for aoJR » 1 Nm~2. To determine Pa and h, we consider a water-in-oil film (as in o/w emulsions) and used the angle 6, which is measured inside the film. Hence for an oilin-water film (in the w/o emulsions), the angle (180°-6>) must be substituted instead of the angle 9 when using Eqs. (19-21, 24-27). Then, the combined solution of Eqs. (19) and (20) or Eqs. (27) and (20) gives the capillary pressure isotherm. Examples of such isotherms for aow = 10 mN m"1, R = 100 nm and contact angles 10° and 60° are shown in Fig. 5. For these values of 9, the capillary pressure in the aqueous solid-stabilised film increases during the film thinning (dPa- Idh < 0). This is equivalent to the condition of the stable mechanical equilibrium. The pressure increases to its maximum value Pa,mm corresponding to the critical thickness hcr of the film (at a = 0°). Further thinning would be accompanied by the spontaneous pressure decrease {dPaldh > 0) and the film would become unstable. Fig. 6 demonstrates the theoretical capillary pressure threshold in the bilayer solid-stabilised film for R = 100 nm and <JOW = 10 mN-nT1. The magnitude of Pa.max depends on the contact angle <9and on the packing type of the solid particles in the interfacial layer covering the droplet. In accordance with the minimum surface curvature principle, at hexagonal packing of the particles with the wetting angle 9 < 50° (o/w solid-stabilised emulsions) and 6>13O° (w/o emulsions), the maximum capillary pressure in the emulsion film is determined by the MSP-method and Eqs. (22, 24-27). For 9 in the range of 50°-130°, the SM-method and Eq. (19) determine the value of Pa [28,29]. At cubic packing of the particles, Pa is determined by the MSP-method for 6< 35° and 9 > 145°; and by the SM-method - for 9= 35-145°. The emulsion becomes unstable if the capillary pressure in the film is close to zero. This occurs at angle 6 close to 90° or at very low interfacial tension (Tow as it follows from Eq. (19). However, in order to estimate the emulsion stability, it has to be taken into account that cubic packing elements and some defects (for example, pore formed by 5 and more particles) can probably always be contained inside the close packed layers, even for monodisperse particles. Polydispersity of the particles rises the extent of the packing imperfections and consequently reduces the threshold pressure in the film. Aggregation (flocculation) of the solid particles also affects the structure
Emulsions Stabilised by Solid Particles
653
Fig. 5. The calculated capillary pressure isotherms Pdh) for 0= 10° (curve 1) and 6= 60° (curve 2) at i?=10 nm and
Fig. 6. The dependence of the maximum capillary pressure Pamux in the solid-stabilised emulsion film on the contact angle 9 for R = 100 nm and atm = 10 mN-nf1. Solid lines are for hexagonal packing of the particles in the interfacial layer. The dashed lines are for cubic packing type. 1,2 — the pressure in the aqueous emulsion films; 3, 4 — in the hydrocarbon emulsion films.
of the interfacial layer, and therefore Pa. During emulsification, the aggregation of the nanoparticles often precedes the adsorption stage of aggregates at the
654
P. Kruglyakov and A. Nushtayeva
droplet surface [33]. Eqs. (19) and (20) can be used to calculate Pa and h for emulsions stabilised by the particle aggregates (assuming a dense structure and a quasi-spherical shape of the aggregates). In this case the mean radius of the aggregates (Ragr) must be substituted for the single particle radius R in the equations. Besides, there is an additional factor, which can influence the real threshold pressure in the emulsion and model emulsion films stabilised by solid particles. It is the rapid local stretching of the film that leads simultaneously to increasing of the distance between the particles in the interfacial layer and to decreasing of the film thickness (respectively to Pa change). 3.1.1. Capillary pressure change at the solid-stabilised film stretching (the capillary component of the film elasticity) The emulsion film stabilised by the solid particles in the presence of a surfactant-modifier has the equilibrium and dynamic elasticity of surfactant adsorbed layers (Gibbs-Marangoni effect). At the same time, at the contact of the opposing interfacial particle layers, the film stretching causes a capillary pressure change in the film and an additional elasticity component appears: Yk=-^-
d\nSf
=^ S
dSf
f
,
(28)
J
where Yk is the capillary component of the solid-stabilised film elasticity [34] and Sf is the film surface area. Then at a distance u between the particles (Fig. 7), the capillary pressure is equal to (e.g., at the hexagonal packing)
Fig. 7. Scheme of the solid-stabilised film stretching by assuming constant liquid volume in the film and the uniform distance u increasing.
655
Emulsions Stabilised by Solid Particles
p
2aowcos(6 + a) a
U ( l . l 5 - c o s a ) +0.58M"
We examined the film stretching (with the area Sf and the thickness h) at constant film volume (V = h Sf= constant). It is assumed that an additional adsorption of the particles does not take place, i.e. the particle number (N) on the film surfaces does not change during the stretching. Then the capillary pressure becomes a function of the two variables (film thickness and distance between the particles). Each of them depends on the film area change dSf. The change of the function PJ^a, u) upon stretching is related to the film area change through the intermediate functions h(a), h{Sj) and u(Sj). For given 0, R and a, the function PJ^a, u) can be differentiated for 0°< a< 90° and all positive u. The total differential of the function is
*A%\
*, + f^| *.
(30)
The partial derivatives are equal to dP ] ^ — =-2aow\ da )u —— du
sin (9+a)\R(l. 15-cosa)+0.58w]+i?sinacos (0+a) [R(\A5-cosa)+0.58u]z
= —1.16cr ow -
>a
-.
,
(31)
(32)
[R(l.\5 - cosa) + 0.58M]
The function h(a) can be differentiated for a in the range from 0° to 90°: dh , D sina(l - sin(6» + a)) (1.15-cosa)(l-sin(0 + a ) ) ] .__, =ZK- cosa 1 (Jj) da film thickness decreases cos(6» + aproportionally ) costhe (9 + a) area increase (h = The with film
VISf). The area belonging to one spherical particle attached to the oil/water interface (at hexagonal close packing of non-aggregated particles) is equal to So =3.46i?2 (Fig. 7). Then the film surface area containing iV identical particles on each side at the distance u between the particles is
656
P. Kruglyakov and A. Nushtayeva
Sf = N 3.46 ( * + - ) •
(34)
The derivatives of the functions h{Sf) and u{Sj) for Sj > N 3.46 R2 are
- ^ =- A
(35)
^-(S^S,)-"2.
(36)
Substituting the partial derivatives of the intermediate functions into Eq. (30), we obtain the capillary pressure change in the bilayer solid-stabilised film at indefinitely small stretching: dPa _ dPa dSf da J
dh I dh 1 dS f da u
J
I
J
dPa) du du j dSf "
J
Fig. 8. The theoretical curves of the capillary component Yk of the film elasticity modulus depending on the film thickness h (related with the particle radius R). The contact angle 9 is equal to 0° (curve 1), 30° (2), 60° (3) and 80° (4).
Emulsions Stabilised by Solid Particles
657
This calculation gives the sign ("+" or "-") and the change rate of Pa at the stretch depending on the angle 6 and the film thickness. Fig. 8 presents the curves for the Yk(h) dependence at different contact angles. The curves have a maximum at hIR ~ 2.3-2.8, and the sign changes at hIR ~ 1.9-2.5, for 8= 0°-70°. If Yk > 0 corresponds to the initial stage of film thinning, the capillary pressure increases the overall film elasticity caused by the Gibbs-Marangoni effect. At Yk < 0, the calculation predicts that the elasticity modulus decreases due to the capillary pressure of the film. Most likely, the Pa decrease (in the case of the film micro-stretching by a local disturbance at hIR < 1.9-2.5) can explain the strong sensitivity of the "dry" films and high-concentrated emulsions (stabilised by solid particles) against local disturbances. The interfacial tension increase and/or the particle radius decrease lead to proportional growth of both the capillary pressure and the capillary component Yk. The magnitude of Yk does not depend on the film area. Comparing the Yk{h) curves in Fig. 8 and the capillary pressure isotherm PJJi) (for example at 9 = 60°, see Fig. 5), we find that the maximum elasticity corresponds to a capillary pressure (0.2-03)PlTmax, and the elasticity sign change region is (0.5-0.6)Pamax. Therefore, the solid-stabilised emulsion film sensitivity to external influence must reduce the threshold capillary pressure and, on the contrary, rise the critical thickness hcr. 3.2. Investigation of Pa for a macroscopic model of the interfacial layer of solid particles A series of were carried out to determine the capillary pressure as a function of the contact angle and the model film thickness. The macroscopic model of the interfacial layer is presented by a monolayer of glass spheres with radius i? = 3.15±0.1 mm forming a hexagonal close packing on a porous plate with a thickness of about 1.5 cm (Fig. 9(a)) and placed in quartz cuvette full water solution. The liquid phase of the model is the water solution that fills the voids between the spheres. The external water level in the cuvette is gradually adjusted to decrease, hence the thickness of the model film (i.e. of the liquid phase of the model) also decreases and the capillary pressure in the space between the spheres increased. The pressure is measured by the capillary rise of the solution in the void space between the spheres. The thickness of the model film in the range between h = R{\ + COS61) and hcr= R- rc, is measured using a cathetometer with an accuracy of 0.01 mm. The refraction by the glass sphere and the gravity effect are taken into account [28,29]. The maximum pressure is also determined in the space between three glass rods coming into contact (Fig. 9(b)) because the meniscus curvature at a ~ 0° between spheres is equal to that between rods of the same diameter. The mean radii of the rods in the different experiments are 1.15 ± 0.03 mm and 3.35 ± 0.05 mm. In the experiments with
658
P. Kruglyakov and A. Nushtayeva
Fig. 9. Macromodel of capillary space of the interfacial particle layer: a) monolayer of glass spheres with radius 3.15 mm on a glass porous plate with thickness 1.5 cm in quartz cuvette containing water solution; b) three contacting glass rods in a cylinder filled with water and oil phase: wa, oa, ow are interfaces of water/air, oil/air and oil/water respectively; H is the capillary rise height.
Fig. 10. The experimental data for dependence of the normalised maximum curvature Cmax-R of the meniscus on the contact angle 0 in the porous space of the macroscopic model. The curve is theoretical. Dark triangles — in experiments with spheres; the open circles are values obtained in experiments with the rods (1—4 - the water advancing angles 0W; others - the water receding angles (90).
Emulsions Stabilised by Solid Particles
659
Fig. 11. The experimental capillary pressure isotherm (triangles) obtained for the macroscopical model with spheres with a radius of 3.15 mm, contact angle 6W = 0° and the water/air interfacial tension aw = 32.8 mNm"' (0.5% solution of OP-10).
the rods placed into a cylinder filled with water solution (or with water and oil), Pa is measured at the interface of water solution/air and water solution/oil (octane, Diesel fuel) at water receding and advancing contact angles. To ensure the complete wetting of the spheres or the rods (&w = 0°) we use 0.5% solution of commercial, non-ionic surfactant (12-oxyethyleted octylphenol — OP-10). The various contact angles are achieved by hydrophobisation of the glass surface with solutions of cationic surfactants of various concentrations. The cationic surfactants are dodecylamine hydrochloride (DAHC) and cetylthreemethylammonium bromide (CTAB). The hysteresis contact angles are determined by the depth of immersion (in water) of a single glass sphere with radius R = 4.9 ±0.1 mm at the interface of oil/water or air/water. Moving of the sphere from the water phase through the interface gives the water receding angle 0w. Moving the sphere from the oil (air) phase across the interface yields the water advancing angle Oo [29]. The interfacial tension is measured by weighing of the platinum frame with and without the emulsion film [12]. Fig. 10 shows the dependence of the maximum normalised curvature of the meniscus on the contact angle CmaxR (0) and the collected data for the experiments with both spheres and rods. The contact angle values range between 0°-50° (receding of water), and they are equal to 50°-70° (advancing of water). Fig. 11 shows the capillary pressure isotherm PJh) for 0W = 0° (for the spheres). As the reported results show, the experimental values agree well with the prediction of the calculations by the suggested formulae. Some deviations of
660
P. Kruglyakov and A. Nushtayeva
the predicted values are due to the contact angle hysteresis that always occurs for incomplete wetting. 3.3. Experimental investigation of isolated solid-stabilised films Isolated films are a convenient model that allows for the examination of various properties of emulsions because they express many characteristics of liquid films separating drops in highly concentrated emulsions, especially the properties that are difficult or impossible to study in the emulsions themselves. We obtain and examine macroscopic (4.6 and 6.5 mm in diameter) vertical isolated emulsion films stabilised by particles of silica or aluminium stearate [35]. Three types of silica (SiO2) are used in this study: a) 41% sol Ludox-HS (L-l) with the particle diameter 15 nm, containing 0.168 gl~' Na2O and 2.04 gl~ ' NH3; b) Aerosil 200 (A-2) with the particle diameter 12 nm; c) Stober silica particles [36] with the particle diameter 540±90 nm (S-3). For hydrophobisation of the silica surface (necessary to stabilise the emulsions), we use a CTAB solution at the initial concentration [CTAB] = 10~6-10~J M. The contact angles that differ due to the surfactant adsorption are measured on a glass sphere surface modified together with a silica sample [34,35]. Fig. 12 shows the hysteresis angles measured in absence of electrolyte. At the presence of 0.1 M potassium chloride, the angles increase: 0W from 10°-30° (in the region of [CTAB] = 510"5-10"*) to 35°-37° due to the reduction of the surface charge density of the silica (and glass surface). Aluminium stearate forms at the interface due to a chemical reaction of an aqueous dispersion of A1(OH)3 and an oil solution of stearic acid [37]. The dispersion of A1(OH)3 (more precisely, of alkaline salts of aluminium) is obtained from 0.01% aluminium chloride solution by adding 0.1 M NaOH up to pH = 7.5. The degree of hydrophobicity of A1(OH)3 increases with the stearic acid concentration [HSt] in the oil phase. The hysteresis contact angles measured by the clasped drop method on the interfacial layer (transferred to solid substrate) [37] are shown in Fig. 13. The stable aqueous emulsion films are obtained from silica at the angles 0W = 15°-55° (and 0O = 75°-120°), at which the o/w emulsions are also stable against coalescence (Fig. 12(b)). At 0W = 55° and 0O = 120° both aqueous and hydrocarbon films (and both emulsion types) are obtained depending on the direction of the frame transfer. For aluminium stearate, the stable aqueous films and hence, the o/w emulsions, are formed at the angles 0W = 35°-45° and 0O = 35°-104°; the stable hydrocarbon films (and w/o emulsions) at 0W = 43°-46° and 0O = 56°-158° [37] (Fig. 13(b)). The region of instability of the w/o emulsions, due to very high hydrophobicity of the particles, is not reached in the experi ments with aluminium stearate. In the case of silica, the region of instability of both the o/w emulsions and the w/o emulsions is not reached.
Emulsions Stabilised by Solid Particles
661
Fig. 12. The hysteresis contact angles 8W, 0o without electrolyte (a) and lifetimeTof the emulsion with [KC1] = 0.1 M (b) depending on [CTAB] for 3% S-3 silica (solid line), 0.5% A-2 (broken line) and 0.05% L-l (dashed line). The oil (heptane) volume fraction is equal to 0.33 in the o/w emulsions and to 0.66 in the w/o emulsions.
The single solid particles formed aggregates after a surfactant-modifier adsorption. The large silica particles S-3 (with the radius ~270 nm) formed aggregates at 9W > 30° (0O > 95°). The dispersions of little silica particles (L-l and A-2) and the particles of A1(OH)3 always consisted of aggregates. At contact with the oil/water interface, in an experimental cell (Fig. 14), the aggregates spontaneously adsorbed to the interface, forming a frosted, "wrinkled" interfacial layer similar to the interfacial layer made at protein (albumin) adsorption in the microphotograph reported by Hatschek (in [5]). We form the isolated emulsion films by the way of displacement of the
662
P. Kruglyakov and A. Nushtayeva
Fig. 13. The contact angles (a) and the emulsion lifetimer(b) depending on the stearic acid concentration [HSt] for [A1(OH)3] = 0.01% (at evaluation in A1C13) and pH = 7.5. The oil (mixture of octane and CCU) volume fraction is equal to 0.5 in all emulsions prepared by adding a disperse phase to the continuous one [37]. platinum frame from one phase (for example, water) into the other phase (oil) through the interfacial layer. The formed films are opaque, frosted and noninterfering.
Using these films on the rectangular frame, we measure the oil/water interfacial tension. For the film on the round frame we investigate the film thinning under the gravitational field or applied reduced external pressure influence. 3.3.1. Spontaneous thinning of aqueous emulsion films stabilised by solid particles in the gravitational field The platinum frame construction used in this experiment is shown in Fig. 14. The internal electrode radius, including the meniscus, is equal to r\ = 0.28±0.05 mm; the external electrode radius - r2 = 2.25±0.15 mm. An important
Emulsions Stabilised by Solid Particles
663
Fig. 14. Conductive method for investigating the film thickness: a) scheme of the cell and the electrical chain; b) the platinum frame with the film. 1 - the frame from the platinum wire; 2 - crook of the twisted platinum wire; 3 — conductometer; 4, 5 are external and internal electrodes respectively.
element of the frame is a crook touching the lower aqueous phase that allows water flowing out of the film under the gravitational field. The film electroconduction (aey) change during the thinning is monitored (Fig. 15, emulsifier is silica) with conductometer OK-102/1 at [KC1] = 0.1 M in water. The electroconductivity of the water phase is sejp = 128.810 Scm (at 25°C). It is clear that at the moment of formation (the frame is displaced quickly - for 5-10 sec), the film consists of two interfacial particle layers and the water layer between them and the water layer begins thinning. This process is quicker for large particle S-3 aggregates (Ragr = 4-5 \xm, curve 1 in Fig. 15) and for widely polydisperse aggregates A-2 {Ragr about 10-20 |im, curve 2). The electroconduction Xj reaches equilibrium value for 3-10 min for S-3 and A-200. The equilibrium value of ae^ remains constant during its entire lifetime (sometimes above 60 min). Then a hole appears in the frozen film and its growth (i.e. the film rupture) could be observed for a minute. The films stabilised by relatively small aggregates of L-l silica (Ragr = 70110 nm, curves 3, 4) continuously drain over a few hours. A similar emulsion microfilm was obtained by Taylor et al [38] from asphaltens at high concentration. The adsorbed asphaltens aggregates form a "skin-like" interface of the film, and this microfilm thins very slowly: its thickness does not change appreciably during an hour. The slower thinning in the case of L-l silica can probably be explained by the presence of the aggregates within the liquid layer of the film
664
P. Kruglyakov and A. Nushtayeva
Fig. 1 5. Dependence of the isolated aqueous emulsion film electroconduction ce/on time rof spontaneous thinning in the gravitational field. The solid emulsifiers are 0.5% A-2 (1), 0.5% S-3 (2), 2% L-l (3) and 0.5% L-l (4) silica particles at [CTAB]=10"4 M, [KCI] = 0.1 M. Oil phase is decane.
because the water phase in this experimental cell is a turbid sol. In the case of A2 and S-3, the water phase under the interfacial layer is transparent after the aggregates settle to the cell bottom. The equilibrium film thickness is determined from the electroconduction value corresponding to the horizontal part of the curve as/(r) by the formula
K,=
f
I2
Vf
,
(38)
2TTXyd
vf vL
where nf = — =
vf — vf-vp
is the expansion ratio in the solid-stabilised film, i.e.
the ratio of the film volume VfXo the liquid part volume VL (Vp is the volume occupied by the particles), n^-is between 2.5 and 4.0, depending on the film thickness [35]; B is an experimental coefficient equal to 1.1-2.0, depending on the angle 9 [3 5]. Table 1 demonstrates the equilibrium thickness heq of the isolated aqueous films in oil obtained from silica particles.
Emulsions Stabilised by Solid Particles
665
For the case of A-2 and S-3 silica, the thickness heq is equal to 14-22 |im. For L-l silica the lowest experimental thickness of the film is 0.52-0.55 um. The theoretical heq value for Ragr =110 nm and 0W = 26° is equal to 0.42 |j,m and it is not reached in the experiments probably because of the observed high sensitivity of the film to accidental perturbations that are impossible to avoid at a large surface areas of the films and such long time of the thinning. For aluminium stearate: {[A1C13] = 0.01%, pH = 7.5, [KC1] = 0.1 M; [HSt]= 0.001% in octane + CC14 mixture (1:1)} - the equilibrium film thickness is equal to 35-40 (am in our experiments. Table 1 also presents the aggregates radii: the effective radius Re/r and the radius Ragr determined by one of the traditional methods (from turbidity or from rate of the turbid dispersion/water boundary displacement). The effective radius R^. is calculated from the equilibrium film thickness using the equation
Rfagr
^
.
(39) V
2(l + cos0 w )
'
Here we assume a combined packing type of the aggregates in the film: hexagonal packing inside the interfacial layer, but cubic packing between the opposite interfacial layers [35]. Table 1 Equilibrium thickness hea of the isolated aqueous films in oil stabilised by silica particles aggregates with effective radius ReJ [CTAB], SiO2 type Angle Oy,, ° moll1 [KCI1=O.1M io-4 L-l (0.5%) 26 A-2 0.5%
S-3 3%
hCq, ^im
Ragr, urn
0.55-0.58*
0.11
5 10 6
35
15.8
4.4
-
5
5-10"
35
21.9
6.1
-
Iff4
40
15.9
4.3
-
5-Iff4
55
15.5
4.6
-
5-10'5
35
20.2
5.6
4.4
35
13.9
3.9
3.3
2-10-"
42
21.3
7.0
5.6
5-KT4
53
14.7
4.6
4.3
io-
4
The smallest experimental thickness of the film.
666
P. Kruglyakov and A. Nushtayeva
For S-3 silica, the mean radius Ragr = 4.3+1.2 (am is similar to the mean effective radius Re/gr= 5.3+1.3 u.m calculated by Eq. (39) considering the combined packing type (while Re/gr= 7.4±2.7 \±m - for completely hexagonal packing). 3.3.2. Capillary pressure isotherm for the aqueous emulsion films To investigate the film thinning due to external reduced pressure, we use a cell made of a glass porous plate with thickness near 1.5 mm and the pore diameter 16 jam (it is shown in Fig. 16). The reduced pressure in air space communicated with the porous cell is due to very slow flowing of water (drop by drop) from the funnel (number 2 in Fig. 16). The pressure difference AP (i.e. the reduced pressure in the porous cell in comparison with the atmospheric) is measured by a U-shaped water gauge manometer. The pore diameter of the cell allowed to maintain AP below 5xlO3 Pa. The capillary pressure in the film is considered to be equal to AP. At the same time, the film electric conduction is measured. The electrode radii including the menisci are r\ = 0.28±0.05 mm and r2 = 3.2+0.2 mm. Fig. 17 shows the capillary pressure isotherms PJJi) (the theoretical curves and the experimental values) in isolated aqueous emulsion films stabilised by aluminium stearate (a), and S-3 silica particles (b). The theoretical curves are drawn for entirely hexagonal (dotted lines) and combined packing (continuous lines) of the aggregates using Eqs. (19) and (20). For aluminium stearate {[A1C13] = 0.01%, pH = 7.5, [KC1] = 0.1M; [HSt]= 0.001% in octane + CCL, mixture (1:1)} the effective radius is Rfgr = 9.4±1.5 urn, the angle 9W = 43° and the interfacial tension aow(p) = 17.2 mN m~' (&ow(P) is tension of the oil/water interface containing adsorbed solid particles) are used to draw the theoretical curves PJJi). The density of the oil phase is equal to 1.14 g cm"3 (mixture of octane and carbon tetrachloride), so it is heavier than the water phase. The cell is lowered through the oil/water interface for the formation of the film (as it is shown in the inset of Fig. 17(a)). For S-3 silica particles {[SiO2] = 3%, [CTAB] = 2x10^ M, [KC1] = 0.1 M; decane} - the curves in Fig. 16(b) are drawn for values of Ragr = 5.6±0.1 urn, the angle 0W = 42° and the tension <7ow(p) = 13.8 mN-irT1. The water phase is heavier than the oil. At the AP increasing up to 5 kPa, the film lifetime decreases from 30-60 min to 1-2 min. However the film ruptures already at AP that is less than the calculated maximum magnitude (Pa,max= 6.5 kPa for aluminium stearate and PCTmax = 9 kPa for silica). This fact can be explained by the high sensitivity of the solid-stabilised films to mechanical perturbations due to Yk, the large surface area of the film and by incomplete wetting of the cell by water.
Emulsions Stabilised by Solid Particles
667
Fig. 16. Scheme explaining the investigation of the film thinning influenced by applied external reduced pressure. 1 - glass porous cell; 2 - dividial funnel; 3 - conductometer; 4 - watergange manometer.
Fig. 17. The capillary pressure isotherms Po(h) for the isolated aqueous emulsion films stabilised by solid particles: a) aluminium stearate - system {[A1C13] = 0.01%, pH = 7.5, [KC1] = 0.1 M; [HSt]= 0.001%, octane + CCL, (1:1)}; b) silica S-3 - system {[SiO2] = 3%, [CTAB] = 2-10 4 M, [KC1] = 0.1 M; decane}. The solid line is the theoretical curve drawn in consideration of the combined packing of the aggregates in the film. The dashed line is the theoretical curve for completely hexagonal packing. The circles are experimental points.
The surfactants used for the particle modification simultaneously lead to hydrophobisation of the cell.
668
P. Kruglyakov and A. Nushtayeva
It is possible that microscopic films would be more stable at high capillary pressures. However techniques for investigation of such opaque microfilms containing solid particles have not yet been developed. The arrangement of the experimental values for PJ^h) in Fig. 17 allows to conclude that at increasing pressure the aggregates packing type becomes denser and a transition, from film thickness corresponding to cubic packing of the opposite interfacial layers to a thickness corresponding to hexagonal packing, occurs. This transition may give rise to film rupture. 3.3.3. Oil/water interfacial tension the presence of the adsorbed solid particles Values of the interfacial tension aow(p) of the system containing particles SiO2 or A1(OH)3 and the tension aow without particles in the system are shown in Table 2 for temperature t = 25°C [35]. Here aow{p) is the tension of the interface that has the adsorbed particle layer. The tension is determined by the method of platinum frame weighing or by the drop volume method (in the case of L-1 silica). For L-1 silica particles forming a stable turbid sol, the tension aow is measured at the oil/CTAB solution interface. In the other cases aow is measured at the oil/water interfaces, which are in equilibrium with the particle interfacial Table 2 The oil/water interfacial tension crtm{p) at the presence of the silica or aluminium stearate particles and the tension anw without the particles (/ = 25°C) Emulsifier
[CTAB], rnolf1
L-1
10-6
27.4
34|0
2%
10-5
20.7
23 ft decane
4
uOW(p), mN-m
10.2
lJ.O
A-2
5-10
6
21.3
42)5
0.5%
510" 5
18.9
42.7 > octane
5-10"4
21.2
40.1 J
5-10"5
23.4
4510, f~ heptane 41.5 J
10-
S-3
16.6
3% 2-10^
13.8
5-10^
16.7
I decane 31.9-J
17.2
22.1 (octane + CCU)
A1(OH)3
[HSt], %
[A1C13] = O.O1%
0.001
[A1C13] = 0.02%
0.1
6.9
17.0 (Diesel fuel)
Emulsions Stabilised by Solid Particles
669
layer (the liquid phases are taken from the cell with the interfacial layer). Hence the water phase (for silica) and oil phase (for A1(OH)3) contains a relatively small amount of surfactant residual after the adsorption. Besides, in the case of A-2 and S-3 silica the aqueous solution, after the surfactant adsorption is substituted for distilled water. As seen in Table 2, in presence of particles the interfacial tension decreases approximately twice. The reason for the interfacial tension decrease by the particles adsorption is not fully clear. Levine et al [10] analysed the change of the Gibbs energy after the particles adsorption. They considered single particles only. From the Gibbs energy analysis follows that the ratio crow{p}/<7ow must depend on the contact angle 6 [10]. Our experimental result does not confirm this conclusion. We consider that the aggregates can reduce the interfacial tension more strongly than the single particles. Also the interaction forces between the adsorbed particles (Van der Waals, electrostatic etc.) have an effect on the interfacial tension. We used measured values of crow(P) in researching the P<j in the emulsion films. 3.4. Experimental investigation of the lifetime vs. Pa dependence for solidstabilised emulsions The emulsions for these studies are prepared from modified silica particles by adding small portions of the dispersed phase to the continuous phase (1:1 oil/water volume ratio) during shaking of the mixture in a shaker. All investigations are carried out 24 hours after the emulsion preparation. This period is enough to establish equilibrium polydispersity since the Ostwald ripening in solid-stabilised emulsions proceeds most intensively during the first several hours [39]. The stability (lifetimes) of the emulsion thin layers on a porous plate are measured. In the thin layer, the experimental time is sufficient for establishing of the equilibrium pressure and the conditions for rupture correspond to the ones for isolated emulsion films. A ceramic porous plate with the pore diameters 1.4 um is used (Fig. 18). Small amount of the emulsion (volume -0.02-0.03 ml) is placed on the plate. The small glass plates (number 3 in Fig. 18) regulated the thickness of the emulsion sample layer (0.6 mm). The space under the porous plate is filled with buffer into which a defined under pressure is maintained. The pressure difference AP is measured by a water gauge manometer (up to AP = 7 kPa) with an accuracy of ±0.05 kPa. For higher values (up to AP = 60 kPa) we used a vacuum-meter with an accuracy of ±2.5 kPa. To decrease evaporation of oil, the emulsion sample is covered with an upturned beaker. We observed the emulsion drops through the bottom of the beaker on dark background (black paper) in reflected light with a microscope at magnification 8.1-77.4. The air space inside the glass is not fully sealed because it is necessary to hold the constant
670
P. Kruglyakov and A. Nushtayeva
Fig. 18. Scheme of the cell for investigation of the emulsion thin layer stability at the high capillary pressure. 1 - ceramic porous plate (filter); 2 — covering glass plate; 3 - plates for regulation of the emulsion layer thickness; 4 - glass turned upside-down; 5 - microscope; 6,7 - electrodes; 8 - vacuum-meter (or watergauge manometer).
Fig. 19. Dependence of the emulsion lifetime ron the applied pressure difference AP: a) For the different aggregates sizes. There is [CTAB]=10~4 M. 1 — the emulsion system {[SiO2] = 3% (S-3), [KC1] = 0.1 M; heptane} with Ragr = 4.5 |0.m; 2 -{[SiO 2 ] = 3% (S-3); heptane} with radius R = 0.27 |im; 3 -{[SiO 2 ] = 3% (L-l); octane} with Ragr=50 nm. b) For different the continuous phase receding contact angle 8. The emulsifier is S-3 silica at [KCl]=0.1 M. 1 - o/w emulsion {[CTAB] = 5-10 5 M; octane}, 0W = 35°; 2 - o/w emulsion {[CTAB] = 2-10^ M; octane}, 6W = 43°; 3 - o/w emulsion {[CTAB] = 5-10"4 M; decane}, $, = 55°; 4 - w/o emulsion {[CTAB] = 5 10 3 M; decane + CC14 (1:1)}, 9n = 115°. The pointers show the threshold capillary pressure calculated for cubic packing of the aggregates with i?ugr=4-5 |am in radius.
Emulsions Stabilised by Solid Particles
671
(atmospheric) pressure above the porous plate. The pressure Pa in the emulsion layer is calculated by Eq. (17). The lifetime of the emulsion thin layer vs. pressure difference r{AP) is compared with the lifetime r0 of the same sample on the porous plate in the absence of AP. The time r0 depends on the hydrocarbon volatility. For a layer thickness of 0.6 mm and emulsion area -0.25-0.5 cm2 r0 increases in the direction: heptane (1 h) —» octane (2-4 h) —> decane (3-4 h) for the o/w emulsions. The w/o emulsions visibly broke more rapidly (r0 = 10-30 min). Fig. 19 shows typical curves for dependence z(AP) in emulsions. Note that the emulsion layers life-times decrease sharply even at very little pressure difference (1-2 kPa) although the calculated threshold capillary pressures in the films are equal to 100-300 kPa for some systems. This fact is related to the velocity of the continuous phase flowing process, reaching lower values of the films equilibrium thickness and rapid structural rearrangement of the films during transformation of the emulsion to a polyhedral one. At this time, at maximum possible pressure difference at which the dispersion drops still have not penetrated through the porous plate (up to 20-90 kPa in dependence on the interfacial tension), the full rapid ("avalanche-like", typical for foam at critical capillary pressure [40]) destruction of the total emulsion volume is not observed in our experiments. In Fig. 19(a) are shown the results for emulsions stabilised by S-3 silica aggregates {[SiO2] = 3%; [CTAB] = 10"^ M, [KC1] = 0.1 M; heptane}, by non-aggregated S-3 particle (the same system without KC1) and by L-l silica aggregates {[SiO2] = 2%, [CTAB] = 10"4 M, [KC1] = 0.1 M; octane}. The radius of the aggregates (or particles) decreases from curves 1 to 3. The emulsion layer is destroyed slower at the high capillary pressure if the radius of the aggregates is smaller. It is interesting to compare curves 1 and 2 - the emulsions with aggregated and non-aggregated particles. Even in absence of AP the lifetimes of these emulsions are different. The first emulsion is broken completely for 70 min, but the second emulsion - for more than 3 hours. The lifetimes of both emulsions decrease with increasing the pressure difference and reach 7-10 min (this lifetimes are equal to the establishment-time x' of the equilibrium pressure on the porous plate surface) at AP ~ 6 kPa, when the particles are aggregated and at AP~ 13.5 kPa, when the emulsion is formed by the non-aggregated particles. Taking into consideration the drop mean diameter, the corresponding values of the capillary pressure are 8 kPa (for aggregates) and 15.5 kPa (for non-aggregated particles). In the first case (Rag, = 4.5 [xm, aow{p) = 16.6 mN-m"1, 6W = 35°) this pressure is 0.5 times the calculated value Pamax =17 kPa (for close cubic packing of the aggregates). In the second case (Rp = 0.27 |im and 6W = 28°) the experimental pressure is ten times less than Pa<mca = 300 kPa, calculated for cubic packing. One of the reasons for the critical pressure
672
P. Kruglyakov and A. Nushtayeva
decrease in the experiments can be the presence packing defects. Besides, the non-aggregated (and hence charged) particles do not form a compact packing at the drop surface and a certain distance remains between the particles due to electrostatic repulsion. It leads to further decrease of the critical capillary pressure in the film. Hence these experiments have shown that the emulsion drops covered with compact layers (or poly-layers) of the non-aggregated particles are more stable relative to the applied AP than those covered with fragile layers of aggregates. Fig. 19(b) also shows the influence of Po-on the thin layer stability for the emulsions stabilised by silica S-3 with various surface hydrophobicity (in presence of KC1). The mean radii of the drops are equal to 40-100 um in those emulsions. The hysteresis contact angle value increases gradually from curve 1 to curve 4. Pointers show the magnitudes of the threshold capillary pressure calculated for cubic packing of the particle aggregates. The mean radius of the silica particle aggregates dispersed in water (curves 1-3) or in oil (curve 4) is equal to Ragr= 4-5 um. As it is seen in Fig. 19b the pressure corresponding to emulsion life-time of 10-15 min (needed to establish the equilibrium pressure on the porous plate surface) has been in the range of 0.3 (0W = 35°) to 0.85 {6W = 43°), only a part of the calculated one. Beside the effect of the packing defects, as mentioned above, another factor reducing the threshold capillary pressure in the film can be the "fragility" of the films due to the capillary contribution to the film elasticity that decreases sharply near the critical value film thickness. Fig. 20 shows the lifetime dependence of the emulsion thin layer on the quasi-equilibrium contact angle 9e = (0w +80)/2 of the solid particles at certain
Fig. 20. The lifetime of the o/w emulsions formed by octane and 2% water dispersion of Aerosil ([CTAB] = 5-KT6-2-1(T4 M) depended on the quasi-equilibrium contact angle ft. (without electrolyte) for some AP values. The layer thickness is equal to 0.6 mm.
Emulsions Stabilised by Solid Particles
673
applied pressure difference. As the capillary pressure increases, the emulsions of octane in aqueous 2% dispersion of modified A-2 silica (without electrolyte), become more stable relative to coalescence at intermediate contact angles. For example at the quasi-equilibrium angle 0e in the range of 55-70° (or at the hysteresis angles 6W= 20-40° and 0o= 85-100°). These results agree with data for emulsions formed from aqueous 14% dispersion of glass powder modified by octadecylamine. The stability of those emulsions in a centrifugal field has a maximum for angles 6e~ 60-80° [41]. Theoretically, the solid-stabilised emulsion instability in the region of small contact angles (0 -> 0° for o/w emulsions) or very large ones {6 —> 180° for w/o emulsions) is explained by the low work value of particle wetting by the continuous phase. As the contact angle approaches 90°, and hence the wetting work (by continuous phase) increases, the emulsion stability increases accordingly. The optimal wetting works correspond to 6= 90°. The other factor determining the emulsion stability is the threshold capillary pressure in the emulsion film and it decreases with the contact angle change from 0° to 90° (or from 180° to 90°). At 6= 90°, the pressure Pamax is equal to zero and the emulsion drops become unstable against coalescence. Therefore the theory and experiments confirm that the most stable (against coalescence) emulsions, stabilised by the solid particles, can be obtained at those contact angles at which the values of the particle wetting work and the threshold capillary pressure in the film are large enough. 4. PHASE INVERSION IN EMULSIONS STABILISED BY SOLID PARTICLES The phase inversion in emulsions can be induced by the variation of hydrophile-lipophile balance of the solid particles (transition inversion). A typical picture of the change in the stability of a solid-stabilised emulsion with increasing hydrophobicity of glass and graphite particles, according to our research data [17,18,41] is shown in Fig. 21. It means that o/w emulsions become sufficiently stable (the lifetime exceeds twenty-four hours) when the minimum specific work of the particle wetting by water is equal to
^L = aow(l-cosewf^5-\0
mJm"2.
TTR
In this case the wetting work by the oil is very large. This case corresponds to water receding angles about 17-40° and water advancing angles about 53-62°, for non-polar hydrocarbons (heptane, toluene, benzene and others) with the oil/water interfacial tension about 27-47 mN m~' [12]. For octanol and unde
674
P. Kruglyakov and A. Nushtayeva
Fig. 21. Typical pattern for the dependence of type and stability of emulsions on HLRS . 0
canol as oil phases with low oil/water interfacial tension 6.5-7.5 mN-irf1, the contact angles are equal to 6W = 42-50° and 60 = 90-105° [12]. However the magnitude of the energy barrier associated with the liquid film collapse is more significant. At contact angles below 90° the derivative dPaldh < 0 and the capillary pressure (analogous to the disjoining pressure in the thin films stabilised by surfactants) prevent the film from thinning and collapsing. Hydrocarbon films with 6 < 90° and dPa/dh > 0 thin spontaneously down to collapse. On the contrary for aqueous films with 0< 90° and dPjdh > 0 are stable. The existing of o/w emulsions at HLR^< 1 (above 0.98) and of w/o emulsions at HLR^> 1 (below 1,12) [12] is explained by the hysteresis of the contact angle. A tendency for sharp change in the emulsion stability is observed at examination of the dependence on the contact angle (i.e. onHLR^) for the high capillary pressure (see Fig. 20). The stability decreases sharply at 0e < 50° and at 6e —> 90° when the energy of the particle desorption from the interface is large. Therefore in the case of hydrophobicity change the transition phase inversion is due to decreasing the capillary pressure in the emulsion films down to zero at HLRg = 1 and to change of the derivative sign dPaldh. Some aspects concerning the analogy between solid-stabilised and surfactant-stabilised emulsions and their relation to the hydrophilic-lipophilic balance change have been discussed in our paper [42]. 5. CONCLUSIONS 1) The stability of solid-stabilised emulsions is determined by the particle wetting work on the one hand and by the capillary pressure in the emulsion films on the other hand. Expression for P o in the film stabilised by spherical
Emulsions Stabilised by Solid Particles
2)
3)
4)
5)
675
particles (or the quasi-spherical aggregates) as a function of the interfacial tension, the contact angle and the film thickness has been obtained. The experiments with close packed glass spheres (macroscopic model of the adsorbed particles interfacial layer) confirm the calculated values of P^,max(0) and PJ.h). The capillary component Yk of the solid-stabilised film elasticity has been calculated. Theoretical analysis showed that sign and rate of Pa change with the film stretching depend strongly on the angle 6 and ratio hIR. The experiments show that solid-stabilised emulsions and the isolated macroscopic emulsion films breakdown at lower pressure than the calculated Pa,max which could be explained by defects in the interfacial layer of close packed particles (aggregates), different size or/and shape of the aggregates and the capillary component of the elasticity. Hence, the experimental critical pressure in the films is more often equal to (0.2-0.6)-Paiinax (calculated for cubic packing of the aggregates). The transition phase inversion in the solid-stabilised emulsions is due to the change of derivative dPaldh sign at HLR^ ~ 1. REFERENCES
[1] W. Ramsden, Proc. Roy Soc. (London), 72 (1903) 156. [I] S.U. Pickering, J. Chem. Soc, 91 (1907) 2001. [3] T.R. Briggs, Ind. Eng. Chem., 13 (1921) 1008. [4] P. Finkle, H.D. Drapper, J.H. Hildebrand, J. Am. Chem. Soc, 45 (1923) 2780. [5] W. Clayton, The theory of emulsions and their technical treatment. Fourth edition, London, 1943. [6] N. Yan, J. Masliyah, Colloids and Surfaces, A.: Physicochemical and Engineering Aspects, 96 (1995) 229. [7] B.P. Binks, A.K.F. Dyab, P.D. Fletcher, Proc. Third World Congress on Emulsions, Lion, France (2002) Theme 1-JVsl (CD-Rom), 10 p. [8] P.A. Rehbinder, K.A. Pospelova, Vstupitel'naya stat'ya v knige Claytona Emulsii, Moscow, lnostrannaya literatura, 1950, p. 11 (Russian). [9] Th.F. Tadros, B. Vincent, Encyclopedia of Emulsion Technology. P. Becher (Ed), New York, Marcel Dekker, 1983, v. 1, p. 129. [10] S. Levine, B.D. Bowen, S. Partridge, Colloids and Surfaces, 38 (1989) 325. II1] R. Aveyard, J. Clint, J. Chem. Soc, Faraday Trans, 91 (1995) 2581. [12] P.M. Kruglyakov. Hydrophile-lipophile balance of Surfactant and Solid Particles, Amsterdam, Elsevier, 2000. [13] R. Aveyard, J.H. Clint, T.S. Horozov, Phys. Chem. Chem. Phys., 5 (2003) 2398. [14] J.H. Clint and S. Taylor, Colloid Surf., 65 (1992) 69. [15] J.H. Schulman, J. Leja, Trans. Faraday Soc, 50 (1954) 598. [16] A.F. Koretski, A.B. Taubman, Abh. deutsch Akad. Wiss. Berlin, Kl. Chemie, Geol., Biol., 68(1966)576. [17] A.F. Koretski, P.M. Kruglyakov, Izv. SOAN SSSR, Ser. khim. nauk, M> 2 (1971) 139. [18] P.M. Kruglyakov, A.F. Koretski, Izv. SOAN SSSR, Ser. khim. nauk, Xs 9 (1971) 16.
676
P. Kruglyakov and A. Nushtayeva
[19] R. Aveyard, J.H. Clint, J. Chem. Soc, Faraday Trans., 92, 1 (1996) 85. [20] A.V. Rusanov, U.Y. Golger, Kolloid. Zh., 32 (1970) 616. [21] P.M.Kruglyakov, S.M. Selitskaya, T.V. Mikina, Izv. SO AN SSSR, Ser. khim. nauk, Mil (1983)40. [22] A. Sheludko, B. Radoev, A. Fabricant, God.Sof. Univ., 63 (1968/1969) 43. [23] B.D. Summ, Y.V. Gorunov. Physico-chemical basis of wetting and spreading, Moscow, Khimia, 1976. (Russ.) [24] A. Scheludko, B. Toshev, D. Platicanov, in: The Modern Theory of Capillarity, To the Centennial of Gibbs' Theory of Capillarity, ed F.C. Goodrich and A.I. Rusanov, Academie-Verlag, Berlin, 1981. [25] D. Duncan, D. Li, J. Gaydos, A.W. Neumann, J. Colloid Interface Sci., 169 (1995) 256. [26] P. Garrett, in: Defoaming - Theory and Industrial applications, P. Garrett (ed.), chapter 1, Marcel Dekker, New York, 1993. [27] N.D. Denkov, I.B. Ivanov, P.A. Kralchevsky, D.T. Wasan, J. Colloid Interface Sci., 150 (1992)589. [28] A.V. Nushtayeva, P.M. Kruglyakov, Mendeleev Commun., 6 (2001) 235. [29] A.V. Nushtayeva, P.M. Kruglyakov, Colloid Journal, 65, JVa 3 (2003) 341. [30] G. Mason, N.R. Morrow, J. Chem. Soc, Faraday Trans., 80 (1984) 2375. [31] R.P. Mayer, R.A. Stowe, J. Colloid Interface Sci., 20 (1965) 893. [32] H.M. Princen, J. Colloid Interface Sci., 30 (1969) 60. [33] H. Hassander, B. Johansson, B. Tornel, Colloids and Surfaces, 40 (1989) 93. [34] P.M. Kruglyakov, A.V. Nushtayeva, N.G. Vilkova, J. Colloid Interface Sci. (in press). [35] A.V. Nushtayeva, P.M. Kruglyakov, Colloid Journal (in press). [36] W. Stober, A. Fink, E. Bohn, J. Colloid Interface Sci., 26 (1968) 62. [37] P.M. Kruglyakov, S.M. Selitskaya, T.V. Mikina, Izv. SOAN SSSR, Ser. khim. nauk, JSTa 1 (1983)40. [38] S.D. Taylor, J. Czarnecki, J. Masliyah, J. Colloid Interface Sci., 252 (2002) 149. [39] N.P. Ashby, B.P. Binks, Phys. Chem. Chem. Phys., 2 (2000) 5640. [40] D. Exerowa, P.M. Kruglyakov, Foam and Foam Films, Amsterdam, Elsevier, 1998. [41] P.M. Kruglyakov, V.V. Sviridov, Kolloidn. Zh., 55 (1993) 18. [42] P.M. Kruglyakov, A.V. Nushtayeva, Advances in Colloid and Interface Sci. (in press).
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 17
Brownian dynamics simulations of emulsion stability G. Urbina-Villalba, J. Toro-Mendoza, A. Lozsan, and M. Garcia-Sucre Centro de Fisica, Institute Venezolano de Investigaciones Cientificas (IVIC), Apartado 21827, Caracas 1020-A, Venezuela 1. INTRODUCTION One of the main advantages of simulations is the freedom to incorporate processes and change their parameters at will. This characteristic makes them particularly useful in those cases where several variables are involved and multiple time-dependent processes simultaneously occur. This is just the case of macro-emulsions, where the predictive power of the theory is challenged from the very moment of their preparation. In these systems the effect of each formulation variable is complexly mingled, concealing its intrinsic contribution to the final observable result. The progressive incorporation of effects by means of simulations allows decomposing the experimental evidence into its partial contributions. This information is otherwise inaccessible and provides further understanding of the phenomenon under consideration. Emulsions result from a forced blending of two immiscible phases (generically denominated "oil" and "water") due to the action of an impeller. During emulsification, the increase of interfacial area (AA) augments the enthalpy of the system in the magnitude yAA, where y is the interfacial tension. In the absence of a surfactant, this free-energy contribution cannot be counterbalanced by the increase of configurational entropy resulting from the formation of drops. Thus, liquid dispersions are unstable and tend to separate into its former phases. Even in the simplest situation in which no surfactant is added, and the oil phase is composed of a single insoluble alkane, the Drop Size Distribution (DSD) resulting from emulsification is the product of the competition between the opposite processes of phase breaking and drop coalescence. The breakage of the phases depends on the geometry and velocity of the impeller, the viscosity of the phases, and the value of the interfacial tension. Coalescence on the other hand, depends on the diffusion of the drops, their surface properties, and the process of drainage of the intervening liquid between flocculated drops.
678
G. Urbina-Villalba et al.
Fortunately, the influence of mechanical variables on emulsification can be often summarized in empirical relations like [1-3]: i0.6
d32 = 0.058
7
3 "f 'imp ^imp 2
(1 + 5Act>)Dimp.
(1)
Where: rf32 is the Sauter diameter, df the fluid density of the continuous phase, <> j the volume fraction of internal phase, Vimp the impeller speed, and Dimp its diameter. Independently of the specific form of Eq. 1 for a particular system, the mean diameter of an emulsion is expected to decrease with the velocity and diameter of the impeller. Similarly the drop size should diminish, as the energy required for forming the drops (interfacial tension) decreases. The fact that several non-thermodynamic variables are required to describe the initial DSD [4], is a consequence of the kinetic stability of this system. Since emulsions are not in equilibrium, the number of variables required for the characterization of their thermodynamic states is always larger than the one predicted by the phase rule. This is a fancy way to asseverate that the initial state of an emulsion depends on its method of preparation. When a pure surfactant is added to one of the phases prior to mixing, all variables related to coalescence and breakage change as a function of surfactant addition. Since surfactants are known to lower the interfacial tension considerably, they should decrease the mean diameter of the drops according to Eq. 1. It is found that a large excess of surfactant molecules is usually able to produce a single stable phase, while small amounts (ordinarily between 1 % and 5% w/w) are able to stabilize liquid/liquid dispersions. Unfortunately, surfactant partition is a time-dependent process, function of the diffusion of these molecules in each bulk phase and the interface. Diffusion on the other hand, depends on the total surfactant concentration, the mechanism of surfactant adsorption, and the available interfacial area. Since the interfacial area constantly changes during emulsification, chemical effects caused by surfactant addition are difficult to account for during this process. Despite these difficulties, the emulsion type (water-in-oil (W/O), or oil-inwater (O/W)) resulting from a mixture of similar volumes of water and oil (ij) ~ 0.50), is generally determined by the surfactant location at the moment of emulsification (Bancroft rule) [5-6]. Davis [6] justified this experimental behavior in terms of the stability of the oil and water drops, and its connection with surfactant adsorption and chemical structure. A dynamic interpretation of the Bancroft rule is also available [7]. It is based on the difference of surfactant diffusion in each phase, and its relation to the reduction of interfacial tension gradients during coalescence [7]. However, the role of surface diffusion on the
Brownian Dynamics Simulations of Emulsion Stability
679
stability of thin liquid films is a matter of constant debate [8]. On the other hand, the connection between the rheological properties of the interfacial layer and its stability towards coalescence is well-established [9], but its association with the Bancroft rule is not evident. Further experimental support on the relation between surfactant diffusion and emulsion stability is given by the fact that the initial emulsion type may change with time if a substantial surfactant migration occurs after emulsification. This may happen whenever the surfactant is formerly dissolved in its lesssoluble phase, or when one of the formulation variables is systematically changed [10]. In those cases, surfactants will move from one phase to the other favoring a phase inversion. Thus, the initial DSD resulting from emulsification may evolve with time in different ways depending on the original distribution of matter. It is clear that the spatial distribution of molecules can vary amply depending on the mixing conditions. Consequently, comparable DSDs may diverge significantly as a function of time, depending on their initial surfactant distribution. Once formed, a typical emulsion destabilizes through six major mechanisms: a) creaming, b) sedimentation, c) flocculation, d) coalescence, e) phase inversion [10-11], and f) Ostwald ripening [12]. All these mechanisms are severely affected by the presence of surfactant molecules, and constitute a subject of study in their own right. Here we present an overview of emulsion stability simulations. The differences that distinguish these calculations from standard Brownian Dynamics simulations of dispersions will be outlined. Our results will mostly concentrate on the process of flocculation with special emphasis on surfactant effects. A review of the most relevant theoretical aspects on this subject was recently published by Petsev [13]. An updated examination of experimental results on coagulation kinetics is not available, but a very useful compendium on the subject was published back in 1987 by Sonntag and Strenge [14]. 1.1 Flocculation kinetics In order to provide an analytical description of the simulation data, Smoluchowski formalism [15] will be employed. This theory was formerly used to describe the behavior of sols, and proved to be very successful to explicate the dynamics of aggregate formation in dilute dispersions of polystyrene latex [1617]. Through the years, several improvements of the original expressions had been forwarded [18-20]. Some of them account for coalescence [18-19] while some others consider the possible loss of particles due to sedimentation and creaming [20]. However, Smoluchowski expressions are believed to be more convenient for the present purpose, since they are able to describe the dynamics of fairly concentrated emulsions (<j) < 0.20), and can be easily fitted to the simulation data. As will be shown, the use of one-parameter equations clarifies
680
G. Urbina-Villalba et al.
the effect of different variables on flocculation, avoiding the ambiguity that results from the fitting of more complex expressions with several parameters. According to Smoluchowski [15], the total number of particles per unit volume of a dispersion (n), decreases with time as a consequence of flocculation, following Eq. (2): B =J^.
(2)
1 + £f not Where, n0 is the initial particle density, t the time, and kf the flocculation constant. According to this expression, the time required for the initial number of particles n = n0 to drop to Vi n0, tm = t{n0l2), is equal to 1/ kf n0. In the absence of interaction forces, the flocculation constant is solely the result of particle diffusion, and can be expressed as: kf
f
= nntxn = 0l/2
= iirRiDo
.
(3)
4kBT
Here Do stands for the diffusion constant of the particles at infinite dilution:
oirrjR j
Where kB is the Boltzmann constant, /?,- the particle radius, and r\ the viscosity of the solvent. The variation of the number of i-particle aggregates as a function of time can also be obtained, and is given by a simple modification of Eq. 2:
_ "b[*fV]'~' +|
(5)
'~[i+wr The above equations are deduced calculating the number of collisions per unit time of one stationary particle, subject to the influence of the Brownian motion of its surrounding neighbors. This can be achieved solving the diffusion equation for a given collision radius in the absence of convection. The collision frequency is found to be proportional to the number of colliding aggregates per unit volume. Using this result, population balance equations can then be formulated for each floe size. It is assumed that: a) the initial DSD is monodisperse, b) every collision leads to flocculation, and c) the relative diffusion constant is equal to the sum of the diffusion constant of each particle.
Brownian Dynamics Simulations of Emulsion Stability
681
Thus, the total number of aggregates of each size k (nk) existing at time / can be expressed as a difference between two contributions. The number of aggregates of this size produced by the flocculation of smaller clusters of size /' and j , and the loss of £-size floes due to collisions with other clusters:
-^r = \ E
k nn
u i j-
"* £ *,•*«/
i<^
(6)
j=k-i
The group of rate constants {ky}, constitute the kernel of the set of coupled differential equations (Eq. 6) for each particle size. In order to obtain a differential equation for the total number of particles, the collision frequencies of different particle sizes have to be related. After some arithmetic it can be shown that ky - 4nDlkRik which is approximately equal to 8nDR = kf, where D and R are the diffusion constant and the radius of the initial particles. Summing all the expressions for different particle sizes (Eq. 6) produces a simple differential equation for the total number of particles, n (= Y,nk)> which can be easily solved. Both n and nx are found to decrease monotonically with time (Eqs. 2 and 5) while all other /-size floes (/ > 1), present maximum values at t = (i - 1) /2kf n0 . The approximate variation of the average radius can be estimated for the DSD described by Eq. (5), if it is supposed that the hydrodynamic radius of an iparticle aggregate is equal Rt = i Ro: < R (t) > = RQ (1 + kfn01). More reasonable variations of the hydrodynamic radius like Rj -^[i Ro, do not produce an analytic expression for
682
G. Urbina-Villalba et al.
allowed to coalescence instantaneously as soon as their separation becomes equal to the sum of their radii. Analytical expressions for more involved cases including some sort of repulsive potential between particles, can be deduced following a similar mathematical procedure [24-25]. They require the inclusion of an additional flux of particles in the diffusion equation, which result from the effect of the repulsive potential. As a result of this treatment, the collision frequency was found to decrease by a factor W, conveniently defined as a stability ratio: W=^L.
(7)
Here kfml corresponds to the value of k{ in the absence of the barrier, and kslow is equal to the effective value of kf in the presence of the potential barrier. The analytical form of W in terms of the interaction potential between two particles of radius Rt is given by Eq. (8):
2R.
h
Where V is the total potential of interaction between two spheres, and h the distance between them. As shown by Verwey and Overbeek, a stable colloid that lasts for a month prior to coagulation (t{ > 106 s), requires a stability ratio greater than 105 for diluted and 109 for very concentrated dispersions [25]. Furthermore, since the integrand in Eq. (8) is maximum in the region where V is maximum (V = V max ), it can also be shown that Vmax has to be approximately equal to 15 kT for W = 10 5 ,and25kTfor J^=10 9 . 1.2 The Derjaguin-Landau-Verwey-Overbeek (DLVO) potential The potential of attraction between two drops of radii Rj and R2 is given by [26]:
r.= -£ -^— 1^ x +xy + x
+ ?,„ f+Xy
+^ ^ x^+xy^x
+y
+X
\W
x + xy + x + y
Here A is the Hamaker constant, x = d /2R/, y = R2 I Ri, and d is the shortest distance between the drops. This energy of attraction is limited by the value of A typically varying between 10'2' J and 10'19 J [27]. The magnitude of the potential increases as the distance diminishes until contact. At this point the potential goes
Brownian Dynamics Simulations of Emulsion Stability
683
to ~oo. As shown by Lu et al. [28], the singular behavior of Eq. (9) at very short distances results from considering point molecules instead of finite molecular sizes in the computation of the dispersion interaction. However, more accurate formulas are very involved, and the correction with respect to Eq. (9) is expected to be minor when d > Ri + R2. The DLVO potential between two particles results from a combination of their attractive interaction (Eq. (9)), with a repulsive electrostatic energy (Ve). In the case of emulsions the repulsive potential is generally the result of surfactant adsorption but can also be caused by the preferential adsorption of protons or hydroxyl ions. The referred electrostatic potential is usually calculated as a particular solution of the Poisson equation, which expresses this potential Q¥) as a function of the distribution of charge density (p) in the system [29]: A¥ = - ^ £ .
(10)
e
Where s is the dielectric constant of the medium. Since p(x,y,z) can vary amply, the solutions of Eq. (10) can be very involved [25,29-30]. In the simplest DebyeHiickel approximation [29] for instance, the typical form of the electric potential at a point r outside a spherical particle of radius i?,is given by: R. ~rJr-R.) * = ^0—-e x ''. r
(11)
Where To is the surface potential of the particle and 1/K is the Debye length: 9 9
8impe z A
2
(12) ekBT Here, ne is the number of ions of valency ZA per unit volume, and e is the unit of charge. The exponential decay of T as a function of distance is produced by the organization of counter-ions and co-ions in the vicinity of the charged interface. The 1/r-dependence of the potential is rather related to the spherical symmetry of this particular problem. In the presence of several electrolytes, K is more generally related to ionic strength of the solution: K
=
I = \ZnAz2A. 2 A
(13)
684
G. Urbina-Villalba et al.
In this equation nA is equal to the value of ne for specie A. As the ionic strength increases, both the range (1/K) and the value of the electrostatic potential decreases (Eq. (11)). In all cases, solution of Eq. (10) leads to a relationship between the particle charge (q) and the magnitude of the electrostatic potential at its surface. In the referred Debye-Hiickel theory, the following relationship is found: q=Rie(l + KRi)V0.
(14)
The calculation of the interaction potential between two charged particles in an electrolytic solution is a difficult problem, closely related to the analytical form of the interaction between their ionic clouds (double-layers). Different equations are obtained if it is supposed that either the surface charge or the surface potential of the particles is fixed as the particle approach. For a complete discussion of this problem the reader is referred to Ref. [25]. In the present work we use the analytical expressions of Sader [31-33] in order to account for the electrostatic potential between drops. Sader [31] recently provided a convenient form for the far-field potential between two particles in the presence of a symmetric z:z electrolyte. It was calculated using Pade approximants to connect the limiting forms of the surface potential and surface charge for small and large values of KRJ. The resulting expressions are accurate for most surface potentials of practical interest (< 200 mV). In this formalism, the repulsive potential (Ve) between two charged spheres separated by a distance r,j between their centers of mass, is given by: AR(i\RI i\
Ve(rg)/kBT =
AB(1)B(J)
Wje
na(r..—R.—R.) K * > J>
(15)
r
ij
AJ^Mpa
(16)
chn +4^QKRja
B (i) = -2
-
(17)
-
l+QnRjCt
7=tanh(^/4)
Q=^p~^'.
(18)
Brownian Dynamics Simulations of Emulsion Stability
685
Where So is the permitivity of vacuum, Rk the radius of particle k in units of a particle radius of reference (a), and
= -^T-
(19)
For that purpose, the following relation between a and(/)p is used: ae
,
K,ekBT
p
4>p Karj
/«»}[2sinh(> / 2 ) - 0 ] 2 4 tanh (> IA) — <j)p— Karf [2 sinh ( ^ / 4) — ^ ] '
The total potential resulting from adding Eq. (9) and (15) is illustrated in Fig. 1. The average radius of the drops and their ^-potential (~ ^o) correspond to the experimental data of Salou et al. "E3"-emulsion [34]. This emulsion was prepared pouring hot Bitumen (150 °C) into acidified water, stirring the mixture with an ultra-Turrax turbine for 5 minutes. The resulting emulsion has an average drop radius of a = 3.9 um, a surface potential of+115 mV at pH 2.9, and a very high ionic strength, Ka = 1536.
Fig. 1. Derjaguin-Landau-Verwey-Overbeek (DLVO) potential for two 3.9-^im Bitumen droplets. The parameters correspond to the E3 emulsion reported by Salou et al. [34]. (a) Repulsive barrier, (b) Secondary minimum of the same potential.
686
G. Urbina-Villalba et al.
Except for the considerable height of the repulsive barrier, the potential of interaction shown in Fig. 1 presents the typical characteristics of a DLVO potential [25]. The size of the potential barrier is a function of the size of the particles, their surface charge and the ionic strength of the electrolytic solution. Due to the unusually large surface potential of Bitumen drops, its repulsive barrier can reach maximum values higher than 10,000 kBT [34-35]. As shown in Fig. la, the attractive contribution prevails at very short separations forming a primary minimum, which corresponds to irreversible flocculation. At longer distances (1 - 50 nm), a secondary minimum also appears (Fig. lb). The depth of the secondary minimum is typically few kBTs, so it generally allows reversible flocculation. 2. MESOSCOPIC SIMULATIONS The DLVO theory [25] previously outlined was originally intended for (solid/liquid) dispersions. In this case the surface charge of the particles is fixed. The final state of the system basically depends on the repulsive barrier. The size of the barrier determines de rate of aggregation, but it also distinguishes between reversible and irreversible flocculation. If the barrier is low, the particles may access the primary minimum. At such short distances, the van der Waals forces are strong enough to hold the particles together irreversibly. The aggregates are then expected to grow considerably, until they are finally subjected to gravitational separation. If the barrier is high enough, the particles will only flocculate in the secondary minimum. Since this minimum is generally few kBTs deep, the thermal fluctuations of the solvent are able to disunite the floes, making the dispersion stable for a long time. For a typical (micron-size) emulsion drop, the dispersion forces are significant within a few hundred nanometers. At § « 10"4, the lowest value of the Hamaker constant (A = 10"21 J) produces a flocculation rate similar to one obtained in the absence of attractive forces (kf~ k0 = 5.49 x 10'18 m3/s) [36-37]. Rising A to 10"19 J can augment the aggregation velocity in one order of magnitude [38]. Higher values can be obtained increasing the volume fraction and the polydispersity of the dispersion [36,38]. Unlike solid dispersions, the natural charge of liquid drops is generally low. Polystyrene latex for instance, shows sizable electrostatic potentials due to the occurrence of natural charges on their surfaces [39]. Oil drops on the other hand, usually present a small negative charge in the absence of surfactants, caused by the ionization and selective adsorption of the surrounding water molecules [40]. The adsorption of ionic surfactants to liquid drops can easily increase their charge generating surface potentials higher than 30 mV. As shown in Refs.
Brownian Dynamics Simulations of Emulsion Stability
687
[34,37], surface potentials around 44-50 mV are able to stabilize bitumen emulsions of 3.9-um interacting with a Hamaker constant of A = 1.24 x 10~'9 J. Unfortunately, liquid interfaces cannot be "prepared" in advance, but are created during the process of emulsification. Thus, the repulsive barriers between flocculating particles evolve with time as a function of surfactant adsorption. It is clear also that the repulsive barrier is a function of surfactant structure that can vary amply. Furthermore, the surfactant concentration at the surface (surface excess) depends on the partition of these molecules among the immiscible phases. Due to its ambivalent nature, surfactant behavior is appreciably influenced by most formulation variables like temperature, salinity, oil composition, etc. Additionally, surfactant solutions show a rich phase behavior [41], along with a significant variety of mechanisms of adsorption to O/W interfaces [42-44]. Further differences between dispersions and emulsions regard the deformable nature of liquid drops. Unlike solid particles, liquid drops are able to deform and coalesce. During the process of flocculation, the intervening liquid is squeezed between the drops. This generates an outward flux, which perturbs the interfacial layers of the drops causing their deformation [45-46]. At very short distances the frontal face of each drop flattens forming an O/W/O film. This thin film may break and lead to coalescence, or it may remain stable. The final result depends on the properties of the film, which are deeply related to the surfactant excess. It is clear from above that the theoretical prediction of the stability of emulsified systems requires the appropriate consideration of the surfactant behavior and chemical structure. In order to achieve this goal we had found convenient to divide the problem into three levels of statistical description (microscopic, mesoscopic, and macroscopic), which correspond to three different kinds of theoretical evaluations. In principle, the microscopic information (surfactant charges, rotational barriers, interaction energies, etc) could be forwarded by empirical, semiempirical, and ab-initio calculations leading to a molecular description of surfactant properties [47-48]. Molecular simulations of model interfaces also fall within this category [49-52] even though they may resemble the behavior of a considerably larger number of molecules. In the best scenario, the possible outcomes of this type of simulations may include interfacial properties like surface entropy, interfacial viscosity, and elasticity. At the opposite extreme of the statistical scale, macroscopic calculations have to account for the behavior of a large number of drops, typically: 1012 1016 drops/m3. Those numbers are extremely large even for modern computers. Consequently, this task is more properly achieved employing analytical expressions instead of simulations [53-58]. In particular, the works of Reddy,
688
G. Urbina-Villalba et al.
Melik and Fogler [55-56] already forwarded a very convenient macroscopic description. It is based on a set of integro-differential equations, which account for the population balance of a number of drops Ns contained in each section "s" of a macroscopic container of size y = L. For the case where only flocculation and creaming are present, this set of equations take the following form:
1=1
-Ns(y,t) E PfsN^t)
^
s = \,2,X
(21)
J = S+1
Here Us is the average creaming velocity, and the P coefficients are inter- and intra-sectional flocculation rates. Those coefficients are related to the frequency of collisions, or in more simple terms, to the collision kernel {ky} previously referred [14,55-56]. As shown by Reddy et al. [55-56] the predictions of these equations closely reproduce the experimental variation of the DSD if appropriate forms of the rate constants are introduced. Mesoscopic simulations are aimed to build a bridge between microscopic calculations and macroscopic ones. They should be able to incorporate interfacial properties resulting from microscopic simulations, and produce an adequate output for the macroscopic characterization of emulsions. The required program should be versatile enough to incorporate experimental parameters as input in the absence of calculated ones. This will allow independent progress in the comprehension of the stability problem. Additionally, mesoscopic simulations should provide a reasonable set of flocculation rates {ky} so that macroscopic calculations can be suitably carried out. If this three-level scheme succeeds, the final outcome would be a prediction of the DSD of an emulsion in terms of surfactant parameters. 2.1 Brownian Dynamics Simulations of Colloidal Dispersions Brownian Dynamics has been used for more than two decades in order to reproduce the behavior of dispersions [59-64]. In this technique, the large number of degrees of freedom corresponding to the molecules of the suspending medium are transformed into two effective hydrodynamic contributions to the particle movement: a diffusion matrix, and random kick. The diffusion matrix D accounts for the diffusive motion of a particle and its hydrodynamic interaction
Brownian Dynamics Simulations of Emulsion Stability
689
with the surrounding neighbors. The random kick results from the thermal interaction between the dispersed phase and the continuous one. It summarizes the effect of the large number of collisions occurring between each suspended particle and molecules of the solvent during a finite time At. In the algorithm of Ermak and McCammon [65], the equations of motion of a set of N particles in a liquid medium are deduced from a Fokker-Planck [66] description of the evolution of this system in the phase space. The displacement (He ~rk^ °fa Brownian particle during a time At results from three contributions: (a) the gradient of the diffusion tensor, (b) the diffusive motion of the particles, and (c) the thermal contribution of the solvent:
rk = 4
L
+ T/^^At+Z^
/
dr{
^^t
i
+ RG (At).
(22)
kBT
In the original article [65], the coordinates of the particles are arranged in a 3N vector (lx,ly,lz,2x,..,Nx,Ny,Nz). One subscript is used to identify each entry in that vector. Subscript k has the information of the particle number and its coordinate. Thus, k might be equal to lx, ly, 5z, etc. In Eq. 22, rk stands for the position a given particle along one coordinate axis. F® is the total force acting on the particle in direction /, RG (At) is a random displacement with a Gaussian distribution, and variance 2Dj-/ At, where subscripts k and / span all directions (x,y,z) and particles. Superscript "0" indicates that the value of the variable is evaluated at the beginning of the time step (At). The general form of the diffusion matrix components is given by [67-69]:
Dti = Doi + Do 53 {A^r^+B.ir^i-f^]}
(23)
^ =A,H(r^4-+5c(^)[/-^]}-
(24)
Subscripts i and j in Eqs. (23) and (24) are particle labels, and Dn is the diffusion (Stokes) coefficient of a solid particle at infinite dilution (Eq. (4)). Vector *ij=rij/rij 1Sa u m t vector in the direction rij=ri-rj . The scalar functions As, B s , Ac, B c are mobility functions. Their analytical expressions are usually calculated supposing pairwise additive interactions [70-72]. They can be usually expressed in terms of the relative position of the suspended particles, so that the gradient of the diffusion matrix (second term in the r.h.s. of Eq. (22)) vanishes. Since the displacement of each suspended particle perturbs the surrounding liquid, the motion of each particle is not independent. Whenever
690
G. Urbina-Villalba et al.
Hydrodynamic Interactions (HI) are accounted for, the random term %(Af) couples the random displacement of each particle. The evaluation of D and RG(&() are by far the most time-consuming steps in a conventional BD simulation. A brief account of former simulations on flocculation previous to the appearance of the Brownian Dynamics technique can be found in Ref. [37]. Ermak and McCammon originally used their BD algorithm to study the motion of bead-and-spring polymers [65] and the behavior of hard-sphere poly-ions in a solution of ionic species [73]. Probably the first application of this methodology to the study of colloids is that of Bacon et al. [60]. These authors studied the motion of floes of two and three particles (Rt = 1 um, A = 5 x 10'21 J) around the secondary minimum of a DLVO potential. HI were accounted for using the Oseen tensor [67]. The depth of the secondary minimum was changed as a function of the ionic strength, presenting values of 0.87, 4.68 and 10.67 kBT for electrolyte concentrations of 10, 50, and 120 mol/m3. Among other interesting results, it was found that the particles could jump out the secondary minimum in a matter of seconds. One of the major limitations of BD in the simulation of colloidal systems was already observed in the seminal work of Bacon et al. [60]. In a typical colloidal system the width of the repulsive potential barrier is of the order of a few nanometers. As a consequence, the simulations require a very small time step in order to sample the potential appropriately. Due to computational restrictions these authors employed time steps (At) between 20 and 50 us, requiring a value lower 1 us in order to prevent the possibility of particles being "pushed" over the potential barrier. The authors circumvented the problem artificially holding the repulsive force constant at inter-particle distances shorter than a convenient threshold (~ 2 um). The selection of a small time step on the other hand, limits the application of the BD technique, and questions the validity of the specific algorithm of Ermak and McCammon [65] for the study of colloidal systems. This algorithm is accurate for sufficiently long times, when the particle displacement is proportional to tm (
Brownian Dynamics Simulations of Emulsion Stability
691
(Rj = 0.25 nm, A = 1.94 x 10~19 J) was selected. That distance defined clusters of particles, which interacted internally with the Oseen tensor. The coagulation time was estimated, as the time required for the central particle of the cluster to collide with another particle. It was found that the coagulation time depended on cluster size. The effect of "switching on" the HI was to reduce the coagulation rate by a factor of 2.5 at <> ) = 0.155, and 3.0 at <> j = 0.303 [59]. Doublet dissociation times in good agreement with experimental observations had also been published. In the case of 2-um-diameter particles, values of 0.258 s, 1.49 s, and 140 s, were calculated for well depths of 0.87, 4.7 and 10.7 kT [62]. Coagulation times t\n and ?49 corresponding to a decrease in the initial number of particles (no), to either Vi no or to 1 particle, are also reported for a 49-particle system [76]. Times t\a vary between fractions of milliseconds to a few milliseconds depending on the potential and the volume fraction. Values of tw between 520 - 700 ms (<> | = 0.307), 110 - 140 ms (<> | = 0.401), and 15 - 71 ms (<|) = 0.545), were also observed depending on the inter-particle potential. Recently Hutter et al. [77] studied the coagulation rate of dispersions of A12O3 particles in water (Rt = 0.5 urn, A = 4.76 x 10'20 J) by means of BD without HI. These authors noticed that in a diffusion-controlled scheme of aggregation, the coagulation rate depends on the mean free path between particles. Hence, the average time between collisions should be roughly equal to t = I2/Do, where i is the mean surface to surface separation between two nearest neighbors in a randomly distributed configuration:
t = 2R.^SL-\ .
( 25)
V0 Here >m is the maximum volume fraction, which depends on the type of packing arrangement chosen. Accordingly, tl/2 was supposed to be equal to: I
' l1^2
'i/2=«i ?Pr--l
•
(26)
For a vanishing value of the surface potential [77]: a, = 3.3 x 10"2 s, a 2 = 2.23, <> | m = 0.479. Eq. (26) successfully correlates BD data corresponding to a wide variety of conditions and repulsive barriers between particles. Furthermore, it also fitted flocculation data corresponding to monodispersed systems of 3.9-(im particles, subject to van der Waals attraction (A = 1.24 x 10"19
692
G. Urbina-Villalba et al.
J) and/or thermal interaction [36]. In this case cci = 1.4 x 10+2 s, a 2 = 2.00, and <> | m = 0.51 (cubic packing). Simulations on aggregate structure in concentrated dispersions were published shortly after the BD algorithm was developed [62,76,78]. These were motivated by the knowledge that isolated aggregates made by fast irreversible coagulation are fractal with a dimensionality dfr defined by:
RjR,=nXld« .
(27)
Where Rn is the radius of gyration of an n-particle aggregate. Two-dimensional 49-particle simulations showed dfrvalues between 1.76 - 1.80 (<> | = 0.307), 1.83 - 1.86 (<> | = 0.401), and 1.94 - 1.97 (<> | = 0.545), which depend on the particle potential [76]. Three-dimensional random walk evaluations suggested a value of dfr= 1.75 - 1.8 in the limit of large clusters [78-79]. Additionally, the number of clusters of size "s" at time t, ns(t), was found to be represented -in two dimensions- by a dynamic-scaling function of the form:
"S(f) = \f{s/t2).
(28)
Where function f{x) has a power law behavior (f(x) ~ x5) for x « 1, and f(x) « 1 for x » 1. Recently, a different kind of Brownian Dynamics simulation based on the probabilistic calculation of particle encounters was reported [80-84]. This makes use of the fact that the probability (P) of finding a particle at position r after a time t, is a Gaussian function which depends on the interaction forces [85]: 2
__ _ . Wb»'o) = -
1
r-rQ ^exP
-FextAt/~ff TT^
•
(29)
Where Fexl comprises external and inter-particle forces, and yf is the friction coefficient. Since P is Gaussian, a standard Box-Muller [86] algorithm can be employed to produce random walks compatible with Eq. (29). This methodology had been used with success in the evaluation ku [80], to study the phenomena of homo and hetero-aggregation [81-83], and the effect of steric interactions on the coagulation rate [84].
Brownian Dynamics Simulations of Emulsion Stability
693
2.2. Basic modifications of the BD algorithm for emulsion stability studies. 2.2.1. Hydrodynamic corrections In dilute systems, tensorial expressions (Eqs. (23) and (24)) along with mobility functions of distinct levels of accuracy reproduce the observed behavior of colloidal particles [87-91]. However, those expressions fail in concentrated emulsions, and might even lead to incorrect results in dilute systems whenever significant aggregation occurs. In the case of two-particle systems, the use of an average diffusion constant instead of a diffusion tensor had proved to be successful [16,59,80,92]. Honig et al. [92] provided a simple formula for an effective diffusion tensor based on a series expression deduced by Brenner [93]:
Where d / /?, = u, and: „/ x A«)=
6M2
+ 13M + 2 , * . •
(31) '
V
6M +AU
Function p takes values of 1.08, 2.03, 7.83, at u = 18.14, 1.09, 0.091, respectively, going as high as 202.5 at u = 0.0025. The original rational function [93] can reproduce the exact mobility functions with an accuracy of four figures for u > 0.1, and three figures for 0.01 < w < 0.1 [92]. However, usual mobility functions are deduced from pair contributions and do not consider the screening of the hydrodynamic interactions between two particles due to the presence of other particles in between. As Bacon et al. demonstrated [60], the use of the Batchelor tensor in concentrated systems can generate negative diffusion constants above a critical volume fraction (equal to 0.45 for hard spheres). Similarly, it can predict negative sedimentation speeds for § = 0.27 [60, 94]. The existence of such screening was already observed in suspensions of charged silica particles [95]. Furthermore, it was recently shown that this overestimation of HI favors the occurrence of multiple collisions in BD calculations, abnormally increasing the value of the flocculation rate [96]. In order to account for many body hydrodynamics, Heyes [97] proposed a configuration-dependent friction coefficient #,-(R) = kBT/D0 :
C(R) = e o l + C E A ( ^ f
.
(32)
694
G. Urbina-Villalba et al.
Where £0 is t n e friction coefficient at infinite dilution, C = C (<))) is a constant which depends on the volume fraction and is fitted with experimental data, m is an arbitrary exponent, a the hard-sphere diameter of the particles, and X (~ kBT) a constant that is used to set the energy scale. Eq. (32) can be very useful to account for HI but it is troublesome to implement since it depends on several empirical parameters. Recently our group proposed a combined methodology [96] that makes use of the accuracy of Eqs. (30) and (31) for a binary collision, along with an effective diffusion coefficient, D(§), for longer distances. The analytic form of £>(<)>) is based on light scattering results [98-100], but can be obtained averaging mobility functions over spatial distributions of particles [101-105]: Z)(<0) = D o [1-1.830 + ...]
(33)
D(
+ ..]
(34)
£>(0) = D O [1-1.730-O.930 2 + 1.8O02 +... ].
(35)
In Eq. (35) the third term is obtained using two-particle distribution function, while the fourth term employs a three-particle one. The magnitude of these ())2 contributions suggests that this series converge slowly. For § < 0.30, the two-body term appears to be more reliable [99]. At higher volume fractions, the trend is adequate but sizable differences are observed [98-100]. The methodology used in our simulations for the calculation of the local volume fraction is illustrated in Fig. 2. First, an internal (Rjnt) and an external (Rext) radius are defined around each particle. These radii divide the space into three sections. Particles in the outermost region d > Rext (d = r-y - R, - Rj) do not contribute to the HI of the central particle. Particles that fall in the intermediate region (Rint < d < ReXt ) contribute with the fraction of their volume inside that region to the calculation of a local volume fraction around the central particle. Once the local volume fraction <>j has been calculated, the diffusion constant of the central particle can also be evaluated by means of Eq. (35) up to second order in the volume fraction <j). Notice that -according to Eq. (35)- the diffusion constant tends to a minimum value of 0.14 Do at <> j ~ 1.0. However, Honig et al. [92] showed that the diffusion constant could be reduced to less than 0.005 Do during a binary
Brownian Dynamics Simulations of Emulsion Stability
695
Fig. 2. Calculation of the Hydrodynamic Interaction (HI) for a central sphere C (dotted circle). In the present model [96] particle 1 does not contribute to the HI of C. Particle 2 adds its inner volume (R;nt < d < R ex t) to the calculation of a local volume fraction <>| around C (Eq. (35)). If some particles enter the inner region (particle 3), the closest one is used to calculate the diffusion constant of particle C according to Eq. (30).
collision. Hence, the present methodology uses a different diffusion constant whenever a particle enters the internal region of a central particle (d < Rint ). In this case, the central particle is assumed to diffuse according to Honig et al. formula (Eq. 30). As shown in Ref. [96], the described methodology reproduces the behavior of the exact diffusion tensor (Batchelor [71-72] whenever § < 0.1. At higher concentrations, the model overcomes the overestimation of HI producing reasonable flocculation rates at all volume fractions studied (10"6 <(|) £0.40). Using the approximations of the previous paragraphs, the equation of motion of the particles (Eq. (22)) is considerably simplified:
n
(t + At) = rt (t) + Di &:dlFAt
+ RG (A (*,d)) •
(36)
Here r, (t) is the position of particle / at time t, and F is the sum of interparticle and external forces acting on \.D{
Eqs. (35) and (36) reproduce the correct limit value of the diffusion constant (Do) at infinite dilution. According to the Einstein relation [67-69,106107] the product of the friction and the diffusion tensors of a suspended particle should be equal to kBT. In dispersion, the same relation holds for the product of
696
G. Urbina-Villalba et al.
complete friction and diffusion matrices comprising all particles. This basically establishes that the random displacements of the particles cannot be arbitrary, but have to be consistent with the thermal energy of the suspending medium. However, kBT is the time-average value of the thermal energy, which cannot be conveniently evaluated during the course of a simulation, especially in those cases in which the number of particles changes as a function of time. According to the theory of Brownian movement [85,106] and experimental measurements [107], the diffusive motion of each particle complies to a Gaussian distribution with zero mean and variance ^j6Dt. Due to the statistical properties of a Gaussian distribution, the absolute value of the random displacement should be higher than ^J6Dt, 63.4% of the time. With these statistics it is easily confirmed that the square displacement of a particle comes out to be proportional to the time with a slope of 6D when an appropriate random number generator is employed. In Ref. [36] the effect of the thermal interaction on the flocculation rate was evaluated implementing a routine that calculates the kinetic energy gained by the suspended particle after a random kick. For this purpose, the velocity of a particle in each axis was calculated dividing its random displacement along that direction, by the time step. The value of the kinetic energy was then compared to a pre-selected threshold. Random displacements were only accepted if they came out to be lower than the given threshold. When this routine was applied to the movement of one particle in the absence of hydrodynamic corrections, it became clear that in order to reproduce the experimental value of the diffusion constant, high values of the thermal threshold should be considered (UT ~ 10,000 kT!). For this case, the average value of the thermal energy was found to be of the order of several hundreds kBTs. Inclusion of HI through the use of the effective diffusion constant D{(/>,d) lowered the magnitude of the kinetic energy transfer. In this case, average values of 3.6, 5.1, 40.5 and 8.8 kBT were obtained for § = 0.1, 0.2, 0.3, and 0.4, respectively. These values were obtained in the presence of a considerable van der Waals attraction (A = 1.24 x 10"19 J), and in the absence of any repulsive force. Furthermore, the value of &f calculated with only 125 particles (<)) = 10'4), comes out to be very close to the theoretical value predicted by Eq. 3 [36]. 2.2.2 Use of a double time step for the evaluation ofkjin dilute dispersions Up to our knowledge no experimental technique is available for the evaluation of flocculation rates in concentrated systems. This is partially due to the multiple types of collisions that may simultaneously occur, and the different types of the flocculation rates ky involved [17]. Typical evaluations of ku and k{ [14,16-17] correspond to low volume-fraction systems ((j) < 1 x 10"4). These volume fractions cannot be simulated with a fixed time step. A small time step is generally too short for a significant diffusion of the particles at these dilute
Brownian Dynamics Simulations of Emulsion Stability
697
concentrations. On the other hand, long time steps do not allow the appropriate sampling of a nano-metrical potential. Thus, conventional BD simulations can only calculate the flocculation constants of dilute systems in the absence of a repulsive barrier. To overcome this problem, the implementation of an algorithm with two time steps is required. This technique has been used in the past to speed up simulations of dilute systems [108]. In its present version, the range of the interaction potential is used as input. For these simulations {R, - 3 . 9 urn), it was set equal to 50 nm. The value of the short time step was taken from previous calculations [36-38,96,109-111], in which the same DLVO potential was employed. The time step selected in those cases (Ats = 3.40 x 10"7 s) preserved the initial number of particles in the presence of a large barrier height, and in the absence of the random displacement (RG = 0 in Eq. (36)). The longer time step could be set arbitrarily high, except for the fact that the larger its value the smaller the accuracy of the flocculation constant evaluated. In the present simulations: 1.36 x 10"6 s < AtL < 5.44 x 10"2 s. Once AtL and Ats are selected, a double time step calculation begins employing the value of AtL in Eq. (36). The minimum separation between the particles is calculated in every iteration. If this distance is smaller than twice the pre-selected potential width, all particles are returned to their previous positions, and the shorter time step is used instead. Following, the particles move at this lower time step for AtL/Ats iterations in order to sample the interaction potential adequately. When this inner cycle finishes, the particles had moved for a space of AtL seconds, going back in phase with the longer time step formerly used. With this methodology, coalescence can only occur within this inner cycle. A typical calculation enters the inner cycle from time to time when the particles are close, traveling large distances at AtL steps in between particle encounters. 2.2.3 Time dependent surfactant adsorption. As previously discussed, usual stability considerations start from the evaluation of Eqs. (7) and (8), assuming a constant surface charge. Such reasoning comes from our previous knowledge on the behavior of dispersions. In the case of emulsions, the same reasonable assumptions tacitly imply an extremely fast surfactant adsorption along with instantaneous relaxation of the adsorbed layer. Similar considerations are usually implicit in the calculation of the double-layer interaction between two charged drops, even in the case in which these drops are not flexible and a constant electrostatic potential is assumed [25]. However, surfactant adsorption is a time-dependent process depending on the equilibrium between the surfactant concentration at the subsurface and the amount of surfactant adsorbed [42-44]. These considerations are especially important during the process of emulsification, where surfactant partition defines the type of emulsion obtained [6-8,112]. They are also relevant
698
G. Urbina-Villalba et al.
in the coalescence of deformable drops where the interfacial area constantly changes [2-4,7]. As shown by Ivanov and Dimitrov [7], the velocity of film drainage is substantially slower when the surfactant is dissolved in the continuous phase. This is an indication of the pronounced influence of surfactant diffusion on emulsion stability. In the case where ionic surfactants are used to stabilize dispersions of nonpolar oils in water, the surface charge density (a) of the drops is essentially the result of surfactant adsorption. Hence, the charge of the drop in Eq. (19) is equal to: Z,- = Ai zs r ( 0 .
(37)
Where A, is the total area of drop /, zs is the effective charge of one surfactant molecule at the oil/water (O/W) interface, and F(t) is the surfactant surface excess at time t. The value of zs can be deduced from the ^-potential of an oil drop at maximum surface coverage. For this purpose, the equilibrium value of the excess, r eq is needed. This can be approximated from the slope of the Gibbs adsorption isotherm by plotting y as a function of log (cs) [5]. In most cases, 1/Feq falls in the range 20 - 100 A2 [113]. In order to compute the surfactant charge for the present calculations, it was assumed to be 50 A 2 [113], an average value for alkyl ammonium surfactants in 0.1 MNaCl solutions [111]. Once zs has been estimated, the surface charge at time t can be calculated using Eqs. (37) and (19), if F(t) is known. The value of the surface potential
Brownian Dynamics Simulations of Emulsion Stability
699
solution is longer than that of adsorption, the total adsorption process is regarded to be diffusion-controlled. In this case [114]:
T{t) = 2\^2\ct^-fQCsub{z)d{t-z)V2\
(38)
Where C and Csub are the surfactant concentration in the bulk phase and the subsurface, respectively, z is a time variable ranging from 0 to t. Whenever back diffusion can be neglected the second term of Eq. (38) can be omitted. As a result: r(/) = 2 [ Z > > ] 1 / 2 C f l / 2 .
(39)
As shown by Liggieri et al. [115] the equations describing surfactant adsorption under mixed adsorption kinetics can be re-formulated into an effective diffusion-controlled formalism. As a result, Eq. (39) can be used to describe a wide variety of systems when Ds is substituted by an apparent diffusion constant, Dapp. This allows mimicking the dynamics of adsorption corresponding to different isotherms (Freundlich, Langmuir, etc) through Eq. (39). Additionally, the logarithmic form of Eq. (39): log(0 = 21og(r(/)/C) + log(7r/4Da/7p).
(40)
is compatible with the experimental findings of Hua and Rosen [116-117], which allows the estimation of Dapp from experimental data. According to these authors, a typical plot of dynamic surface tension vs. time, can be divided into four regions: (I) induction region, (II) rapid-fall region, (III) meso-equilibrium region, and (IV) equilibrium region. This behavior can be represented by a simple equation:
7o-7^^V l + [t/t*}n
(41)
In Eq. (41): y0, ym, and yt stand for the surface tension of the system at the beginning of the adsorption process, at meso-equilibrium, and at any time t. t* is defined as the time required for the surface pressure (FI = y0 - y t ) to reach onehalf of its value at mesoequilibrium. Eq. (41) allows the definition of a set of complementary variables for the characterization of surfactant behavior. One of the most relevant ones for the present studies is Cm, which is defined as the
700
G. Urbina-Villalba et al.
minimum surfactant concentration for which the surface tension shows little further change with the increase of either time or surfactant concentration. It is basically the concentration that guarantees the attainment of maximum adsorption in a minimum time. This variable was found to be related to the mesoequilibrium time (tm) through an equation similar to Eq. (40): l o g ( 4 ) = a + blog(C^).
(42)
Where a and b are negative constants, and t*m is the minimum time required to obtain ym (a mesoequilibrium tension that does not change much with an increase of surfactant concentration). According to these authors [116-117], similar relations are found to hold between the log (t) and the quotient (r(t) / C). In particular, when Feq and t* are used as particular values of F(t) and t:
Iog/* = a 1 +6 1 log(r e ? /c).
(43)
Here a\ and b\ are constants. This equation is equivalent to Eq. (40), and allows the calculation of Dapp from experimental data [116]. In the present work, we will span a wider range of diffusion constants varying Dapp parametrically between 10"9 and 10"12 m2/s, and employing Eq. 39 for the calculation of the surface excess. 3. USE OF EMULSION STABILITY SIMULATIONS FOR THE CALCULATION OF FLOCCULATION RATES 3.1 Flocculation Dynamics in the Absence of Surfactant Molecules 3.1.1. Effect of the volume fraction on the value ofkj Figures 3 and 4 show typical ESS results for the variation of the number of particles per unit volume as a function of time. In these calculations only 125 drops of radius 3.9 |j.m were simulated. The hydrodynamic corrections specified in section 2.2.1 were employed along with the double time step technique (section 2.2.2) for dilute cases. The particles moved as the result of their thermal interaction with the solvent and their van der Waals attraction given by Eq. (9) with A = 1.24 x 10"21 J. At low volume fractions, <j) = 10"5 (Fig. 3a), n is found to follow Eq. (2) reasonably well. The value of the flocculation constant for this particular calculation comes out to be k{= 5.42 x 10"18 m3/s (r2 = 0.9898), close to the
Brownian Dynamics Simulations of Emulsion Stability
701
Fig. 3. Change in the number of particles per unit volume as a function of time for § = 10"" (a) and <j) = 0.15 (b). At the beginning of the simulations the drops were randomly distributed in a cubic cell keeping a minimum separation of 2.1 radii. Only van der Waals forces (Eq. (9)) were employed in this case (A = 1.24 x 10"21 J ). The particles were allowed to coalesce as soon as they overlap causing a progressive decrease in n as a function of time.
702
G. Urbina-Villalba et al.
Fig. 4. Change in the number of particles per unit volume as a function of time for <>| = 0.40.
theoretical prediction (Eq. (3)) for thermal interaction only. It is usually observed that all factors that disfavor multiple collisions enhance the accuracy of the \ln vs. t fitting. Additionally, large number of particles, and small values of AtL, usually produce good regression coefficients. As the volume fraction increases (Fig. 4), Eq. (2) becomes less suitable for describing the dynamics of the system. At (j) = 0.40, the curve of n vs. t changes its shape due to the appearance of a lag time previous to the occurrence of the first coalescence. A strong hydrodynamic coupling along with an ordered arrangement between neighbor particles causes this drastic reduction of the flocculation rate for a finite period of time. As soon as the first coalescence occurs, the flocculation rate increases considerably causing a sharp decrease in the number of particles during a few seconds. The referred lag time is consistent with the decrease in the creaming velocity observed in similar emulsions (/?, = 0.86 urn) for § > 0.40 [118]. It could also be related to the variation of the average droplet size in hydrocarbon oil-water emulsions stabilized with (3-lactoglobulin [119-120]. In this case, it was found that the droplet size of a <>| = 0.10 emulsion may remain stable after three hours of continuous shearing, increasing drastically afterwards, going from 0.36 um to 4.5 um in the following 20 minutes. Figure 5 shows the variation of the flocculation constant as a function of the volume fraction for two extreme values of the Hamaker constant. The flocculation rate changes exponentially as a function of the volume fraction for 10'5 < <>| < 0.40 [38]:
Brownian Dynamics Simulations of Emulsion Stability
kf = (9.5497x1(T'8)e +'5M5^
703
m3/s
(A = 1.24 x 10"'9 J )
(44)
kf = (6.4103xl(T 1 8 )e + 8 0 0 2 4 ^ m3/s
(A = 1.24 x 10"21 J )
(45)
The marked dependence of the flocculation rate on the volume fraction is basically due to the decrease of the mean free path between particles. Notice that despite the short range of the van der Waals potential, large differences in the flocculation rate are observed. These result from the acceleration of the particles at very close separations. It is clear from inspection of Fig. 5 and the coefficients of Eqs. (44) and (45), that a large difference in the values of kf may result as a consequence of the magnitude of the Hamaker constant. Those differences diminish as the volume fraction decreases, but according to the simulations, they are still noticeable as § -> 0. The value of £ffor A = 1.24 x 10"'9 J comes out to be -1.5 times larger than the one corresponding to A = 1.24 x 10"21 J, in the absence of retardation effects [27]. As discussed in Ref. [25], several experimental determinations in dilute systems had suggested values of the flocculation constant as low as 50% of the
Fig. 5. Volume fraction dependence of the flocculation rate for A = 1.24 x 10 J (•), and A = 1.24 x 1(T21 J (o). The lines represent the fittings of Eqs. (44) and (45) to the simulation data.
704
G. Urbina-Villalba et al.
value theoretically predicted (2.6 x 10"18 m3/s) [121]. Such discrepancy had always been ascribed to hydrodynamic interactions. In the present simulations the effect of solvent displacement due to particle movement has been correctly accounted for [96]. However, it was assumed that at infinite dilution the value of the diffusion constant was that of a solid particle: D{
m3/s
(A = 1.24 x 10"'9 J )
(46)
Brownian Dynamics Simulations of Emulsion Stability
kf =(6.4118xl0- 1 8 )e + 1 9 7 6 6 ^
m3/s
(A= 1.24 x 10"21 J )
705
(47)
As shown in Fig. 6 the regression coefficient of Eq. (47) is substantially better than the one obtained for the larger Hamaker constant (Eq. (46)). This might be an indication of the standard deviation of the calculations. These results were obtained departing from one initial configuration of 125 particles. The monotonous tendency obtained using one particular configuration indicates that the mean free path does not vary appreciably between different configurations with a fixed value of <j). Figure 7 shows the dependence of the half life time of the emulsions (ti/2 t(no/2)) as a function of the volume fraction, along with the predictions of Eq. (26). The data corresponding to the lower Hamaker constant appears to fit Eq. (26) better. Again, the trend is significant considering that only one initial configuration was computed for each volume fraction. In order to fit the data to Eq. (26), the values of a 2 and §m were held constant at 2.0 and 0.51, respectively. The value of a! came out to be 375 s for A = 1.24 x 10'2' J, and 125 s for A = 1.24 x 10"'9 J ones. As expected the larger value of oti corresponds to the weaker attractive force. The value of this parameter for the higher Hamaker constant is similar to the one formerly obtained (140 s [36]).
Fig. 6. Variation of the flocculation constant as a function of the inverse of the mean free path for A = 1.24 x 10"19 J (D), and A = 1.24 x 10'21 J (o). The dashed lines correspond to exponential fittings.
706
G. Urbina-Villalba et al.
Fig. 7. Half life time of the emulsions as a function of the volume fraction for A = 1.24 x 10"" J(D), andA= 1.24 x 10"21 J (o). Notice that significant deviations from the theoretical expression occur at <) = 0.15 and <j) = 0.20 for the case in which A = 1.24 x 10'19 J. However, it should be remarked that multiple collisions usually cause an uncertainty in the determination of t\a at high concentrations, which cannot always be accurately evaluated employing simulations with a small number of particles. 3.1.2. Effect of hydrodynamic interactions on the evaluation of the stability factor W. Figure 8 illustrate the effect of HI on the calculation of the stability factor (Eqs. (7) and (8)) in concentrated systems. In both pictures (a and b), the solid line represents the variation of n vs. t for the case in which the thermal interaction is the only source of particle movement (term RG in Eq. (36)). The small dashes show the variation of n when the van der Waals attraction is included but no HI corrections are made. The flocculation rate increases as expected for A = 1.24 x 10"19 J (Fig. 8a) and A = 1.24 x 10"21 J (Fig. 8b), producing a curve that falls below the thermal counterpart. When the hydrodynamic interactions are included (patterns • and o), the flocculation diminish approaching the thermal interaction for A = 1.24 x 10"19 J, but producing a fcf slower than the thermal one in the case of A = 1.24 x 10"21 J. It is clear from above that care must be taken in the selection of the rapid flocculation constant for the evaluation of Eq. (7), which in general should
Brownian Dynamics Simulations of Emulsion Stability
707
include the van der Waals attraction as well as proper hydrodynamic corrections. Consequently, the use of Smoluchowski's kf constant (Eq. 3) as an estimation of the rapid flocculation rate should be avoided in theoretical evaluations of the stability ratio.
Fig. 8. Variation of n vs. t at <j) = 0.15 for the cases of: thermal interaction only (solid line), thermal plus van der Waals attraction (slash), thermal plus vdWs with HI corrections (symbols). In Fig. 8a, A = 1.24 x 10"19 J (Q), while in Figure 8b, A = 1.24 x 10"21 J (o).
708
G. Urbina-Villalba et al.
In more complex systems where surfactants and/or electrostatic forces are present, the situation can be even more involved. This anomaly illustrates the usefulness of ESS for the study of dispersed systems. The simulations allow separating the effect of each contribution over the flocculation rate. As shown in Fig. 8 the value of kf is the result of two opposing forces: an attractive van der Waals force which increases the flocculation rate, and a frictional hydrodynamic force which opposes the acceleration of particles, decreasing the value of kf. 3.1.3 Polydispersity The effect of polydispersity can be studied generating a common spatial distribution for a set of particles, and changing their radii appropriately to produce distinct DSD. In former simulations only thermal and van der Waals interactions were considered (A = 1.24 x 10"19 J) with no account of hydrodynamic corrections [36]. For this purpose a polydispersity index (P.I.) was defined as: pj
-=T,iN'\Ri-R\
(48)
Here, TV, is the number of drops with radius i?,, in a DSD with an average radius of R . In those calculations, P.I. was varied between 0 and 0.6, using 64, 125 and 216 particles. It was found that in the absence of HI, polydispersity can increase the value of the flocculation rate in one order of magnitude, taking as a reference the value of kf for a monodisperse DSD of equal volume fraction. The referred variations are considerably larger than those previously reported. According to Lee [124], polydispersity could be simply accounted for, multiplying the value of the flocculation constant by a = 1 + exp(ln
Average Flocculation Rate (m3/s)
Range of Rvalues (m3/s)
Thermal
4 .06 x
io-"
3.01 x
io-" -•
6.00 x
io-"
Thermal + vdW
1.49 x
io-' 6
8.67 x
io-' 7 -•
3.01 x
io-' 6
Thermal + vdW (*)
1.26 x
io-16
7.42 x
io-' 7 -• 1.95 x io-16
Thermal + vdW + (HI)
5 .83 x 1 0 "
4.25 x
1 0 " " •• 8.32 x
(*) These calculations include a very weak electrostatic potential. (HI) Hydrodynamic Interactions.
10""
Brownian Dynamics Simulations of Emulsion Stability
709
Similar corrections were reported by Wang and Friedlander {a - 1 + 1.1289 cr) for systems considerably diluter than the ones previously calculated by our group [125]. At the present time, the data available corresponds to d> ~ 0.15 only. Table 1 presents the average flocculation rates resulting from nine of these calculations, with 0 < P.I. < 0.35, 0.1511 < <)| < 0.1597. The entries from the first three rows of Table 1 illustrate the considerable increase of the flocculation rate as a function of polydispersity in the absence of hydrodynamic interactions. However, both the average value of kf and its range of variation should decrease under the action of hydrodynamic corrections. The effect of the number of particles on these results was not studied. Interestingly, the average value of kf in the presence of a large electrostatic barrier ( » kBT) comes out to be 6.71 x 10"'7 m3/s (r2 = 0.9678) [126]. 3.2 Flocculation Dynamics in the Presence of Surfactant Molecules In a system with one unique interface, surfactant diffusion will lead to the rapid accumulation of these molecules at that boundary, and eventually to a stable surfactant distribution. However, the behavior of surfactants under the influence of many interfaces is less evident. Surfactants can be introduced in the system in many possible ways, generating a wide variety of initial distributions. Additionally, the everlasting strive of emulsions towards phase separation manifest itself through several destabilization mechanisms, which continuously change the scenario for surfactant adsorption. If it is assumed that the degree of surfactant adsorption depends on its local concentration, this leads to conclude that different drops may acquire distinct amounts of surfactant molecules during emulsification depending on their spatial location. The period of time required for the homogenization of the initial surfactant distribution will depend on the reversibility of the adsorption process. Ci4SO4Na molecules for example, are known to interchange between micelles in hundreds of picoseconds ((is), while micelles form and disintegrate in a matter of milliseconds (ms) [127]. However, the time required for surfactant exchange between two drops remains unmeasured. Additionally, surfactant distribution can be highly inhomogeneous. In the case dispersions, Velegol et al. [128-129] demonstrated that polystyrene spheres generally present a nonuniform surface charge, which causes their rotation under the influence of an electromagnetic field. Such in-homogeneous distribution of charges may have persisted during their synthesis or could have been originated after their subsequent chemical treatment. On the other hand, if adsorption is assumed to depend on the total surfactant concentration of the system and its available interfacial area, surfactants will be simultaneously attracted toward several interfaces located at distinct points in space. This will decrease their directional diffusion towards one specific site, and consequently, this might delay the adsorption process.
710
G. Urbina-Villalba et al.
In order to cover most possible cases, nine different forms of surfactant distributions were implemented in our program. For a complete description of these options we refer the reader to Refs. [110-111]. Some relevant cases comprehend: 1) Instantaneous and irreversible adsorption with uniform distribution among all droplets, 2) Time-dependent adsorption with homogeneous surfactant distribution, 3) Non-homogeneous surfactant distribution according to a gradient of surfactant concentration that persists during the course of the simulation, and 4) Random distribution of surfactant molecules at each time step (simulating a quick adsorption/desorption process). In all cases, care is taken to prevent a crowding of molecules at the interface beyond the minimum interfacial area per surfactant molecule (r^ 1 ). Since non-homogeneous surfactant distributions favor the collisions between drops insufficiently covered by surfactant molecules, they show an enhancement of the coalescence rate with respect to their homogeneous analogs. Furthermore, they lead to the maximum destabilization of the emulsion. Fig. 9 shows the dependence of the final interfacial area of Bitumen emulsions as a function of surfactant concentration for cases 3) and 4) of the previous paragraph.
Fig. 9. Final interfacial area (IA) as a function of surfactant concentration (Cs) for two types of inhomogeneous surfactant distributions. Instantaneous irreversible adsorption was assumed. ( O ) A gradient of surfactant concentration was maintained during the course of the simulation. ( A ) Surfactants were randomly distributed at each time step.
Brownian Dynamics Simulations of Emulsion Stability
711
According to the results of Fig. 9 [110-111] the total interfacial area (IA) remaining after the coalescence process increases as a function of surfactant concentration. When Cs = 0 M, the number of drops decreases until only one remains. If the value of Cs is enough to entirely cover the initial IA up to maximum packing, all drops are preserved. Homogeneous surfactant distributions produce similar IA vs. Cs plots (Fig. 9). However, the values of Cs required for the preservation of a given amount of interfacial area are considerably lower. The total number of drops was found to remain constant when the initial value of Cs was able to produce an initial surface potential of at least 50.7 mV in all drops [37]. 3.2.1 Instantaneous surfactant adsorption with homogeneous surfactant distributions Figure 10 shows the typical variation of the number of particles per unit volume as a function of time for two different DSD with C s = 8.0 x 10"5 M. The surfactant concentration is evenly distributed among all drops. However, the selected concentration is insufficient to stabilize the initial interfacial area of these emulsions, causing a decrease of n with time. As time evolves, the interfacial area diminishes as the product of coalescence. Consequently, the excess of surfactant molecules is re-distributed among the surviving drops, causing their electrostatic potential to increase.
Fig. 10. n vs. t for two different DSD: (o) (j> = 0.295, IA = 938 a2. (O) (j> = 0.300, IA = 1571 a . (a = 3.9 (im). The polydispersity index is shown in the inset.
712
G.Urbina-Villalbaetal.
After a while, the potential barrier between all pair of drops is high enough to avoid further coalescence. In the absence of other perturbations, the number of drops does not change further, although they may still flocculate into the secondary minimum. If the electrostatic potential between drops is large enough from the very beginning of the simulation, the dynamics of the system does not follow Smoluchowski's aggregation scheme (P.I. = 0.50 in Fig. 10). Otherwise, the flocculation constant can be measured during the sharp decrease of n with t. In this second case, the value of kf usually comes out to be similar than the one evaluated in the absence of the repulsive force. Due to the large value of the ionic strength (0.014 M) in this case, the maximum width of the repulsive potential does not exceed 50 nm. Thus the particle move under the action of the random force until very close separations. Consequently, the stability ratio calculated with respect to the value of kf -including van der Waals attraction and HI- rarely exceeds a few units [126]. 3.2.2 Time-Dependent Surfactant Adsorption In a recent article [130] the effect of the temporal dependence of surfactant adsorption was studied. For this purpose Dapp substituted Ds in Eq. (39). Values of the apparent diffusion constant in the range 10" m /s - 10" m /s were used in order to calculate F(t). The total surfactant concentration was varied between 10"3 M and 10"5 M, and proper hydrodynamic corrections for particle movement were included. As previously discussed, whenever surfactant adsorption occurs through a complex mechanism, its apparent diffusion constant might be too slow (Dapp ~ 10"12 m2/s) to prevent coalescence. In this case the amount of surfactant that will reach the available interfaces prior to flocculation will be insufficient to build up a significant repulsive barrier. At C = Cs = 10"4 M, the number of surfactant molecules is not enough to prevent the coalescence of 125 drops of radius a = 3.9um («0 = 8.05 x 10+M drops/m3), even in the case of instantaneous adsorption. Thus, when Dapp = 10"12 m2/s and Cs = 10"4 M the number of drops diminishes steadily during the course of the simulation as shown in Fig. 11. However, if the surfactant concentration is increased to Cs = 5 x 10"4 M, F(t) augments more rapidly and the coalescence is prevented after 1.5 seconds. This causes a reduction of the initial interfacial area from 1571 a2 to 1250 a1. From that point on, the surface excess is high enough to avoid further destabilization. If the diffusion constant of the surfactant shows the typical order of magnitude of molecular diffusion (Dapp = 10'10 m2/s), the initial number of drops could be preserved depending on concentration. At Cs = 5 x 10"4 M for example, the referred diffusion constant allows covering of the available interfacial area with enough time to prevent destabilization (Fig. 11). However, a smaller
Brownian Dynamics Simulations of Emulsion Stability
713
Fig. 11. n vs. t for a monodisperse <j) = 0.20 simulation, (o) Dapp = 1 x 1 0 " m /s, C s = 1 x 10" 4 M. (+) Dapp = 1 x 10" 12 m 2 /s, C s = 5 x 10"4 M. ( A ) Dapp = 1 x 10"'° m 2 /s, C s = 1 x 10"4 M. ( O ) Dapp = 1 x 10" 10 m 2 /s, C s = 5 x 10"4 M.
surfactant concentration (Cs = 1 x 10"4 M) cannot stop the decrease of the initial interfacial area until it reaches 1466 a2. According to Hua and Rosen [116-117] a surfactant concentration of at least 5 x 10"4 M is required for most air/water-solution interfaces in order to achieve a 1 -second surface tension that does not change much with time even in the presence of larger surfactant concentrations (Cm). This finding might have important consequences to emulsion stability if the surfactant is insoluble in the oil phase. As is shown in Fig. 11, a typical surfactant (10"10 m2/s - 10"9 m2/s) is generally fast enough to prevent coalescence in an emulsion with (j) < 0.20. Figure 12 shows similar results for the case in which (j) = 0.40. According to these results, a concentration of 5 x 10"4 M may possibly stabilize emulsions with higher volume fractions (see pattern ). As before, slow Dapp values (10"12 m2/s) are unable to prevent complete destabilization in the first two seconds after preparation. As shown in Fig. 12, the flocculation rate of emulsions with different surfactant concentration might be similar to the one exhibited in the absence of a repulsive barrier. This situation holds until the electrostatic potential is significant or large enough to prevent coalescence.
714
G. Urbina-Villalba et al.
Fig. 12. n vs. t for a monodisperse i/ = 0.40 simulation, (o) Dapp = 1 x 1 0 " 4
2
2
4
9
m /s, C s = 1 x 10 2
M. (+) Dapp = 1 x 10"' m /s, C s = 5 x 10' M. (-) Dapp = 1 x I0" m /s, C s = 1 x 10"4 M.
( • ) Dapp = 1 x 10' 9 m 2 /s, C s = 5 x 10' 4 M.
It is clear that the algorithm for a time-dependent surfactant adsorption can be also very useful in the cases where competitive adsorption occurs. This is the usual situation in food emulsions, where natural proteins can be displaced from the surface by either monoglycerides or commercial surfactants [120]. Similar simulations can also be a valuable tool for studying the stabilization of negatively charged drops under the influence of cationic surfactants [131]. 4. FINAL REMARKS At the beginning of Section 2 we outlined the value of mesoscopic simulations for closing the gap between molecular properties and macroscopic behavior. The results of Section 3 further demonstrate that this type of simulations can be useful in their own right for the study of emulsion stability, and in particular for the calculation of flocculation velocities and/or coalescence rates. Despite the double time step implementation, the present type of simulations is very demanding in computational time. This fact set a limit to the maximum number of particles of the simulations. One could either concentrate in the effect of a particular variable carrying out simulations with a large number of particles, or sacrifice part of the statistical significance of the calculations in order to establish the combined effect of several variables. In what respect to
Brownian Dynamics Simulations of Emulsion Stability
715
flocculation rates, we have not observed mayor differences between 125 and 1000 particle calculations, although a considerable improvement in the regression coefficient of the fittings is obtained. In the absence of HI and under the action of van der Waals forces (A = 1.24 x 10"'9 J), a (j) = 0.15 calculation produces a flocculation constant of 1.70 x 10"16 m3/s. The same simulation with 1000 particles generates a kf value of 1.40 x 10"16 m3/s. Much larger deviations had been observed in polydisperse systems. On the light of previous results, and considering different techniques for the fitting of the simulation data, we have recently estimated an average error of ±1.4 units for flocculation rates [38]. Several improvements of the present technique are required in order to reproduce the behavior of real systems (Fig. 13). Some of these modifications are possible, while other seem extremely complicated for Brownian Dynamics. Incorporation of steric potentials, and deformable droplets seems possible. In this last case for instance, the detailed simulation of interfacial deformation is not viable, but it is feasible to calculate the deformation energy from the differences in the analytical expressions of the interaction energy corresponding to pairs of deformed droplets and spherical ones [45-46]. Thus, the resulting code will be blind to transient effects, but it will still be able to keep track of deformed drops based on pair interactions. In this sense it should be noticed that numerical algorithms for processes as complex as Ostwald ripening are currently available [132].
Fig. 13. Variation of the Drop Size Distribution (DSD) as a function of time for a (j) = 0.15 emulsion under the action of van der Waals forces without HI. The broken line indicates the initial monodisperse distribution with radius a - 3.9 um. (o) DSD at t = 4.8 s. ( O ) DSD at t = 68.5 s.
716
G. Urbina-Villalba et al.
The research presented has focused on the treatment of surfactant effects. This methodology is foreseen to be very valuable for the treatment of complex diffusion phenomena like Gibbs-Marangoni. Furthermore, the algorithms are believed to be transportable to more sophisticated types of simulations with only minor modifications.
REFERENCES [I] [2] [3] [4] [5] [6] [7]
[8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29]
Y. Mlynek, and R. Resnick, A.I.Ch.E. J., 18 (1972) 122. F.H. Verhoff, S.L. Ross, R.L. Curl, Ind. Eng. Chem. Fundam., 16 (1977) 371. P. Walstra, Formation of Emulsions, in: P. Becher (Ed.), Encyclopedia of Emulsion Technology, Marcel Dekker, NewYork, 1983. Vol. 1, p. 57. T. Tobin, and D. Ramkrishna, Can. J. Chem. Eng. 77 (1999) 1090. A.W. Adamson, Physical Chemistry of Surfaces, John Wiley & Sons, Inc., New York, 1990. J.T. Davis, Proc. 2nd International Congress on Surface Activity, 1 (1957) 426. I.B. Ivanov and D.S. Dimitrov, Thin film drainage, in: I.B. Ivanov (Ed.), Thin liquid films. Fundamentals and Applications. Surfactant Science Series, Marcel Dekker, Inc., New York., 1988. Chapter 7, p. 379. S.D. Stoyanov, and N.D. Denkov, Langmuir 17 (2001) 1150. A. Yeung, K. Moran, J. Masliyah, J. Czarnecki, J. Coll. Interf. Sci. 265 (2003) 439. J.L. Salager, N. Marquez, R.E. Anton, A. Graciaa, J. Lachaise, Langmuir 11 (1995) 37. I. Mira, N. Zambrano, E. Tyrode, L. Marquez, A. Pena, A. Pizzino, J.L. Salager, Ind. Eng. Chem. Res. 42 (2003) 57. A. Kabalnov, J. Disp. Sci. Tech. 22 (2001) 1. D.N. Petsev, Mechanisms of Emulsion Flocculation, in: A. T. Hubbard (Ed.) Encyclopedia of Surface and Colloid Science, Marcel Dekker, Inc., 2002. p. 3192. H. Sonntag, and K. Strenge, Coagulation kinetics and structure formation, VEB Deutscher Verlag der Wissenschafiten, Berlin, 1987. Smoluchowski, M., Z. Phys. Chem. 92(1917)129. H. Holthoff, A. Schmitt, A. Fernandez-Barbero, M. Borkovec, M.A. Cabrerizo-Vflchez, P. Schurtenberger, and R. Hidalgo-Alvarez, J. Coll. Interf. Sci. 192 (1997) 463. H. Sonntag, V. Shilov, H. Gedan, H. Lichtenfeld, C. Diirr, Coll. Surf. 20 (1986) 303. M. van den Tempel, Recueil 72 (1953) 433. R.P. Borwankar, L.A. Lobo, D.T. Wasan, Coll. Surf. 69 (1992) 135. K. D. Danov, I.B. Ivanov, T.D. Gurkov, R. Borwankar, J. Coll. Interf. Sci. 167 (1994) 8. J.C. Brown, P.N. Pusey, J.W. Goodwin, R.H. Ottewill, J. Phys. A 8 (1975) 664. P.N. Pusey, J. Phys. A 8 (1975) 1433. P.N. Pusey, J. Phys. A l l (1978) 119. N.A. Fuchs, Z. Phys. 89 (1934) 736. E.J.W. Verwey, and J. Th. G. Overbeek, The Theory of Lyophobic Colloids, Elsevier, NewYork, 1948. H.C. Hamaker, Physica (Amsterdam) 4 (1937) 1058. J. Israelachvilli, Intermolecular and Surface Forces, Academic Press, London, 1992. J.X. Lu, W.H. Marlow, V. Arunachalam, J. Coll. Interf. Sci., 181 (1996) 429. F. Evans, H. Wennerstrom, The Colloidal Domain: Where Physics, Chemistry, Biology and Technology meet, VCH Publishers, New York, 1994.
Brownian Dynamics Simulations of Emulsion Stability
[30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62]
[63] [64] [65] [66] [67]
717
G.M. Bell, S. Levine, L.N. McCartney, J. Coll. Interf. Sci. 33 (1970) 335. J.E. Sader, J. Coll. Interf. Sci. 188 (1997) 508. M.R. Oberholtzer, N.J. Wagner, A.M. Lenhoff, J. Chem. Phys 107 (1997) 9157. M.R. Oberholtzer, J.M. Stankovich, S.L. Carnie, D.Y.C. Chan, A.M. Lenhoff, J. Coll. Interf. Sci. 194(1997)138. M. Salou, B. Siffert, A. Jada, Coll. Surf. A, 142 (1998) 9. M.A. Rodriguez-Valverde, M.A. Cabrerizo-Vilchez, A. Paez-Duenas, and R. HidalgoAlvarez, Coll. & Surf. A, 222 (2003) 233. G. Urbina-Villalba, M. Garcia-Sucre, J. Toro Mendoza, Molec. Simul. 29 (2003) 393. G. Urbina-Villalba, M. Garcia-Sucre, Langmuir 16 (2000) 7975. G. Urbina-Villalba, J. Toro-Mendoza, A. Lozsan, M. Garcia-Sucre, J. Phys. Chem. -in press (2004). J.M. Peula, A. Femandez-Barbero, R. Hidalgo-Alvarez, F.J. de las Nieves, Langmuir 13 (1997)3938. Y. Gu, and D. Li, J. Coll. Interf. Sci. 195 (1997) 343. M. Bourrel, C. Chambu, R. Schechter, W. Wade, Soc. Pet. Eng. J. 22 (1981) 28. C. Chang, E.I. Franses, Coll. & Surf. A, 100 (1995) 1. F. Ravera, M. Ferrari, L. Liggieri, Coll. Surf. A, 88 (2000) 129. R. Miller, Coll. Surf. 46 (1990) 75. K.D. Danov, D.N. Petsev, N.D. Denkov, R. Borwankar, J. Chem. Phys. 99 (1993) 7179. K.D. Danov, N.D. Denkov, D.N. Petsev, I.B. Ivanov, R. Borwankar, Langmuir 9, (1993) 1731. G. Urbina-Villalba, I. Reif, M.L. Marquez, E. Rogel, Coll. Surf. A, 99 (1995) 207. G. Urbina-Villalba, E. Rogel, M.L. Marquez, I. Reif, J. Comp. Aided Molec. Design 8 (1994)273. M. Garcia-Sucre, G. Urbina-Villalba, L. Araque-Lameda, R. Parra, Ciencia 5 (1997) 51. G. Urbina-Villalba, R. Martin-Landrove, J.A. Guaregua, Langmuir 13 (1997) 1644. G. Urbina-Villalba, I. Reif, Coll. Surf. A, 106 (1996) 175. G. Urbina-Villalba, Rev. Soc. Venez. Quim. 20 (1997) 9. A. Suzuki, N.F.H. Ho, W.I. Higuchi, J. Coll. Interf. Sci. 29 (1969) 552. F. Gielbard, J.H. Seinfeld, J. Coll. Interf. Sci. 68 (1979) 363. S.R. Reddy, D.H. Melik, and H.S. Fogler, J. Coll. Interf. Sci. 82 (1981) 116. S.R. Reddy, and H.S. Fogler, J. Coll. Interf. Sci. 82 (1981) 128. M.J. Hounslow, R.L. Ryall, V.R. Marshall, AIChE J. 34 (1988) 1821. I. Vinckier, P. Moldenaers, A.M. Terracciano, N. Grizzuti, AIChE J. 44 (1998) 951. J. Bacon, E. Dickinson, R. Parker, Faraday Discuss. Chem. Soc. 76 (1983) 165. J. Bacon, E. Dickinson, R. Parker, J. Chem. Soc. Faraday Trans. 2, 79 (1983) 91. E. Dickinson, J. Chem. Soc. Faraday Trans. II, 75 (1979) 466. E. Dickinson, Computer simulation of the coagulation and flocculation of colloidal particles, in: D.M. Bloor, and E. Wyn-Jones (Eds.), The Structure, Dynamics and Equilibrium Properties of Colloidal Systems, Kluwer Academic Press, The Netherlands, 1990. P. 707. M. Whittle, E. Dickinson, Molec. Phys. 90 (1997) 739. E. Dickinson, C. Elvingson, S.R. Euston, J. Chem. Soc. Faraday Trans. 2, 85, 891 (1989). D. Ermak, J.A. McCammon, J. Chem. Phys. 69 (1978) 1352. H. Risken, The Fokker Planck Equation, Springer Verlag, Berlin, 1984. J.K.G. Dhont, An Introduction to Dynamics of Colloids, Elsevier Science B.V., Amsterdam, 1996. Chapters 3 and 5.
718
G. Urbina-Villalba et al.
[68] T.G.M. van de Ven, Colloidal Hydrodynamics, Academic Press Limited, Padstow, 1989. Chapters 1-2. [69] W.B. Russel, D.A. Saville, W.R. Schowalter, "Colloidal Dispersions", Cambridge University Press, Cambridge, 1989. [70] J. Rome, S. Prager, J. Chem. Phys. 50(1969)4831. [71] G.K. Batchelor, J. Fluid Mech. 119 (1982) 379. [72] G.K. Batchelor, J. Fluid Mech. 74 (1976) 1. [73] D.L. Ermak, J. Chem. Phys. 62 (1975) 4197. [74] M.P. Allen, D J . Tildesley, Computer Simulation of Liquids, Oxford University Press, 1987. [75] B. U. Felderhof, Physica 89A, (1977) 373. [76] G.C. Ansell, and E. Dickinson, Chem. Phys. Lett. 122 (1985) 594. [77] M. Hiitter, Phys. Chem. Chem. Phys. 1 (1999) 4429. [78] P. Meakin, T. Vicsek, F. Family, Phys. Rev. E 31 (1985) 564. [79] M. Kolb, R. Botet, R. Jullien, Phys. Rev. Lett. 51 (1983) 1123. [80] A.M. Puertas, A. Fernandez-Barbero, F.J. de las Nieves, Comp. Phys. Comm. 121 (1999)353. [81] A.M. Puertas, J.A. Maroto, A. Fernandez-Barbero, F.J. de las Nieves, Phys. Rev. E 59 (1999) 1943. [82] A.M. Puertas, J.A. Maroto, A. Fernandez-Barbero, F.J. de las Nieves, Coll. Surf. A 151 (1999)473. [83] A.M. Puertas, F.J. de las Nieves, J. Coll. Interf. Sci. 216 (1999) 221. [84] M. Romero-Cano, A.M. Puertas, F.J. de las Nieves, J. Chem. Phys. 112 (2000) 8654. [85] S. Chandrasekhar, Rev. Mod. Phys. 15 (1943) 1. [86] G.E.P. Box, and M.E. Muller, Annals Math. Stat. 29 (1958) 610. [87] K. Zhan, J.M. Mendez-Alcaraz, G. Maret, Phys. Rev. Lett. 79 (1997) 175. [88] J. Meiners, S.R. Quake, Phys. Rev. Lett. 82 (1999) 2211. [89] M. Kollmann, G. Nagele, Europhys. Lett. 52 (2000) 474. [90] J.C. Crocker, J. Chem. Phys. 106 (1997) 2837. [91] P. Bartlett, S.I. Henderson, S.J. Mitchell, Phil. Trans. R. Soc. Lond., A 359 (2001) 883. [92] E.P. Honig, G.J. Roebersen, P.H. Wiersema, J. Coll. Interf. Sci. 36 (1971) 97. [93] H. Brenner, Chem. Eng. Sci. 16 (1961) 242. [94] A. B. Glendinning, W.B. Russel, J. Coll. Interf. Sci. 89 (1982) 124. [95] D. O. Riese, G. H. Wegdam, W. L. Vos, R. Sprik, Phys. Rev. Lett. 85 (2000) 5460. [96] G. Urbina-Villalba, M. Garcia-Sucre, J. Toro Mendoza, Phys. Rev. E 68 (2003) 061408. [97] D. M. Heyes, Molec. Phys. 87 (1996) 287. [98] P.N. Pusey, W. van Megen, J. Physique 44 (1983) 285. [99] A. van Veluwen, H.N.W. Lekkerkerker, C.G. de Kruif, A. Vrij, Faraday Disc. Chem. Soc. 83 (1987) 59. [100]W. van Megen, S. M. Underwood, R. H. Ottewill, N.St. J. Williams, P. N. Pusey, Faraday Discuss Chem. Soc. 83 (1987) 47. [101]C.W. J. Beenakker, and P. Mazur, Phys. Lett. A 91 (1982) 290. [102]C.W. J. Beenakker, and P. Mazur, Physica A 126 (1984) 349. [103]B.U. Felderhof, Physica A, 89 (1977) 373. [104]B.U. Felderhof, J. Phys. A, 11 (1978) 929. [105]M. Medina-Noyola, Phys. Rev. Lett. 60 (1988) 2705. [106] A. Einstein, Investigations on the theory of Brownian movement, Dover Publications Inc., New York, 1956. [107] J. Perrin, Atoms, Ox Box Press, Woodbridge, 1990. [108] A.M. Puertas and F.J. de las Nieves, J. Phys. C 9 (1997) 3313.
Brownian Dynamics Simulations of Emulsion Stability
719
[109]G. Urbina-Villalba, M. Garcia-Sucre, Coll. Surf A, 190 (2001) 111. [110]G. Urbina-Villalba, M. Garcia-Sucre, Interciencia 25 (2000) 415. [111]G. Urbina-Villalba, M. Garcia-Sucre, Molec. Simul. 27 (2001) 75. [112] W.D. Bancroft, J.Chem. Phys 17 (1913) 501; ibid 19 (1915) 275. [113]Rosen, M.J., Surfactants and Interfacial Phenomena, John Wiley & Sons, New York, 1989. Chapters 5, 8 and 9. [114] A.F.H. Ward, L. Tordai, J. Chem. Phys. 14 (1946) 453. [115]L. Liggieri, F. Ravera, A. Passerone, Coll. Surf. A, 114 (1996) 351. [116]X.Y. Hua, and M.J. Rosen, J. Coll. Interf. Sci. 124 (1988) 652. [117]X.Y. Hua, and M.J. Rosen, J. Coll. Interf. Sci. 141 (1991) 180. [118] R. Chanamai, D. J. McClements, Coll. Surf. A 172 (2000) 79. [119] E. Dickinson, J. Chem. Soc. Faraday Trans. 94 (1998) 1657. [120] J. Chen, E. Dickinson, and G. Iveson, Food Struct. 12 (1993) 135. [121] W. Hatton, P. McFadyen, A.L. Smith, J. Chem. Soc. Faraday Trans. 70 (1974) 655. [122]M. Boussinesq, Ann. Chim. Phys. 29 (1913) 349. [123] J.G. Rarity, K.J. Randle, J. Chem. Soc. Faraday Trans. 81 (1985) 285. [124]K.W. Lee, J. Coll. Interf. Sci. 92 (1983) 315. [125]C.S. Wang, S.K. Friedlander, J. Coll. Interf. Sci. 24 (1967) 170. [126] G. Urbina-Villalba, M. Garcia-Sucre -to be published-. [127]E.A.G. Aniansson, S.N. Wall, M. Almagren, H. Hoffman, I. Kielman, W. Ulbricht, R. Zana, J. Lang, C. Tondre, J. Phys. Chem. 80 (1976) 905. [128]D. Velegol, J.D. Feick, L.R. Collins, J. Coll. Interf. Sci. 230 (2000) 114. [129]D. Velegol, S. Catana, J.L. Anderson, S. Garoff, Phys. Rev. Lett. 83 (1999) 1243. [130]G. Urbina-Villalba, Langmuir-in press- (2004) [131]K.G. Marinova, R.G. Alargova, N.D. Denkov, O.D. Velev, D. N. Petsev, I.B. Ivanov, R. P. Borwankar, Langmuir 12 (1996) 2045. [132]H.W. Yarranton, J.H. Masliyah, J. Coll. Interf. Sci. 196 (1997) 157.
This page is intentionally left blank
Emulsions: Structure Stability and Interactions D.N. Petsev (editor) © 2004 Elsevier Ltd. All rights reserved.
Chapter 18
The rheology of emulsions H. A. Barnes Institute of Mathematical and Physical Sciences, University of Wales Aberystwyth, SY23 3BZ, UK 1. INTRODUCTION Rheology is, in elementary terms, the study of the flow properties of liquids, either simply in terms of viscosity or also relative to various viscoelastic properties. There are obviously many areas where the rheology of emulsions is very important. Examples of these range over Foods - mayonnaise, salad creams, dairy products, Personal care products - creams, lotions, Pharmaceuticals - creams, ointments, anaesthetics, Paints - alkyd resin based products, Medicine - blood, Agrochemicals - pesticides. Other examples where emulsion rheology is important in industrial situations have been cited by Tadros [1], while Miner has specifically reviewed the area of Cosmetics and Toiletries, [2]. Another interesting example of the importance of emulsion rheology is in the production of emulsion polymers, where polymer monomer droplets are formed and them polymerised to give particles, see Summers [3]. Then, the rheology of food systems, either directly or indirectly, has been seen as a determinant of the flavour release of food emulsions, either in processing or consumption, and de Roos [4] has just reviewed this area, 2. PREVIOUS PUBLICATIONS ON EMULSION VISCOSITY Studies on the effect of phase volume, shear rate and droplet-size distribution on viscosity go back to at least to 1911 [5] with papers, for example, [6], and
722
H.A. Barnes
conference proceedings on the subject appearing regularly thereafter, for instance as far back as 1935 [7] and 1949 [8]. Previous reviews of the rheology of emulsions have been produced by Richardson [9], Sherman [10], Darby [11], Pal et al [12], Tadros [13] and Barnes [14], this present review being an expansion of the latter. 3. EMULSIONS AS PRODUCTS The flow properties of an emulsion are obviously among some of its more important physical attributes in either technical or aesthetic terms. Hence the ability to measure, adjust, and, if possible, predict such properties is very important. This is therefore a challenging area for the physical and the consumer scientist (as well as the engineer) to work in. As to technical matters encountered in manufacturing, such as the mixing, pumping, filling or packing of emulsions, all require a good knowledge of the flow properties to assess mixing efficiency, power consumption, pump ratings etc. The physical behaviour of emulsions as products as they are used needs also to be understood, such as pouring or extrusion from packs, draining off surfaces, and levelling in containers. As to the many consumer-perceived attributes of a commercial emulsion, the visual and sensory properties are among the most important. Consumers have various expectations of emulsion creaminess, body and consistency for instance, and these dictate their buying preference of different products. Manufacturers have to be aware of this and should seek to meet these needs — once again an important reason for understanding the flow properties. With care, these sensory measurements can be related to quantitative physical measurements obtained on rheometers. For non-consumer emulsions such as bitumen emulsions, the flow properties as a function of temperature are obviously important, given the physical limits placed on the emulsion because it has to be spread. There is therefore an upper limit on the viscosity, but equally, the ability to minimise the viscosity for a given concentration or, vice versa, to maximise the concentration for a given viscosity is very important. In qualitative terms, emulsions range from low viscosity, milk-like Newtonian liquids through thicker shear-thinning liquids, right up to thick, cream-like materials with apparent yield stresses. In terms of microstructure, the one common element is the possession of a liquid continuous phase together with a dispersed phase of liquid of - at least - deformable drops. The interface between the drops as well as being the source of stability of these otherwise unstable systems can in many ways shield the internal liquid from velocity gradients in the continuous phase liquid. Some situations can be envisaged here
The Rheology of Emulsions
723
the relative effective 'stiffness" of the interface or else the interior phase viscosity is so high that the system behaves like a dispersion of solid particles. This is more often the case with small, sub-micron droplets. 4. WHAT IS AN EMULSION IN RHEOLOGICAL TERMS? As Miner [2] has said 'An emulsion in its simplest form is a two-phase system containing two immiscible liquids, one of which is dispersed in the other in the form of microscopic or sub-microscopic droplets. The two phases are usually oil and water, producing either an oil-in-water (o/w) or a water-in-oil (w/o) emulsion ... The presence of an emulsifier as a stabiliser extends the inevitable separation of the phases and frequently is a major contributor to the flow properties of the entire system'. Plenty of examples are available of oil-in-water, water-in-oil and even oilin-oil emulsions. However, the widest possible definition of an emulsion that makes sense Theologically is that of a dispersion of deformable particles in a continuous liquid phase. This will include, at the one extreme, systems where the interior of a disperse particle is simply a low-viscosity liquid - possibly lower than that of the continuous phase - with the surface stabilised with a suitable surfactant - as essentially described by Miner above. Then the continuous phase in such a system could be a low-viscosity Newtonian liquid such as water. At the other extreme, we might consider a system with multilamellar dispersed particles, in a non-Newtonian continuous phase, where the non-Newtonian properties could simply be due to an excess of the stabiliser. In between we could consider multiple emulsions; barely deformable particles, and systems with very strong interactions between the particles that are for instance highly flocculated. Following this wide definition, blood is a special case of a water-in-water emulsion where the interest centres on the form of the emulsifying layer and its interaction with the layers of neighbouring blood cells. All these systems find their application in natural and manmade systems, and their rheology is very often important - if complicated! Emulsions may be the desired end product in manufacture, or they might form important intermediary states. Emulsion polymerisation is a popular route in polymer particle manufacture, and the state of the emulsion, including its rheology-dictates the polymer particle formation. Emulsions might be the starting point, e.g., dairy products, then desired rheological properties are sought such a creaminess on concentration, and particle formation in spray drying. All these have rheological ramifications. In other cases, as in blood, emulsion properties are of direct interest.
724
H.A. Barnes
Sometimes, the word emulsion is a misnomer, as for instance in so-called emulsion paints, which are actually dispersions of solid particles. In very complex products such as skin lotions, emulsion droplets only form part of the product that might contain polymers, particles and concentrated surfactant structures. Nevertheless an understanding of their particular role is helpful in understanding the overall rheology, especially if the droplets change over time by coalescence or disproportionation (see below). Of course, in complex emulsions, the composition of the continuous phase may not be known in advance. Even in what appear to be simple emulsions, the excess emulsifier present in the continuous phase might itself be associated in micellar-type structures that can have a profound effect on the overall viscosity. In really complex emulsions, with added ingredients partitioned between the continuous phase and the disperse phase, and possibly being present in multilayered emulsifiers, the only way to convolute such complexity and establish the composition of the phases is to centrifuge or ultracentrifuge the emulsion. This will make it possible to establish the composition of the continuous phase. Further work is needed to establish the nature of the droplets and the emulsifier layer, but this can be often affected using microscopy. 5. SPECIFIC TYPES OF FLOW There are two basic kinds of flow with relative movement of adjacent particles of liquid; they are called shear and extensional flows. In shear flows liquid elements flow over or past each other, while in extensional flow, adjacent elements flow towards or away from each other, see Fig. 1 for illustrations of shear and extensional deformation and flow respectively, as seen in terms of emulsion droplets. For a given velocity, the resulting force increases when the viscosity is increased, whereas for a given force, the velocity is reduced when the viscosity is increased. As we have seen in Fig.l, simple shear flow is the continual movement of emulsion droplets over or past each other, while extensional (or elongational, or stretching) flows are where droplets move towards or away from each other (see below). In Fig. 2 we visualise shear flow alternatively as the movement of hypothetical layers sliding over each other. In the simplest case the velocity of each layer increases linearly with respect to its neighbour below, so that layers twice the distance from any stationary edge move at double the speed, etc. The gradient of the velocity in the direction at right angles to the flow is called the shear rate (sometimes called the velocity gradient or strain rate), and the force per unit area creating or produced by the flow is called the shear stress. In our simple example, the shear rate is V/h and is described by the symbol y (usually
The Rheology of Emulsions
725
Fig. 1. Emulsion droplet motion in shear and extensional flows.
Fig.2. Hypothetical layers in shear flow. referred to as 'gamma dot'), while the shear stress is given by F/A and is given the symbol a (sigma). The shear rate has the units of velocity divided by distance, i.e. metres per second / metres, leaving us with the units of reciprocal (i.e., one over) seconds, or s"'. Shear stress—force per unit area—has the units of newtons per square metre, N m"2, but in the SI system, stress, like pressure, it is given in units of pascals, Pa. The approximate value of the shear rate encountered in a wide variety of circumstances found in situations important for emulsions is shown in Table 1. Readers may relate these numbers to their own field of interest by simply dividing a typical velocity in any flow of interest by a typical dimension. An example of this is the average velocity of a liquid flowing in a pipe divided by the pipe radius, or the velocity of a moving sphere divided by its radius.
726
H.A. Barnes
Units that you will come across most frequently in rheology are shown in Table 2. Using the same diagrams we can also define deformation rate, where now the deformation S(delta) is continuous and V= dyldt, and y = dV/dh, and is called the shear rate. The dot above the deformation symbols is used to signify differentiation with respect to time, and follows Newton's calculus convention ('Newton's dot'). Hence y is an amount of shear, and y is the rate of shear, or shear rate
y , (usually referred to as 'd/by dt'), which is sometimes called the
rate of strain or strain rate. Note that although it is good practice to write shear viscosity to distinguish it from extensional viscosity (see below), it is usually referred to simply as viscosity. The Greek symbol y (gamma) represents deformation in simple shear, this is illustrated in Fig. 3, where shear stress a (sigma) is also defined. Table 1 Typical shear rate ranges of various physical operations in emulsions.
Situation
Shear Rate Range / Examples s"1 Slow creaming of fine 10"6- 10"J Long-term storage of emulsion droplets emulsions Levelling due to 10"2-10"' Creams in open surface containers Tension Lotions on hands Draining under gravity 10"'-10 1 Spooning, pouring IO'-IO2 Creams 3 General manufacturing Mixing and stirring io'-io 3 10°-10 Pumping suspensions Pipe flow 1 0 3 - 104 Skin creams Rubbing Spray drying of milk Spraying 105-106 Table 2 Some commonly used rheological quantities and their units. Note that the units of viscosity are Pa.s and not Pasecs!
Quantity Shear Shear rate Shear stress Shear viscosity
Symbol y (pronounced gamma) y (pronounced gamma dot)
Units s"1 Pa Pa.s
The Rheology of Emulsions
727
Fig. 3. Definition diagram for shear flow.
Fig. 4. Definition diagram for extensional flow.
Similarly, for uniaxial extension the proper measure of deformation is ^(epsilon) and stress (sometimes called tension) <je as shown in Fig. 4. For Newtonian liquids in simple shear flows we can write
(1)
For uniaxial extensional or stretching flows, the equivalent symbols are ae sand rje and then for Newtonian liquids, tfe = TL£.
(2)
728
H.A. Barnes
We have given the description of extensional flow, but we shall not pursue the matters further because few emulsion systems show exceptional behaviour in this form of flow. 6. SHEAR FLOW PROPERTIES The measurement of flow properties of any liquid is divided into linear and nonlinear behaviour (see Barnes et al, [15]) for a detailed description of this subject). The former is measured within that range of stresses, strains and shear rates where the rheological properties of the emulsion being measured are not a function of these imposed variables, but are only functions of frequency and time (as well as of course temperature and to some extent pressure). These measurements include the linear viscoelastic moduli such as the frequencydependent storage and loss moduli, G' and G" (or the various combinations of these such as complex modulus G* and the phase lag — usually presented as tan S), and the time-dependent creep compliance as the more commonly measured parameters. These properties are usually confined to that range of stresses and strains below any perceivable visual movement during measurement. Although these linear properties are not generally applicable to large strain and stress applications such as in-use situations, they are nevertheless useful in assessing the microstructure and even possibly the long-term stability of an emulsion. On the other hand, the non-linear properties are a function of the applied stress, strain or shear rate. Although they include the non-linear versions of the moduli and compliances mentioned above, the parameters we are most interested in are the viscosity and possibly normal stress differences although as we shall see that the latter are usually small for most emulsions. Emulsions in common with other dispersed systems can show various artefacts when being measured in rheometers and viscometers, including those due to inertia and slip, and before we can conclude that the measured data are unambiguous we must eliminate or account for these effects. Also, care has to be taken that data are taken in a situation where the shear rate is constant in the flow geometry. This is only found in small (less than 4°) cone-and-plate apparatus or narrow-gap concentric-cylinder geometries with the gap-to-radius ratio no less than 0.95. As the latter system is not always used, intermediate calculations have to be made before the true viscosity—shear rate relationship is found. Much use is still made in industry of geometries that violate this principle but happen to be easy to use, with the often-used Brookfield-type rotating bob or disc immersed in a beaker of liquid being a typical example. Traditionally, measurements in the non-linear range have been made on strain-controlled instruments such as the Rheogoniometer or the viscometers of the Haake range, but the introduction of the stress-controlled rheometers during the last decade or so has revolutionised the measurement of dispersed systems.
The Rheology of Emulsions
729
7. DETAILED CONSIDERATION OF TYPICAL VISCOMETERS Viscometers are instruments that can either apply a force and measure a speed, or apply a speed and measure a force, to or from a simple geometry. This might be as simple as a U-tube viseometer that measures the time taken for a gravity-driven flow to move from one vertical position to another, or a similar situation for flow from the hole at the bottom of a carefully manufactured cup called a flow-cup, see Fig. 5. In both these situations, the shear stress experienced by the liquid being tested varies • in time as the vertical driving-force reduces; • in space because the flow is quite different at different places, and • in kind because the flow-cup is not a simple shear flow with the liquid being 'squeezed' into the hole as an extensional flow. These cups are perfectly adequate for Newtonian liquids—once a correction has been made for the effect of fluid density—but because of the variations mentioned, the sensible measurement of viscosity for non-Newtonian liquids becomes impossible. However, if we want to apply a more well-defined situation, such as a fixed shear stress or shear rate everywhere in the liquid, we need more control. Originally, such controlled viscometers were based on an applied shear stress which was generated by a weights-and-pulleys arrangement, as shown for instance in Fig. 6. Limitations were soon found with this type of instrument in terms of attaining low shear rates—due to bearing friction—as also were the limitations of
Fig. 5. A schematic diagram of a simple U-tube viseometer and a flow cup.
730
H.A. Barnes
Fig. 6. A schematic diagram of the 1912 Searle controlled-stress, concentric-cylinder viscometer.
prolonged shearing at high shear rates—the finite length of the string! These methods were then superseded by electrically driven motors and the (what we now call) controlled-strain instruments had arrived, largely replacing the controlled-stress variants. These became more and more sophisticated, and eventually with logarithmic mechanical gear boxes were able to span a shearrate range typically from ~ 10 "4 s"1 to ~10 4 s"1 (e.g. the Weissenberg Rheogoniometer). 8. COMMERCIAL CONTROLLED-STRESS VISCOMETERS In the mid-1970s, a new generation of controlled-stress rheometers began to appear, as opposed to the controlled-strain type of viscometer commonly used in laboratories, see Barnes and Bell, [16]. Jack Deer and colleagues at the London School of Pharmacy, who used air bearings and an air-driven turbine to provide the torque, had developed the first of the modern instruments. It was claimed that 'with such an instrument the application of a known and controlled stress provides the rheologist with vital information of this critical region in the form of a "CREEP" curve'. These claims could be made because they had introduced a 'specially designed air bearing ... and an air turbine drive system for the application of torsional stress that is independent of rotational speed ... throughout the operational range'. (Air bearings and air turbines had been and still are used for dentists' drills, but of course at much higher speeds, hence the whine!) Using a pre-commercial version of the air-turbine rheometer in the creep mode, it was found that the 'creep' (i.e. zero shear) viscosity of wool fat BP (an emulsion component) at 25 °C over a shear rate range of 1.76 - 6.14 x 10" 6 1 s" was constant (to within a small experimental error) at ~ 2 x 106 Pa.s.
The Rheology of Emulsions
731
Then around 1980, commercial versions of the new generation of electrically driven controlled-stress rheometers began to appear, still based on air bearings that greatly reduced friction, but now using so-called drag-cup electrical motors that allowed controlled stresses to be more easily applied, but still independent of rotational speed. Along with these features came new ways of measuring smaller and smaller rotation and rotation rates. The use of the latest optical-disc technology now means that rotation rates as low as 10"8 rad/s (~ 1 revolution in 20 years!) can be measured. Access to these ultra-low shearrate regions is now called creep testing, by analogy to the testing of solids under similar low-deformation-rate, long-time conditions; albeit solids creep testing is usually performed in extension rather than in shear. 9. EVOLUTION OF VISCOMETERS
COMMERCIAL
CONTROLLED-STRESS
Table 3 compares the specifications of some typical controlled-stress viscometers as they have evolved over the last thirty years or so. It is clear that an amazing increase in sensitivity and range has been brought about, especially in the lower-and-lower minimum torques (and thus stresses) and rotation rates (and thus shear rates) achievable. Table 3 Some typical commercial controlled-stress rheometer specifications. Date Typical Instrument Torque (N.m) Min. Max. Resolution
-1970 Air Turbine Rheometer
-1978 Deer Rheometer
- early 1980s - l a t e 1980s Carrimed Carrimed Mkl C S L 100
-1999 TA Inst. AR 1000
io- 4
10 b
io- 6
2
2
2
10"6 10" 2
10"' 0.1
io- y
io- 9
IO
10"
10"4
10" 10"'
Ang. Veloc. (rad/s) Min. (creep) Max. Resolution
-
-
-
-
I0' 8
50 -
50 -
50
50
lO' 2
lO"4
100 -
Creep (strain) Resolution
2 x 10"2
2.5 xlO" 3
2.5 x l O ' 4
10° -
6.2 x l O " 7 1300
Max.
-
-
-
732
H.A. Barnes
10. NON-SIMPLE VISCOMETER GEOMETRIES There are a number of popular viscometer geometries used for emulsion measurement where the shear rate is not the same everywhere. In order to convert the basic experimental data into unambiguous viscosity /shear-rate data, an intermediate calculation step is needed, see Barnes [17]. This uses an assumption about the liquid, usually that at any particular value of shear rate, the local viscosity/shear-rate data can be described by a power-law-type behaviour, where the slope of the log/log curve is given by n'. (For true power-law liquids, this is the same as n, the power-law index.) The following are the necessary equations for wide-gap concentric cylinders; the parallel-plate geometry and tubes used as viscometers. In each case the viscosity data is related to a certain shear rate calculated at some fixed point in the geometry, and n' is related to the basic measured parameters. 10.1. Wide-gap concentric cylinders The software provided by some viscometer manufacturers for calculating viscosities in concentric-cylinder geometries only makes use of the narrow-gap approximation (you should check this in your manual). However, this only applies to gaps of at most a few percent of the outer-cylinder radius, and assumes that the shear rate and shear stress are approximately constant throughout the gap. This is not the case in many practical cases, and if for instance the inner cylinder is being rotated, the value of the shear stress decreases as the inverse square of distance from the centre of rotation as one moves away from the inner cylinder. If non-Newtonian liquids are being measured and the narrow-gap approximation is used, the data from different-sized gaps does not coincide. However, the data can be corrected by the use of a number of lengthy analytical methods. However, a simple correction of the data can be made as follows, using an empirical correction factor worked out by the author, for the viscosity at the inner-cylinder, uncorrected shear rate: Tl = T W [6+H 1 (l-&)]
(3)
where rj is the correct viscosity at the inner-cylinder shear rate, r/ng is the calculated viscosity based on the narrow-gap approximation, n' is the local powerlaw index at that shear rate at which the viscosity is being evaluated, calculated as dlog77dlog&> where T and co are the torque and rotational speed respectively, and, b is the gap ratio (inner-cylinder radius divided by the outer-cylinder radius). When this formula is used, data for different gap ratios, b, which would have
733
The Rheology of Emulsions
originally appeared as a set of different curves if calculated using the narrow-gap approximation now lie on a single curve. Table 4 shows the viscosity correction factors predicted by our simple equation for a range of possible values of b and n. Complete equations for the evaluation of the viscosity in the wide-gap situation should be used for very shear-thinning liquids (say n' < 0.3), i.e., i=
2u
2 ,
(4)
n(l-b") and
,= ^#i
(5)
4TT^ZLO
where co is the rotation rate of the inner cylinder, C is the couple on the inner cylinder, and Tj is the radius of the inner cylinder of length L. 10.2. The rotating parallel-plate viscometer In a parallel-plate viscometer geometry, the upper plate is usually rotated and the couple C is like-wise measured on the upper plate. The shear rate varies from zero at the centre to aco/h at the edge of the plates, where a is the plate radius, h the gap and co the rotation rate in rad/s. The shear rate at the edge is given by (6>
^ = IT ' h Table 4
Viscosity correction factors for non-Newtonian emulsions measured in wide-gap concentric cylinder viscometers.
Values of b 0.9 0.7 0.5
n' = 0.8 0.98 0.94 0.90
«' = 0.7 0.97 0.91 0.85
«' = 0.5 0.95 0.85 0.75
734
H.A. Barnes
and the viscosity (evaluated at the edge shear rate ya) is given by 3Ch 11
o < (, ^ 1 d\ogC)' 2-TW2 UJ 1 H
{
—
3 JloguJ
For a power-law liquid, dlogC/dloga> is simply n. 10.3. Pipe/tube viscometer Tube or pipe viscometers take many forms, but they should all be able to give the pressure-drop P as a function of flow rate Q for situations where the tube is long enough to be able to neglect entrance and end effects, say L/a > 50. In this case we can calculate the viscosity as a function of the wall shear rate, ^w, which is given by
X
"^3(4
4JlogP •
(8)
The viscosity r/ is then given by T,
=
^
T
(9)
dlogP) where, for power-law liquids, dlogQ/diogP is simply IIn. 11. SLIP AND OTHER WALL EFFECTS IN RHEOMETERS One serious problem that can occur in any flowing system with concentrated, dispersed entities, such as emulsion droplets, is that because of physical depletion of droplets near a smooth surface, the resulting lower viscosity in the immediate vicinity of the smooth surface results in effective slip at the wall [18]. This depletion arises simply because droplets cannot penetrate solid walls, and are displaced out from the walls. This leaves an effective lubricating layer, and this in turn means that the overall indicated measured viscosity is lower than it should be, but the amount of error depends on the size of the viscometer gap. For small gaps the problem is larger.
The Rheology of Emulsions
735
Fig. 7. Haake viscometer SV geometries, showing the six-bladed vane (SV2FL), and the corresponding serrated (SV2P) and smooth (SV2) cylinders.
A further complication, even for dilute emulsions, arises because droplets moving near walls always tend to move away from the wall [18]. Special attention has been given by Gallegos and Franco [19] to slip effects in the rheology of rheology of food, cosmetics and pharmaceuticals. All these matters are discussed and expanded upon by Barnes [18]. One way of eliminating the problem for large gaps is to use roughened or serrated geometries, or even using a vane geometry. Fig. 7 illustrates a typical solution, where a range of modified cylinders is used in the Haake range of commercial viscometers, see www.thermohaake.com. The other solution to the problem is more long-winded, but effective. This is to use a number of geometry sizes and extrapolate the results to an infinite size gap, see Barnes [18] for details. 12. TYPICAL COMMERCIAL LABORATORY VISCOMETERS Standard laboratory viscometers have a shear rate range ~ 0.1-1000 s"'. At the moment these include the Brookfield DV3 HaakeVT7, VT500/550, RV1 Bohlin Visco 88 Rheometrics RM 180, 265 Reologica ViscoCheck PhysicaMCl
736
H.A. Barnes
Typical output: viscosity/shear-rate as a function of time of shearing for the shear rate specified, so that the general flow curve and thixotropy can be studied for a range of temperatures, usually in the range -10 to 100 °C. 13. TYPICAL COMMERCIAL VISCOMETERS
CONTROLLED
-
STRESS
The following are typical examples of controlled stress viscometers that are capable of producing viscosity data at low shear-stress/ shear rate. Bohlin - CS, CVO 50/120, DSR CSR, CVR Brookfield - RSCPS Haake-RSI, RS 75/150 Physica - DSR 4000, UDS 200 Reologica - STRESSTECH/HR, VISCOTECH Rheometrics - SR5/ 2000/5000, ARES (controlled-rate instrument using feed back) TA(Carri-med) - AR 1000, CSL2 500 14. THE SHEAR VISCOSITY OF EMULSIONS We will now consider the typical forms of flow curves found using the above viscometers, when any artifacts have been eliminated our accounted for. The simplest flow curve is found for dilute emulsions, where if we plot the basic measured parameters of shear rate and shear stress, we obtain a linear relationship, as illustrated in Fig. 8. This data may be replotted as viscosity (shear stress divided by shear rate) as a function of shear rate, and of course the viscsoity is independent of the shear rate, i.e., the emulsion of interest is Newtonian, Fig. 9. If we now increase the oil-phase concentration for our model emulsion, we see non-Newtonian behaviour setting in. First if we repeat our first plot, we still see a kind of linear response if we plot data over the range -300 - 3500 s"', but now the curve is displaced vertically, and wew see what appears to be an intercept on the stress axis. This is called Bingham behaviour, and is charaterised by a yield stress (the extrapolated intercept) and a plastci viscosity, the slope of the curve - see Fig. 10. If we repolt the data for our higher concentration emulsion in a slightly different way, we can extent the straight-line/intercept behaviour down to lower shear rates. This is done by the simple stratagem of plotting the square root of shear rate againt the square root of shear stress - this is called a Casson plot, see Fig. 11.
The Rheology of Emulsions
Til
Fig. 8. Shear rate versus shear stress for 30, 40 and 50% respectievly for the model emulsion, plotted on a linear basis: Newtonian viscosities are 0.0030, 0.0044 and 0.0095 Pa.s respectievly.
Fig. 9. Shear rate versus shear stress for 30, 40 and 50% respectievly for the model emulsion, plotted on a logarithmic basis: the Newtonian viscosities are 0.0030, 0.0044 and 0.0095 Pa.s respectively.
The data over a limited range of shear rates also fits another well-known flow law - this is called the so-called power-law, and is illustrated in Fig. 12, where the shear rate is plotted against the shear stress, with both axis plotted on a logarithmic basis.
738
H.A. Barnes
Fig. 10. Shear rate versus shear stress for the 75% model emulsion, plotted on a linear basis. The straight line shows the best fit for the Bingham model, which fits reasonably well over for selected data over the range ~ 500 s" to ~ 3,300 s" , with a plastic viscosity of 0.0477 Pa.s and a Bingham yield stress of 25 Pa.
Fig. 11. Square root of shear stress plotted against squart root of shear rate, showing the goodness of fit of the data for the 75% model emulsion to the Casson model for data over the shear rate range ~ 80 s"' to ~ 3,500 s"'. The Casson yield stress is 6.25 Pa and the viscosity parameter is 0.04 Pa.s.
If we now plot the complete data set of flow curve points obtained over a wide shear rate range, we find that the behaviour looks quite different, see Figs. 13 and 14. Now, although we find that the emulsion shows Newtonian behaviour at very low and very high shear rates, there is a transition in between, where non-Newtonian behaviour is observed.
The Rheology of Emulsions
739
Fig. 12. Shear rate versus shear stress for selected data for 75% model emulsion, plotted logarithmically, with the solid line as k = 2.54 Pa.s" and n = 0.3285.
Fig. 13. Shear rate versus shear stress for the 75% model emulsion, plotted on a logarithmic basis, showing all the data. The Newtonian asympotes represent viscosities of 75 and 0.055 Pa.s respectively.
This transition shows up clearly on a plot of viscosity versus shear rate - this behaviour is called 'shear thinning'. Yet another way of showing the data is to plot viscosity as a function of shear stress - see Fig. 15. This is quite important for predicting the behaviour of emulsions under the action of a given force, e.g., gravity or an imposed pressure.
740
H.A. Barnes
Fig. 14. Viscosity versus shear rate, plotted on a logarithmic basis, showing all the data. The Newtonian asympotes represent viscosities of 75 and 0.055 Pa.s respectievly.
Fig. 15. Viscosity versus shear stress for the 75% model emulsion, showing all the data. The Newtonian asympotes represent viscosities of 75 and 0.055 Pa.s respectievly.
15. THE OVERALL STEADY-STATE FLOW CURVE Given that we can measure the flow curve of an emulsion, how do we use it? It is possible to associate various parts of the flow curve with particular operations in emulsion use: Fig. 16 shows these areas in terms of shear stress, which may also been related to the pertinent (approximate) shear-rate ranges as -
The Rheology of Emulsions
741
Shear stress, a (log scale) Fig. 16. Graphic representation of the various parts of the flow curve of a typical emulsion categorised in terms of the various end-use operations. . . . • .
separating out levelling out squeezing out pouring out rubbing out
10"6 - 10~4 s"1 10" 2 -10"'s"' 10° — 102 s"1 10 — 102 s"1 102 - 1 0 4 s " '
Shear stress, a / Pa Fig. 17. The viscosity/shear-stress curves of various consistencies of real emulsions, showing the viscosities relative to the viscosity of water at 10" Pa.s.
742
H.A. Barnes
Fig. 18. Typical plot of the linear viscoelastic moduli versus frequency for emulsions.
The various kinds of concentrated oil-in-water emulsions possible may be seen by consumer as a lotion, a cream or a paste, and the kind of flow curves these follow is shown in Fig. 17. The mathematical forms of these curves are manifold, and Barnes has given an outline of these in ref. [17]. 16. LINEAR VISCOELASTIC PROPERTIES The frequency-dependent storage and loss moduli, G' and G", for typical emulsions usually give the kind of behaviour depicted in Fig. 18, with the storage modulus usually lower than the loss modulus, but with the possibility of it being greater for higher frequencies and higher concentrations of dispersed phase, to give typical gel-like behaviour. For a specific example, van Vliet et al. [20] cite data on an oil-in-water emulsion stabilised by a polyelectrolyte. Although of little or no relevance to real flows, the linear properties are very useful in assessing the degree of interaction of droplets. 17. NORMAL FORCES Normal forces arise because in liquids with a strong elastic component, a tension arises along the streamlines. Al-Hadithi et al [21] have shown that the viscoelastic properties of disperse systems shows significantly less overt elastic properties that polymeric systems. Among other things this means that such extravagant features such as extrudate swell, rod-climbing, etc. are not usually seen in emulsions. However, it is possible to have systems that are technically emulsions, but which may be essentially polymeric in nature, such as droplets dispersed in concentrated polymeric solutions or even melts, where the dispersed drops are sometimes phase-separated entities.
The Rheology of Emulsions
743
Normal forces can be measured, with care, in emulsions, but they are always small (unless the continuous phase is itself very viscoelastic). Thus for all practical purposes normal stresses—and thus overt elastic effects like die swell and the Weissenberg effect—are irrelevant for most emulsions. 18.THE DEVELOPMENT OF VISCOSITY IN EMULSIONS 18.1. General comments The development of viscosity in dispersions arises from the deflection of streamlines caused by the suspended entities in flowing liquids. They can rotate, and they can interact, but the fact that they are deformable does not become important until the droplets are large (> 1 micron) and the phase volume is quite high. In the dilute state the fluid forces cause deformation of the droplets and induce internal flows in the droplets. When the phase volume is large, then the particles themselves cause mutual deformation, and the phase volume can be much higher than the equivalent solid-particle dispersion. The rheology of emulsions merges with that of dispersions (of solid particles) when the emulsion droplets are very small, and are not very deformable, and especially when the phase volume is small and the particles are flocculated. If we have small droplet size and flocculated droplets, the emulsion
Fig. 19. Two examples of commercial mayonnaise, probably flocculated: data from Jason Stokes, Unilever Research, Colworth House, UK.
744
H.A. Barnes
is very shear thinning, and in the linear viscoelastic region, the behaviour is very solid-like. These two factors are themselves strong evidence for flocculation. A good example of this situation is shown in Fig. 19 which shows the flow curve of two commercial mayonnaises. 18.2. The effect of specific parameters The basic rheology-determining parameters of an emulsion are (i) the continuous phase rheology; (ii) nature of the particles, size distribution, deformability, internal viscosity, concentration: (iii) the nature of any particle colloidal interactions. 18.3. The role of the continuous phase All theories that seek to describe the effect of a disperse phase on the overall viscosity of an emulsion, rj, show that the outcome ends up as a modification of the continuous phase, as — = /(parameters)
(10)
where rjc is the viscosity of the continuous phase. One obvious implication of this is that any change to the continuous-phase viscosity will have the same effect on the overall viscosity. So, doubling the viscosity of the continuous phase will—all else being equal—double the overall viscosity. Thus, controlling the continuous phase viscosity is very important. The effect of temperature on the viscosity of an emulsion has to be looked at in this light, so that again—all else being equal—any change in viscosity of the continuous phase brought about by temperature will reflect in exactly the same way in the change of the overall viscosity. Of course, temperature will also change other parameters that might cause extra changes. These include changes in the internal viscosity of the droplets, the deformability of the droplets, and specific colloidal interactions. 18.4. Thickening the continuous phase Quite a number of Theologically active materials such as polymers (artificial or natural), inorganic thickeners (clays and silicas) and surfactants can be used to thicken the continuous phase of emulsions. A useful list of these is shown in table 5, following Barnes [17]. This kind of thickening will give an emulsion more 'body' and will slow down and possibly eliminate any creaming. The thickening potential of various 'natural' thickeners is shown in Fig. 20. As is obvious from this figure, the greatest difference between them - as assessed at a given weight percent and in water only - is the thickening potential at very low
The Rheology of Emulsions
745
Fig. 20. The steady-state flow curve of various aqueous solutions of 1% by weight of'natural' polymers [Barnes, 17].
shear rates, i.e. in the region regulating creaming. Care has to be taken in adding these kinds of material, lest accidental separation takes place via the kind depletion flocculation mentioned elsewhere in this chapter. 18.5. Surfactant-rich emulsions If an emulsion is physically stable when excess surfactant is present—so no depletion flocculation, see later—then the surfactant molecules will often associate, even at relatively low concentration, and form structures where they increase the viscosity far more than expected from mere phase volume considerations. The kind of structures possible are shown in Fig. 21. Many personal cleaning products now have emulsified oils as an added benefit, and in this case the emulsion droplets are added to the surfactant. The ensuing complex mixture might be difficult to analyse and Centtfifugation may have to be undertaken to separate the continuous and dispersed phases for analysis. The websites of the suppliers cited in Table 5 are BASF, www.basf.com Cabot, www. cabot-corp. com/cabosil Dow, www.dow.com FMC, www.avicel.com Noveon, www.carbopol.com/ Hercules, www.herc.com/foodgums Kelco, www.cpkelco.com/ Laporte, www.scprod.com Rheox, www.rheox.com
746
H.A. Barnes
Table 5 Thickeners for water-based products - note that in the table 'PP' is short for personal products, 'pharm.' is short for pharmaceutical. Thickener Systems 'Naturals' Alginates Carrageenan Guar gum Locust bean gum Xanthan gum Celluloses Methylcellulose
Example/ supplier
Particular Limitations
Typical Product areas
Comments systems
Kelcosol* Kelco Viscarin® FMC Jaguar® Rhodia Meyprodyn Rhodia Kelzan® Kelco
Electrolytes, Cations K. or Ca for gelling -
Food
Anionic, calcium! Anionic
avoid
Food, pharm., Nonionic PP cationic Food, PP Nonionic
or
-
Not good for Food, pharm., PP, household cationics
Carboxymethyl cellulose (CMC)
Methocel ~ Dow Aqualon Hercules
Hydroxyethyl cellulose (HEC)
Natrosol Hercules
®
Hydroxypropyl methylcellulose Organics Associative thickeners
Benecel Hercules
® -
Polyacrylic acid Polyethylene oxide Polyvinylpyrrolidone Inorganics Clays Fumed silica Organoclays
Food, pharm.
-
Food, pharm.
on 1 1
Anionic
Nonionic
Food, pharm., Sodium salt good thickener PP, household -SCMC Food, pharm., Good PP, household electrolyte tolerance Food, pharm., Nonionic PP, household
Aery sol® Rohm and Haas Carbopol Not good for Noveon cationics Polyox ® ShearUnion sensitive, Carbide stringy Kollidon® Yellowish BASF on heating
PP, household, Certain grades industrial good for bleaches/acids Pharm., PP, Sensitive to household electrolytes Pharm., PP, Nonionic household, industrial Pharm., PP, Cross-linked by household strong alkali
Laponite® Not good for Laporte cationics Cab-O-Sil ® Cabot Bentone® High shear Rheox to disperse
Pharm., household Pharm., household Pharm., household
Good for bleaches PP, Thickens nonaqueous liquids PP, Some thicken organic liquids
PP,
The Rheology of Emulsions
1A1
Rhodia, www.food.us.rhodia.com Rohm and Haas, www.rohmhaas.com Union Carbide,www.dow.com/merger/'specpoly.htm The spherical micelle form causes a little more thickening of the continuous phase than would be expected from the amount of surfactant, but the worm-like micelles act as 'living polymers' and cause considerable thickening due to their intertwining. The lamellar phase sheets or droplets can also cause significant thickening, the former due to the extended shape of the sheets, and the latter particularly due to the excess water trapped between the successive surfactant layers. These facts are often forgotten in trying to understand the viscosity of surfactant-rich emulsions. 19. THE ROLE OF THE DROPLET PHASE IN DETERMINING OVERALL VISCOSITY 19.1. The effect of phase volume The simplest way to proceed is to limit ourselves to the situation where we have essentially an emulsion made up of non-interacting, monodisperse, spherical solid droplets. Small droplets with a thick, stiff stabiliser layer and a high internal-phase viscosity best approximate to this simplification. The rheology of such systems has been well researched and will serve as a good basis for progress. We might state here that progress on emulsion research is at
Fig. 21. Various forms of surfactant 'superstructure'.
748
H.A. Barnes
present hindered by the paucity of data on well-defined monodisperse emulsions. The viscosity of a dispersed system of effectively solid droplets is well described by a simplified form of the so-called Krieger-Dougherty equation: n = Tici-f-
(ii)
where again 77 is the viscosity of the emulsion (usually defined at a shear rate); rjc is the viscosity of the continuous phase (usually but not always constant); <j> is the phase volume of the dispersed phase and <j>m is the maximum phase volume when the viscosity diverges to infinity. The term [rj\ is called the intrinsic viscosity, and accounts for particle shape effects. This relationship shows that (i) the sensitivity of the viscosity to that of the continuous phases is multiplicative not additive, so that the effect of, for instance, temperature is, all else being equal, pro rata, see above for a more detailed consideration, (ii) the sensitivity to phase volume becomes very important for
ri = TIC l - - £ -
(12)
This approximation has been given the name MPKQ model by Barnes [22]. The non-Newtonian nature of dispersions can be simply accounted for by a change in microstructure. This is accounted for by the flow-induced movement from a three-dimensional, isotropic, random distribution of particles at rest towards a more two- dimensional ordered distribution at higher shear rate, with the latter being often a system of strings and layers of particles. The former has a lower <j>m, being around 0.62, compared with just above 0.7 for the latter. This seemingly small difference has a very large effect in dispersions of commercial interest. The effect of the shear stress on \l
The Rheology of Emulsions
(i/Q-(i/
i
=
(l/^,o)-(l/^)
749
n3) m
l + {Bo)
where the subscripts on \l<j)m relate to the zero- and infinite-stress asymptotes respectively; B has the dimensions of inverse stress to make its multiple with (the shear stress) dimensionless and m is a dimensionless number. The restoring force for a dispersion to return to a random isotropic situation at rest is either Brownian (thermal fluctuations) or osmotic, the former being most important for sub-micron particles, and the latter for larger particles. These are contained in the B term in the above equation. Changing the flow conditions changes the structure, and this can take time, whether going from higher to lower, or lower to higher, shear rates/shear stresses, and this accounts for thixotropic effects, which are especially strong in flocculated systems (see later). As (/>m essentially represents the way the particles fill space, we may here consider the other space-filling considerations. The first has to do with the droplet size distribution: anything that widens the distribution increases m and thus decreases the viscosity. Bimodal and trimodal distributions of particles of a given size have been frequently studied, and some studies have been reported of the effect of continuous size distributions (such as log—normal) and a few about the effects of mixing continuous distributions. As we said, anything that makes the droplet size range wider and more efficient as regards space filling decreases the viscosity. Particle deformability can be accounted for in >m that the packing becomes more efficient as particles distort to accommodate their neighbours. One can easily envisage the situation where very deformable particles can he completely space filling. 19.2. Effect of the internal phase viscosity Theory covering very dilute model emulsions—with no emulsifier layer— shows that predicated viscosity depends on the viscosity ratio of the dispersed to the continuous phase, r/r, and is of the form 1+ - T J
-^ = 1+ ilc
2__ ^.. .
(14)
1 + r\r
In the limit, this shows that very viscous increases when rjr becomes is large (> 10), the solid particle, 'Einstein' situation pertains, and the inner term in brackets approaches 5/2. On the other hand, if 77,. is very small, as in a dilute
750
H.A. Barnes
suspension of air bubbles, the inner-bracket term approaches unity. Although these limits have proved realistic in predicting ideal situations, they have little relevance in predicting real situations. Also, the presence of a stabilising layer often over-rides the inner effect, and it is often more appropriate to model dilute emulsions as elastic spheres, see a more detailed view of these points in Macosko's textbook [23] and in a review by Nadim [24]. In more practical cases, the typical droplet size of emulsions, together with a thick stabilising layer means that viscosity is often identical to that of the equivalent suspension of solid spheres, and only at the point where the concentration is very high, and the particles crowd and compress each other will large deviations be seen from this simple picture. For flocculated emulsions, the picture is even more realistic, since such high concentrations are rarely reached. 19.3. Role of the emulsifier Although we do not need to be concerned here with the precise mechanisms that control the mechanism of emulsification, we do need to think about the results of the process as it produces an emulsifier layer with a certain thickness and elasticity. The former will affect the effective phase volume of the emulsion because the hydrodynamic radius is increased, and the latter because the overall droplet elasticity will affect the droplet deformability at rest and during in flow. Theoretically, the dynamic deformability can change when all the other emulsion parameters are identical, viz., the droplet size distribution and the viscosities of internal and external phases. The effective increase in the droplet radius due to the emulsifier layer can be gauged by measuring the viscosity of a very diluted emulsion, taking great care to dilute with the continuous phase, not just water or whatever the major disperse phase component is, see Fig. 22. The elasticity (hence the dynamic
Fig. 22. A schematic representation of adjacent emulsion droplets stabalised by a single surfactant layer.
The Rheology of Emulsions
751
deformability) of the droplet is more difficult to measure, but obviously strongly affects the maximum packing fraction and the viscosity and linear viscoelasticity at high phase volume. As emulsifiers can range from simple molecules, through surfactants and polymers to liquid crystal multilayers (and even solid particles in Pickering emulsions), the effect on the droplet size can vary considerably. The form of the emulsifier also determines the possibility of flocculation. If insufficient repulsion exists between neighbouring droplets, then aggregation takes place caused by the ever-present van- der-Waals attraction. Also, coalescence and Ostwald ripening (disproportionation) is largely controlled by the form of emulsifier. This in turn will alter the droplet size distribution, leading to a change in rheology. As there is usually a move to larger particles, this normally leads to a reduction in viscosity and linear elastic properties. A concise introduction to surface rheology with application to dilute emulsions of viscous drops, see Nadim [24]. 19.3.1. Hydrocolloids as emulsifiers Specific examples of complex stabilisers include the often-used, so-called hydrocolloids, and are often simply called gums. Hydrocolloids are highmolecular-weight polysaccharides. Their use has been summarised by Garti [25]. Gum Arabic is an example of a good hydrocolloid emulsifier - adsorbing onto oil-water interfaces and producing steric stabilisation between droplets. Galactomannans, xanthans, pectins, etc, are also used as emulsifying agents. Like all such polymers, they are usually used to excess, so that, not only do they have a stabilising effect, but they also thicken the continuous phase. The surface activity of gum arabic derives from the anchoring ability of the hydrophobic proteineous moieties (attached to the polysaccharide backbone) onto the oil phase. Adsorption can be induced by a salting-out effect, resulting in semi-solid interfacial layers. Hydrocolloids can form thick, birefringent, gel-like mechanical barriers at the oil-water interface of emulsion oil droplets. It is believed that gum arabic adsorbs strongly and effectively onto the oil droplets via its proteinaceous moieties. Guar gum and locust bean gum (LBG) adsorb weakly, but effectively precipitate onto the oil surface, and form birefringent layers of the polymer oriented with its hydrophobic mannose backbone facing the oil. In food systems, protein-hydrocolloid complexation at interfaces can be associated with bridging flocculation or steric stabilisation. As well as controlling the rheology, the presence in solution of a non-adsorbing hydrocolloid can affect creaming stability by inducing depletion flocculation [26].
752
H.A. Barnes
20. DROPLET EFFECTS 20.1. Droplet size It has long been known that decreasing the droplet size of emulsions increases the viscosity (see for instance Richardson [9]). This is due to . Effect of emulsifier layer • Effect of Brownian motion • Possible change effects . Other particle interaction effects. What is not often pointed out is that the actual form of the distribution is also very important. As droplet size is decreased in processing apparatus, the width of the distribution usually—but not always—decreases also, so that the emulsion tends towards being mono-modal. Narrow droplet-size distributions means poor packing, while a wide droplet-size distribution means good packing, with the latter giving a much lower viscosity for the same disperse phase volume. This arises from the packing constraints on the maximum phase volume, see below. The droplet size distribution itself is of course a function of the type and level of stabiliser used and also the type and level of agitation used. As far as the latter is concerned extensional flow is more efficient than shear flow and turbulent flow has a significant component of extensional flow in it. Usually size effects are due to significant colloidal interaction between particles, that is to say when the particles are considerably smaller than I/an. However, many emulsions have sizes in excess of this, and any charge effects that might produce similar effects are negligible. Two reasons might be put forward: first that particle deformability decreases with droplet size and secondly the width of the droplet size distribution usually also decreases with droplet size, something which is rarely, if ever, mentioned in this context. However, in the study of disperse systems of powders it is well known that increasing the width of the distribution increases the maximum packing fraction, which in terms of viscosity means a decrease in viscosity (see Ref. [27] for an example). Typically, emulsions obey a log—normal form of droplet size distribution [28] allowing us to reduce the description to two parameters. Another effect we could imagine, especially for small particles with thick stabilising layers, is the effect of the layer thickness of surrounding the particle, as shown above in §19.3. This does not change much with droplet size a, but becomes more significant for very small particles, such that the effective phase volume scales as (1 + 51a) Sherman [29] gives a good example of this phenomenon for an oil-in-water emulsion stabilised by a thick lipoprotein layer.
The Rheology of Emulsions
753
Decrease in droplet size means increase in surface area and thus an increase in the amount of stabiliser needed. If the stabiliser is not in vast excess, then reduction in droplet size can lead to poorly stabilised particles, and possible (perhaps temporary) flocculation. For flocculated systems anyway we expect a large effect of droplet size since the number of particles, p, in a floe plays such a major role in determining viscosity (see above). Last of all there is a subtle effect because of the Brownian or osmotic pressure. Consider a simple emulsion of monodisperse particles; the degree of movement from random to more ordered will, for the same phase volume, be resisted for longer in terms of shear rate from smaller particles than from large ones, owing to the stronger nature of the Brownian or osmotic effects. Thus, as shown in Fig. 23 at any given shear rate in this range, the viscosity increases with droplet size, even though the viscosities at very low and very high shear rates, respectively, are unchanged. 20.2. Droplet-size-distribution effects The droplet size distribution of any particular emulsion system depends on the amount of energy used in its manufacture. The energy ranges from low values in simple paddle mixers, through to the high values obtained using highpressure homogenisers or ultrasonic devices.
Fig. 23. Flow curves for decreasing droplet size of an emulsion.
754
H.A. Barnes
Fig. 24. The effect of droplet deformability on maximum phase volume. Average droplet sizes in the 1 - 1 0 0 /an range are produced in typical colloid mills and in-line mixers, while sub-micron droplets are possible using highpressure homogenisers or ultrasonic devices.
20.3. Droplet deformability At very high phase volumes, the importance of droplet deformability increases, particularly when the phase volume is such that particles begin to touch. Depending on the flow conditions and the droplet size distribution, this is when the phase volume is around 70%. The end result of any increase in deformability is an increase in the maximum phase volume, as shown in Fig. 25. 21. SPECIFIC COLLOIDAL INTERACTIONS 21.1. General considerations The form of the interaction between adjacent emulsion droplets where there are large droplet-droplet interactions are shown in Fig. 25 below. 21.2. Overall attraction between droplets Particle interaction is most pronounced in emulsions when we have an overall attraction between the particles, with the potential to form floes. Although the presence of a stabilising layer is often such that the primary particles of an emulsion are held far enough apart such that the van-der-Waals attractive force is ineffective, nevertheless situations arise when this is not so. One of the strangest found is where the surfactant used to stabilise the
The Rheology of Emulsions
755
Fig. 25. The kind of force-distance curves that are possible between adjacent emulsion droplets. The ever-present van-der-Waals attraction is modified by the presence of divers repulsive surface forces such as steric or charge effects.
drops is present in excess, and where this excess exists above its critical micelle concentration in the form of rod-like particles. The presence of these particles causes depletion flocculation of the emulsion particles; in fact the inclusion of very small droplets of essentially the same emulsion prepared separately and later added can produce this effect (see for instance Pal's work on surfactant flocculated and polymer-flocculated emulsions [30]. Other reasons for flocculation of otherwise stable emulsions might be a very thin stabilising layer that does not protrude beyond the van-der-Waals attraction distance. The structure of typical floes is such that they can be described as fractal bodies, see Fig. 26, that is to say the relationship between the number of particles in a floe, p, their radius a and the radius of the enclosing sphere R is given by
Fig. 26. The various parameters for a fractal emulsion floe, showing the possible extremes of a closely packed floe and a linear floe.
756
H.A. Barnes
p ~ lR/)D• ^
\/aj
05) v
'
The size of suspended floes usually decreases with increase in shear stress, often to revert to single particles at very high stresses, in such a way that R =
^°
(16)
(see Ref. [13] for a more detailed study of the effects of shear on fractal structures). The effective phase volume that described by the enclosing spheres — is then given by 4 v = 4> P ° ~ X •
(17)
This in turn can now be fed into the Krieger-Dougherty equation given above, together with the stress-dependent value of p. The result again is that the viscosity decreases rapidly with shear rate or shear stress, often from a very high viscosity given by a low phase volume of primary particles. It is also possible that the continuous phase of an emulsion is nonNewtonian owing to the presence of dissolved polymer, or other thickening agents. Although the emulsion droplets increase the viscosity further, because the local shear rate in the continuous phase is higher than the average and even more so as the phase volume increases. The increase in viscosity is not as large as expected from the considerations above. 21.3. Depletion flocculation One particular example of overall droplet attraction is that caused by depletion flocculation. The presence of entities of a certain size - be they polymers or particles - can under certain circumstances bring about flocculation. These entities are used to produce thickening of an emulsion, but in some situations, they often do so by this depletion mechanism rather than by the normal process of increasing the viscosity of the continuous phase, and this is not always realised. This can often lead to problems due to instability. A good review of this area is provided by McClements and Chanamai [31]. They reviewed the influence of droplet concentration (0 to 67 vol%), droplet radius (0.1 to 1.0 mum) and droplet flocculation (depletion) on the rheology, creaming stability and optical properties of monodisperse oil-in-water emulsions.
The Rheology of Emulsions
757
21.4. Overall repulsion between charged droplets There are some situations where the large surface charge on droplets causes repulsion between the droplets. This would lead to an increase in the viscosity of a neutral emulsion because of the extra energy that needs to be disippated to move droplets to move past each other. Gradzielski and Hoffmann [32] have investigated such a system, and have accounted for the increase in viscosity by a term that includes the electrostatic repulsion. 22. HIGH-CONCENTRATION EMULSIONS We may view high-concentration emulsions, or high-internal phase emulsions (HIPEs) are a collection of viscous, highly deformed, viscous elastic bags, with the rheology being a very complicated function of all the parameters concerned. One extreme version of this class of materials is a foam, where the internal phase material is a virtually inviscid gas, often air. These materials are highly viscous and usually very elastic. In this case, the linear viscoelastic parameters are quite different from a lower-concentration emulsion. 23. Complex emulsions Although the main ingredients of a commercial product may be technically an emulsion, it is possible that other components are present. Such a system is described by Klein [33], who comments on the addition of highmelting point waxes and fatty alcohols which increase the viscosity of the product and thus enhance its physical stability. Of course the addition of some additives can cause depletion flocculation if their concentration and size is critical. 23. SOLIDS ADDED EMULSION
TO
THE
CONTINUOUS
PHASE
OF
AN
Pal et al [12] have made a special study of the effect on the viscosity of adding solids to an emulsion. The basic effect of simple, non-interactive solid particles depends of course on their shape. The thickening potential of various shapes is shown in Fig. 27. In the simplest case with small and large particles or droplets, then the effect smallest (whether particles or drops) entity acts to thicken up the continuous phase. The effect of solid particles added to a small-droplet-sized emulsion can be understood using the approach of the present author [22], who showed that the thickening effect can be simply predicted by reducing the flow curves to viscosity versus shear stress, and then a simple vertical shift alone is needed.
758
H.A. Barnes
Fig. 27. The thickening potential of variously shaped, non-attractive entities dispersed in a liquid.
24.CONCLUSIONS In qualitative terms the rheology of emulsions is well understood in terms of the effects brought about by various formulation and processing variables that result in a change of dispersed phase volume, size distribution and particle - particle interaction as well as particle deformability. What we do not have, however, is a very good quantitative understanding - at least compared with dispersions of solid particles - because of the lack of good model systems ranging from monodisperse to known and controlled degrees of polydispersity, together with an independent change in particle deformability. Until this happens, emulsion rheology will continue to be the study of multifunctional effects with best guesses at effects found. For instance we shall not be able to distinguish between the effects of particle deformation and particle-size-distribution width effects (see for instance Refs. [34] and [35]. Diat and Roux [36] describe some interesting possibilities for the production of mono- disperse multilayer vesicles of controlled droplet size which could form the basis of such an approach.
REFERENCES [1] [2] [3] [4] [5] [6]
T.F. Tadros, Adv. Colloid Interf. Sci., 46 (1993) 1 - 4 7 . P.E. Miner, Rheological Properties of Cosmetics and Toiletries, Ed. D Laba, Dekker, New York, chapter 9: Creams and lotions 1993. J.W. Summers, J. Vinyl & Additive Tech., 3 (2) (1997) 130-139. K.B. de Roos, Inter. Dairy Journal, 13(8) (2003) 593-605. E. Hatschck, Kolloid-Z., 8 (1911) 34. J.O. Sibree, Trans. Faraday Soc, 26 (1930) 26; 27 (1931) 161.
Marcel
The Rheology of Emulsions
[7]
[8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25]
[26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36]
759
A. Harvey, Technical aspects of emulsions, Papers read at a symposium held in London, 7th December 1934 under the auspices of the British Section of the International Society of Leather Trades' Chemists, London, 1935. V.G.W. Harrison, Nature, 165 (1950) 182. E.G. Richardson, 'Emulsions', chapter 2 in 'Flow properties of disperse systems, ed. J.J. Hermans, North-Holland Publishing House, Amsterdam, 1953. P. Sherman, 'Emulsion Science', Academic Press, Oxford, 1968. R. Darby, Surfactant Sci., Ser., 6 (1984) 45-110. R. Pal, Y. Yan, and J. Masliyah, Advances in Chemistry Series, 231 (1992) 131-170. T.F. Tadros, Colloids and Surfaces A-Physicochemical And Engineering Aspects, 91 (1994)39-55. H.A. Barnes, Colloids and Surfaces A: Physicochemical and Engineering Aspects 91 (1994)89-95. H.A. Barnes, J.F. Hutton and K. Walters, An Introduction to Rheology, Elsevier, Amsterdam, 1989. H.A. Barnes and D. Bell, Korea-Australia Rheology Journal, 15(4) (2003) 187-196. H.A. Barnes, 'Viscosity', The University of Wales Institute of Non-Newtonian Fluid Mechanics, Aberystwyth, ISBN 0-9538032-2-8, 2002. H.A. Barnes, J. Non-Newtonian Fluid Mech., 56 (1995) 2 2 1 - 2 5 1 . C. Gallegos and J.M. Franco, Current Opinion in Colloid & Interface Science, 4(4) (1999)288-293. T. van Vliet, J. Lykienia and M. van den Tempel, J. Colloid Interface Sci., 65(3) (1978) 505. T.S.R. Al-Hadithi, K. Walters and H.A. Barnes, Proc. 10lh Int. Congr. on Rheology, Sydney, Ed. P.H.T. Uhlherr, Vol. 1, pp. 137-139, 1988. H.A. Barnes, 'The rheology of filled viscoelastic systems: A review', in Rheology Reviews 2003, British Society of Rheology, ISBN 0-9538904-8-1, pp. 1-36, 2003. C.W. Macosko, 'Rheology: Principles, measurements and Applications, Wiley - VCH, New York, 1994. A. Nadim, Chem. Eng. Comm., 150(1996) 391 - 407. N. Garti, D. Reichman, H.A.CM. Hendrickx, E. Dickinson, L.K. Jackson and B. Bergenstahl, Food Structure, 12(4) (1993) 411-426; N. Garti, Journal of Dispersion Science and Technology, 20(1-2) (1999) 327-355; N. Garti and M.E. Leser, Polymers for Advanced Technologies, 12(1-2) (2001) 123-135. E. Dickinson, Food Hydrocolloids, 17(1) (2003) 25-39. R.J.M. Tausk, and P.N. Wilson, Colloids Surfaces, 2 (1981) 71. E.S. Rajagopal, Kolloid-Z., 162(2) (1959) 85. P. Sherman, Food Technol., 15(9) 394 1961. R. Pal, Colloids Surfaces, A71 (1993) 173; Chem. Eng. Commun., 98 (1990) 211. D.J. McClements and R. Chanamai, Journal of Dispersion Science and Technology, 23(1-3) (2002) 125-134. M. Gradzielski and H. Hoffmann, Adv. Colloid Interface Sci., 42 (1992) 149-173. K.Klein, Cosmetics and Toiletries, 99(1988) 121-126. E.S. Rajagopal, Z. Phys. Chem. Folge, 23 (1960) 342. A.K. Das, D. Mukesh, S.V. Doble, D.D. Dilip and P.K. Ghosh, Langmuir, 8 (1992) 2427. O. Diat and D. Roux, J. Phys. II, 3 (1993) 9.
This page is intentionally left blank
SUBJECT INDEX
Adhesion 33, 513, 549-552, 641 Adsorption 8, 17, 23, 30, 33-35, 37, 43, 44, 45, 46, 51, 55-57, 61-64, 67-70, 73, 74, 76, 79, 83, 84, 358, 388, 479, 578, 623, 625, 627, 641, 646, 660, 678, 683, 686, 687, 697-700, 709-712, 714, 751 - energy of solid particles 646 - isotherm 43, 44, 45, 46, 67, 68, 147, 155, 157-159, 161,624,698 - kinetics 76, 699 - negative 123, 161 - of ions 92, 93, 122, 123, 131, 136, 137, 140, 150, 153, 154, 168, 317, 683 - of solid particles 642, 643, 646, 653, 661,663,666,668,668,675 Bancroft rule 216, 240, 613, 678, 679 Bending elasticity constant 329, 491, 495, 498, 503 Bending mode 232, 233, 247 Bond number 423 Brownian aggregation (coagulation, flocculation) 335, 339, 352-355, 356 Brownian droplets (emulsions) 332, 336, 339, 351, 353-358, 391, 482, 494 Brownian dynamics simulations 487, 494, 501, 502, 677, 679, 681, 688, 690, 692, 715 Brownian motion 252, 332, 345, 353-357, 479, 632, 680, 752 Capillary 64, 73, 344, 729 - force 183, 188, 199, 246, 260, 409, 410,561 - length 456
- number 392, 394, 404, 405, 410, 413, 416, 418, 419, 423, 430, 433, 434, 561, 562 -pressure 183, 185, 187, 190, 199, 207, 246-248, 255, 323, 330, 344, 552, 569, 671-675 --, critical 672 --, threshold 670-673 - waves 39, 50, 52, 64, 66, 67, 73, 138 Coalescence 62, 186, 216, 250, 252, 255, 336, 337, 339, 342, 343, 345, 346, 368, 376-378, 380, 382, 383, 386, 388, 414, 420, 424, 445, 479, 508, 519, 537, 558, 573, 584, 585, 587-590, 593, 596, 600, 607, 608, 610-617, 619-628, 630-632, 637, 642, 660, 673, 677-679, 682, 687, 697, 698, 701, 702, 710-714, 724, 751 -, efficiency 404, 405 - frequency 615, 617, 620-622, 625, 630,631,637 -, kinetics of 351,357, 388 -, probability of 384 - rate 351, 352, 378, 380, 383, 384, 518,615,620,624,627,628 -time 351, 359, 363, 378, 380 Collision frequency 680-682, 688, 689, 691,693-696,702,704,710 Collision time 402, 403, 702, 704 Compressibility - of an interface 63, 66, 74, 76, 78 - intrinsic 66-68, 74, 76 -, osmotic 482 Computer simulations 8, 49-51, 225, 262, 269, 272, 302, 309, 423, 687 Contact angle 74, 189, 198-202, 205, 208, 210, 212, 213, 323, 331, 525, 550, 551, 641-647, 649, 650, 652, 653, 656-662, 669, 670, 672-675
762
Subject Index
Contact zone 183, 184, 196-201 Correlation function 1-3, 6, 7, 28, 163, 175,269,302,482 -, density 15,269 -, direct 16, 17, 22-24, 262, 272, 281, 485-487 -, fluctuations 15, 247, 253 -, pair 6, 26, 273, 482, 487, 489 -, partial 281 -, particle 41, 175, 266, 268-272, 274, 282,331,485 -, total 6, 262, 272, 281, 282, 300, 307, 485 Correlation length (range) 2, 18, 22, 174 Creaming 314, 353, 388, 436, 530, 537, 549, 559, 583, 600, 611, 614, 617, 679, 688, 702, 721-723, 726, 741, 742, 744, 745,751,756 Critical angle 35, 462 Critical distance (thickness) 342, 343, 612, 622, 652, 657, 672 Critical droplet radius (size) 194, 198, 201, 355,356 Critical micellar concentration 66, 755 Critical point 17-19,30 Critical volume fraction 511, 693
- controlled ripening 608, 618, 622 -flux 234, 235, 357, 360 - of water 523 -, surface 241, 242, 678 - tensor 336-339, 688, 689, 693, 695, 696 Disjoining pressure 183-185, 188, 192, 194, 195, 199,201-203,211,217,245, 247-249, 323-325, 344, 398, 614, 621, 674 -, electrostatic component 183, 211, 213,245,300,301 Distribution functions 11, 12, 16, 21, 251 -, atomic 29 -, Gaussian 689, 696 -, molecular 123,679 -, pair 5-7,15, 9, 20, 21, 27, 436, 483 -, particle 694, 704, 708 -, radial 5, 8, 15, 21, 262, 481-487, 490, 494-500, 502, 503, 507 - of ions 92-94, 96, 98, 103, 108, 110, 113, 114, 122, 125, 127, 129, 132 DLVO theory 183, 245, 246, 250, 260, 261, 300, 304-307, 313, 357, 492, 550-552, 682, 683, 685, 686, 690, 697
Debye - and Huckel theory 99, 683, 684 - constant (see Debye parameter) - length (see Debye parameter) - parameter 115, 117, 119, 120, 126, 144, 166, 168, 174,211,317,492, 683 Density distribution 303-307, 680, 681 Derjaguin approximation 197, 269, 274, 300, 326, 327 Diffusion 64, 66, 67, 315, 331, 332, 354, 388, 559, 580, 583, 591, 593, 596-600, 607, 609, 610, 611, 619, 677-680, 696, 698,709,712,716 - boundary layer 237, 240 -, Brownian 353, 677 - coefficient 234, 335, 339, 360, 387, 609, 680-682, 689, 690, 693-696, 699, 700,704,712 - controlled adsorption 67, 69, 76, 699 - controlled aggregation 691
Elasticity -, dilational 62-66, 73, 79, 84, 614 -, film 654, 656, 657, 672, 675 -, shear 84-86 - surface 76, 79, 614, 615, 687, 750 - modulus of 67, 76, 524-526 Electric double layer 126, 127, 131-133, 135-137, 139, 143-145, 149-153, 162-164, 168, 260, 261, 292, 294, 307, 356 - vectors 38 Electrocapillarity - curves 151 -, equation 147-149, 151 Emulsion droplets 35, 40, 42, 43, 47, 48, 184, 198, 217, 219, 246, 248, 308, 353, 479-482, 487, 491, 502, 504, 505, 557. 558, 569, 570,, 575, 581, 587, 600, 627, 642, 646, 647, 669, 672, 673, 686, 724-726, 734, 743, 745, 750, 754-756 -, deformable 199, 217, 230, 247, 249, 250, 329, 332, 336, 339, 341, 342,
763
Subject Index
345, 358, 391-393, 436, 442-444, 495, 496, 498, 499, 547, 550-552, 610, 621, 630,647,649,687,698,715 -, interactions between 343, 344, 481 Energy -, activation (barrier) 612, 613, 616, 617,623,627,632,635 -, bending 328, 330, 480, 491, 552, 589 -, interaction 330, 332, 335, 340, 358, 359, 363, 373, 386, 387, 480-482, 486-488, 490-495, 499, 625, 682, 715 -, interfacial 19, 329, 343, 502, 506 Equipartition theorem 506 Film -, adsorbed 37, 53 -, colloidal 302, 304-306, 308 - curvature 48 - drainage 225, 250, 255, 414, 630, 698 - equations 414-417, 419 -flow 81, 240 -, interfacial 10, 562, 581 -, macroscopic 30 -, microscopic 253 - optical 38 - rupture 246, 216, 250, 253, 376, 378, 380, 383, 384, 386, 387, 389, 558, 613, 616,620,642 -, thin liquid 29,37,62, 183-185, 199, 202-204, 211-213, 227, 246, 248-250, 351, 353, 359, 363, 403, 414-417, 419, 424, 445, 489-495, 502, 504, 511,518, 525, 526, 550, 551, 570, 571, 593, 596, 607, 611, 613-615, 616, 620-623, 647, 648, 652-654, 657, 659-661, 664, 666, 667, 669, 673-675, 679, 687 -- stability 62, 364, 365, 625, 627, 646 - wetting 30, 43 Flocculation 300, 358, 479, 508, 652, 708, 709, 712-717, 723, 743-745, 749-751, 753, 755-757 Force -, attractive 22, 161, 705 -, between droplets 512, 550-552, 680, 692, 704, 708 -, between interfaces 343, 344
-, Brownian 333,339-341,354,712, 749, 753
-, capillary 561 -, colloidal 183, 184, 445, 492, 507, 692, 695 -, contact 394, 395, 397, 399-401, 416 -, depletion 260, 273, -, drag 218, 228-233,244, 245 -, electromotive 147 -, electrostatic 307, 357, 398, 575, 609, 708 -, gravitational 449, 642, 644, 645 -, hydration 614 -, hydrodynamic 405, 491, 708 -, image 123, 129, 160-164, 166, 168, 169 -, interatomic 23 -, intermolecular 2, 66, 405 -, lubrication 400 - measurements 306, 308 -, potential of mean 490 -, repulsive 480, 550, 609, 690, 696, 712 - see also Brownian -, structural 300, 302 -, surface 183, 184, 186, 188, 195, 197, 199, 245, 255, 356-358, 376, 614, 637, 642, 645, 755 -, total 689 -, van der Waals 30, 343, 351,352, 357,392,686,701,708,715,754 -, viscous 457, 724, 725, 729, 739, 741-743,749 Free energy -, adsorption 157 - change 10 -, electrostatic 113, 124, 125, 161-163 -, excess 14,23,25,28, 186, 189, 193, 202-206, 677 -functional 1,2, 11, 13-16 20-22, 25, 27, 29 -, Gibbs93, 106, 137, 139, 157, 161, 169, 171, 178, 576-578, 642, 643, 646, 669 ~ , hydrophobic contribution 137 -, Helmholtz 11, 13, 16, 20, 21, 24, 25, 27, 193 -, interfacial 313
764
Subject Index
- of interaction 313, 320, 321, 492, 535 -, surface 8, 9, 10, 19, 322, 324, 326 Gel (gelation) 314,511-513,517-519, 521, 530, 537, 539, 543, 545, 547, 548, 552, 583-585, 591, 600, 632-636, 645, 742,746,751 Gibbs - dividing surface 10 -elasticity 329, 491, 614, 637 - equation 43 - isotherm 43, 44, 68, 147, 698 Hamaker - constant 211, 284, 300, 320, 325, 331, 336, 375, 398, 402, 404, 482, 493, 499, 551, 682, 686, 687, 702, 703, 705 - method 284, 318, 327, 328, 492 Hard spheres 17, 21-26, 28, 29, 42, 126, 163, 164, 289-291, 294, 295, 300-302, 438, 479, 486, 495-498, 500, 501, 637, 690, 693, 694 Hard wall 22, 24, 29, 30, 164, 172, 175, 176,295 Hydration 51, 123, 308, 313, 322, 357, 491, 573, 591, 592, 594, 599, 600, 614 Hydrophilic lipophilic balance, 480, 481, 516,521 Interaction -, adhesion 641 -, anisotropic 26, 27 -, attractive 42, 330, 336, 682 -, between adsorbed particles 669 - between droplets 194, 329, 335, 339, 341, 342, 344, 345, 358, 359, 429, 431, 434, 436, 439, 443-445, 488, 489, 491, 495, 502, 504, 505, 507, 512, 550-552, 715,742,754 -, colloidal 265, 300, 304, 345, 491, 494, 507, 623, 625, 744, 752, 754, 758 -, depletion 26, 276, 279, 307, 308, 330 -, dipole 170, 178, 265, 291, 295-302 -, electrostatic 480, 49, 492, 499, 507, 609, 669, 697 -, hydrodynamic 261, 330-332, 335, 345, 425, 479, 480, 695, 704, 706, 708, 709
-, hydrophobic 322, 357, 535, 535, 552, 576, 610, 612, 620, 623, 627-629, 645, 659,660,667,672-674,751 -, induced break-up 425 -, intermolecular 1, 30, 63 -, ionic 161, 164, 169 - parameter 47, 51 -, particle 14, 359, 360, 482, 483, 485, 723, 752 - potential 697 -, repulsive 300, 302, 304, 330, 336, 487, 492, 499, 535, 575, 576, 609, 621 -, soft 26 -, specific 170, 178 -, symmetric 27 -, thermal 689, 692, 696, 700, 702, 706, 707 -, van der Waals 27, 30, 245, 248, 260, 284-286, 313, 316, 318-320, 326-330, 341, 343, 345, 376, 391, 394, 398-402, 405, 415, 417, 424, 445, 491, 495, 507, 551, 623, 625, 669, 708, 751, 754, 755 Interface 17, 463, 642-647, 661, 663, 668, 687 -, electrochemical 261, 299, 307 -, fluid 20, 22, 23, 26-30, 53, 55, 56, 61, 62, 66, 70, 75, 80, 86, 91, 93-96, 98, 101, 104, 106, 108, 110-112, 115-117, 122-129, 131-144, 146-149, 151, 153-170, 177, 183, 315, 323, 327, 331, 333, 343, 464, 466-469. 473-475, 479, 482, 491, 495, 500, 507, 552, 687, 704, 709, 710, 712, 713, 723 -, oil/water 562, 585, 609, 614, 621, 627, 628, 630, 641-643, 647, 649, 661, 666,668,687,698,751 -, polarizable 93, 140-142, 149, 683 -, solid 23 Interfacial curvature 329, 502 Interfacial deformation (flexibility) 329, 339, 393, 479, 480, 487, 492, 500, 502, 715 Interfacial fluctuations 507, 614 Interfacial profile (region) 8, 17, 18, 277, 293, 297, 299, 307 Ionic Strength 99, 148, 336, 482, 683-686, 690,712
Subject Index
Laplace 1 - equation 160, 236, 242, 243, 323 - pressure 359, 407, 584, 607, 609, 610,621 - transform 270 Lennard-Jones 8, 22, 27, 283, 283 Light - absorption 366, 386 - reflectivity 35 - scattering 386, 484, 498, 617, 681, 694 Metastability 186, 306, 331, 570, 607, 608, 614-616,622,645 Monodisperse particles (droplets) 41, 279, 308, 309, 343, 344, 353-356, 364, 386, 387, 430, 479, 494, 525, 560, 568, 570, 571, 600, 609, 611, 616-618, 622, 631, 642, 649, 652, 680, 691, 708, 713-715, 747, 748, 753, 756, 758 Monolayers 35-37, 43, 45, 48-57, 67, 68, 76, 83, 84, 98, 153, 155, 156, 164, 302, 303, 306, 329, 512, 513, 520, 522, 571, 612-615,620,623,625 - of solid particles 644, 647, 657, 658 Micelles 23, 40, 42, 47, 56, 259, 305, 306, 308, 309, 321, 521, 522, 528-537, 542-545, 563, 596, 609, 612, 709, 724, 747 -, reversed 573, 575, 578, 587, 591-594,599,600 Microemulsions 1,10, 34, 43, 47, 48, 53, 115, 116, 260, 321, 345, 480, 481, 491, 494, 495, 497-505, 507, 508, 512-514, 516, 519-524, 552, 563-568, 599, 600, 612 Microinterferometric technique 216, 260, 344 Miniemulsions 331, 345, 357, 479, 481 Mobility -, between phases 579 -, interfacial 333 - function 689, 693, 694 -, surface 250, 254, 255, 331, 335, 337, 359 -, water 590 -, within double emulsions 583
Neutron - reflectivity 33-35, 40, 42, 44, 46-51, 53,55,56 - scattering 35, 484, 614 - wave vector 36 Offset parameter -, critical 403, 424-426, 429 Osmotic pressure 322, 482, 483, 500, 501, 552, 563, 571, 577, 584, 587, 590, 591, 593,599,609-612,749,753 Ostwald ripening 315, 356, 607-610, 618-620,622,669,679,715 Partition - coefficient 93-95, 100, 102, 104-106, 112, 113, 116, 117, 137, 173, 178-180,572 - function 5, 10 Partitioning 48, 98, 678, 687, 697, 724 Phase - coexistence 8, 17, 18,20,313, -, contiguous 215, 218, 221, 223, 228, 233, 255 - diagram 484, 514-518, 520, 522, 536-538, 544, 548, 564, 580, 581 -, droplet 220, 221, 225, 235, 238, 240, 335, 479, 498, 648, 747, 752 -, emulsion 48 - equilibria 24, 484, 516, 528 - inversion 514, 519, 560, 573, 574, 617,673,675,679 -, lamellar 34, 55, 83, 563-566, 614, 747 -, liquid 8, 29, 92, 94, 114, 151, 184, 192, 199,215,217,226,484,512, 516, 521, 523, 544, 557, 722-724, 728, 736, 742-758 -, micellar48, 612 -, microemulsion 514, 516, 521 -, oil 48, 179, 344, 513-516, 518523, 530, 531, 535-537, 560, 562, 566-569, 572, 575, 578, 580, 583-585, 589-591, 593, 596, 598, 599, 617, 624-626, 642-645, 658, 660, 664, 666, 669, 674, 751
765
766
Subject Index
- separation 1, 30, 436, 531, 542, 543, 545,581,632,709 -, solution 57 - space 3 -transition 13,433,519,532 -, water 572, 584, 585, 589, 623, 624, 643, 645, 648, 663, 664, 666, 669 Polymers 28, 29, 61, 284, 528, 543, 611, 632, 690, 724, 742, 744, 745, 747, 751, 755, 756 -, adsorption of 33, 34, 321, 576-579 -, amphiphilic 578, 580 -, hydrophobic 579 Potential - distribution 94-96, 98, 99-101, 103-105, 112, 113, 115-121, 142, 143, 146, 151, 161, 163, 168, 169 -- theorem 489 -, electrostatic 164, 168-170, 291, 292, 294, 307, 697, 708, 711-713 -, particle 691, 692, 697 -, repulsive 690, 698, 712 -, steric715 -, surface 691, 698, 711 -, van der Waals 703 -,zeta 211,212, 698 Probability function 339, 342, 692 Rheology -, bulk 721-723, 726, 735, 743, 744, 747,751,756-758 -, interfacial 61, 62, 80, 82, 84, 679, 751
- measurements 584 Rigidity -, bending 10, 480, 481, 491, 499, 507 -, interfacial 330, 500 SANS 34, 35, 40, 42, 47, 48, 614 Size distribution 278, 279, 404, 479, 560, 569, 570, 600, 608-610, 617-619, 677, 708, 715, 721, 744, 749-754, 758 Solvation interactions 267, 287, 507 -, electrostatic 124, 136, 161, 162, 170-176 Stability - factor 335, 343, 706 -, kinetic 313, 678
-, numerical 407, 408, 413, 419, 431, 452, 467 - of colloidal dispersions 260, 300, 307,309,436,491,690 - of emulsions 33, 47, 53, 79, 215-217, 233, 245-250, 253, 255, 351, 352, 359, 364, 386, 393, 439-442, 481, 523, 531, 558, 560, 562, 563, 570, 572, 574-576, 578, 579, 581, 583-587, 590, 591, 596, 598-601, 607, 608, 610, 614, 615, 623, 630, 637, 652, 669, 670, 672-674, 677-679, 687, 688, 690, 693, 697, 698, 700, 713, 714, 722, 728, 751, 756, 757 - of flows 475 - of foams 62, 79 -ratio 682, 707, 712 Steric interactions 213, 246, 321, 357, 358. 491, 492, 507, 535, 552, 574-578, 580, 583,600,692,715,751,755 Strain 84, 86, 430, 431, 525, 526, 541-543, 724, 726, 728, 730, 731, 752 Stress 63, 80, 82-84, 86, 407, 413, 526, -, electrostatic (Maxwell) 294 -, normal 398, 420, 430-434, 452, 463, 469, 728, 743 -, osmotic 611 -, shear 225, 229, 232, 233, 430, 434, 473, 474, 569, 724-732, 736-741, 749, 756, 757 -, tangential 397, 398, 452, 458, 464, 469 - tensor 406, 430, 442, 457-460 - vector 218 , yield 722, 736, 738 Structure - factor 16, 18, 484, 494, 495-498 -, monolayer 53, 55, 56, 63, 80, 82 - of cusps 424 - of droplet aggregates 314, 329, 692 - of droplet doublet 331 - of emulsions 481, 482, 494, 507, 511, 512, 515, 516, 518, 519, 522-524, 526, 528-533, 535, 537, 543, 545, 549, 632, 634, 722, 724, 728, 745, 748, 749, 755, 756 -of fluids 1,2,4,5,8, 12, 13, 16,24
767
Subject Index
- of interfaces 8, 29, 33, 35, 53, 56, 61, 98, 122, 139-151, 156, 160, 163, 169-171,482 - of mesophases 563-565, 573, 578 -ofmicelles23,29, 563 - of particle monolayer 652, 654 -, oscillatory 272, 276 -, solvent 170, 171, 174 -, surface 37-39, 48, 52, 54 -, surfactant 52-54, 56 Surface charge 92, 113, 132, 133, 144, 146, 149, 150, 154, 162, 164, 168, 169, 660, 684-687, 697, 698, 709, 757 Surface curvature 341, 342, 345, 652 Surface extension (deformation) 345, 491, 495 Surface layers 284, 303-306, 698 Surface of fluids 7, 20, 62, 184, 255 Surface potential 331, 336, 364, 375-377, 492, 499, 683-687, 691, 698, 711 Tension -, bulk 727, 742 -, interfacial 8, 17, 19, 29, 201, 329, 330, 332, 345, 479-481, 486, 488, 491, 494, 495, 499, 500, 503, 505, 507, 524-526, 551, 552, 643, 644, 657, 659, 666, 668, 669, 673-675, 677, 678 -, line 183, 201-204, 206, 209, 212, 213,646 -, surface 8-10, 19, 33, 43, 70, 73, 74, 80, 122-124, 126-128,130, 131, 135-138, 177, 178, 184, 185, 201, 217, 246, 247, 254, 423, 436, 442, 452, 455, 457, 499, 589, 609, 614, 632, 699, 700, 713,726 ~,dynamic 62 Van Der Waals - approach 1, 20 - perturbation theory 20-22 see Disjoining pressure, Interactions and Forces Viscosity -, droplet 632, 636, 677, 704 -, emulsion 617, 636, 637, 721-724, 728-730, 732-741, 743-754, 756-758
-, extensional 726 -, interfacial 80-82, 687, 704 -, shear 726 -, surface 233, 234 Wave -, critical 36, 248 -, surface 250-252
This page is intentionally left blank