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!
are separated into local and non-local contributions: / =/loc
+
j-nloc
ft,=
ft^oc+ft;nloc
[1.3.4]
For the local contribution to the free energy density the bulk expression (1.2.2] is used but with the bulk concentration
INTERACTION BETWEEN POLYMER LAYERS
^3 floe
—kT
1.7
„
= — \n
) + ym(\ -
{z)) (the energy is 'non-local'), but we disregard this refinement here. When this form of u (which contains no ;fsterm) is used, [1.4.5] applies to all lattice layers except z = 0 , where the train segments are located. For z = 0 , u(z) in the SF model does include xs a s a n e x tra term. It is easily shown that when the unit increments As = 1 and Az = C in [1.4.5] are allowed to approach zero (which implies that u for a unit segment has to be replaced by II As for a 'fractional segment'), [1.4.5] reduces to the continuum form [1.4.1]. We do not discuss here the more difficult question of how to relate c in [1.4.3] to xs i n the lattice model; for this we refer to the literature 1'. 11
G.J. Fleer, J. van Male, and A. Johner, Macromolecules, 32 (1999) 825; ibid, 32 (1999) 845; A.A. Gorbunov, A.M. Skvortsov, J. Van Male, and G.J. Fleer, J. Chem. Phys. 114 (2001) 5366; G.J. Fleer, F.A.M. Leermakers, in Coagulation and Flocculation: Theory and Applications, 2nd ed., Surfactant Science Series, B. Dobias and H.J. Stechemesser, Eds., M. Dekker (2004).
1.10
INTERACTION BETWEEN POLYMER LAYERS
Equations [ 1.4.4] are also used in the lattice model, obviously in a discretized form, see [II.5.5.9-11 ]. There is a minor difference in [ 1.4.4b] as compared with [II.5.5.9] (the factor Gj(z) in the latter equation, which corrects for double-counting of the joining segment, is not present in the continuum version), but that is only an unimportant detail. Much more important is the fact that the SF model gives a generalized expression for
the Helmholtz
energy density in a gradient, which includes
the non-local
contributions. This generalization of [ 1.2.2] is e3 f 0 -^- = —lncp+
(1 -
(ph, [1.4.7] gives a negative value for co (provided v > 0 , % < 0.5 ). We can write [ 1.4.7] in terms of the osmotic pressure profile by introducing a local osmotic pressure n°s (z), defined by [ 1.2.3] but with
[1.4.17]
0
The integral may be written as Jgdg' where g' = dg/dz. Integration by parts gives gg'\0 -Jg'dg = -gg' z_0 - j(dg/dz) 2 dz , where we used g' = 0 at z = h and g'dg = (dg/dz)(dg/dz)dz . The term gg'| _ follows from [1.4.12] as -cg^ = -ccpo/a2 . Hence, [ 1.4.17] transforms into
^=-^0+^[f^fdz kT
[1.4.18,
6 r ° 6£ 1 {dz)
Inserting this result into i2a = ^ o c +i2° loc with C2l°c = \coXoc dz , we find for the grand potential per unit area:
MO =_^. kT
r[^i + £ir^ a L
&r° J [ kT
[L4.19]
6f {dz) \
which has the same form as [1.3.7]. Apparently, yQ in [1.3.8] may be identified with ckTI&t. With g = Jip / a, the square gradient term may be written as (24^r)" 1 (d^/dz) 2 , which shows that K in [1.3.7] is inversely proportional to
loc{(pm)j
P.G. De Genncs, Macromolecules 15 (1982) 492.
[1.5.1]
INTERACTION BETWEEN POLYMER LAYERS
1.15
Using [1.4.15] for <3loc , one may choose a certain value of
=6eg = 0, which must imply e = 0 since
9b ~ V^b~ i s non-zero. Hence, [ 1.4.13b] reduces to the simple form (p=fhg2
[1.5.2]
This result, used already in [ 1.4.14], shows that in the bulk solution g = \ (which also follows from [1.4.10] with e = 0 and Gb = 1). Inserting [1.5.2] into the expanded version [1.4.2c] of the segment potential (neglecting the quadratic term) gives the relation between II and g : u = vcpb(g2-l)
[1.5.3]
With this form of II, [1.4.11] reduces to ^ 2 d 2 g/dz 2 = 6u^ b (g 3 -g) . We may write this as
where Z is the spatial coordinate in units d;, and £ is the correlation length (or blob diameter). This is the natural length scale in semidilute solutions and scales for meanfield good solvents as #>b1/2 • Note that [ 1,5.4c] also gives the prefactor. As discussed in sec. II.5.2d, £, — ^ ~ 3 / 4 when chain swelling is taken into account. Upon multiplying [1.5.4a] with dg/dZ (dgldZ)
2
4
2
- g +2gr
we find d[(dg/dZ) 2 -g4 +2g2\/dZ = 0, so
is constant across the slit. The constant equals 9^a~'^9m since
dg/dZ = 0 in the middle of the slit. Hence,
[ft-
21
2
[1-5-51
1.16
INTERACTION BETWEEN POLYMER LAYERS
Note that this equation is a special case of [1.5.1] or [1.4.20] with a2 = (pb and (p =
1
2 8( 0, a» 8 ) with an exact result for the contact potential Ga(0) derived by Eisenriegler11. This result is Ga (0) aV {z) = JG at (z,s)G fr (z,JV-s)ds, where now the end-point distribution function Gfr for a free walk is also needed. For c < < 0, this function is given by Gfr(z,JV) = erf(Z) ~ ~cp2 . These deviations become more noticeable when for fixed chain length the grafting density is increased. This increases the segment concentration in the brush and the expansion of the logarithm becomes less accurate. Wijmans et al. showed51 that if the logarithmic form of the potential is maintained, the analytical approach remains accurate up to very high segment concentrations. 11 0, the solution shows a discontinuity at a charge parameter x = 1: l-x the chemical potential of the solvent //j and the osmotic pressure 77 turns out to be c = exp(l / Q. Multiplication with this "degeneracy" leads to a better value for the "reaction constant" K, i.e. K = a>c exp(-AGec / kT). Typically this correction has a minor effect on the size distribution. For very long micelles, the persistence length remains a relevant quantity as it may determine the overlap concentration of micelles and other properties. Another point of criticism is that this phenomenological model is not expected to be accurate for very small linear micelles. As we will discuss below, very small dumbbell- = 0.5. (b) Projection onto the T-(S plane of the phase prism. Within this projection, the trajectories display a sigmoidal shape extending from the lower critical endpoint (cep™ ) at Tj on the water-rich side to the upper one ( cep a ) at Tu on the oil-rich side. (Redrawn from .) j = 0.5), one clearly sees water-rich and oil-rich domains, which are mutually intertwined in a sponge-like fashion. Furthermore, many saddle-shaped structures can be observed. Typically, the two principal curvatures appear to be almost equal but of opposite sign, i.e. cx ~ -c 2 . As a consequence, the mean curvature J of the amphiphilic film may be near 0 and the Gauss curvature K is everywhere negative. In other words, the structure is water-continuous and oil-continuous at the same time, a situation which is called bicontinuous or, more tongue-incheek, plumber's nightmare. As we will see below, NMR-self-diffusion and electric conductivity measurements on related samples of the same system confirm this type of structure. 0 at low temperatures to J < 0 at high temperatures. In between, at the phase inversion temperature the mean curvature passes zero realized by the bicontinuous structure. While qualitative insight into the manifold structural properties is nicely obtained from such pictures, a more quantitative determination of parameters, such as the mean length scale, is difficult to infer in this way. 5.3c Scattering techniques Complementary information on the statistical properties of the microstructure of microemulsions can be collected using scattering techniques. In contrast to transmission electron microscopy, where the liquid mixtures have to be fixed through rapid freezing-scattering techniques are non-invasive and usually do not influence the mixtures. Moreover, these experiments performed in reciprocal space average over comparatively large volumes, ignore defects and irrelevant idiosyncracies, which can dominate a real space TEM picture. In each kind of scattering experiment, two physical properties play an important role, the length scale and the contrast, the latter being related to the difference in "power scattering" of the components forming the different structural domains in the samples. For the length scales, the fundamental variable is the scattering vector q, which, for elastic and quasi-elastic scattering, is defined as q =—nsinA =0.8) and in bulk contrast, that is by deuterating only one solvent. Again, the composition of the samples was located as close as possible to the X -point, i.e. at wc < wc < wc + 0.015 , to have a finite temperature interval for preparation. The scattering curves obtained are plotted in a double-logarithmic graph, where the scattering curve at <j> = 0.1 is given on an absolute scale and, for clarity, the other curves are displaced by factors of 5, for better visibility. Figure 5.19. SANS curves of microemulsions of the ternary mixture D 2 O-n-C g H 1 8 -C 1 2 E 5 prepared in bulk contrast as a function of cj>. All samples are measured near the respective X -point. The scattering curve at <j> = 0.1 corresponds to absolute intensities while the others are separated by factors of 5. As an example, for the curve at (j) = 0.\ the fit of [5.3.14] to the large-q part is shown. Each scattering curve can be equally well described by the Porod decay for diffuse profiles . The peak of the symmetrical bicontinuous microemulsion ( 0 = 0.5) is well fitted by the Teubner-Strey formula . = 0.9, the Teubner-Strey formula does not provide a satisfactory description of the scattering curve. Furthermore, the intensity at the large q part of the scattering can be used to determine the specific area of the internal interface41 — =0g^£. v V c wc m . In many studies, this isotropic channel, which increases temperature-wise from the water-rich to the oil-rich 11 , ), very similar curves of sigmoidal shape are observed. In nearly perfect agreement, both methods locate the transition from a branched tubular water network to discrete aggregates around an oil-to-water plus oil volume fraction of , the sponge-like, bicontinuous structure is found around the mean temperature of the three-phase body Tm , that is the p.i.t., if 0.2 < 0 < 0.8. On the oil-rich side, dispersed water droplets are found at high temperatures, which transform into a branched tubular water network with decreasing temperature. Accordingly, a gradual change in the mean and Gauss curvature of the amphiphilic film with temperature and composition is the overall observation. 5.3h Length scales From a quantitative evaluation of the type and length scale of the microstructure the mean curvature of the amphiphilic film can be deduced. An appropriate way to provide these numbers is detailed analysis of the SANS curves. i) Maximum length scale Starting at the optimum point of these systems, i.e. at the X-point (X-point at ) A/V . Curve a, elastic shear modulus for a homodisperse system; b, same for a polydisperse system; c, yield stress for a polydisperse system. Approximate results observed for emulsions. . Above a 2 \vm. ; the width of the droplet size distribution is substantial. The advantage Is that the amount of mechanical energy that is dissipated into heat can be substantially smaller than in most other methods. Current research may lead to further improvement. 3. Aerosol to liquid. One may atomize a liquid in air by forcing it at very high velocity through a nozzle. A spray of fine droplets can be obtained and be directed to hit another liquid, to form an emulsion. This is somewhat akin to injection, but the jet leaving the opening has a high Re, and drop formation is due to Inertial forces. The droplet size distribution tends to be quite wide, and the method has no real advantage over others. A spray of fine droplets can also be obtained by applying a high electric potential (up to +10 kV) to a liquid of high dielectric constant (water, ethanol, etc.). A receptacle then should be placed not far away from the spraying point, and have a negative potential. Small quantities of reasonably fine and homodisperse w o emulsions can be obtained In this way11. 4. Ultrasonic emulsification2\ An ultrasound generator causes great pressure fluctuations and can thereby produce cavitation in a liquid that has a substantial vapour pressure. For emulsiflcation, ultrasound of frequency 20 kHz is generally used. At high sound intensity, many very small cavities are formed. Such a cavity soon collapse again and it does so over a very short time (e.g., within 10 ns), thereby inducing a shock wave in the liquid. The wave concentrates, and loses energy, at sites where the compressibility changes, hence at a phase boundary. The interface can be disrupted, so drops are formed and disrupted again into smaller ones. The mechanisms involved are comparable to those in the TI regime discussed below. Small drops can result, but the size distribution tends to be quite wide, and is often bimodal. Upon prolonged exposure to ultrasound, a unimodal distribution of very small drops (down to d 32 = 0.15 |j.m ) can result, although at the cost of much energy. Cavitation can also occur in other machines, e.g. in some high pressure homogenlzers. Intense cavitation may also degrade high polymers and even induce chemical reactions. 5. Agitation. When a liquid is intensely agitated, hydrodynamic forces deform and 11 21 0.8 were placed between parallel plates, one being moved laterally, whereby a virtually constant Vu value is obtained; the slit width L was about five times the droplet size obtained, and Vu values up to 10 3 s~ 1 were applied. Almost homodisperse droplets were obtained, down to 5 urn in diameter. In another study41, Couette flow was used (Vu also constant) with larger L and smaller = 0.5, viscous forces take over. The further decrease in d max with increasing 0 ; - the concentration of the material of the disperse in the continuous phase is everywhere the same (corresponding to the solubility of the material in drops of radius a = a cr , explained below), except near the drop surfaces where a local concentration gradient exists. Ostwald ripening then results after some time in a number frequency distribution of the drop radius given by u 2 exp[3/ ( 2u-3)] (u + 3) 7/3 (1.5-u) 11/3 . The main problems are that the values of
dilute
— u(#>2 - cp^)
semidilute
2
-— JV
[1.6.2] vm2
dilute
2
In the dilute version of [1.6.1], we replaced e=-N 'ln^ b by C2 l&d2, according to [1.5.14], Note that the signs of co and « Ioc are different for adsorbing polymer (tp>
INTERACTION BETWEEN POLYMER LAYERS
1.21
the distal length) in dilute solutions, and about £/4 (with £ the correlation length) in the semidilute case. However, the presence of end points leads to a correction which is repulsive, with a decay length twice that of the attraction. As a result, Ga(h) has a (weak) repulsive barrier at relatively large separations and a (strong) attraction at short separations. 1.6b Attraction due to bridge formation For an isolated surface, the volume fraction profile in semidilute solutions is given by q>=
'
/7
mp
=
-«m-
We are interested in the sum of these two contributions to the disjoining pressure:
These discrete changes can be translated into continuous language in the following way. For the adjustment part we have h
/7 ad dh = - J (d«)dz 0
and for the midplane contribution we can write
11
P.G. De Gennes, Macromolecules 15 (1982) 492.
[1.6.4]
1.22
INTERACTION BETWEEN POLYMER LAYERS
% d h = - fflm dh
[1.6.5]
where a>m = co{
is
repulsive, because com is negative. The adjustment part /7 ad is attractive. We will see below that the net effect is still attraction, /7 ad + /7
< 0 . We note that the alternative
route to obtain the disjoining pressure, as used by De Gennes starting from a>loc (z), also gives attraction. To give a quantitative estimate of the attraction, we need the concentration profiles of the polymer within the slit. We start with the semidilute case. Upon substitution of [1.6.2a] into [1.6.4, 5] we obtain
0
L
J
For the concentration profile we use the linear superposition [1.5.9], which we rewrite as: — = coth 2 x + c o t h 2 y - l
x =Z +P
y = -Z + 2H + P
[1.6.7]
Note that in this case the left surface is situated at z = Oand as before Z = z / | , H = h/%, and P = p/£. For the evaluation of /7 a d , we need the differential quotient dip2 16H = lipdipl6H . From [1.6.7] we have d(ipl(ph)/dH = 4cothy (1 -coth 2 y) = -16e" 2 y cothy . It is sufficient to integrate over half of the slit. We take the left half where cothy = 1 and coth2 y - 1 « coth 2 x . We thus find 4
= -32e- <
H+p
2
2
)e * cosh x.
The
in first
integral
of
order 2
dip2 / &H ~ -32e~2y cosh 2 x 2
e * coth x
is
icothx(e2x-3) +
2(x + lnsinhx). Taking the leading terms at both boundaries, we obtain as a first approximation
^
/7 ad = -8v
[1.6.8]
Figure 1.5. Numerical (discrete) example of the function -OJ{Z) at separation h = 9 (closed diamonds) and at separation h + dh =10 (open circles). Areas I and II are two contributions to the disjoining pressure FI(h) . The bar diagram at the bottom gives -d
INTERACTION BETWEEN POLYMER LAYERS
1.23
With
[ 1 6 9 ]
We conclude that the leading contributions of 77 (repulsive) and /7 ad (attractive) exactly compensate each other! This illustrates the subtle balance between the various contributions. We also see that -77ad is larger than 7 7 ; the adjustment of the a>profile overcompensates the repulsive contribution of 77 . Integrating the disjoining pressure to obtain the Gibbs energy of interaction ([1.3]) is simple. We cast the result in the general exponential form G a (h) = -Ae~ 2/1/L
[1.6.10]
where L/2 is the decay length of the Gibbs energy of attraction. In this case L = £/2 or Ga - e~4h/%. The reason that the decay length equals £/4 is that a> is dominated by
1.24
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.6. Configurational entropy gain occurring when a walk starting from the left surface (black contour) reaches the midplane. It can follow some path (light gray), which returns to the same surface to make a loop, or it may follow with equal probability its mirror image (dark gray) to form a bridge.
exactly the same as that of making a bridge connecting to the opposite surface (dark gray), without any energetic penalty. The entropy is thus increased, which leads to attraction. Therefore, our statement above that there is attraction between loops has to be modified: the attraction arises because loops can transform into bridges. 1.6c Repulsion due to tails
In the previous section we considered only ground-state dominance for 'average segments', which for single surfaces describes only loops; tails are neglected. In this situation we only have attraction between two polymer layers in close proximity, both for adsorption and for depletion. For adsorbed polymer layers, end effects do play a role. This has been known since the 1980s due to the numerical work of Scheutjens et al. 1 '. Analytical theories describing tails became available only after 1995 and are largely due to Semenov et al. 2 The starting point is that, whereas the overall (loop) concentration is proportional to g2 , the end-point concentration is proportional to g (see [1.4.13]). From [1.4.13a] we have pe = N(pe / (pb = eeNg . Here, cpe = ^e(z) = (p(z,N) is the volume fraction of segments with ranking number JV . The total volume fraction of end points is then 2(pe . With £ = 0 and g = coth(Z + P) in semidilute solutions, and (f>he£N ~ 1 and g ~ l/sinh(Z + P) in dilute solutions, we may approximate the excess of end points with respect to the bulk solution as pe -1 = big -1) ~ coth 2 + P -1 pe=bg
semidilute dilute
sinh
[1.6.11a] [1.6.11b]
— d
11
J.M.H.M. Scheutjens, G.J. Fleer, J. Phys. Chem. 84 (1980) 168; G.J. Fleer, M.A. Cohen Stuart, J.M.H.M. Scheutjens, T. Cosgrove, and B. Vincent. Polymers at Interfaces. Chapman and Hall, 1993. 21 A.N. Semenov, J. Bonet Avalos, A. Johner, and J.-F Joanny, Macromolecules, 29 (1996) 2179; A.N. Semenov, J.-F. Joanny, and A. Johner. Polymer adsorption: Meanfield Theory and Ground state Dominance Approximation in: Theoretical and Mathematical Models in Polymer Research. Academic Press 1998.
INTERACTION BETWEEN POLYMER LAYERS
1.25
The factor b is a normalization constant, which ensures that the excess number of end points in the adsorbed layer equals the excess number of chains; it is defined through b\(g-l)dz
= Ug2 -l)dz for semidilute solutions and similar (without the term -1) for
dilute systems. We do not consider the precise value of b , suffice it to note that b is of order unity. In order to see the effect of chain ends on the interaction force, the GSA description as discussed in the previous sections has to be extended to include the end points. This was done by Semenov et al. 1 '. The disjoining pressure /7e(h) due to the ends was related to the end-point concentration
treatment led to attraction, the Semenov extension gives an additional contribution /7e(h) ~ +{
.
The most important implication is the sign: end effects give rise to repulsion between adsorbed polymer layers. This repulsion may be interpreted as the repulsion originating from an ideal gas of end points. When the two surfaces are brought closer, the concentration of chain ends increases, the 'gas' pressure increases, and the plates repel each other. In order to make a quantitative estimate, we approximate gm - 1 in semidilute solutions
as
gm - 1 ~ coth(H + P ) - l = e-^h+p)/^
por
dilute
solutions,
we find
gm ~ l/sinh(H + P) ~ e~' h + P' / d for the dilute case. The result, written in the same way as [1.6.10], is Gae(h) = R e - h / L
[1.6.12]
As before, L equals £/2 in semidilute solutions and d / 2 in dilute solutions. The decay length L for the repulsion in [1.6.12] is twice the decay length L / 2 for the attraction in [1.6.10]. The reason is that the overall profile (~g2),
which determines
the attraction, decays two times faster than the end-point profile ( ~ g ).
1.6d Total Gibbs energy of interaction Upon combining [1.6.10] and [1.6.12], we obtain for the total Gibbs energy of interaction Ga =-Ae~2h/L
+ Re~h/L
where L is the decay length of the repulsion,
[1.6.13] L / 2 that of the attraction, and
A (attraction) and R (repulsion) are the amplitudes.
We use a dimensionless form
Ga = G a ( 2 IkT . A sketch of Ga[h), according to [1.6.13], is given in fig. 1.7. The curve has a maximum Gm at h = hm and passes through zero at h = hQ.
11
A.N. Semcnov, J.-F Joanny, and A. Johner,toecit.
1.26
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.7. Sketch of Ga(h) according to [1.6.13]. The dashed curves are the attractive (loop/bridge) and repulsive (tail) contributions.
When L , A and R are known, these characteristic values follow from G
m
R2 = ^
2A hm = Lln—
ho=hm-Lln2
[1.6.14]
Exact numerical SCF calculations also provide a dependence Ga(h) as sketched in fig. 1.7. From such curves, the values of Gm , h m and h 0 follow directly for any bulk concentration
k
°
A = Gme2h™/L
R = 2Gmeh™/L
[1.6.15]
In the next section we will see that exact numerical results can indeed be interpreted in this way. 1.7 Numerical examples for adsorption In this section, we present some numerical results for the Gibbs energy of interaction between two surfaces with adsorbed polymer layers In full equilibrium. We first show in section 1.7a that Indeed the dominating effect is attraction, but that there is a weak repulsion at large separation. The interaction curves for weak overlap can be analyzed along the lines of [1.6.13] and fig. 1.7, using the parameters Gm (height of the maximum), h m (position of the maximum), and L (decay length of the repulsion or twice the decay length of the attraction); this will be discussed in section 1.7b. In section 1.7c, we will show how the bridging attraction found at strong overlap depends on the governing parameters. 1.7a Interaction curves
In fig. 1.8 we give the whole Interaction curve for N = 1000 and q>h = 10"2 (which is in the semidilute range). The linear plot of fig. 1.8a shows an enlarged view of the region of weak overlap around the maximum. The symbols in fig. 1.8a are the num-
INTERACTION BETWEEN POLYMER LAYERS
1.27
Figure 1.8. The Gibbs energy of interaction Ga=Ga(2/kT as a function of the plate separation 2h (in units ( ) for N = 1000, ^ = 10~2 , ^fs = 1, and % = 0 . In diagram (a) the region near the maximum is plotted with linear scales. The maximum G m , the corresponding position h m , and the position /1Q where Ga = 0 are indicated. The symbols are the SCF results, the curve is the fit with the two exponentials according to [1.6.13], using A = 1.41 10~ 2 , i? = 3.11 10" 3 , and L = 2.95 . The attractive part of the interaction curves for the range h < /IQ (solid curve) is plotted in panels (b) and (c) on semi-logarithmic and doublelogarithmic coordinates, respectively. In these graphs, the dashed curve is the extrapolation of the solid curve in diagram (a). The dotted curve was calculated from numerical integration of the superposition profile according to [ 1.5.9].
erical SCF data. The maximum is found at hm =7.0 and has a value of Gm = 1.71-10"4. The point where the interaction energy is zero corresponds to hQ = 4.96. From these data, [1.6.15] gives A = 1.4110~2, R = 3.11 10~3 , and L = 2.95 . The solid curve in fig. 1.8 is [1.6.13] with these parameters. In this region of weak overlap the fit is quite good, which shows that near the maximum the interaction curve can indeed be described as a difference between two exponentials with a single interaction length L , which is the decay length of the repulsion; that of the attraction is L / 2 . For h
1.28
INTERACTION BETWEEN POLYMER LAYERS
For higher separations the discrepancy with the numerical SCF results is larger because then the end-point repulsion (not accounted for in [1.5.9]) starts to play a role. The numerical data cross over smoothly from the dotted curve at low h (which constitutes an overestimation of - G a ) to the dashed (double-exponential) curve at higher h. Even though we have no accurate analytical description over the full range, this crossover from one analytical limit to the other is gratifying. From fig. 1.8b,c we see that the attractive part of the interaction curve has a complex structure. It is neither a true exponential nor a true power-law curve. In fig. 1.8c it would be possible to force a power law through the data in the middle range with an exponent not far from -3 . For very small slit widths an exponential decay of the interaction is found in fig. 1.8b, which is not present in the GSA result. GSA is not expected to hold when the train layers start to overlap.
Figure 1.9. (a) The position hm of the maximum in the interaction curve, the position h 0 where the interaction is zero, and the interaction length L as a function of the polymer concentration tp^. For comparison, the distal length d and the correlation length £ are also plotted. All lengths are in units { . (b) The maximum G m in the interaction curve as a function of the polymer concentration cp^ . Both graphs are for JV = 1000, xs = 1 anc> X = ° •
1.7b Weak overlap In the following sets of graphs we discuss the interaction for weak overlap. One example was shown in fig. 1.8a. We now present the dependencies of Gm , its position hm, the position h0 where Ga = 0 , and of the characteristic interaction length L on
INTERACTION BETWEEN POLYMER LAYERS
1.29
L goes through a maximum, following the distal length d in dilute solutions but the correlation length £, in the semidilute range. In fig. 1.9a, d and £, are also indicated (dotted curves). These two lengths cross at a point well into the semidilute regime. The maximum in L, however, is situated around the overlap concentration <pov . In the dilute range, L is close to d/2 and in the semidilute region to £/ 2, as expected from section 1.6b. The value of Gm is a rather strong function of the polymer concentration, roughly as Gm « Jip^ for low
Figure 1.10. (a) The position h m of the maximum in the interaction curve, the position h 0 where the interaction is zero, and the interaction length L as a function of the chain length JV. For comparison, the distal length d and the correlation length £ are also plotted, (b) The maximum G m in the interaction curve as a function of the chain length JV. Both graphs are for ^ = 1 0 - 3 , j s = 1 and x = 0 •
The interaction also depends on the chain length (fig. 1.10). In this graph the polymer concentration was fixed at
1.30
INTERACTION BETWEEN POLYMER LAYERS
or less logarithmic dependence with N is found. On the other side of the maximum, the drop may be explained from the fact that the adsorbed amount goes down with decreasing chain length. Thus, when one is interested to use homopolymers to stabilize particles, one should choose an intermediate degree of polymerization. Very long polymers are not effective because they do not have enough ends; short polymers do not adsorb strongly enough.
Figure 1.11. (a) The position h m of the maximum in the interaction curve, the position h 0 where the interaction is zero, and the interaction length L as a function of the (effective) adsorption energy A%s . For comparison, the distal length d is also plotted. All lengths are in units t . (b) The maximum G m in the interaction curve as a function oiAx . Both graphs are for p,j = 10" 3 , N = 1000 and x = 0 •
Figure 1.11 shows the effect of A%s = Xs~ Xsc o n the characteristics of the maximum. The adsorption energy does not dramatically influence the spatial characteristics of the maximum in the interaction curves, at least for relatively strong adsorption. Only when the adsorption energy is close to xSc ^° w e °t> serve that the position of the maximum shifts to higher h values. This signals the approach to the critical region. The height of the maximum increases monotonically with the adsorption energy. This can be understood from the fact that the adsorbed amount and, hence, the number of end points is an increasing function of the effective adsorption energy. To stabilize particles with homopolymers one should, therefore, use strongly adsorbing polymer. 1.7c Strong overlap In figs. 1.12,13 a number of plots is collected similar to fig. 1.8c under various conditions. The attractive part of the interaction energy (h < hQ ) is plotted on a doublelogarithmic scale for three values of the bulk concentration cph (fig. 1.12a), the chain length JV (fig. 1.12b), the effective adsorption energy Axs (fig- 1.13a), and the solvency X (fig. 1.13b). The first thing to notice is that for narrow gaps only the adsorption energy has an effect; for small h the curves for different
INTERACTION BETWEEN POLYMER LAYERS
1.31
Figure 1.12. Numerical SCF data for the attractive part of Ga in the range h
whereas there Is an effect of ipb , N, and %. All the curves start at 2h = 1, which corresponds to one (lattice) layer of solution between the plates. In this situation, the attraction Is linear In the effective adsorption energy. For larger plate separations, the attraction becomes weaker in all cases. For h close to h0 (where the loop/bridge attraction and the tall repulsion just compensate each other), the curves become approximately independent of %s (fig. 1.13a), but they do depend on the other parameters. The interpartlcle separation where the curves bend upwards is determined mainly by the value of h , which increases with decreasing
Figure 1.13. Numerical SCF data for the attractive part of G a , in the range h < h0 , for three values of A%s (a) and x (b)- I n diagram (a) JV = 1000, (p^ = 10" 3 , / = 0 , and in (b) JV = 1000, ( 9 b - 1 0 - 3 , Xs = 1-
1.32
INTERACTION BETWEEN POLYMER LAYERS
The solvent quality affects to a very large degree the range of the attraction, as shown in fig. 1.13b. Going from a good ( % = 0 ) to a theta solvent, the attractive component becomes stronger because the attraction between (loop) segments reduces the excluded volume effect; at x = 0.5 , the second virial coefficient vanishes. The repulsive effect of the chain ends is then relat less important. Apparently, GSA is qualitatively correct in this case and the interaction curve for ^=0.5 in the double-logarithmic coordinates of fig. 1.13b extends to very large h. Indeed, for large h the interaction is purely exponential (not shown). At shorter plate separations (2 < 2h < 15), we again find a power-law-like dependence of the interaction force. Now the exponent is to a good approximation equal to -2. 1.8 Depletion at flat surfaces 1.8a Introduction In the previous sections we have mainly considered adsorbing polymers, where the segment adsorption energy ~xskT is the driving force for accumulation of polymer near the surface. When xs i s at>ove the critical value xSc • t n i s adsorption energy overcompensates the entropy loss experienced by the chains near an impenetrable surface, so that there is positive adsorption. When such a driving force is absent, only the unfavorable entropy remains, and a depletion zone develops where the concentration is lower than in the bulk solution. A schematic picture is given in fig. 1.14. This figure is a variant of fig. II.5.9, and gives half the profile in a wide slit (fig 1.1 top right). The characteristic width 8 of the depletion zone is of the order of the radius of gyration a g in the dilute solutions, but becomes smaller as, at higher concentrations, the osmotic pressure in the solution pushes the chains closer to the surface. Beyond overlap of the coils in solution (
Figure 1.14. Volume fraction profile for depletion at a flat wall. The continuous profile is indicated by the solid curve, the dashed lines represent a step profile with width 5, which is the depletion thickness. The hatched areas left and right of z = S are equal.
INTERACTION BETWEEN POLYMER LAYERS
1.33
The depletion thickness may be related to the adsorption F , which is the excess of material per unit area. A general definition of F was given in [II.2.1.2]. For depletion, the excess of polymer is negative, so F < 0. We will use a dimensionless form 0 ex for the excess of polymer, F = Bex/t2 (on a segment basis). In a lattice model, 0ex is found by summing (f>(z)-
'--^-Jo-'V'T-Je-*)'? rb
o
o
Hence, the depletion thickness can be computed from either the end-point concentration profile or the overall concentration profile. As in the previous sections, we can either use the Edwards equation (or some GSA equivalent) or the numerical SCF lattice model. We will show that in all cases the overall profile is well approximated as p = tanh 2 z/S . When two surfaces, each with a depletion layer, approach each other to an interparticle distance 2h , which is of order 25, the layers overlap, see fig. 1.1. We could try to solve for the profile along the lines discussed in section 1.5, and from that compute the Gibbs energy of interaction as was done in sec. 1.6 for adsorbing polymer. We would then get quantitative analytical information for weak overlap, but no easy analytical solutions for strong overlap. We shall not follow that route, as for depletion the step-function approach gives a convenient alternative. We then replace the continuous profile for each surface by the dashed lines in fig. 1.14. The idea is sketched in fig. 1.15. In this picture the interaction is zero for 2h > 25 ; the region of weak overlap is completely neglected. However, we can rather accurately describe the region of strong overlap: the attraction is roughly given by G[h) = -J7gsVov(h), where G{h) = Ga[h)A is in Joules, /7gs is the osmotic pressure in the bulk solution, and Vov(h) = 2(8-h)A . Dividing by the area A gives Ga(h) = 2/7g s (h-5), which is negative for h<8 and
Figure 1.15. Step-function approach for the overlap of two depletion layers. The overlap volume Vov is 2(5 - h) times the area A per plate.
1.34
INTERACTION BETWEEN POLYMER LAYERS
linear in h . We will see that this simple approach works quite well, but that for an accurate description of Ga the parameter S (thickness of a single depletion layer) has to be replaced by an interaction distance Si, which is slightly larger.
1.8b Depletion thickness and the profile at a single surface (i) Dilute limit. In dilute solutions, the field u in [1.4.1,2] is essentially zero, and the Edwards equation can be solved exactly. Using the boundary conditions G(O,s) = 0 and G(oo,s) = 1, Eisenriegler
derived
p e (z) = G(z,JV) = erf-^2a g
[1.8.2]
for the end points, and a slightly more complicated expression for the overall profile p(z). In the dilute limit, a g = t^N/G is the only relevant length scale. By inserting pe into [ 1.8.1 ], the depletion thickness is found as S0=~a sin 6 We use the subscript
[1.8.3]
0
to indicate that this result applies to the limit of zero
concentration. Tuinier et al.2) showed that the rather complicated exact expression for the overall profile is approximated rather accurately by p=tanh2—
[1.8.4]
fit) Semidilute limit. Above the overlap concentration, the chain-length dependence disappears and the relevant length scale is the blob diameter £,. According to [1.5.4c] it is defined as £2 I^ = 3Uh/kT
(with U b given by [ 1.4.2b]), which for good solvents
2
may be expanded to give t I ^ = 3v(pb. In GSA, the profile follows from [1.5.5] with g m = l so that dZ/dg = l-g2
. Here Z = z/£
and the minus sign was taken for the
square root of [1.5.5] because gf < 1 and dg/dZ>0.
With the boundary condition
g(0) = 0 , the solution is Z = arctanhg , or g = tanhZ . From p = g2 , /7=tanh 2 ^
[1.8.5]
We see that the coth, which followed from adsorption in the semidilute regime (see [1.5.7]), is replaced by a tanh in depletion. Note also the close analogy with [ 1.8.4]. The only difference is the length scale, as So = S0(N) in [1.8.4] is replaced by £, = ^(pb,x) in [1.8.5].
11 E. Eisenriegler, J. Chem. Phys. 79 (1983) 1052; E. Eisenriegler, Polymers near Surfaces, World Scientific (1993). 2) R. Tuinier, G.A. Vliegenthart, and H.N.W. Lekkerkerkcr, J. Chem. Phys. 113 (2000) 10768; G. J. Fleer, A.M. Skvortsov, and R. Tuinier, Macromolecules 36 (2003) 7857.
INTERACTION BETWEEN POLYMER LAYERS
1.35
Inserting [ 1.8.5] Into [ 1.8.1 ] gives the well-known De Gennes result 8=$
[1.8.6]
which expresses that in the semidilute case the depletion thickness (an interfacial property) is equal to the blob diameter (which is a solution property). It is possible to generalise the expression t2 IE,2 = 3txpb to poorer solvents by inserting in t2 IE} = 3Ub / kT the full expression [ 1.4.2b] for U b
^ = ^f
=
-3H1-^)
+ ^b]
[1-8.7]
This form reduces to the known limits in a good solvent (I2 IE,2 = 3vq>b) and in a theta solvent (t2 IE} =3cp2>/2). Moreover, it describes the transition region (£ = 0.4-0.5, relevant for many practical situations) as well. A plot of E, as a function of
Figure 1.16. The correlation length E, (in units C ) in semidilute solutions as a function of
As shown by Fleer et al.11, with this generalized form of E,, [1.8.5] describes the depletion profile in good solvents very accurately and in theta solvents quite reasonably (in the latter case [1.8.4] gets a slightly different mathematical form). Similarly, 8 = E, is accurate in good solvents and a reasonable approximation in a theta solvent. (ill) Generalization. The two limits 5 0 = <50(JV) (dilute) and £ = |(
[1.8.8]
£•£•? According to [1.8.6], the depletion thickness 8 in semidilute solutions equals the bulk solution correlation length E,, which in this case does not depend on N. Khoklov and Grossberg21 introduced a generalized (JV-dependent) correlation length <^(JV), which is 11 21
G.J. Fleer, A.M. Skvortsov, and R. Tuinier, toe. cit. A.Y. Grossberg and A.R. Khoklov, Statistical Physics of Macromolecules, AIP Press, 1994.
1.36
INTERACTION BETWEEN POLYMER LAYERS
defined through C2 /^ 2 (AT) = 3[1/ N + ((3/kT)dnos I dip]. Here, the second term is the inverse of the osmotic compressibility. Using the mean-field version of d/7 os / dip, given in [ 1.2.6], it can be shown that 8 , as defined in [ 1.8.9], is essentially the same as £(JV). This generalizes the statement "depletion thickness equals bulk solution correlation length," first formulated by De Gennes as 8 = E, (where E, does not depend on JV), to more dilute systems where 8 = §(JV), which includes the JV-dependence of 5 and £, .
Figure 1.17. The depletion layer thickness 8 (in units t ) at a flat surface, according to [1.8.9] (solid curves), as compared with exact SCF results (symbols), for two chain lengths (N = 1000, top, and N = 100, bottom) and for three solvencies. The value of xs in the lattice model was chosen as Xs = — (x + l)/6 , which is the value for which
(iv) Comparison with exact lattice results. Figure 1.17 shows <5(cpb) as calculated from 11.8.9] for three solvencies and two chain lengths, and compares the results with numerical SCF data. For % = 0 and 0.4, the agreement is quantitative, whereas for ^ = 0.5 in semidilute solutions [1.8.9] slightly overestimates the exact lattice result. The reason is a different mathematical form for the profile under theta conditions. It can be shown that the simple form 8 - % overestimates the thickness of the depletion zone by some 10%. A correction is possible11, but we do not go into details. For
G.J. Fleer, A.M. Skvortsov, and R. Tuinier, loc. cit.
INTERACTION BETWEEN POLYMER LAYERS
1.37
Figure 1.18. Depletion profiles (with z in units C ) at a flat surface for four different volume fractions in the bulk solution (indicated), at / = 0 and JV = 1OOO. The curves were computed with [1.8.8,9]; symbols are SCF results for £ s = - 1 / 6 .
lengths. Even in those cases where the depletion thickness is somewhat overestimated by [1.8.9] (as for % = °- 5 i n fig 1-17), the tanh2-profile fits the numerical data accurately, provided the width (= zeroth moment) of the profile (which may be too high by some 10% using [ 1.8.9]) is taken from the numerical data. 1.8c Interfacial Gibbs energy for a single surface It is useful to have an expression for the polymer contribution to the interfacial Gibbs energy y-y* = Qa (where Qa=Q/A is the grand potential per unit area), because this quantity is needed for the Gibbs energy of interaction between two surfaces. According to [1.1.1], Ga (h) = 2[i3 a (h)-f2 a Hl = 2[y(h)-yH] • In this section we consider only the value grand potential £2a (<x>), which we denote simply as Qa . We will find that Qa is positive, which implies a negative surface pressure (c.f. chapters III.3 and 4 for monolayers). There are two obvious ways to find Qa . The first is given in [1.4.9], Qa = I^bS ~ noslz)]dz , where the bulk osmotic pressure 77bs and the local osmotic pressure f7os(z) were defined in [1.2.3] and [1.4.8]. We denote this route to find Qa as the osmotic method. The second way is based upon integration of Gibbs' law dy = -Fdji or, in terms of 0ex and Qa , C2dQa = -9exd)i. Here F and /i are defined per segment; /i is given by [1.2.5]. A similar integration was used in [II.1.1.7] and [III.3.1.2]. Because in this method we need the (excess) adsorbed amount during addition of polymer to the system (from pure solvent to the final value
QaH=
r
?nos
J S(l-
[1.8.10]
where
1.38
INTERACTION BETWEEN POLYMER LAYERS
the system. For 5 , we can take [1.8.9] and dnos/dcp
is given by [1.2.6]. This equation
has a form which is quite different from [1.3.2], Qa = J w d z / ( . Yet, when numerically exact values for 81 C = - 0 e x /
to give
according to [1.8.8], the three terms may be integrated
Qa = {q>h8 / t)Q. / N + 2vcpb / 3 + 23
2
[1.8.3],
1/AT =
2
(2/3n)(C15 0 ) , and [1.8.7], [C/Q = 3wpb +(3/2)
=^
|
[1.8.11]
This equation may be considered as a generalisation of the exact expression derived by Eisenriegler et al.
and Louis et al.21 for ideal chains in the dilute limit, which is
[1.8.11] with <50 instead of 5. Equation [1.8.11] extends this expression to finite concentrations.
1.8d Gibbs energy of interaction For calculating Ga[h), we need not only Qa(°°) for a single plate, but also iia{K) per plate when the plate separation is 2h . Again, we can use either the osmotic method or the adsorption method. With the osmotic route we calculate Qa = J (/7gs - J70S(z))dz , where /70S(z) depends on the concentration profile in (half) the slit, which has to be computed. In the adsorption method we can start from [1.8.10], but 0 ex = -
dn°s
y
Qa(h)= J h(l-cp)——d
0
[1.8.12]
For the Gibbs energy of interaction we subtract [1.8.10] from [1.8.12]. We write the difference 0 ex (h)-0 ex (°°) as (
where H(x) is the Heaviside step-
function, which is zero for x < 0 and unity for x > 0 . Hence, 9
r
dnos
Ga(h) = -2j [8-h](l-q>)?—H[8-h]dcp o ^
[1.8.13]
When, at a given iph, a constant value <5 = <5(ipb) would be used in [1.8.13], Ga(h) 11 21
E.E. Eisenriegler, A. Hanke, and S. Dietrich, Phys. Rev. E. 54 (1996) 1134. A.A. Louis, P.G. Bolhuis, E.J. Meijer, and J.-P. Hansen, J. Chem. Phys. 116 (2002) 10547.
INTERACTION BETWEEN POLYMER LAYERS
1.39
would be strictly linear in the range 0 < h < <5(
interaction Ga(h) = Ga(h)C2/kT as a function of the plate separation 2h (in units t ), for IV = 1000, ^ = 0.4, and four bulk concentrations
We conclude that [1.8.13] works quite well. However, this equation still requires a numerical integration. The basic feature of fig. 1.19 is a linear Ga{h) dependence over most of the range. A simple, explicit analytical approximation for this linear part is obtained by applying the step-function approach to Ga(h) = 2[Qa{h) - Qa{oo)] with Qa (h) given by [ 1.8.12] and fla(°°) by [ 1.8.11 ]. For h -> 0 there is no polymer in the slit, so Ga(h = 0) = -2£2a(~) = -(4 / 3n)(pb 18 . For narrow slits the gap also contains only solvent, so /7os(z) = 0 and Q{h) = /7gsh according to [1.4.9]. Hence,
Ga(h) = 2ng s h-2r2 a H = -2f3aH i-Aj
[i.8.i4]
with Qa{oo) = (2/3ji)
[1.8.15]
Intuitively, one expects 5j (the zeroth moment of the osmotic pressure profile) to be of order 8 (the zeroth moment of the volume fraction profile). Indeed, from Qa = lkT/C3)
1.40
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.20. The characteristic quantities Ga (0) = Ga (0)C2 / kT and 8{ as a function of
G a (h) = - 2 H ° s ( 5 i - h )
[1.8.17]
is correct provided the proper interaction distance is taken. This distance is not the depletion thickness 5 at a single plate (as presupposed in fig. 1.15), but 54 as defined in [1.8.15], which is slightly higher than 8. With ng s 5j = Qa , [1.8.17b] is the same as [1.8.14] 1.9 Depletion around a sphere 1.9a Volume fraction profile In section 1.8b we discussed the zero-field solution of the Edwards equation [1.4.1] in flat geometry, which led to pe = erf(z/2a ) for the end-point distribution. For spherical geometry, with V2 in spherical coordinates, such a zero-field solution also exists11. We consider a sphere of radius a, with a radial coordinate r from the center of 11
E.E. Eisenriegler, A. Hanke, and S. Dietrich, loc. cit.
INTERACTION BETWEEN POLYMER LAYERS
1.41
the sphere. The (radial) distance from the sphere surface is then z = r - a . As for a flat plate, the surface is situated at z = 0 . The end-point distribution is now p e = [z/a + erf(z / 2a g )]/(z/a+ 1), which reduces to the flat-plate limit for a —> °°. There is also an exact expression for the overall distribution p , but we do not give that here. We restrict ourselves in this chapter to the approximation
p = [z/a + t a n h z / 5 0 ] 2 /
(z/a + 1)2, which reduces to [1.8.4] for a —> °° , with 5 0 = 2a g /V?r
according to
[1.8.3]. In a similar vein as the transition from [1.8.4] to [1.8.8] for flat plates, we may generalise this to finite concentrations by replacing 5 0 by 8 = S{N,(ph,%) as given by [1.8.9]: - + tanh — ^ - +1 \ a )
p= ^
[1.9.1]
Figure 1.21 gives an example of the profile predicted by [1.9.1] as compared with the numerical SCF results. In this example (semidilute solution, relatively good solvent) [1.9.1] works almost perfectly. In other cases (theta solvent) deviations may occur, which are related to the slight overestimation of S in a theta solvent as shown in fig. 1.17. In good solvents the predictions of [1.9.1] are very accurate for large a / 5 , but the width of the profile is slightly overestimated when a~8l}.
However, in all cases the trends are predicted very
well so that [ 1.9.1 ] is a useful approximation.
2.9b Depletion thickness around a sphere In [1.9.1], 5 is the depletion thickness at a flat plate given by [1.8.9]. This is not equal to the depletion thickness 8S around a sphere, because of geometrical reasons. Analogous to [ 1.8.1 ], 8S may be related to the negative adsorption
where p is given by [ 1.9.1 ] and 0| x is the total depleted amount of segments in a shell
Figure 1.21. Depletion profiles (with z in units I) around a sphere of radius a = 20f for four chain lengths, at
11
G.J. Fleer, A.M. Skvortsov, and R. Tuinier, loc. cit.
1.42
INTERACTION BETWEEN POLYMER LAYERS
with thickness <5S around the sphere. It is equal to Ts (the excess per unit area) times the area of the sphere 0| x =4rea 2 r s
[1.9.3]
Substitution of [1.9.1] into [1.9.2] and carrying out the integration gives the relation between 8S and 5 as (1 + 8S la)3 = l + 35/a + (7T2 /4)(5/a) 2 . This is nearly the same expression as the exact solution derived by Aarts et al. for the dilute limit (8 = So). The only difference is that the numerical factor n2 I4 is replaced by 3n 14 . In order to get agreement with the dilute limit, we use this factor 3^/4
fl + ^ f = , + 3 * 3 f * f {
a)
a
[1.9.4,
4 {a)
Figure 1.22 gives a double-logarithmic plot of <5s/5 as a function of a/8 . For large particles 8S = S as expected, but for small a/8 , Ss/a scales as (a/8) . Combination of the three equations [1.9.2-4] gives, for the excess amount per unit area, s
C3 V 4 a J
For a —> °° , this reduces to r = 0ex/f2 =-(pb8/t3 , which is the flat-plate excess as given by [ 1.8.1 ]. The second term of [ 1.9.5] is the curvature correction.
Figure 1.22. The ratio Ss/S between the depletion thickness <5S around a sphere of radius a and the depletion thickness 5 at a flat plate as a function of the ratio a/5 . The dashed line corresponds to the asymptotic behaviour for small a/8: Ss/8=
(3K/4)1/3(O/5)1/3.
1,9c Interfacial Gibbs energy of a sphere
In [ 1.9.5] we saw that the adsorption Ts per unit area around a sphere may be split up into a flat contribution, F = -
[1.9.6]
The flat contribution Qa was discussed in section 1.8c, and we can directly use the 11 D.G.A.L. Aarts, R. Tuinier, and H.N.W. Lekkerkerker, J. Phys. Condens. Matter 14 (2002) 7551.
INTERACTION BETWEEN POLYMER LAYERS
1.43
result as given in [1.8.11]. In this section we consider only the curvature contribution Qc . In the SCF model, it is obtained as Qs - £>a , where both terms have to be evaluated numerically. In the adsorption method, we start from dQc = -[Fs - F)dn , where Fs - T = -(n / 4)((j>b / C3 )82 I a according to [1.9.5]. Analogous to the derivation of [1.8.10], this leads to <2c=^fs2(l-
[1.9.7]
The integrand depends only very weakly on cp. Let us consider two limiting cases. In dilute solutions 5 2 = 5g ={4/n)N/6 and d/7 o s / dip = 1 / JV, so Qc = < p b / 6 a . In semidilute solutions in a good solvent 82 ~ t;2 - (Suip)"1 and dnos / dq> ~ vq>, so Qc ~ [K/2)(ph/6a. Hence, Qc~
fi
--!^=iHr
[1 9 81
--
for two reasons. The first is that this form coincides with the exact solution11 Qc / kT = [
11 21
A.A. Louis, P.G. Bolhuis, E.J. Meijer, and J.-P. Hansen, J. Chem. Phys. 116 (2002) 10547. CM. Wijmans, F.A.M. Leermakers, and G.J. Fleer, Langmuir 10 (1994) 4514.
INTERACTION BETWEEN POLYMER LAYERS
1.44
Figure 1.23. The overlap volume Vov (gray area) between two spheres of radius a with a depletion layer of thickness 5S at separation 1h between the sphere surfaces.
The geometry is sketched in fig. 1.23. From elementary analytical geometry the overlap volume is easily derived: V
oV(h) = ^-{8s-h?{3a
+ 2S
s+h)
0
[1.9.9]
where 8S is given by [ 1.9.4] Now the Gibbs energy of interaction can be written as an extension of [1.8.13): G
'r' 3/7°s V 1 a( )=-] ov( -
0
[1.9.10]
^
where Vov is a function of
aj =-Cct
C = 4wln2 = 9.71
[1.9.11]
For h = 0 , Vov(0) in [1.9.9] reduces to 2na82 for 8S « a . In this case 8S = 8 (see fig. 1.22), which equals 8Q = (2/4n)a g for cpb -> 0 . Hence, Vov = 8 a • a | . With d/7 o s /dip = kT/NC3 , [ 1.9.10] reduces to [ 1.9.11 ] with C = 8 , which is very close. According to [1.9.11] with a | IN I2 = 1/6, in the dilute limit and for a / 5 s > l , G a (0) depends only on the product a
11
E. Eisenriegler, Phys. Rev. E 55 (1997) 3116.
INTERACTION BETWEEN POLYMER LAYERS
1.45
These features show up in fig. 1.24, which gives Ga[h) for cph = 0 . 0 5 , a/£ = 10, N = 1000 and three solvencies. Indeed, the contact potentials are nearly independent of solvency, whereas the slope is higher and, hence, the range of attraction smaller for better solvents. We note that Ga(h) is roughly proportional to a so that curves for different particle sizes have the same general shape. When
Figure 1.24. The Gibbs energy of interaction Ga (in units kT per site) as a function of the slit separation 1h (in units () according to [1.9.10], for a/C = 10,
1.10
Adsorption versus depletion
In this section, we briefly discuss how adsorption or depletion phenomena influence a colloidal system and how one can experimentally recognize whether adsorption (loop/ bridge) or depletion (overlap of depletion zones) is the relevant attractive contribution in the system. This task is, as we will see, not trivial, the more so because the polymer contribution to the forces between colloids is rarely the only one; Van der Waals forces, electrostatic forces, undulation forces, etc. may mask the observations. We have seen that the depletion mechanism contributes a relatively weak interaction force to the system with an associated length scale proportional to the bulk correlation length (semidilute) or the coil size (dilute solutions). We may anticipate that, in the absence of other interactions, depletion may give rise to a phase transition in the system where there is a phase rich in polymer and dilute in particles (colloidal gas) in equilibrium with a phase dilute in polymer but rich in particles (colloidal liquid or solid), see IV.5.7. It is known11 that the ratio 8i/a between the range of interaction and the particle size determines whether the phase rich in particles remains fluid-like or is a solid (crystal). The latter occurs for 5{ la « 1, whereas a liquid phase is formed when 5j / a > . l . In sec. 1.8b we saw that 5j [~ 5) decreases with increasing polymer concentration. This complicates the theoretical description of such phase behaviour. However, the quantitative dependence 8{N,
M.G. Noro, D. Frenkel, J. Chem. Phys. 113 (2000) 2941.
1.46
INTERACTION BETWEEN POLYMER LAYERS
The relevant length scale in loop/bridge attraction is also strongly related to the bulk correlation length, but the strength of the bridging attraction is usually much stronger. We illustrate this in figure 1.25 where we present the depth Ga (0) of the attraction (at 'contact') as a function of the adsorption energy Axs with respect to the critical point Xsc for a homopolymer with chain length 1000 in good solvent (% - 0 ) and for two values of the polymer concentration; near overlap (
Figure 1.25. Numerical SCF results for the dimensionless Gibbs energy of interaction Ga (0) between two plates at 'contact' as a function of the adsorption energy Axs = xs ~ XSc ^or a w 'de range of adsorption energies (a) and for the region near the critical value for the adsorption energy (b). The dotted line in (a) is a linear extrapolation according to Ga (0) = 0.2 - Axs • Parameters N = 1000, cpb = 10~3 and 10~ 2 ,and x = °In fig. 1.25a we see that the loop/bridge attraction, which is found for positive values of Axs, is much stronger than the depletion interaction. On the scale of this figure, we can hardly see the attractive interaction caused by depletion in the range Axs < 0 . For high adsorption energy the contact interaction depends linearly on Axs • The dashed line, drawn according to Ga(0) = 0.2 - A%s, describes the data well for Axs > 0.3 . It is interesting to note that this linear dependence breaks down near the critical condition. To examine the critical region, as well as the depletion region, we also present an enlarged view (-0.1 < Axs < 0.1 ) of the same data in fig. 1.25b. Note the difference in absolute values of the interaction energy in the two figures (more than a factor of 100). Only on this scale is the attraction in the depletion range visible. In the first approximation we expect the depletion minimum for strong repulsion to follow G a (0) = 2[Qa(0)-Qa{oo)] =-2i2 a (~)~
INTERACTION BETWEEN POLYMER LAYERS
1.47
1.13a in sec. 1.7c. Obviously, for very large negative A%s (strongly repelling surface), Ga(0) becomes independent of Ax • However, the dependence of Ga(0) around Axs = 0 may be important for depletion interactions in experimental situations. A very small remaining adsorption energy 0 < xs < Xsc =0.18 may seriously influence the strength of depletion interactions. The standard depletion theories all start from the assumption of a strongly repelling surface, which may not always correspond to the real experimental conditions. Near the critical condition one would intuitively expect that the depth of the interaction curves becomes zero. However, in this region GSA is not applicable because the chains near the surface have conformations in which there is a significant ranking number dependence. (Another case where this ranking number dependence shows up so that GSA fails is for brushes, as will be discussed in the next section). In other words, tails are important in these cases. As a result, the contact Interaction becomes positive in the critical region! Note that the absolute value of the positive contact energy is of the same magnitude as the maximum contact interaction found for depletion. Figure 1.25 may be used to elaborate on the issue of the introduction to this section, i.e. how to distinguish between adsorption and depletion. For very strong adsorption the situation is obvious in most cases. All particles will be covered by polymer, and it is not difficult to verify that experimentally, e.g. by dynamic light scattering. Standard techniques will also allow the measurement of the amount of polymer adsorbed on the particles (e.g. with the classical bulk depletion technique or by reflectrometry). However, when the adsorption energy is not far from critical, one may wonder whether the attraction due to the polymers has a loop/bridge or a depletion origin. The key idea here is to use the repulsion found near the critical region to discriminate between bridging and depletion. It is well-known that the effective adsorption energy may be tuned by the addition of a so-called displacer1'. If, by adding a displacer to the system, the attraction first reduces in strength and then increases again, one must have crossed the critical region. The original system must have been in the bridging regime. This trick may become more problematic if addition of a second solvent also influences the effective solvent quality. In section 1.12 we shall discuss some kinetic effects, and at the end of that section one may find additional guidelines to distinguish adsorption from depletion in experimental situations. Such guidelines are useful since many examples can be found in the literature where one presupposes a depletion mechanism without sufficient justification. Not only depletion at a hard wall, as discussed in the present section, is important, but also 'soft depletion' is often encountered. We shall return to this issue in sec. 1.1 lh. The key idea here is to use the repulsion found near the critical region to 11 M.A. Cohen Stuart, G.J. Fleer, and J.M.H.M. Scheutjens, J. Colloid Interface Sci. 97 (1984) 515, 526.
1.48
INTERACTION BETWEEN POLYMER LAYERS
discriminate between bridging and depletion. It is well known that the effective adsorption energy may be tuned by addition of a so-called displacer11. If by adding a displacer to the system the attraction first reduces in strength and then increases again, one must have crossed the critical region. The original system must have been in the bridging regime. This trick may become more problematic if addition of a second solvent also influences the effective solvent quality. In section 1.12 we shall discuss some kinetic effects, and at the end of that section one may find additional guidelines to distinguish adsorption from depletion in experimental situations. Such guidelines are useful since many examples can be found in the literature where one presupposes a depletion mechanism without sufficient justification. Not only depletion at a hard wall, as discussed in the present section, is important; 'soft depletion' is often also encountered. We shall return to this issue in sec. 1.1 lh. 1.11
Tethered polymers and polymer brushes
1.21a Introduction In this section we consider polymers that are terminally attached, or tethered by one end to flat surfaces. Such systems are easily made experimentally and provide an important tool to control colloidal stability. The combination of insights into the behaviour of end-grafted chains with those for unconstrained homopolymer at interfaces (sees. 1.1-9) provides the basic tools to understand the equilibrium characteristics of most, if not all, interfacial polymeric systems. Some aspects of end-tethered polymer chains were already discussed in sec. III.3.4 on Langmuir monolayers. In the previous sections we investigated what happens when homopolymers floating freely in solution are exposed to a solid surface. It was shown that the excess adsorbed amount is one of the important measurable quantities. The number of interfacial chains is determined on the one hand by the solution properties (concentration, chain Figure 1.26. The Huygens equal-time-of-flight oscillator (top) and the corresponding polymer conformations (bottom) in a brush (gray background). On top, a mass suspending from a string with four arbitrary 'starting' positions with zero velocity is indicated in various gray levels; the oscillation time is the same for all cases. At the bottom, the corresponding typical conformations are drawn. The local stretching of the chain increases towards the surface, which translates into the increase in velocity of the mass when it accelerates towards the bottom of the parabolic potential. Note that the relative frequency of each conformation in the brush is not the same for each starting point.
11 M.A. Cohen Stuart, G.J. Fleer, and J.M.H.M. Scheutjens, J. Colloid Interface Sci. 97 (1984) 515.
INTERACTION BETWEEN POLYMER LAYERS
1.49
length, solvent quality) and on the other by the surface properties (adsorption energy). The equilibration of the interfacial chains with those in solution may lead to several subtle effects. An important result is that the layer thickness of adsorbed chains never exceeds the coil size in solution, and for strong adsorption it is even considerably smaller. In this case, the ground-state approximation (GSA) can be successfully applied, because to first order all segments have the same distribution. Directly linked to this symmetric behaviour is the attractive loop/bridge contribution to the colloidal stability. We also saw that tails, i.e. chain fragments with an asymmetric segment distribution, give a repulsive contribution to the pair potential. However, for adsorbing polymer this is a small effect: the main features are captured by GSA, leading to attraction. If equilibration of the polymer chains is restricted, for example by imposing a confining constraint on one of the segments of the chains, the system properties change dramatically as compared with the unrestricted case. One important issue is that the number of chains on the surface can be varied independently, and does not depend on solution properties. Due to the constraint, the condition of equal chemical potentials of end-tethered chains and unconfined chains in solution no longer applies. The second key point is that each segment along the chain automatically gets its own spatial distribution, distinctly different from all others. This is easily seen when, e.g. the first segment s = 1 of the chain with ranking numbers s = 1, ..., N is tethered to the surface, the second segment has just a few degrees of freedom and can be distributed in a region 0 < z < f (where t is the bond length). The third and fourth segments can distribute more freely, and the free end s = N can be in the region 0< z<[N-1)1. The upper limit only occurs when the chain Is fully stretched, which is an unlikely event unless the chains are extremely densely grafted. From these ranking, number-dependent distributions, we must anticipate a complete failure of GSA. Indeed, this is the case, and as we will see GSA is replaced by an astonishingly simple theoretical approach based upon the equal-time-of-flight principle first suggested by Semenov1', the so-called strong, stretching limit. When the number of restricted chains is high, such that at the surface a polymer film with high (semi-dilute or higher) concentration exists, the layer becomes easily (much) thicker than the radius of gyration. This implies that the individual chains are typically stretched as compared with their random coil configurations. A layer of strongly stretched chains is called a polymer brush. The breakdown of GSA does not imply that the Edwards equation [1.4.1] cannot be used. This equation remains valid and the Semenov approach gives a remarkably simple functional form for the selfconsistent field u . The key element is that each chain, starting from its free end, has to reach the 11 A.N. Semenov, Zhur Eksp. Teor. Fix. 88 (1985) 1242; transl. as Sou Phys. JETP 61 (1985) 733.
1.50
INTERACTION BETWEEN POLYMER LAYERS
surface ( z = 0 ) in JV steps, where JV is chain length; the number of steps is the same, irrespective of the starting positions. There is a clear analogy with an harmonic oscillator where the time (the number of 'time steps') to reach the bottom (z = 0 ) is the same for any starting position of the oscillator (see fig. 1.26). Dating back to the days of Huygens, we know that such an equal-time-of-flight oscillation corresponds to a parabolic potential well; the oscillation time of an harmonic oscillator only depends on the length of the oscillator and the pending mass, but not on the amplitude it is given. In a polymer brush a similar phenomenon is present. For any position of the free end (equivalent to the amplitude), after JV steps (equivalent with time) the chain has to reach the surface. Consequently, the self-consistent potential energy field must have the equal-time-of-flight property and, therefore, must be parabolic. In good solvents we know that u ~ cp and thus the brush has as a parabolic concentration profile (more details to follow below). It is necessary to mention the approximation made in this equal time-of-flight argument. From the analogy employed, it is obvious that we only consider those conformations that have the property that segments with higher ranking numbers must be further from the surface than segments with a lower ranking number. 'Loops' in a conformation, where a segment with a lower ranking number happens to be at a higher coordinate than a segment with a higher ranking number, are neglected. This approximation is reasonable when the chains are strongly stretched (at high grafting density), but less so when loops ('fluctuations') become important at low grafting density. It is possible to solve numerically the Edwards equation (or, equivalently, use the numerical SCF model) and check the accuracy. From this exercise, we know that the predictions from the strong-stretching approximation are accurate in a wide range of relevant conditions (see fig. 1.30 below). When describing a polymer brush, one could say that it is composed of a set of polymer tails. The overlap of tails leads to an increase in local concentration. The high segment concentration, in turn, gives a high osmotic pressure, which is unfavourable to the system. Hence, repulsion is expected when two brushes in a good solvent are forced to overlap. Indeed, because there are so many chains on the surface, the expected forces are large and end-tethered polymer layers are an excellent means to stabilize colloidal systems. 1.11b Where do polymer brushes occur? End-tethered chains can be produced in several ways. One way is to chemically graft (existing) chains to a surface. In this approach, it is possible to make a homodisperse brush when the starting chains are all of the same length, but for obvious reasons it is not easy to make a dense brush. Another method is to grow new chains from a surface in contact with a solution. In this method it is easier to generate a dense brush, but it is more complicated to control the polydispersity. There are
INTERACTION BETWEEN POLYMER LAYERS
1.51
several examples in the literature 12 ' 3>4) (see also sec. III.3.8f) showing that one can indeed tune the control parameters, such as the chain length and the grafting density, to physically relevant values. Such systems are being used to test existing theories, and to control surfaces in nanotechnology applications. Polymer brushes, where each chain is fixed with one end at a surface, represent an extreme case. However, similar conformations are found in many situations as the constraints imposed on a chain may also originate from correlations within the macromolecules themselves. Brushes will typically occur when copolymeric materials are adsorbed (or grafted) on the surface. Although we will not treat these systems in any detail, it is worthwhile to briefly mention such systems. If we consider a copolymer composed of two types of segments (A and B), there are several scenarios. Let us first assume that the solvent is not selective and that the solvent quality is good for both segment types, i.e. for the whole chain. There are many ways to distribute the segments A and B along the chain. Let us consider the extreme case of a diblock copolymer A n B m . In this case there are n segments A chemically grafted to m segments B. When both n and m are large, and when the average adsorption energy is sufficiently above a critical value, the polymer chains will adsorb onto the surface already at very low polymer concentration in the bulk. Even the slightest disparity in surface affinity (xsA * # sB ) will cause one of the blocks to be preferentially on the surface because, typically, WxsA - m ^ s B | » 1 because n and m are large. The plateau of the adsorbed amount is now found when the adsorbing block (let us assume this block is composed of segments A) covers approximately all adsorption sites, such that the surface layer is nearly filled. Obviously, the B-block then protrudes into the solution, similar to end-grafted chains on the surface, but with the starting point of the B-block displaced somewhat from the surface and distributed through space to some extent. Assuming that the system is at full coverage, the grafting density is inversely proportional to the length of the A-block and the thickness of the brush is given by the degree of polymerization of the B-block. A much better scenario presents itself when the solvent is selective, i.e. one block is soluble whereas the other is not. An extra driving force (escape of the insoluble segments from the solvent) is now present to bring the insoluble chain parts together and form a brush composed of the soluble entity. When n and m are large, one could qualify such polymers as polymeric surfactants. In many polymeric surfactant systems we can observe brush formation. The corona of a micelle, irrespective of the 11 E.P.K. Currie, F.A.M. Leermakers, M.A. Cohen Stuart, and G.J. Fleer, Macromolecules 32 (1999) 487. 21 E.P.K. Currie, W. Norde, and M.A. Cohen Stuart, Ado. Colloid Interface Sci. 100-102 (2003) 205. 31 J.H. Maas, M.A. Cohen Stuart, A.B. Sieval, H. Suilhof, and E.J.R. Sudholter, Thin Solid Films 426 (2003) 135. 41 J.H. Maas, G.J. Fleer, F.A.M. Leermakers, and M.A. Cohen Stuart, Langmuir 18 (2002) 8871.
1.52
INTERACTION BETWEEN POLYMER LAYERS
geometry (spherical or cylindrical micelles or flat lamellae) is always brush-like. Polymeric surfactants are excellent substances to cover surfaces with brushes (adsorbed micelles, adsorbed monolayers and bi-layers). Again, the grafting density of the corona chains, as well as the length of these brushes, may be tuned by choosing the values of n and m. Some more details were discussed in chapter III.4. Polymer brushes may also form in systems that feature macromolecules with a more complex internal topology. As an example, we mention the generic properties of comb-like copolymer. The insight here is that in most cases there is some chemical disparity in the macromolecule; the teeth usually have different chemical groups as compared with the backbone. This disparity will generate a structure with some local asymmetry. The organization may in part be driven by unfavourable enthalpic interactions, but is also helped by a tendency to reduce the local osmotic pressure (chain crowding). In solution, a comb-like copolymer with a sufficient number of teeth may be seen as a 'bottle-brush' with a structure similar to that of worm-like micelles (in this case with a quenched 'aggregation number'). Like most macromolecules, they may also be surface active. When they have an adsorbing backbone with nonadsorbing teeth attached to it, they may form very robust brushes with high grafting densities. In conclusion, one can safely say that a brush is often a key structural element of polymers at interfaces and that the physical consequences are substantial and important. 1.11c Thermodynamics In most of section 1.11 we will be concerned with systems in which there are no chains in solution. For such a system, the number of chains at the surface is also automatically the excess number of chains. We then consider brushes in pure solvent. However, in section 1.1 lh we will briefly visit a system with a brush in the presence of an unrestricted polymer in solution. We first discuss the macroscopic thermodynamics of such systems, including the possibility of additional (free) components. The end-grafting of some chains prevents their equilibration with chains in the bulk. Due to this grafting, the chains are distinguishable from any (otherwise chemically identical) mobile ones. Even when the chains are chemically linked to the surface, we may still assign them a chemical potential, albeit they do not have the usual translational degrees of freedom. We will use the symbol fi for the chemical potential of grafted chains; ft does not have a translational part, unlike all other chemical potentials /i{. The Helmholtz energy of the system is F(iVp, {JV,}, A, V, T) = -pV + MpNp +X i / / i JV i + YA • where V is the volume of the system and where all mobile components are indicated with the index i; the grafted polymer (index p) is written explicitly. The differential form is dF = -SdT - pdV + MpdN + ^ //jdJVj + ydA . At least in a thought experiment one should be able to determine the change of the Helmholtz energy by changing the
INTERACTION BETWEEN POLYMER LAYERS
1.53
number of grafted chains, while keeping the remaining degrees of freedom fixed to obtain the chemical potential p of the grafted chains. The surface tension y is found by changing the surface area at a fixed number of grafted chains. This quantity is experimentally accessible when the chains are 'grafted' on a mobile (e.g. liquid-air) surface. On a homogeneous flat surface one typically expresses the extensive quantities per unit of surface area. The relevant thermodynamic potential that determines the interaction forces between two surfaces covered by a brush is then the semi-grand potential per unit area. Because of the constraint imposed on the grafted chains, their number per unit area is fixed (although their chemical potential as a function of h varies) and the number of solvent molecules (and all other mobile components) is variable (at constant chemical potential of the solvent). As the macroscopic pressure is also fixed, we prefer to work in a constant pressure, rather than a constant volume ensemble and thus N
p
QSO
fl
=
" "^- A"^
M i
"Z
V +P
A
= (7
~7*) + /iPa
[1 11 11
- -
where so again denotes semi-open, and where we introduced the number of grafted chains per unit area a = N / A , often denoted as grafting density. When the chains are permanently grafted to a solid substrate, one can defend the point of view that the chains (with conformational degrees of freedom) are an inherent part of the surface and thus is it possible to include the chemical potential term of the grafted polymer pa in the surface tension contribution. However, in a simple brush system where segments of the grafted chains do not interact with the surface, one may argue that the notation of [1.11.1], where the surface tension and the chemical potential of the grafted chains occur, is useful. This is so because the main contribution of interest is then pa and we can assume y -j* , independent of a . This Ansatz is used in most of the analyses in the following subsections. However, in section 1.1 lh the full equation [1.11.1] will be used and it must be understood that the numerical SCF model more easily gives access to Q^°, rather than to the surface tension or the chemical potential of the grafted chains. We mention again that Q^° [1.11.1] is the central quantity that must be computed as a function of the separation 2h between two surfaces. Typically, it is calculated per surface (i.e. over half of the slit). As before, the Gibbs energy of interaction is found as Ga(h) = 2(i2|0(h)-r2|°(oo)), where the factor 2 is needed to account for the two surfaces, as in [1.1.1]. 1.1 Id Isolated grafted chains on a single surface: mushrooms In general, segments of grafted chains interact with the substrate. They are either repelled by or attracted to the surface. For high-grafting density o~, a brush develops in which the majority of the segments cannot find a place on the surface so that the
1.54
INTERACTION BETWEEN POLYMER LAYERS
segment-surface interaction, as expressed by the term y — y* in [1.11.1], hardly makes a difference for the macroscopic properties, such as the brush height or the way brushes affect the colloidal stability; entropic (crowding) effects dominate In this regime we may use [1.11.1] in the form £2|° = jx a because y = y* . When o" is low and no substantial crowding occurs, the repulsion and attraction situations are very different, leading to so-called mushrooms for repulsion and pancakes for attraction (see fig. 1.27).
Figure 1.27. Illustration of a mushroom on a repulsive surface (top) and a pancake on an attractive surface (bottom). In a mushroom the conformation of the chain resembles that of a chain in solution, with length scale a«. In a pancake the thickness is 1/c (see fig. 1.2), which is of order C .
To describe mushrooms or pancakes, we can start again from the Edwards equation [1.4.1] and note that, for this dilute case u = 0, like in the Henry region of the adsorption isotherm or in the dilute depletion limit (section 1.8b(£)lConfigurations with one end on the surface are characterized by the function Gat (z, IV) (where the subscript a t denotes 'attached') for which the boundary conditions are quite different from those in the previous sections, Gat(0,0) = 1 (the grafted end) and Gat(°°,IV) = 0 . As before, we can express the segment-wall interaction by the parameter c as defined in [1.4.3] and fig. 1.2. (Note, however, that in this case G = 0 in the bulk 'solution'). The exact solution for terminally attached configurations, for arbitrary value of c and expressed in the scaled variable Z (in units of twice the radius of gyration a ? ) is available" (p=Gat(z,N)=-^— + cY(z-ca«)e-z2 & •Jna \ > _
&
J
Z= — 2a
[1.11.2]
&
Here
11
Y. Lepine, A. Caille, Can. J. Phys. 56 (1978) 403.
INTERACTION BETWEEN POLYMER LAYERS
e~z2
Gat{z,N) = -^
1.55
[1.11.3]
The overall distribution of segments
[1.11.4]
which is the same as [ 1.8.2] With this form of the composition law, the overall concentration profile turns out to be
6VJra N
s- erf (2Z) - erf (Z)\ L
[1.11.5]
J
Figure 1.28 shows the two distributions (pe{z) and ip{z), according to [1.11.4 and 5], respectively. Both distributions are normalized to unity. Clearly, the behaviour at large z is reminiscent of a Gaussian, but near the surface both distributions are more like linear functions of z, starting at the origin. One also sees that the average distance of free ends (the first moment of
l.lle
Brushes on a single surface
(i) General features. Tethered chains can be considered mutually non-interacting as long as their inverse surface density CT"1 is smaller than the area r02 = Nt2
Figure 1.28. End-point distribution
1.56
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.29. Artist impression of a polymer brush. The height L and the grafting density a are indicated. Note that not all end points are in the periphery of the brush; the end-point concentration increases with increasing distance from the wall. The overall density decreases monotonically.
occupied by an isolated mushroom. Hence, this condition roughly amounts to (jNC2 ~ 1 (for ideal chains or theta solvent). When the grafting density is substantially higher, the chains interact strongly, and we have a polymer brush. We sketch such a brush in fig. 1.29. Since the segment concentration and, hence, the osmotic pressure in a brush is high, brushes tend to swell substantially (compared with a ), particularly in good solvents, thereby reducing their concentration. However, all chains have one end on the surface, so the segment density can only be lowered by stretching the brush in the direction normal to the surface. Stretching a chain (increasing the average distance between grafted and free end) will lower the conformational entropy (increase the Helmholtz energy) and, therefore, give rise to a restoring elastic force between the ends. In contrast, a brush in a poor solvent tends to contract to dimensions smaller than the ideal Gaussian coil. This contraction also reduces the number of available conformations and the associated entropy, and thus also produces an elastic force. It is possible to calculate density profiles for brushes on the basis of the Edwards equation [1.4.1] or with the numerical SCF scheme (see section 1.4b), but a simple scaling analysis of brushes, pioneered by Alexander11 and De Gennes21, already provides much insight. In this analysis it is assumed that all chains are stretched to the same extent, so that it is enough to consider the balance between stretching and osmotic pressure for a layer of homogeneous density
S. Alexander, J. Phys. France 38 (1977) 983. P.-G de Gennes, Macromolecules 13 (1980) 1069.
INTERACTION BETWEEN POLYMER LAYERS
1.57
is given by the well-known Gaussian formula: _3r^
G(r,JV) = Ce
2r
o
r$ = Nt2 = 6 a |
[1.11.6]
where C is a normalization constant and rQ the unperturbed, averaged end-to-end distance. The integral of G(r,JV) over all r is the single chain partition function. Hence, G(r, JV) can be looked upon as a partial partition function (namely, that part of the total partition function for which the end point is at position r). Quite often in the literature, the characteristic function of this partition function is given the letter F . In fact, it is the elastic stretching part of the chemical potential of the chain, which in [1.1.11] was denoted by jx . Here we choose a compromise and use the notation F e l . The elastic Helmholtz energy for stretching a free chain (with unperturbed end-to-end distance r0 ) to a distance r follows as:
^l^lni^L^t-XV-D kT G(ro,JV) 2{rg J 2 l
[1.11.7]
which is positive for r > r 0 . We have defined an expansion (deformation) coefficient a by: a =— r
[1.11.8]
0
which is larger than unity when the chain is stretched.
Note that the yardstick for
brushes is the end-to-end distance r0 = ^VJV (determining the degree of stretching a ), whereas in describing the solution properties of chains typically the radius of gyration a g ( = r0 / V6 ) Is appropriate. The elastic/orce corresponding to [ 1.11.7] is / e l = -3F el /dr or
^=4=-A a kT
r02
[M1.9]
r0
Chains confined in a slit of width r smaller than r0 undergo compression (a < 1). Also in this case the free energy is increased. One can show1'21 that for strong confinement the partition function G(z,iV) scales as exp[-7r2r02 / 6r 2 ] so that the elastic Helmholtz energy with respect to unperturbed chains is now
which is again positive because a < 1. In order to have a continuous description for any a , an interpolation formula is often used, which is simply the sum of [ 1.11.7] and [1.11.10]: 11
A.K. Dolan, S.F. Edwards, Proc. Roy. Soc. A337 (1974) 509. E.F. Casassa, J. Polym Sci B: Polym Lett. 5 (1967) 773, Macromolecules 9 (1976) 182; E.F. Casassa, Y. Tagami, Macromolecules 2 (1969) 14. 21
1.58
INTERACTION BETWEEN POLYMER LAYERS
— =-(a2 +cr2-2) kT 2
[1.11.11]
with the associated elastic force
l111121
w-H-st)
{Hi) Box model. Let us now consider a brush of homogeneous density tp and thickness L in which all chains have their free end at L, so that the expansion coefficient for any chain equals a =— r o
[1.11.13]
This simple model (step function for
is the number of
chains in the system and A the area. Clearly, the amount of segments per unit area equals JV- JVp / A = No . Since the area per segment is t2 , NG('2 is the number of segments per segmental area or, in a lattice model, per site. We use the symbol 9 for this quantity e =
d = at2
[1.11.14]
In the equations to follow, the dimensionless grafting density a will often occur. The volume occupied by the polymer is N • N £3 , so that the average volume fraction (p in the brush is N • N £3 / AL or (p = (7lV- = (7lV— L ar0
[1.11.15]
As shown above, there is an elastic force per chain fei. unit area, there is an elastic 'pressure' J7el = felo
Since there are a chains per
in the brush. Mechanical equil-
ibrium requires zero (internal) pressure; hence, the osmotic pressure /7 0S needed to balance this is simply J7os=-/elo-
[1.11.16]
Inserting J7OS ~ vq)2 /2 + q>3 / 3 (which is [1.2.4] but without the
where, analogous to [ 1.11.8], the equilibrium swelling coefficient is defined by aeq=Leq/r0 [1.11.18] We can now examine various limiting cases of [ 1.11.17 ]
INTERACTION BETWEEN POLYMER LAYERS In good solvents, l/o:
1.59
is small and we find
implying that the thickness of a brush in a good solvent increases as o"1/3 . This has been verified experimentally (see, e.g. fig. III.3.96 in section III.3.8f). We shall see in the next section that for a more realistic (parabolic) profile the same scaling applies as in [1.11.19], with a slightly higher numerical prefactor. In theta solvents (u = 0), and for low grafting densities (small 6), we have from [1.1.17] simply a = l, but when the grafting density is sufficiently high ( 0 » 3 ) we obtain
Mt)" V 2
¥-(!)"'»
Again, the numerical factor depends on the details of the model. Finally, in poor solvents (v < 0), layers contract; now the term a 3 can be ignored and we find a brush thickness proportional to 6 = dN : e
1
3|u|
C
3\v\
3 \v\
In this latter case, the factor \v\ plays the role of an effective density, which controls the thickness of the collapsed brush. Note that in all three cases, L ~ N according to [1.11.19-21], which is a direct consequence of the dominating lateral interactions in a brush, tending to align the chains. [iv) Parabolic potential profile. The 'box model' assumption that all chains are deformed to the same extent, and that the polymer density in the brush is uniform ignores fluctuations in the distribution of end points and oversimplifies the overall profile. A slightly more sophisticated approach assumes that the total Helmholtz energy F of a chain, i.e. its chemical potential (composed of an elastic and an osmotic contribution) is fixed, but that the way this is realized by a given chain may differ, depending on where its end point is located (see fig. 1.26). Strongly stretched chains have a small osmotic term and a large elastic contribution, whereas the reverse applies for weakly stretched ones. When the total Helmholtz energy is minimized, subject to this constraint of constant F , one finds quite generally that the field u = u(z) in a brush is not uniform but must have a parabolic shape1'21, u = B-C(z/L)2 . Referring once more to fig. 1.26 we may elaborate on the analogy between a particle moving in a parabolic well (harmonic 11
A.N. Semenov, Zhur. Eksp. Tear. Fiz. 88 (1985) 1242, transl. as Sou. Phys. JETP 61 (1985) 733. 21 S.T. Milner, T.A. Witten, and M.E. Cates, Europhys. Lett. 5 (1988) 503; Macromolecules 21 (1988) 2610.
1.60
INTERACTION BETWEEN POLYMER LAYERS
oscillator) and the chain ends in a polymer brush. From classical mechanics, the period x of an harmonic oscillator is given by x2 = 2n2m / C, where m is the mass of the oscillating particle and C/2 the spring constant. Here we are interested in a quarter of the oscillation time, as this is the time used for the particle to travel from its starting point (with zero velocity) to the bottom of the well (z = 0). We translate the particle problem to the brush by considering a unit mass and set N = x 14, exploiting the analogy between steps along the chain and time 'steps.' Consequently, C = In2 / T 2 = n2 /(8JV2). For a not too dense brush, u = vq> +
Z = z/L e q [1.11.22]
L
eq
(4vd)U\r
-r=brj
N
3(n2)1/3
«° = 21TJ
1/3.2/3
v a
The equilibrium thickness scales in exactly the same way as in the box model, compare [1.11.19], but is higher by a factor (24/7T2)173 =1.345 . In a theta solvent (u = 0, u =
Z = z/Leq [1.11.23]
t
% v3 )
Again, the scaling is the same as in [1.11.20], but the prefactor in L is higher by a factor (2/7r)121/4 =1.185 . Of special interest is the distribution of the free end points. In the box model and in the Alexander-De Gennes model, the end points are all located at z = L . Fluctuations of end points are not accounted for. The more exact analysis based on the equal time of flight argument shows that the ends cannot all be located at the edge, because for a given subset of conformations with fixed (free) end position, the concentration profile is an increasing function with the distance. The reason is that the local stretching in each chain is stronger near the surface than at larger z. This applies to all chains with fixed position for the end point. Hence, if all end points were at z = L , the overall profile would increase with increasing z. We know that the overall profile is doing just the opposite; therefore there must be a non-trivial distribution of free ends. In the strong stretching limit it is possible to obtain analytical expressions for the (free) endpoint distribution
E.B. Zhulina, O.V. Borisov, and V.A. Pryamitsin, J. Colloid Interface Sci. 137 (1990) 495.
INTERACTION BETWEEN POLYMER LAYERS
1.61
result is
^ =0
[1.11.24a]
# = 0.5
[1.11.24b]
In this parabolic potential profile, (pe (z) = 0 for z > L . Again, Z = z/ L , where Leq is defined in [ 1.11.22,23] for x = 0 and 0.5, respectively. From these equations it is seen that the end-point distributions are linear in z for X = 0.5 and approximately so for x - 0 (unless z is close to L ). These simple relations for the end-point distributions may be used to show that the fluctuations of the end-points can be decsribed as 5e = ((z e ) 2 -(z 2 ))/JV °= N , where (ze) = Jz
A. Chakrabarti, R. Toral, Macromolecules 23 (1991) 2016. T. Cosgrove, T. Heath, B. van Lent, F. Leermakers, and J. Scheutjens, Macromolecules 10 (1987) 1692. 31 P. Auroy, L Auvray, and L. Leger, Macromolecules 24 (1991) 2523. 41 A Karim, V.V. Tsukruk, and J.F. Doublas, J. Phys. II 5 (1995) 1441. 51 CM. Wijmans, E.B. Zhulina, and G.J. Fleer, Macromolecules 27 (1994) 3238.
1.62
INTERACTION BETWEEN POLYMER LAYERS
0.2
I
1.5
Figure 1.30. Segment volume fraction profiles
In fig. 1.30c,d we present the normalized concentration profiles for the end points. The profiles approach the analytical forms given in [1.11.24] for long chains, but the profiles for the short chains deviate from this asymptotic behaviour. Again, this is not unexpected because the shorter chains are not very deep in the brush regime. In fig. 1.30 the theoretical brush thickness L c*=iVcr1/3 in a good solvent and L °c No112 were used to normalize the abscissa axis. The fact that these scaling coordinates work well and give (approximately) universal curves proves that the numerical profiles obey these scaling laws for the brush thickness very accurately. Numerical evaluation of the thermodynamic properties of the brush, such as the surface pressure, show that scaling predictions for these properties are very accurate as well. Some of these predictions, for example that the surface pressure scales as JV(75/3 , were discussed and compared with experiment in sec. III.3.4J for a liquid surface. We conclude that, especially for systems well in the brush regime and for cases where the polymer concentration in the brush is not too high (long chains), the simple theory is already very accurate. This implies that this simple theory can be
INTERACTION BETWEEN POLYMER LAYERS
1.63
used as a good starting point for the discussion of the interaction between brushes. [vi] Scaling analysis. As an alternative to the mean-field approach sketched above, it is possible to take into account the excluded volume correlations occurring in polymers in a good solvent. This approach distinguishes regions in which swelling due to excluded volume is dominant (so-called 'blobs') and considers the interactions between blobs as hard-sphere interactions. This is pictorially represented in fig 1.31a; blobs were also mentioned in 1.5b in the context of the properties of semidilute solutions. The scaling analysis is mostly relevant for very long and flexible chains in good solvents; these have a relatively wide semidilute concentration regime in which spatial fluctuations are important. Each blob in a chain contains JVb segments and, hence, there are NI Nb blobs of size £ in a chain. We can still use an elastic contribution to the Helmholtz energy of the form F e l ~{L/rQ)2 as in [1.11.7], but now rQ refers to an ideal chain of (swollen) blobs rather than segments, which implies that in [1.11.5b] r^ = Nt2 is replaced by 2 75 according to [H.5.2.8], as discussed also in r 2 = (jV/JVb)£; , with % given by % = fiV^ section 1.2. Eliminating Nb , one finds pel
T2
T7T N$1/3
kT
[1.11.25]
with an associated stretching force, / e l = -3F el / 3L fel
r T7T
[1.11.26]
kT JV£1/3 Using the well-known result £, ~ cp"3/4 , according to [II.5.2.18], where in this case
T«1/4
may
be
compared
/ e l ~ cc/rQ = L/r02 = L/N(2
with
the
mean-field
result
given
in
[1.11.9]
with
. The difference is a factor
1.64
INTERACTION BETWEEN POLYMER LAYERS
unity, when
^ 1 =^ - ^ / 4
[L11.28]
which also differs from its mean-field value nos ~ q>2 , according to [1.2.4], by a factor
[1.11.29]
which, with
[1.11.30]
The result has exactly the same scaling as that obtained from the mean-field argument [1.11.19] for a good solvent. Note, however, that both the osmotic pressure and the stretching force were different; these differences happen to cancel exactly for L . In semidilute solutions the blob size or correlation length scales as £ ~ (p~a , where in excluded volume scaling a = 3 / 4 and for the mean-field case a = 1/2 . In the brush the density is just a function of the grafting density,
= cr 2/3 .
The blob size in the brush is thus only a function of the grafting density, i.e. £ ~ ( j ~ 2 a / 3 , which is proportional to the average distance between grafting points in the excluded volume scaling (E,/i = 6~xl2),
but larger than that in the mean-field
1/3
picture (S,/ f. = <7~ ). As a consequence, the blobs in the mean-field picture overlap as shown in fig. 1.31b. More importantly, the Helmholtz energy per chain, which is found by counting the number of blobs per chain, is not the same for excluded volume chains as for mean-field chains. Therefore, we must expect that the interaction Helmholtz energy for the two approaches will not be the same. In other words, the cancellation just mentioned leading to the same brush height in both approaches is not expected for interaction. Yet, we will see that the predictions of both models are roughly the same. 1.1 If Interaction between surfaces covered with mushrooms (i) Mean field. Compression of a mushroom will lead to an increase of the Gibbs energy and, hence, to repulsion. We consider two flat plates at distance 2h with some mushrooms between them; the segments do not adsorb (fig. 1.32). As discussed in the text above [1.11.10], the elastic deformation work to obtain one strongly confined mushroom can be calculated by the continuous mean-field method and is approximately equal to kTn2 /6a2, where a is now given by a = 2h/r 0 (a < 1). Again, as for a simple brush, we can in [1.11.1] ignore the surface tension term and
11
P.-G. de Gennes, Scaling Concepts in Polymer Physics, Cornell Univ. Press (1979).
INTERACTION BETWEEN POLYMER LAYERS
1.65
Figure 1.32. Compression of mushrooms between two surfaces a distance 2h apart. In this figure, the top chain is grafted to the left surface and the bottom chain to the right surface. As the chains are so far apart that they do not feel each other, it is obvious that the same compression force is felt when the two chains are both grafted on the left surface and are compressed by a bare surface. In the scaling picture, one counts the number of blobs (spheres) per chain to find the interaction Helmholtz energy per chain.
identify jx = F. As we have two plates, each with a chains per unit area, we compress 2aA chains by a factor 1h / r0 . Hence, the total interaction equals Ga(h),^^ = 2 ^ M
2
=^r
,2h
[1.11.31]
The disjoining pressure fj(h) = -dG a (h)/d2h is
[ii) Scaling. A scaling argument can also be given. When a chain In a good solvent Is confined between two parallel surfaces, it can still behave as a self-avoiding walk (SAW) on length scales up to 2h. Since a three-dimensional SAW of size 2h accommodates (2h/^) 5/3 segments, chains longer than this number of segments must spread out sideways. One can consider these flattened chains as two-dimensional arrays of sub chains or 'blobs', each of which takes a diameter 2h (see fig. 1.32 where three such blobs are indicated). The Helmholtz energy required to deform a chain of length JV (which would have an end-to-end distance In solution of r = N3/5( ) would then be given by kT times the number of newly formed blobs, F e l = (IVINb)kT , where JVb is the number of segments per blob. Since the blob size is 2h = £N^5, we have JVb = (2h/(f/3. Similarly, for an unperturbed chain, N = (rQ/()5/S- For one mushroom, this leads to an elastic Helmholtz energy of the form
Since we have a (non-interacting) chains, we obtain the following scaling expression for the interaction Gibbs energy
1.66
INTERACTION BETWEEN POLYMER LAYERS
G a (h)-ff^J
-SJjJ
(2h
[1.11.34]
^
[1.11.35,
The disjoining pressure becomes ^ ~ ^ ) "
8 / 3
Note that these fi-dependencies are slightly weaker than those from the mean-field results [1.11.31,32], namely by a factor (h/f) 1/3 . Compression of mushrooms gives a weak repulsion. Assuming an unperturbed end-to-end distance of 10 nm, and a grafting density smaller than r,j2 (mushroom regime), e.g. 0.005 nm"2, one finds from [1.11.31] at In = 3 nm a compression energy of about 0.18 kT per nm2, which is about 0.6 mN/m at room temperature. This is much less than for brushes, as will be shown below. Nevertheless, this mushroom repulsion is much higher than the tail repulsion for adsorbing homopolymers (typically of order 10"4 fcT/f2) or the attraction due to depletion (of order 10"3 kT/C2). 1.1 lg Interaction between surfaces with brushes (i) From Helmholtz energy per chain to interaction forces. When the grafting density increases and a brush forms, the situation is more complicated. Now, chains do not just interact with the two surfaces, but also with each other. What happens is then determined largely by the extent of interpenetration of the two layers. In good solvents, we have seen that the grafted chains (on a single surface) repel each other and this leads to stretching (a > 1). The swelling is driven by the fact that by doing so the polymer concentration goes down, thereby reducing the osmotic effect. Interpenetration of stretched chains is always unfavourable because it increases the local polymer concentration without reducing the stretching. Therefore, interpenetration is unlikely to occur in good solvents. Instead, the thickness of the brush decreases upon compression. In poor solvents, on the other hand, interpenetration actually increases the conformational entropy. Interpenetration of the chains keeps the local density high and the chains can relax back to less compressed conformations. In a theta solvent, around v = 0, a transition occurs between these two cases. It is possible to calculate Ga{h) exactly in terms of mean-field models, e.g. via the Edwards equation [1.4.1] or the discrete SCF model. However, some extra insight can be obtained by considering the diagram of fig. 1.33. In this figure, we present an example of the Helmholtz energies F e l and F o s m per grafted chain, as a function of the deformation coefficient a = L/r 0 , for JV = 1000 , d = 0.01 and v = 1 (good solvent). The curve for F e l is given by [ 1.11.11 ]; note that F e l = 0 for a = 1. An expression for F o s was not yet given here, but this follows from the Helmholtz energy density [1.2.2] in a straightforward manner, namely by multiplying f[cp) by the volume available per
INTERACTION BETWEEN POLYMER LAYERS
1.67
Figure 1.33. Helmholtz energy F per chain (in units of fcT / fi ) in a brush (upper solid curve) as a function of the dimensionless chain extension a = L/r0 for a chain of 1000 segments in a good solvent {v = 1), at a = 0.01. The constituting contributions, due to deformation ('el') and mixing with solvent Cos'), are shown as dotted and dashed curves, respectively. The equilibrium extension a = 3.89 occurs where F has its minimum. A curve given by AF = F - Fmin is drawn in the range a < a e q ; this curve represents the interaction between two brushes compressed between surfaces in a good solvent.
chain, which is JVf3 /
1 1 j V ( y - l ) + -JVu
fcT
*
2
[1.11.36]
6
The curve labeled 'os' has been drawn according to this expression, but without the constant term N[% -1)
as this does not contribute to either a
equilibrium brush extension a. which is a
=L
or G a (h). The
/r 0 is found at the minimum of F = F e l + F ° s ,
= 3.89 in this example. From our approximate expression for a good
solvent [1.11.19], obtained by omitting the second term of [1.11.17], we find a
=3.75 ; clearly, our simple expression works adequately. Note that the slopes of
pel _ pos
ancj
p represent the forces that we used in section 1.1 le {ill) to find a
.
Although fig. 1.33, as discussed so far, applies to a single chain in a brush, it also gives information on the interaction between two brushes! As long as interpenetration can be neglected, which is the case in a good solvent, the response of two surfaces being pushed together is twice that of a single brush compressed by an inert wall. The repulsion in the latter case ( a chains per unit area) is oAF, where AF is equal to F-Fmin
in fig. 1.33. The curve labeled AF in fig. 1.33 therefore directly gives the
Gibbs (Helmholtz) energy of interaction as a function of the degree of compression a a
^ eq
=
^ / ^ e q '^ o r
com
P r e s s i ° n by an inert surface) or a I a
=2h/2L
(for two
brushes). For finding G a (2h) for two brushes, we have to multiply AF with
la
(= 0.02 in the example of fig. 1.33). The corresponding plate separation is found as 2h = 2 L a / a e q . It is clear from fig. 1.33 that Ga(h) h =L
is high (of order ikT/t2
for
, in this case) and it increases steeply with decreasing h. In [1.11.40,44] we
will find explicit analytical approximations for G a (h). The interaction curve AF = G a / 2d in fig. 1.33 is on a linear scale. Figure 1.34 gives G a = 26AF (upper solid curve) with a logarithmic scale for G a , and compares it
1.68
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.34. The dimensionless Gibbs energy of interaction Ga = Ga (h) 12 / kT as a function of the scaled distance H = 1h 12L between the surfaces for N = 1000, a good solvent (v = 1), and a - 0.01 on a semi-logarithmic scale. The dashed curve is the numerical SCF result; the abscissa axis was scaled with Le_ = 119(, which is the theoretical result of [1.11.19]. The top solid curve is AF from fig. 1.33 multiplied by 26 ; the stretching parameter a of fig. 1.33 was converted into the scaled plate separation using H = a/aeq , with a e q = 3.89 . The curve with label 'mean field' is [ 1.11.40] and the curve with label 'scaling' is [1.11.44] multiplied with an (arbitrary) factor of 2.
with the result of full numerical SCF calculations (dashed curve). For convenience, we introduced a scaled particle separation H, defined as H = h / L , with L given by [1.1.19] (corresponding to L =l\9C ~lOa« in this case). The two curves labeled 'mean field' and 'scaling' will be discussed below. We see that the result of the box model matches rather well with numerically exact results. The box model gives (obviously) no interactions for H above unity, whereas in the numerical SCF theory there is a (small) repulsion for distances h > L . In the full SCF model there is a tail on each profile (compare fig. 1.30a,b), which gives rise to some repulsion when these tails start to overlap. From the comparison we see that the box model slightly overestimates the interaction energy, but the shape of the interaction curves is remarkably similar. This shows that the box model gives qualitatively correct predictions for the interaction curves. In the following subsection we consider analytical approximations for the interaction curves. (it) Analytical interaction curves. To get a simple expression for the disjoining pressure, we use nel =ajel, with Jel =-3a/r 0 (in units kT/ft) according to the approximate expression [1.11.9], J7OS ~ v
[1.11.37]
where fl{h) is the disjoining pressure. The osmotic contribution is repulsive because the segment concentration increases and the elastic term is attractive since the stretching decreases. The osmotic term depends only on
INTERACTION BETWEEN POLYMER LAYERS
(p(h)
=-'^_—
1.69
[1.11.38]
For a good solvent, L /£ = N(va/6)l/3 and
^Mj^fV^l kT
-H]
[H2
{2 )
[1.11.39,
\
The Gibbs energy of interaction is found from [1.1.3] (with the boundary h = °= replaced by h = L ). The result is 6 (h) =
a
(ir)
JVd5/3
[f + H 2 - 3 j
[1.11.40]
The latter result is plotted in fig. 1.34 with the label "mean field.' We see that [1.11.40] slightly underestimates the interaction energy as compared with the numerical analysis with the box model (label 2CTAF in fig. 1.34), of which [1.11.40] is an approximation. Interestingly, in this case the approximate [1.11.40] is closer to the numerical SCF prediction than the numerical solution of the box model. Anyhow, the simple box model, which neglects the interpenetration, rather accurately describes the full numerical SCF data. A similar exercise can be done for a theta solvent, where J7OS =cp 3 /3 (in units kT/e3). One therefore expects an H~3 dependence, according to [1.11.38]. With Leq/lV = (cT/3)1/2(see [1.11.20]), we obtain
which upon integration gives Ga(h) = IV<72p?2- + H 2 - 3 l
[1.11.42]
The disjoining pressure can also be calculated from [1.11.37] using the scaling expressions [1.11.25,28], instead of their mean-field counterparts [1.11.9, 1.2.4]. The result is ^-o»«f-L-H"«l kT
LH9/4
[1.11.43]
J
with the Integrated form G a (h)~ N6n/6\^(H-5'4
- l ) + | ( H 7 / 4 -1)1
[1.11.44]
INTERACTION BETWEEN POLYMER LAYERS
1.70
In fig. 1.34 we also plotted Ga (h) according to [1.11.44] (curve labeled 'scaling'). In the scaling approach the numerical prefactor is unknown. We multiplied [1.11.44] by 2 to bring the interaction energy to the same level as the mean-field prediction [1.11.40] at the maximum value of the interaction energy plotted (i.e. at Ga (h) = 15 ). Comparison of [1.11.44] with [1.11.40] shows that at weak compression the scaling approach gives slightly weaker repulsion than the mean-field model, whereas at strong compression
the opposite
is the case. However, the differences
are small.
Experimentally it is probably difficult to distinguish between the two. It is also interesting to note that the scaling approach gives a good match with the numerical SCF data. The repulsion due to dense brushes is much stronger than that exerted by dilute chains, mainly because the grafting density a can be so much higher. As seen from [1.11.39] the mean-field pressure scales as cr 4/3 , which implies that if a is increased by a factor of 10 the distance between grafting points is reduced by a factor Vfo and the pressure increases by a factor of about 20. As a rule, pressures measured between brushes can be very high indeed: brushes in good solvents are excellent stabilizers.
Figure 1.35. Numerical SCF results for the scaled Gibbs interaction energy as a function of the normalized slit separation H-2h/2L , where L is given by [1.11.19]. In diagram (a) the Gibbs energy of interaction is normalized by the chain length and results are shown for three degrees of polymerization. The grafting density is a - 0.01. In panel (b) the Gibbs energy of interaction is normalized by d 5 / 3 and curves are given for N = 500 and three grafting densities. (iii) Numerical
examples.
According to [1.11.40], G a should be proportional to
JVcr5/3 . In fig. 1.35 we check these dependencies. In fig. 1.35a we have varied the chain length at a fixed grafting density, a = 0 . 0 1 . We normalized the Gibbs energy of interaction by the length of the chains. There is excellent overlap of the interaction curves for small values of the normalized slit width H. For very weak overlap, i.e. near H = 1, differences are noticeable. The shorter the chain length, the more important the tail on top of the parabolic profile is, and the deviations at weak overlap are due to this diffuse periphery of the brush. In fig. 1.35b the interaction curves are presented for a fixed chain length of JV = 500 and for various grafting densities. Here, we normalized G
by d 5 / 3 . Again,
INTERACTION BETWEEN POLYMER LAYERS
1.71
there is quite satisfactory overlap of the interaction curves at relatively high compression (small H). The differences near H = 1 are again attributed to the fluctuations around the most likely trajectories. For low grafting densities these deviations are relatively more important. For very large overlap, i.e. for small values of H, we see an upward deviation for the highest grafting densities. Equation [1.11.40], based only upon the second virial coefficient, predicts a scaling Ga °= H~l in this region. However, when at strong overlap the segment concentration becomes very high in the brush we can no longer neglect the third virial coefficient. As a result, the interaction curves are more repulsive than the Ga °= H"1 curve. This is most noticeable for high grafting densities. In conclusion, the box model provides excellent guidelines to analyze the numerical SCF data; the main trends are well described and the deviations are easily interpreted. (iv) Experimental example. An experimental example for interaction between brushed surfaces is presented in fig. 1.36 where scaled interaction curves, measured with the Surface Force Apparatus, are given for brushes prepared by adsorbing diblock copolymers consisting of an insoluble polyvinylpyridine block (the 'anchor' block) and a soluble polystyrene block (the 'buoy' block) on a mica surface". The data apply to several block copolymer samples with varying lengths JVA of the anchor block and JVB of the buoy block. The value of Ga was normalized by JVo~11/6, according to [1.11.44], and h by L , according to [1.11.19], using a model to relate N (~ NB), 6 , and L to JVA and JVB . Figure 1.36 shows that all data fall nicely on a mastercurve. The scatter at weak overlap is due to limitations in measuring weak forces. The mastercurve should be described by the H-dependent part of [1.11.44]; as shown by the solid curve in fig. 1.36 this is approximately the case. The agreement could be improved by scaling up the experimental H [i.e. scaling down L eq ) by some 10%; we have not done that. We note that the mean-field expression Ga ~ 2/H + H"2 - 3 would work nearly equally well, as shown by the dashed curve. Hence, one cannot experimentally discriminate between the two models. Figure 1.36. Scaled Gibbs energy of interaction Ga(h) as a function H for two mica surfaces carrying brushes of polystyrene in toluene, modified from Patel and Tirrell11. The solid curve is drawn according to the Hdependent part of [1.11.44] (scaling), the dashed one according to [1.11.40] (mean field) and scaled such as to match at H = 0.1.
11
S. Patel, M. Tirrell, Colloids Surfaces 31 (1988) 157.
1.72
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.37. (a) Gibbs energy of interaction between two plates bearing a brush with nonadsorbing segments (JV = 200 ) for 0 = 10 (a = 0.05). Curves are given for three solvency conditions: u = 0, - 0.02, and -0.04 ( x = 0.5, 0.51, and 0.52) on a linear scale. The theta solvent still gives repulsion (apart from a shallow minimum which is hardly visible on this scale), but for v = -0.02 an attractive minimum appears which deepens (quadratically) as v decreases further. See also Fleer and Scheutjens11. Panel (b) gives the theta curve of (a) on a semi-logarithmic scale, and compares it with the prediction of [1.11.421 (dashed curve), which is the box model without interpenetration.
[v] Poor solvent. Two brushes in poor solvents (x > 0.5) attract each other. The reason is two-fold: the solvent film at the edge of the brush can be removed and the chains will interpenetrate when they touch, due to the fact that the compression of the chains can be relaxed. This increases the chain entropy which is favourable. Upon further decrease of the gap width 2h, the Helmholtz energy can be lowered until a homogeneous polymer phase with a density of order |u| is obtained, which happens at distances of order H = 0.5 or h ~ 0<73|u|, as follows from [1.11.22]. The attraction is expected to start for v just below zero (depending somewhat on o"), which implies that, from a practical point of view, the stabilizing effect of a brush is lost as soon as the theta temperature is crossed. In fig. 1.37a we give an example of the interaction between two brushes as v is varied around zero. Three curves are shown in fig. 1.37a. For v = 0 (# = 0.5) the interaction is nearly completely repulsive, albeit relatively soft. Close inspection shows that before the repulsion sets in a very small attraction occurs. For v = -0.02 (£ = 0.51), this minimum is already much more pronounced and the minimum deepens rapidly as v decreases, as the curve for v = -0.04 (% = 0.52 ) shows. The repulsion at small h is due to crowding effects. For the theta solvent we presented the analytical prediction in [1.11.42]. We show the comparison with the numerical SCF result for % = °-5 in fig. 1.37b. As in the good solvent case we see once again that the numerical SCF predictions already show repulsion for 1h > 2L , due to the tail in the profile in the complete SCF profile, caused by fluctuations of chain conformations not included in the box model. For 11
G.J. Fleer, J.M.H.M. Scheutjens, Colloids Surfaces 51 (1990) 281.
INTERACTION BETWEEN POLYMER LAYERS
1.73
H < 1 this box model systematically overestimates the interaction energy. This is because the model does not allow for interpenetration of the chains, which is a poor approximation for a theta solvent. Indeed, a brush in a theta solvent gives a softer repulsion than predicted by the box model. As H = 1 corresponds to 2L ~ 60 , we note that the shallow minimum in the interaction curve shown in fig. 1.37a is outside the range of fig 1.37b. 1.1 lh Soft depletion In sees. 1.8-9 we discussed the effect of free polymer on a dispersion of hardspheres. Quite generally, we found that polymer induces a weak attractive interaction between the colloidal particles, the strength and range of which depend on the concentration of the polymer and the coil/sphere size ratio a la. In figs. 1.34-36 we saw that particles carrying tethered polymer repel each other. We now consider briefly the mixed case of colloidal particles with grafted polymer on their surfaces, dispersed in a solution of free polymer molecules. We suppose that the grafted chains (length JV, grafting density a) and the free polymer (chain length P) are chemically identical. In fig. 1.38 we present a measure for the thickness of a polymer brush immersed in a polymer solution in which P = N = 500 . The thickness is a strong function of both the concentration of the free chains and the grafting density of the tethered chains. Above we have seen that, upon overlap of stretched chains, they relax to less stretched conformations. We can apply the same argument to explain why the brush thickness decreases upon an increase in the concentration of the free chains. The free chains try to penetrate the brush and therefore increase the chain crowding. The response is a collapse of the swollen brush. The symbols in fig. 1.38 indicate the point where the polymer concentration in the (initial) brush (without outside polymer) equals the polymer concentration in the polymer solution. At those points, the osmotic pressure in the polymer solution equals that of the (initial) brush. The fact that the symbols are in the steepest part of the drop in brush thickness strengthens the argument. At very low grafting density, the (initial) tethered polymer layer has no stretched chains and free chains do not alter the thickness. Up to 9 = 10 a brush immersed in the melt,
Figure 1.38. The average position (first moment) of the free end of a brush with chain length JV = 500 and B = dN as indicated, as a function of the polymer volume fraction
1.74
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.39. A set of interaction curves for a brush composed of JV = 500 segments with grafting density a = 0.01 in a good solvent (% = 0) with free identical chains in solution with P = 2000 (solid curves) and P = 125 (dashed curve) for various concentrations (pb of the free polymer. The open symbols are for
INTERACTION BETWEEN POLYMER LAYERS
1.75 Figure 1.40. Depth of the interaction Gibbs energy minimum (in units fcT//2 | for two surfaces immersed in a solution of nonadsorbing polymer (P = 300, ^ = 0.1, X = 0 ) as a function of the amount 9 = Na of grafted polymer [N = 50) on the surface. For high grafting density a linear decay is found, which extrapolates to approximately - 7 10~ 3 , which is the zero grafting density value, indicated by the dot.
by the dot at 0=0. For intermediate grafting densities there is a weakened depletion effect due to the presence of grafted chains. Indeed, this so-called 'soft depletion' effect is seen in fig. 1.40. Already very few grafted chains destroy most of the depletion force because the free chains interpenetrate with the dilute brush. As the brush becomes thicker and denser, the depletion effect passes through a minimum (so that G™'*1, which is negative, goes through a maximum). Upon further increase of a, the free chains are repelled from the brush to an increasing extent, and the attractive well deepens until it becomes comparable with the depletion attraction at a hard and inert wall. For the free chains to enter the brush, a necessary condition is aS2 < 1, where the depletion thickness S is given by [1.8.9] and varies between the value 2ag/%/^ in dilute solutions and £/[3vcpb)l^2in semidilute solutions, see fig. 1.17. Short chains thus have a much better chance to be taken up in a brush than long ones; this effect has been analysed in some detail11. l.lli Concluding remarks The literature on polymer brushes has grown tremendously in the last few years. It is impossible to cover all these developments. The fact that the chains are tethered by their ends makes computer simulations relatively easy. Such computer simulations appear to provide essentially the same picture as that obtained by numerical SCF. Furthermore, we have seen that the analytical route via the strong stretching approximation is a very accurate method, which is very helpful in analyzing brushes. The crudest, but nevertheless useful, analytical approach is the box model. The amazing aspect of this primitive model is that all results obtained scale in the same way as those following from more refined models. Only when the end-point distribution is not strongly peaked at the periphery of the brush is the box model less reliable. It is interesting to note that in recent years polyelectrolyte brushes have also received ample attention. In favourable cases, the electrostatic potential dominates the self-consistent potential. Then, the analogy with the Huygens oscillator directly leads
11
CM. Wijmans, E.B. Zhulina, and G.J. Fleer, Macromolecules 27 (1994) 3238-3248.
1.76
INTERACTION BETWEEN POLYMER LAYERS
to a parabolic electrostatic potential profile in a polyelectrolyte brush; accurate analytical solutions of the polyelectrolyte brush closely follow the numerical SCF predictions. The interaction between polyelectrolyte brushes has received some attention as well. Other interesting developments, which we did not deal with in this chapter, are brushes on curved surfaces (relevant for polymeric micelles and vesicles), brushes with nematic ordering, polydispersity effects, and the uptake of proteins or small micelles in brushes. This is a very active field of research where many new theoretical results and experimental applications are to be expected. 1.12 Non-equilibrium aspects 1.12a Introduction In sees. 1.3-7 we considered the effect of adsorbing polymers on interactions between two surfaces. We took the viewpoint that full equilibrium (a constant chemical potential) between the slit and an external reservoir could be maintained for all molecules in the system, i.e. for the solvent and the polymer. We found that such relaxed adsorbed layers give rise to a weak repulsion at weak overlap, and to a strong attraction related to bridge formation at smaller gap widths. In sec. 1.11 we discussed the case of end-attached polymers between two surfaces, where the conserved property was the number of polymer chains between the surfaces, rather than their chemical potential; only the solvent was allowed to equilibrate with the bulk solution. In this case, we found that polymers in a good solvent give rise to a strong and monotonic repulsion between the surfaces. In the present section, we consider the question as to what extent these considerations can explain the behaviour of various colloid/polymer mixtures as they occur in practical situations. We shall see that adsorbed polymer layers often (but not always) relax slowly and we qualitatively discuss the factors that affect relaxation rates. We then discuss what one may expect when the amount of adsorbed polymer in a gap between two approaching surfaces is supposed to remain constant (an example of restricted equilibrium). We examine experiments in which the distance between two surfaces is varied, for example by approach/retraction cycles, and attempt to draw conclusions on the extent of relaxation in such systems. Finally, we address the issue offlocculation induced by adsorbing polymer and the kinetic effects that are so crucial in this process. We illustrate our argument with a case study on shear-induced flocculation rates where many processes occurring simultaneously in a colloid/polymer mixture (bridge formation, conformational relaxation, stabilization) together determine the overall result. When judging experimental data it is important to emphasize that forces due to polymers almost never appear in their pure form, because other interactions such as electrostatic and Van der Waals interactions generally cannot be eliminated. If such interactions are also operative, it is usually assumed that the energetic contributions
INTERACTION BETWEEN POLYMER LAYERS
1.77
from various types of interaction are additive. We realize that this is a questionable assumption as adsorbed polymer layers may well perturb an electrical double layer, and dense polymer layers contribute to dispersive interactions between surfaces. Nevertheless, we accept it for the mainly qualitative discussion that follows. 1.12b Does equilibrium apply? In order for the equilibrium theory developed in the previous sections to be relevant, the relaxation of an adsorbed polymer layer under variations of the gap width has to be sufficiently rapid; the Deborah number De = r proces / tmeas ( s e e 1-2.3.1) should be much smaller than unity. In dilute solutions, polymer conformations relax at time scales determined by the Rouse time, which is the characteristic time for thermal relaxation of the end-to-end distance ; its value for a chain in water with a contour length of, say, 1000 nm is typically in the millisecond range. In concentrated solutions and melts, relaxations require snake-like movements (reptation) of the chains through the dense systems, which occur on the much slower reptation time scale21. At an interface, a number of segments is bound to the adsorbent as trains, and the mobility of these segments will depend both on the strength of the segment-substrate bond and on the mobility of the substrate itself. If the substrate is a solid and the segment-solid bonds are strong, a polymer molecule may be very effectively immobilized. There is plenty of experimental evidence to illustrate this point. Reduced mobility of polymer segments shows up in NMR experiments as a strongly enhanced relaxation rate of the magnetization, and a virtual disappearance of polymer signals from the usual narrow-line (high resolution) spectra that are obtained for rapidly moving molecules. In several experiments on the kinetics of exchange processes between adsorbed and free polymer molecules, slow surface processes have been detected with characteristic times varying from seconds to days. More often than not, the surface pressure generated by water-soluble polymers adsorbed at water-air interfaces shows substantial hysteresis during compression/dilation cycles (see also sec. III.3.8a). One may therefore wonder, particularly in the case of solid substrates, whether equilibrated (annealed) adsorbed polymer layers exist at all. Yet, we can mention several examples. The surface pressure of polyethyleneoxide (PEO) adsorbed at a water/air interface remains constant at about 10 mN/m under slow surface compressions, showing that this system can equilibrate properly. One would, therefore, expect that PEO adsorbed on the surfaces of a free water film would also equilibrate, so that if the weak repulsion due to tails is overcome, attraction sets in and the film breaks. Indeed, this seems to happen because stable films of PEO solutions have never been reported. Many other water-soluble polymers behave likewise; they are unable to stabilize free water films. Exceptions are found when the 11
P.E. Rouse, J. Chem. Phys. 21 (1953) 1272; M. Doi, S.F. Edwards, Theory of Polymer Dynamics, ch. 4, Oxford Science (1986). 21 M. Doi, S.F. Edwards, Theory of Polymer Dynamics, ch. 6, Oxford Science (1986).
1.78
INTERACTION BETWEEN POLYMER LAYERS
polymer has a more or less pronounced copolymeric character. In order to have equilibration of polymer adsorbed at S/L interfaces, it is probably crucial that the adsorption is very weak. One reported example concerns platelets of nbutylammonium-vermiculite clay". The interparticle spacing in swollen suspensions of this clay has been studied as a function of stress applied along the swelling axis. As a result of these experiments, the relation between gap width and applied pressure is precisely known in this system, so that it can be used as a colloidal 'instrument' to measure disjoining pressures. When PEO of varying molar mass is added to vermiculite dispersions, attraction is found to set in whenever the polymer coil dimensions are equal to, or larger than, the distance between the platelet surfaces. Once contraction occurs, the (negative) disjoining pressure is found to be independent of the PEO molar mass, but proportional to the amount of polymer adsorbed on the platelets. These experiments strongly suggest that this is a case of equilibrated polymer adsorbed between two surfaces. The polymer spontaneously enters the gap between platelets, and the platelet spacing measured does not depend on sample history. Moreover, the saturated adsorbed amount turns out to be low (about 0.1 mg/m2), which indicates weak adsorption. We should note that the case of very weak adsorption (near the adsorption/desorption threshold ^ s c ) was treated only briefly in sec. 1.10 where it was shown that for small /s the GSA approach breaks down. Hence, the expressions derived in sec. 1.6 do not apply to the vermiculite/PEO system. However, if the general arguments discussed in 1.6a still apply, namely (i) that attraction is due to the fact that the concentration in the gap is non-uniform and {ii) that the net effect is attraction, this can qualitatively account for the data. 1.12c Restricted equilibrium The interaction between polymer layers that cannot, or can only partially, equilibrate is likely to be different from that between equilibrated layers, but as yet we do not know to what extent. In order to get an idea of the trends to be expected, one can numerically calculate the interaction between two surfaces with adsorbed polymer layers, assuming that the amount of polymer between the surfaces is strictly constant; only solvent can enter and leave the gap. This may be called 'restricted equilibrium', with a quenched layer. The situation is somewhat analogous to the case of constant surface charge for interacting double layers, as opposed to that of equilibrated surfaces with a constant (electro)chemical potential. The structure of the layers (i.e. the profile) may of course adjust itself, and the starting situation is that of two surfaces at large separation, each carrying an equilibrated, adsorbed layer. This situation is difficult to treat by the analytical methods discussed in 1.3-7, but the numerical SCF scheme can cope quite well with it. 11 J. Swenson, M.V. Smalley, and H.L.M. Hatharasinghe, Phys. Rev. Lett. 81 (1998) 5840. M.V. Smalley, H.L.M. Hatharasinghe, I. Osborne, J. Swenson, and S.M. King, Langmuir 17 (2001)3800.
INTERACTION BETWEEN POLYMER LAYERS
1.79
A survey of this case was carried out by Scheutjens and Fleer1'. The quintessence of their findings can be summarized by the diagrams in figs. 1.40. In fig. 1.40 the Gibbs energy of interaction Ga(h) is plotted both for full equilibrium and for restricted equilibrium, for a good and a theta solvent and JV=1000. For the full equilibrium the bulk concentration was fixed to a value
Figure 1.41. Numerical SCF data for interaction curves for full equilibrium (solid curves) and restricted equilibrium (dashed curves) in a good solvent (a, b) and a theta solvent (c, d). Panels band d are zoomed-in versions with respect to the Gibbs interaction energy (not with respect to the slit separation). In all cases, JV = 1000. In full equilibrium, the polymer concentration in the bulk is q>b = 10~ 3 . In restricted equilibrium, the amount of polymer at 1h = 100 is fixed at the value given by the equilibrium case for that separation; 6 = 1.6191 (both walls) for the good solvent case and 6 = 3.9222 for the theta solvent.
II
J.M.H.M. Scheutjens, G.J. Fleer, Macromolecules I I I (1986)504).
18 (1985) 1882; J. Colloid Interface
Sci.
1.80
INTERACTION BETWEEN POLYMER LAYERS
Upon compression, the (total) amount was fixed. The relevant interaction energy is now Ga{h) = 2(i2 so (h)-i2 sc V)), where the semi-grand potential is the sum of the interfacial energy and the chemical work of the confined chains. The quite general result is that the restricted equilibrium curves are above the full equilibrium curves. In other words, any type of restricted equilibrium is always more repulsive than full equilibrium. This is necessarily so because the relevant thermodynamic potential can relax to lower values when there are more degrees of freedom. By the same token, interaction between diffuse double layers at constant charge gives rise to stronger repulsion than at constant potential (see sec. IV.3.4b). In a good solvent (fig. 1.41a,b), the restricted equilibrium curve is (in this case) fully repulsive, whereas the main trend for complete equilibrium was attraction. The repulsion already sets in at the initial separation 2h = 100, as can be seen in the enlargement of fig. 1.41b. The features of full equilibrium as discussed in sec. 1.6 are completely masked. For a theta solvent, the restricted equilibrium follows the complete equilibrium much better. A very characteristic local minimum in the interaction curve is now found at 2h ~\7. For shorter distances, the extra polymers that are confined and cannot escape lead to repulsion. A more systematic study of this type of restricted equilibrium is found in the work of Scheutjens and Fleer, cited above. These authors showed that the depth of the minimum is a function of the amount of polymer fixed in the slit. The more polymer that is restricted between the plates, the greater repulsion that builds up and the shallower the minimum is. At very high restricted amounts, the minimum even disappears and the interaction curves become completely repulsive. Semi-quantitative agreement with these trends was found in surface force measurements with the surface force apparatus (SFA) for polystyrene adsorbed from cyclopentane on mica11, which was very close to theta conditions. In fig. 1.42, we reproduce these data. In a surface force apparatus, it is very likely that the amount of polymer is restricted because of the large diffusion lengths involved in the setup. In fig. 1.42 the amount of polymer increases from left to right. For the lowest amount of polymer the characteristic minimum is found, and the experimental curve resembles in
Figure 1.42. A few interaction curves for homodisperse polystyrene (M = 2000 K), adsorbed from cyclopentane on mica, under theta conditions. From left to right the adsorbed amount increases as found with the surface force apparatus adapted from Y. Almog, J. Klein".
11
Y. Almog, J. Klein, J. Colloid Interface Sci. 106 (1985) 548.
INTERACTION BETWEEN POLYMER LAYERS
1.81
many aspects the restricted equilibrium curve shown in fig. 1.42d. The correlation between the SFA results and the numerical SCF predictions support the idea that polymer layers in confined spaces may not always be in full equilibrium, because of the difficulty of diffusion out of a very narrow slit with macroscopic lateral dimensions. 1.12d Relaxation phenomena in polymer-induced surf ace forces In a qualitative sense, the difference between quenched and equilibrated (relaxed) adsorbed layers shows up only at large 8 (saturated layers). In that case, relaxed layers have a non-monotonic interaction curve with a weak repulsive maximum and strong attraction at small plate distance (sec. 1.7), whereas quenched, adsorbed layers develop a strong monotonic repulsion (sec. 1.12c). The difference is the strong osmotic effect of the confined polymer, which is not compensated by the subtler entropic effects associated with the evening out of the density profile in the gap due to bridge formation. Hence, when two surfaces with saturated polymer layers are pushed together sufficiently strongly, such that the loop regions can overlap, and if relaxation occurs at experimental time scales, one may expect to observe repulsion at short times, which changes into attraction at longer times. We discuss two examples of such relaxation effects. (t) Indications from AFM experiments. The first example is taken from a study by Giesbers et al. on surface forces between a silica sphere and plate in the presence of PEO. In fig. 1.43 we show the force as a function of distance in a solution of NaCl at pH 8, containing 100 rag/l PEO of molar mass 246 kg/mol. The force is normalized with respect to the radius of the sphere, so that it is given in units of interfacial Gibbs energy (mN/m= mJ/m 2 ). The left diagram (a) refers to the case where the sphere and plate approach each other (ft decreases), the one to the right (b) is for retraction. Three
Figure 1.43 Force-distance curves, measured by AFM between a silica sphere and plate in the presence of polyethyleneoxide (PEO), at pH 8 and three concentrations of NaCl; (a) approach curves and (b) retraction curves. Molar mass of PEO is 246 kg/mol. (Redrawn from Giesbers et al. 1 ').
11 M. Giesbers, J.M. Kleijn, G.J. Fleer, and M.A. Cohen Stuart, Colloids Surfaces A: Physicochem. Eng. Aspects 142 (1998) 343.
1.82
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.44. Approach (dotted curves) and retraction (solid curves), measured with AFM, between a silica sphere and plate in the presence of PEO in aqueous solution. Raw data are given as cantilever deflection A (a linear function of the force) versus piezo position AT. (arbitrary zero) (which is related to the separation 2h). Molar mass of PEO is 246 kg/mol, pH = 6.5, salt concentration 10~3 M NaCl. The approach and retraction curves for a low load force are almost identical, but for a high load force a deep minimum appears due to bridging attraction. Reference as in previous figure.
curves are given in each diagram for different salt concentrations. Clearly, the two sets of curves are rather different; the approach curves are all purely repulsive, whereas the only repulsive retraction curve is for 0.001 M salt. The other two show a clear attraction in the form of a so-called 'pull-off force. The salt dependence in fig. 1.43a must be due to double layer repulsion; at 1 M salt, the double layer force, indeed, has a very small range. Under these circumstances, the retraction curve shows a fairly strong attraction; retraction to about 100 nm is needed before the force comes back to zero. It seems likely that during the brief contact time under compressive load (typically about 0.2 s), the polymer layer was able to form bridges, thus giving attraction. The importance of the load force is shown by fig. 1.44 where the force-distance curves are reported for the same PEO-silica system at pH 6.5 and 1 mM salt. In this experiment, the two sets of curves refer to different maximum compressive load forces. When the maximum load force is small (the contact time is then also shorter), the approach and retraction curves are both repulsive and, moreover, very similar. However, at a larger load force, an attraction appears. This agrees with the idea that in that case the polymer layers relax under compression and then form bridges. Finally, the strength of the attractive pull-off force increases linearly with the amount of polymer between the plates. This is shown in fig. 1.45, where the strength of the pull-off force (the depth of the minimum in the force distance curve) is plotted
Figure 1.45. Normalized pull-off force / o f f / a between a silica sphere and plate due to PEO adsorbed on silica, as a function of the total adsorbed mass F of a saturated layer at pH = 4 and 1CT3 M NaCl. The molar mass of the PEO for each data point is indicated (in kg/mol). Same source.
INTERACTION BETWEEN POLYMER LAYERS
1.83
against the adsorbed mass of the saturated layer (which increases with the molar mass of the polymer). This plot shows that for low adsorbed amounts (below about 0.55 mg/m2) there Is no attraction, probably because these amounts can be mostly accommodated in the train layer so that hardly any loops are formed. The linear behaviour would seem to imply that all extra polymer goes into loops and then forms bridges, proportional to the amount in loops. Hi) Dispersions compressed in a centrifuge. Our next example concerns silica sols (particle radius 15 nm) with adsorbed polyvinyl pyrrolldone (PVP)11. These sols are stable in dilute dispersions, but when compressed by centrifugation, interparticle bonds are formed. Following the compression by centrifugation, the samples were brought into an aqueous medium and tested for their redispersibillty. The results are summarized in the pressure/coverage diagram presented in fig. 1.46. At low adsorbed amounts (0-0.25 mg/m2), particles were always unstable and gels were formed at any centrifugal pressure; these gels could not be re-dispersed in aqueous solution. At intermediate adsorbed amounts (0.25-0.5 mg/m2), small aggregates ('multiplets') were formed without any applied pressure, and at adsorbed amounts above 0.5 mg/m2 particles were stable (singlets) as long as no pressure was applied. By centrifugal pressure, however, the stable singlet particles could be converted Into small aggregates (multiplets); the higher the coverage 9, the higher the required pressure. Likewise, the multiplets could sometimes be converted into gels, particularly for 8 not too much above 0.25 mg/m2. Both the results of the direct surface force measurements and the centrifugation experiments seem to agree with the idea that during prolonged contact between two adsorbed layers (and assisted by applied pressure) some adsorbed polymer will sooner or later escape, allowing the remaining polymer to form bridges that lead to permanent attraction. The mobility of the polymer in the adsorbed state determines the time scale at which this happens. 1.12e Case study: kinetics qfjlocculation (i) Time scales. Polymers are often employed for the purpose of colloidJlocculation,
Figure 1.46. Aggregation states of silica spheres (80 nm diameter) with adsorbed polyvinyl pyrrolidone (PVP) of M w = 1 0 kg/mol as a function of applied centrifugal pressure p c and adsorbed mass F of PVP. Three regions are shown: singlet particles (diamonds), small aggregates (squares), large flocs/gels (triangles).
11
Ph. Ilekti, thesis, Univ. Paris 6, Protection des Surfaces de Silice avec un Polymere adsorbe (2000).
1.84
INTERACTION BETWEEN POLYMER LAYERS
i.e. the formation of aggregates of colloidal particles with the specific purpose of removing them from a dispersion. Obviously, this is possible because of the attraction that polymers can exert between colloidal particles. However, the fact that saturated, quenched layers tend to repel (unless rapid relaxation and sufficient compression occur) implies that it is important to control the adsorbed amount 0 because repulsion sets in and the flocculation stops when G becomes too high. Another factor is also important. At low 0, the adsorbed layer in its equilibrium configuration may be rather thin. If the particles to be flocculated cannot approach each other very closely (e.g. because of electrostatic repulsion), the adsorbed polymers cannot make contact with the other particle and flocculation does not occur. A freshly deposited polymer molecule may be much more extended, though, than an equilibrated one, so that flocculation occurs when the surfaces are brought into proximity before the reconfiguration has taken place. Altogether, we thus have to consider various time scales in a flocculation process: (i) the colloid/colloid collision time, which determines the overall rate of flocculation, [ii) the colloid/polymer collision time, which determines the rate at which the adsorbed amount increases and particles become 'reactive' or 'passive' during a particle encounter, and [Hi) the polymer reconformation time, which determines how long freshly adsorbed polymer molecules remain capable of forming bridges. An early example of these time scales was reported with silver iodide sols flocculated with polyvinylalcohol (PVA)11. With increasing concentration of flocculant the flocculation rate first increased and then decreased again; at a high concentration of PVA, the sol became stable. However, in the latter case there was some initial flocculation in the first five seconds after mixing. Apparently, in these experiments it took about five seconds to build up a stabilizing layer. (ii) Orthokinetic Jlocculation experiment. A case study in which many of these effects are readily seen has been carried out by Adachi et al.2). The chosen colloidal dispersion was a dilute, negatively charged polystyrene latex, which was flocculated with PEO (M = 5000 kg/mol), under conditions where the electrical double layer was entirely suppressed (1.17 M KC1). In order to distinguish the effect of the added polymer, the kinetics of salt-induced flocculation was first carefully characterized. The study focused on shear-induced (orthokinetic) flocculation where the effects of polymer show up most clearly, and the shear conditions were carefully standardized for appropriate comparison. Salt-induced coagulation rate. In the absence of the polymer, the latex coagulates rapidly because of the high KG concentration. The rate of coagulation follows the wellknown second order rate equation
11
G.J. Fleer, J. Lyklema, J. Colloid Interface Set 55 (1976) 228. Y. Adachi, M.A. Cohen Stuart, and R. Fokkink, J. Colloid Interface Set 165 (1994) 310; Y. Adachi, T. Wada, J. Colloid Interface Sci. 229 (2000) 148. 21
INTERACTION BETWEEN POLYMER LAYERS
1.85
^ = -kn 2 dt
[1.12.1]
where n is the (number) concentration of particles, and the rapid coagulation rate constant k is a function of shear rate, particle size and a capture efficiency factor aT , which accounts for the Van der Waals attraction and hydrodynamic interaction between two particles. This capture efficiency factor was calculated by Van de Ven and Mason1' by considering trajectories of a particle approaching another particle, and is given by aT=\
—T
[1.12.2]
In this equation, A is the Hamaker constant, r\ the viscosity, y the shear rate, and a0 the bare particle radius. Adachi et al. used an effective shear rate, which could be accurately controlled by a standardized tumbling method, and for which [1.12.2] remains valid. As follows from this equation, larger particles have smaller capture efficiencies, mainly because there is a severe hydrodynamic drag upon close approach of the particle surfaces. Since the volume fraction
= -k'n
[1.12.3]
and, hence, described by In—
=-fc>J
[1.12.4]
When the total number concentration n = £rij of particles (singlets + doublets + ) is plotted logarithmically against flocculation time, straight lines with a slope proportional to -fc'
"TPC a r e obtained. An example of such a plot is given in
Figure 1.47. Salt-induced orthokinetic coagulation of polystyrene latex in 1.17 M KC1, plotted as particle number concentration n(t) as a function of a scaled time t, for two particle radii a 0 , as indicated. Particle number concentration n0 = 1.7 10 13 m~ 3 . (Redrawn from Adachi et al., loc clt.)
11
T.G.M. van de Ven, S.G. Mason, Colloid Polym. Set 255 (1977) 794.
1.86
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.48. Rate s of rapid orthokinetic coagulation of PS latex with 1.17 M KC1, as a function of the initial particle number concentration riQ (reference as in previous figure).
Figure 1.49. Reduced rate s of rapid orthokinetic coagulation of PS latex with 1.17 M KC1 as a function of the particle diameter a 0 (reference as in previous figure).
As can be seen from [1.12.2], the capture efficiency scales with the bare particle radius as a T ~ a^ 0 5 4 . Moreover,
INTERACTION BETWEEN POLYMER LAYERS
1.87
Figure 1.50. Polymer-induced orthokinetic flocculation plots, presented as lnn(t) as a function of time ( (arbitrary units), for various polymer concentrations c , as indicated. The straight line represents the results for salt-induced coagulation (see fig. 1.46). Polystyrene latex particles (678 nm radius), n 0 = 7.7 • 10 13 m~ 3 , salt concentration 10~4 M KC1. Flocculant: PRO (M = 5000 kg/mol)11.
increases again (fig. 1.50b). We note that several of these effects were also seen in the experiments by Fleer and Lyklema, referred to in the introduction to this section. Let us first consider the initial slope in fig. 1.50. This is higher than that for saltinduced coagulation, indicating a higher collision rate. This cannot be due to a higher sticking probability, because all electrostatic repulsion was eliminated so that the sticking probability is equal to unity at the salt concentration used. Moreover, n 0 was exactly the same. The only explanation is that the effective collision radius of the particles is larger during the polymer-induced flocculation experiment. Apparently the polymer layer on the particle contributes substantially to its size by adding a layer of effective thickness <5e . If we now suppose that polymer layers can form bridges before Van der Waals forces and hydrodynamic effects come into play, the capture efficiency no longer follows [1.12.2]. Instead, the shear-induced flocculation rate is simply proportional to a 3 = (a0 + <5e )3 , whereas that of the bare particles was proportional to C2Q-46 . Hence, by taking the ratio of the two rates we can calculate where the initial rate has its highest value <5e. It turns out that for high polymer concentrations (1 mg/1 and higher), Se is very nearly equal to the diameter of the PEO coil in solution (about 210 nm). This indicates that in these experiments the adsorbed layer contains (some) adsorbed coils, which have very nearly the same size as they had before adsorption; during a particle/particle collision there are always enough freshly adsorbed molecules on the surface that have not had the time to change their conformation. At lower polymer concentrations, the rate gradually decreases, which can be interpreted in terms of a smaller effective collision radius and a smaller <5e. Apparently, the time between successive polymer/colloid encounters is longer, and the particles cannot maintain the 'freshness' of their adsorbed layer. The molecules now have some time to
11
Y. Adachi, T. Wada, J. Colloid Interface Set 229 (2000) 148-154.
1.88
INTERACTION BETWEEN POLYMER LAYERS
Figure 1.51. 12 Effective thickness, Se , of adsorbed polymer during a flocculation experiment, as a function of the characteristic adsorption time r (see text). Flocculant: PEO M = 5000 kg/mol); PS latex particle (678 nm); salt concentration: 10~4 M KC1 (solid points) and 10~2 M KC1 (open points). Reference as in previous figure.
undergo conformation changes leading to a thinner layer, and in these dilute solutions the supply of fresh molecules is too slow. The data thus emphasize the crucial influence of relaxation of adsorbed polymer on the kinetics of flocculation. This is shown in more detail in fig. 1.51 where the value of 5e , as obtained from a series of flocculation experiments, is plotted as a function of the characteristic adsorption time T of the polymer. This adsorption time is found by considering the adsorption flux (number of polymer molecules hitting the particle per unit time) J : J p = 4K(Do + Dp)(ao+ap)np+(27t)l/2Y(ao+ap)3np
[1.12.5]
In [1.12.5], D and a are diffusion coefficients and radii, respectively, and the subscripts 0 and refer to bare particle and polymer coil, respectively; n is the number concentration of polymer molecules in the solution. Note that we consider here the adsorption rate, i.e. particle/coil collisions, whereas above (in the flocculation rate) we used particle/particle collisions. The first term in [1.12.5] is due to Brownian motion and the second one comes from collision due to shear. Taking simply the inverse of the polymer/particle collision rate, we obtain the average time T between two successive collisions between a particle and a polymer molecule. Since J decreases when the polymer concentration is lowered, T increases, and the effective thickness is clearly seen to decrease to very low values, signaling a flattening of the polymer adsorbing from dilute solutions. We finally turn to the leveling-off of n(t) in fig. 1.50, which is due to the appearance of repulsion between particles. As the data show, the final level changes nonmonotonically, being high at low concentration, low around 1 mg/1, and again high at 8 mg/1. Two trends are responsible for this effect: (i) the rate of flocculation itself goes up, allowing larger aggregates to form in a given time, and {ii) the adsorption rate increases, thus shortening the time needed to saturate the surfaces enough to get repulsion. In a subtle way, these two trends conspire to produce optimum flocculation at about 1 mg/1 in this experiment.
INTERACTION BETWEEN POLYMER LAYERS
1.89
1.13 Outlook The equilibrium characteristics of polymers at interfaces are rather accurately described by the Edwards equation, which may be solved routinely by numerical methods up to high accuracy. Due to the large number of identical repeating units along the chain there are many universal properties. These generic effects become clear from accurate analytical descriptions of the interfacial characteristics for various regimes and scenarios. The physics of polymer near an isolated surface is now accurately known, at least on the mean-field level. The extrapolation of these results to colloidal stability issues has its intricacies and we are still far from a complete picture. Nevertheless, there is a leading principle to judge the effect of polymers on the colloidal stability; the key is found in the ranking-number dependent concentration profiles. When all segments along the chain have more or less the same distribution, as for symmetric homopolymers adsorbing from good or even theta solvents with sufficiently high adsorption affinity, the ground-state approximation (GSA) is accurate, and attraction is predicted (caused by depletion when the polymers do not adsorb onto the surfaces, and by bridge formation in the adsorption regime). However, when there is a clear ranking-number dependence, as for segments in a tail, GSA breaks down and repulsion is found. This means that adsorbed polymer layers are repulsive at weak overlap (when tails dominate the interaction) and attractive at strong overlap (when loops can transform into bridges). For end-grafted polymer chains, GSA is replaced by the strong-stretching approximation, which is accurate for densely grafted polymer layers, so-called brushes. Polymer brushes are excellent tools to stabilize colloidal particles because the overlap of the brushed surfaces (each carrying a large number of tails) is very costly due to the compression of each chain in the brush upon approach of the two surfaces. Polymers relax with large characteristic time scales in response to external perturbation. This is one of the most important reasons to select polymers (rather than, e.g. short surfactants) to influence the colloidal stability. Typically, off-equilibrium effects should be superimposed on the equilibrium effects. Very generally, when not all degrees of freedom are relaxed, i.e. when the Deborah number is high because the experimental time scale is short with respect to the chain relaxation time, the interfacial layers are off equilibrium and interfacial forces are necessarily more repulsive than in full equilibrium. For example, in a dilute solution of small particles the time between collisions may be too short to allow bridging. Then there are many collisions necessary before a floe can be formed. In other words, the capture efficiency is low. This effect is probably more important for loop-to-bridge attraction than for depletion. Hence, kinetic aggregation studies could be useful to distinguish adsorption from the depletion mechanism, also when the adsorption is weak. Most of what we have reviewed in this chapter is inspired by numerically solving the Edwards equation or its lattice equivalent in the Scheutjens-Fleer model. In analytical
1.90
INTERACTION BETWEEN POLYMER LAYERS
approximations, such as GSA, the non-local contributions to the self-consistent potential are typically ignored. When the non-local contributions are kept in the analysis one obtains more complicated differential equations with higher order derivatives (up to 3 4 G/3z 4 ). The most important consequence is that one may anticipate non-monotonic density distributions. Indeed, numerical solutions that systematically include these higher order derivatives show these (very weak) effects around the periphery of the density profile; there are concentration oscillations around the bulk concentration. There must, therefore, also be non-monotonic, force distance effects. The magnitude of these effects are, however, typically (at least) an order of magnitude smaller than discussed in this review. In passing, we note that higher order derivatives play an essential role in describing wetting phenomena and liquid-liquid interfaces, which were not addressed in this chapter. A key assumption used throughout has to do with the range of the interactions. In our analysis, we have assumed in all cases that the interactions with the surface are strictly short-range, of the nearest neighbour type. This is reasonable if there are specific (key-lock) interactions between polymer segments and the surface, but not so when the interactions have a dispersive nature. We may consider, for example, the case that the interactions have a power-law decay, such as for Van der Waals forces where the adsorption energy decays as z~3. Such a scenario can be investigated by numerically solving the SCF equations. The main result is that the proximal part of the density profile is affected; this effect could be mimicked by modifying the short-range interaction term, using an effective ^ s • m ^ ne central part of the profile, the selfconsistent potential dominates the (power-law) surface contribution so that the selfsimilar density profile is not affected. Also, the distal part of the density profile decays, as usual, exponentially to the bulk density. The power-law decay of the surface interaction is most fundamentally reflected in the way in which the density profile decays to the bulk density at the very edge of the adsorption profile (at the periphery of the distal part of the profile). The additional contribution has a power-law behaviour with a decay length given by the (long-range) surface energy characteristic. This powerlaw tail may interfere with the oscillatory behavior due to the non-local, self-consistent potential discussed above. Another example of non-trivial long(er) range (surface) interactions is found in systems in which the polymer and the surface are charged. We then should solve the Edwards equation in combination with the Poisson equation, leading to polyelectrolyte theory. Such extensions are feasible and much is known about polyelectrolytes at interfaces. We refer to chapter 2 for more details. In addition to the universal features of interfacial polymer layers, there must also be non-universal aspects. We have assumed throughout our analysis that there are no specific effects on the monomeric length scale; only non-specific effects as expressed in the Flory-Huggins solvency parameter were taken into account. However, when there are complex side groups in a monomer, possibly with mesogenic features, it is
INTERACTION BETWEEN POLYMER LAYERS
1.91
expected that our approximations seriously break down. To treat such systems accurately, one could try to use computer simulations. A semi-atomistic analysis, using a numerical SCF scheme, remains an option as well. However, to account for, e.g. the stiffness of the chain, the structure of the side groups (chain branching), and the possibility of nematic ordering, a much larger set of interaction parameters is needed to describe the various chemical entities and specific interactions. Polymer mixtures and the effects of a molecular weight distribution on the adsorption characteristics have not been discussed in this chapter either. Most of these effects may be deduced from the known behaviour of ideal systems. For example, we know that the critical adsorption energy is a very weak function of the chain length, because the number of bonds is one less than the number of segments. This is one reason for preferential adsorption effects in polydisperse mixtures; it favours the adsorption of short chains. Another feature is that one long polymer chain has less translational entropy than several smaller chains containing the same total number of segments. This implies a tendency towards adsorption preference of long chains. When the first effect dominates, which is the case in a polymer melt, the shorter chains populate the interfacial region. However, in dilute solutions the opposite is true; the long chains displace the short ones from the surface. Very interesting complications arise in solvent mixtures. With a binary solvent one can either modify the solvent quality, tune the effective adsorption energy (for super critical solvent mixtures without a solubility gap), or induce capillary condensation phenomena in slits (when the solvent mixture has a solubility gap). Clearly, a combination of these effects may also occur. Last but not least, we mention that we have assumed that the surface is Ideally flat and homogeneous. We know, however, that surfaces often are chemically heterogeneous (random, patchy), rough (height fluctuations) or flexible (liquid surfaces, membranes). The classical mean-field approximation that the concentrations may be averaged laterally now breaks down. When there Is sufficient symmetry in the surface properties (stripes, spherical patches, etc.), one may use a two-gradient approach of the Edwards equation or the numerical SF equivalent of it. Surface roughness may be implemented in a lattice model using the Lego model. For surfaces with less symmetry, one needs to turn to computer simulations because it is nearly impossible to numerically solve the Edwards equation on a 3D grid. At present, computer simulations are very accurate to model the properties of short chains and relatively small systems. However, we anticipate that one day we will use computer simulations to go beyond the mean-field approximations and give more accurate answers to the many questions encountered in the behaviour of polymers at interfaces and their effects on colloidal stability.
1.92
INTERACTION BETWEEN POLYMER LAYERS
1.14 General References 1.14a Polymers in solution This field is well developed. Numerous textbooks in this area have appeared, but there are a few classical texts which cover most of the field. We list a few books in chronological order. P.J. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca NY (1953). (The 'bible' for polymer chemistry and physics, including polymer statistics and polymers in solution. Old, but still valuable.) C. Tanford, Physical Chemistry oj Macromolecules, Wiley, New York (1961). (Classical, like Flory's book, but with more emphasis on proteins.) P.J. Flory, Statistical Mechanics of Chain Molecules, Interscience, NY (1969). (A more refined and rather mathematical treatment of chain statistics, starting from an atomic level.) P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell Univ. Press, Ithaca, NY (1979). (Scaling description for polymers in good solvents, emphasizing the 'blob' structure as compared to homogeneous 'mean-field' solutions.) H. Yamakawa, Modern Theory of Polymer Solutions, Harper & Row, NY (1971). (Thorough statistical thermodynamic treatment of polymer solution properties.) H.-G. Elias, An Introduction to Polymer Science, VCH (1997). (An elementary text, updated from classical books by the same author.) Water-soluble Polymers: Solution Properties and Applications, A. Amjad, Ed., in: Proceedings Symposium Colloids Surf. Divs. Of the American Chemical Society 1997, Kluwer (1999). (A collection of conference papers.) L. Schafer, Excluded Volume Effects in Polymer Solutions, Springer (1999). (Rather advanced text based on scaling analysis.) K. Kamide, T. Dobashi, Physical Chemistry of Polymer Solutions: Theoretical Background, Elsevier (2000). C. Wohlfarth, CRC Handbook of Thermodynamic Data of Aqueous Polymer Solutions, CRC Press (2004). (Data collection.)
INTERACTION BETWEEN POLYMER LAYERS
1.93
1.14b Polymers at Interfaces This area is newer and still developing. Only a few textbooks are available. P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell Univ. Press, Ithaca, NY (1979). (This text, mentioned above, also pays some attention to polymers adsorbing or depleting from good solvents.) E. Eisenriegler, Polymers near Surfaces, World Scientific (1993). (A valuable text for analytical solutions of the Edwards equation for ideal chains in zero field.) G.J. Fleer, M.A. Cohen Stuart, J.M.H.M. Scheutjens, T. Cosgrove and B. Vincent, Polymers at Interfaces, Chapman & Hall (1993). (The only comprehensive text so far on polymer adsorption, with emphasis on both theoretical developments and comparison with experiment.) Polymer Surfaces and Interfaces: Characterization, Modification and Application, K.L. Mittal, K.-W. Aizawa, Eds., VSP (1996). (Application-oriented, technological text.) Polymeric Systems. I. Prigogine, S.A. Rice, Eds., Adv. Chem. Phys. 94, Wiley (1996). Polymer Interfaces and Emulsions, K. Esumi, Ed., Marcel Dekker (1999). (Relation between polymer adsorption and emulsion stability.) R.A.L. Jones, R.W. Richards, Polymers at Surfaces and Interfaces, Cambridge University Press (1999). (Emphasizes the surfaces of polymer melts, wetting, and adhesion, rather than adsorption from solution.) Apart from these textbooks, several reviews are worth mentioning. The most relevant ones are (in chronological order): E. Dickinson and M. Lai, Adv. Mol. Relax. Interact. Proc. 17 (1980) 1. (A review of simulation methods applied to polymers.) A. Silberberg, in Encyclopedia of Polymer Science and Engineering Vol. I, Wiley (1985)577. I.D. Robb, Comprehensive Polymer Sci. 2 (1989) 733. H. J. Ploehn and W.B. Russel, Adv. Chem. Eng. 15 (1990) 137. M. Kawaguchi and A. Takahashi, Adv. Colloid Interface Sci. 37 (1992) 219.
1.94
INTERACTION BETWEEN POLYMER LAYERS
G.J. Fleer and F.A.M. Leermakers, in Coagulation and Flocculation: Theory and Applications, 2nd. ed., B. Dobias, H.J. Stechemesser, Eds., Surfactant Science Series, Marcel Dekker (2004). (This extensive treatment considers both adsorption and interaction.) 1.14c Effect of polymers on colloidal interactions Even though this is an important field in the context of colloids, not many books are available. We list them chronologically. The Scientific Basis of Flocculation, K.J. Ives, Ed., Sijthoff & Noordhoff (1978). (This collection of papers for a NATO Advanced Study Institute contains several contributions several contributions highlighting practical aspects of flocculation of dispersions by polymers.) D.H. Napper, Polymeric Stabilisation of Colloidal Dispersions, Academic Press (1983). (Comprehensive discussion of steric repulsion between polymer layers, but adsorption as such is not considered in detail.) W.B. Russel, D.A. Saville, W.R. Schowalter, Colloidal Dispersions, Cambridge University Press (1989). (One chapter in this advanced text (chapter 6), deals with forces due to soluble polymer.) Colloid-polymer Interactions: Partlculate, Amphiphilic, and Biological Surfaces, P. Dubin, P. Tong, Eds., ACS Symp. Series 532 (1993). (A collection of conference papers.) Colloid-polymer Interactions: from Fundamentals to Practice, R. Farinato, P. Dubin, Eds., Wiley-VCH, (1999). (A collection of conference papers.) V.A. Hackley, Polymers in Particulate Systems, in: Surfactant Science Series 104, Marcel Dekker (2001). Two recent reviews are worth mentioning: O. Spalla, Nanoparticle Interactions with Polymers and Polyelectrolytes in Current Opinion in Colloid & Interface Sci. 7 (2002) 179. (Emphasizes mixtures of polymers and nanoparticles.) G.J. Fleer and F.A.M. Leermakers, in Coagulation and Flocculation: Theory and Applications, 2nd. ed., B. Dobias, H.J. Stechemesser, Eds., Surfactant Science Series, Marcel Dekker (2004). (This extensive review considers both adsorption and interaction).
2
POLYELECTROLYTES
Martien Cohen Stuart, Renko de Vries and Hans Lyklema 2.1
2.2
Introduction
2.1
2.1a
Physicochemical properties
2.1
2.1b
Chemistry
2.1
Polyelectrolyte electric double layers 2.2a
charge distributed?
2.3
2.4
2.5
2.6
2.7
2.8
2.3
Polyelectrolytes: why are they charged and how is the 2.3
2.2b
Cylindrical electric double layers without added electrolyte
2.2c
Cylindrical electric double layers with added electrolyte
2.13
2.6
2.2d
Proton exchange equilibria of polyelectrolytes
2.16
Polyelectrolyte configurations
2.25
2.3a
Dilute solutions
2.25
2.3b
Semidilute solutions
2.32
2.3c
Grafted polyelectrolytes
2.37
2.3d
Polyelectrolyte gels
2.41
Polyelectrolyte viscosity
2.45
2.4a
General features
2.45
2.4b
Lateral information
2.47
2.4c
The intrinsic viscosity
2.50
2.4d
The dilute and semidilute range
2.53
Polyelectrolytes in electric fields
2.57
2.5a
The issue of electrokinetic binding
2.58
2.5b
Conductivity, backgrounds
2.60
2.5c
Conduction: illustrations
2.64
2.5d
Dielectric properties
2.69
Electrostatically driven complexation and phase separation
2.69
2.6a
Solubility of polyelectrolytes
2.69
2.6b
Polyelectrolyte complexes
2.71
2.6c
Compex coacervation
2.71
2.6d
Polyampholytes
2.77
2.6e
Polyelectrolyte multilayers
2.78
2.6f
Complex coacervate micelles
2.79
Applications of polyelectrolytes
2.81
2.7a
Flocculation
2.81
2.7b
Stabilization enhancement
2.82
General references
2.82
This Page is Intentionally Left Blank
2 POLYELECTROLYTES MARTIEN COHEN STUART, RENKO DE VRIES AND HANS LYKLEMA
2.1 Introduction
2.1a Physicochemical properties Polyelectrolytes are macromolecules that, when dissolved in a polar solvent like water, have a (large) number of charged groups covalently linked to them. In general, polyelectrolytes may have various kinds of such groups. Homogeneous polyelectrolytes have only one kind of charged group, e.g. only carboxylate groups. If both negative (anionic) and positive (cationic) groups occur, we call such a molecule a polyampholyte. These polyelectrolytes will only be briefly discussed at the end of this chapter. Self-assembled structures, such as linear micelles or linear protein assemblies, also often have many charged groups; these structures may have properties very similar to those of polyelectrolytes, but we shall not deal with them in this chapter. Special properties of polyelectrolytes, as compared with uncharged polymers, are their generally excellent water solubility, their propensity to swell and bind large amounts of water, and their ability to interact strongly with oppositely charged surfaces and macromolecules. Because of these features, they are widely used as rheology and surface modifiers. These typical polyelectrolyte properties are intimately related to the strong electrostatic interactions in polyelectrolyte solutions and, hence, are sensitive to the solution pH and the amount and type of electrolytes present in the solution. 2.1b Chemistry As will be explained in more detail in sec. 2.2, polyelectrolytes in solution acquire charges by the adsorption of ions or by the dissociation of surface groups. Especially important are proton exchange equilibria. For macromolecules, there are a limited number of different functional groups that are typically involved in proton exchange equilibria; we list them in table 2.1 below, along with their chemical structure. We also give in this table an indication of their tendency to acquire a charge in terms of a range of values of their deprotonation constant (expressed as its negative logarithm, i.e. pK), which characterizes each group when isolated from other charged groups.
Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
2.2
POLYELECTROLYTES
Table 2.1. Ionizable groups occurring in common polyelectrolytes and their proton exchange properties
Type of group
reaction
intrinsic
examples
10.5-11
poly(allylamine)
weak cationic -NH2 {aliphatic)
-NH3+
-NR2 (aliphatic)
-NR 2 H + ?^ - N R 2 + H +
10.6
poly(dimethylaminoethyl methacrylate); poly(ethylene imine)
-0-NH2 (aromatic)
(*-NH3+ ^ ^-NH2 + H +
4.5-5
poly( vinylaniline)
-Pyridine (= N: )
= N: = NH+ ^ = 1^: + H +
5.3
poly (vinylpyridine)
-
-
poly(diallyl dimethyl ammonium chloride)
-COOH
-COOH ^=i-COO" + H +
4-5
poly(acrylic acid)
-PO 4 H 2
- P O 4 H 2 , =^-PO 4 H~ + H+
0-1
phosphate esters of poly( saccharides)
-PO 4 H 2
=PO 4 H ^- - P O 4 - + H+
0-1
DNA
-
poly(styrene sulphonic acid); carrageenan
- -NH 2 4 -H+
strong cationic -NR+
weak anionic
strong anionic -SO3H
-SO3H
- - S O 3 ~ + H+
As can be seen, the groups are classified into 'weak' or 'strong,' depending on whether or not they have a pK in the range 0-14 so that they can undergo proton exchange reactions at experimentally relevant pH. Polymers with weak groups (0 < pK < 14) can adjust their average degree of dissociation to the physicochemieal conditions, and the charge on individual groups can fluctuate; molecules with such groups are also referred to as 'annealed' polyelectrolytes. Polymers with strong groups (pK < 0; pK > 14) have fixed charges and charge distributions along the chain, whatever the pH, and are therefore sometimes called 'quenched' polyelectrolytes. Many different kinds of backbones occur: all-carbon (vinyl, allyl) chains, polytesters), poly (ethers), poly (saccharides), poly(peptides), as well as poly(imines) where the charged group is itself part of the backbone. Some backbones are hydrophilic, such as those in poly(saccharides), but most are hydrophobic and the corresponding molecules would not dissolve if they would not carry a sufficient number of charges.
POLYELECTROLYTES
2.3
Copolymers, which have more than one kind of monomer (in particular, monomers with and without dissociating groups), should also be mentioned. For this category, the maximum charge that a molecule can acquire is, of course, determined by the chemical composition. Many of the polyelectrolytes discussed here owe their very existence to synthetic polymer chemistry. This is not to say, though, that nature does not produce polyelectrolytes. Many naturally occurring polysaccharides produced by plants, bacteria or animal cells are polyelectrolytes. One of the best-known natural polyelectrolytes Is DNA, which contains in-chaln phosphate groups as chargeable sites, and its close relative RNA is, of course, also a polyelectrolyte. Finally, polypeptides and proteins often carry charged groups and can thus be defined as (very complicated) polyelectrolytes. Since the structure and physicochemical behaviour of many proteins has rather special features, these molecules will receive attention in a chapter especially devoted to them (chapter 3). In this chapter, we shall first discuss how charge and countercharge In a polyelectrolyte in solution are distributed (the electrical double layer), and we consider the proton exchange equilibria from which the charges originate (2.2). We then take a look at molecular shapes (conformations) and the way these are influenced by the charge (2.3) for free chains in solution, end-grafted chains, and gels. Section 2.4 deals with the viscosity of polyelectrolyte solutions and sec. 2.5 with polyelectrolytes in external electrical fields (conductivity, electrophoresis and dielectric properties). Finally, we discuss in sec. 2.6 the solubility of polyelectrolytes and the formation of complexes in mixtures of polyelectrolytes of opposite charge. 2.2 Polyelectrolyte electric double layers 2.2a Polyelectrolytes: why are they charged and how is the charge distributed? In solution, polyelectrolytes can acquire a charge on the chain by two mechanisms: (i) ions from the solution can bind to them and/or (ii) surface groups can dissociate. For instance, as discussed in sec. 2.1, polyamines can pick up protons to form positive -NHg groups and from polyacids, protons can desorb to form negative, say, -COO ~ or -OSOg groups (we ignore other than simple ions; in particular, adsorption of ionic surfactants will not be considered here). For all the charge formation processes, the driving force is of a chemical or entropical origin, determined by the Gibbs energy difference between the bound and free state, just as with the dissociation of low-M molecules. In processes of type (i) the process is entropically unfavourable, whereas for those of type (Ii) it is entropically driven. Electric forces always oppose the charge formation; in fact, all processes come to an end when the counteracting electric forces cancel any further charging. Then, equilibrium is attained. At equilibrium the Gibbs energy is at a minimum and negative. Electric double layers around polyelectrolytes form spontaneously. It is
2.4
POLYELECTROLYTES
typical for polyelectrolytes that the equilibrium state is also determined by the degree of swelling. The spontaneous charge separation leads to the creation of a chain charge and an equal but opposite countercharge. When the solution contains low-M electrolyte, the latter consists of an excess of counterions and a deficit of co-ions. This negative adsorption of co-ions leads to the expulsion of electroneutral electrolyte, the so-called Donnan exclusion. In salt-free solutions the countercharge only contains counterions; in that case, there can be no Donnan exclusion. All of this is basically not different from the charging principles for hydrophobic colloids, see sec. II.3.2. Quantitative differences with double layers on flat plates and macroscopic spheres may come forward because of the way the chain charge is distributed. The issue is connected to the question as to whether it is allowed to treat the double layer charge as smeared out, which would allow for mean-field interpretations, such as the PoissonBoltzmann distribution for the diffuse part and Bragg-Williams type analyses for the non-diffuse part. As long as a polyelectrolyte chain may be considered as a long charged cylinder with macroscopic radius a, one can assign a surface charge density (7° to it and treat the countercharge by one of the models discussed in chapter II.3 for macrobodies. Anticipating the arguments below, it is at least likely that the outer side of the countercharge obeys PB statistics, i.e. it can be described as a diffuse double layer according to Gouy and Chapman. In figs. II.3.14 and 15, we showed that for the diffuse parts cylindrical double layers assume an intermediate position between flat and spherical ones. As intuitively expected, the larger xa, where a is now the radius of the cylinder, the more similar spherical, cylindrical and flat double layers are. However, a is often not so large that this picture is allowed over the entire electric double layer. For further discussion, it makes sense to divide the countercharge into two parts, an outer or diffuse part and an inner, non-diffuse part. The diffuse part is readily defined as that part where ion size effects are negligible, where charges may be considered smeared out, and where binding is solely caused by electrostatic attraction (zey/(r) per ion of valency z at distance r from the axis of the cylinder). In the nondiffuse part these premises no longer hold. Because of the proximity to the chain, specific ( = non-electrical) binding can usually not be ignored. The specific phenomena depend on the nature of the charged groups and on their positions along the chain. The nature effect is nothing else than the specific binding, which gives rise to the familiar lyotropic sequences (sec. IV.3.9i). However, the position effect is new. For instance, for charged groups residing on sidechains (as in polymethacrylates), it is possible for ions to accommodate between the charges whereas polyelectrolytes with all charges on the backbone (as in poly-imines) do not have this option. As positioning of ions between and outside charges can follow different patterns, interesting lyotropic sequences may be expected, beyond those generally observed for macroscopic solid surfaces. According to general experimental experience with macroscopic surfaces, plots of the diffuse (~ electrokinetically or stability-wise active) charge density ad ~ (7ek as a
POLYELECTROLYTES
2.5
function of a° start as a straight line under 45° where a^ = -o° , to level off to a plateau where further increase of a° does not lead to further enhancement of a^ . See, for instance, figs. II.3.63 and 4.13, for f and cf* , respectively. The diffuse charge density rarely exceeds a few |iC cm"2 . The meaning of the plateau is that a situation has arisen in which any further added charge on the surface is compensated by a countercharge in the non-diffuse part. This non-diffuse part was generally called the Stern part of the double layer, charge densityCT1, with a° + a1 + a^ = 0 because of electroneutrality. Obviously, in this respect double layers around polyelectrolytes resemble Gouy-Stern double layers. At this stage it is necessary to introduce the term ion condensation. This term is typical for polyelectrolyte science, and stands for the binding of (counter)ions to the charged chain. In the literature the term is used in various ways, for instance in the interpretation of experiments to account for the fraction of counterions that do not contribute to osmotic pressure, conductivity etc., or, in theory, in which models for the process of condensation are developed. Theories and the various experimental observations may, or may not, agree. The term is well established in the domain of polyelectrolytes and has the proper resonance. It is not appropriate to speak indiscriminately of 'bound' and 'free' ions, because diffuse ions are also bound in a thermodynamic sense. Hence, we shall continue to use the term condensation in this chapter as a general term for ions that are so strongly bound to (or captured by) the chain that they are osmotically, electrokinetically or conductometrically inactive, irrespective of the reason for the condensation (electric and/or specific). According to this definition, condensed ions and Stern ions are identical. History has led to differentiation between the terms, because the founding fathers of polyelectrolyte theory (Fuoss, Katchalsky, Oosawa, Manning...) were unfamiliar with the much older Stern model; on top of this came the independent development of theories. Typically, in polyelectrolyte theory (Oosawa, Manning...) two-state models have been put forward on purely electrostatic principles. We shall return to these in sees, b and c below. Description in terms of Bragg-Williams type approaches, as in Stern theories including specific binding, appears to be virtually absent in polyelectrolyte literature. As a result, in most polyelectrolyte elaborations, there is little room for ion-specificity interpretations. On the other hand, Stern and condensation theories both predict the formation of a plateau in the cfi ( a° ) relationship, as mentioned before. In order to finally arrive at theories for polyelectrolyte configurational statistics, it appears unavoidable to use a coarse-grained description of the polyelectrolyte chain and the electric double layer surrounding it. Following most of polyelectrolyte theory, the complicated polyelectrolyte molecular structure is modelled as a charged cylinder characterized by just two parameters: its radius a and its surface charge density a°. All details at length scales below the radius a are lumped into the effective parameters a and a°. Sometimes an even further simplification is used and the polyelectrolyte chain is modelled as a line charge with a linear charge density of v = 2naa° . Following
2.6
POLYELECTROLYTES
most of polyelectrolyte literature, a Stern layer to account for ion-specific binding is not taken into account. Instead, the double layers are described by the non-linear PB equation only, assuming all counterions are outside of the cylinder of radius a that models the polyelectrolyte chain. As mentioned, even in such purely electrostatic models, there is a plateau in the 0^(0°) relationship. This is the effect that is usually referred to in the polyelectrolyte literature as (counter)ion condensation. The basic idea, stemming from Imai and Ohinshi11, but elaborated by Oosawa21 and Manning31, is that upon charging the chain above a critical chain charge, the resulting potential near the chain becomes so high that any additional charge is compensated for by an electrostatically captured counterion. The simplified two-state models of Oosawa and Manning (based on the PB equation) predict that, in the limit of vanishing concentrations of both electrolyte and polyelectrolyte, o^fcr0) plots (or equivalents thereof, such as plots of the polyelectrolyte osmotic pressure versus the chain charge v) should consist of an initially linear part, which discretely breaks off into a plateau at the critical chain charge. This discrete limiting behaviour is sometimes referred to as the Manning limiting law. Indeed, for this special limiting case, the solution of the full PB equation does exhibit a mathematical discontinuity at a well-defined critical chain charge. In contrast to the simplified models however, it does not predict a distinct 'break' in, for example, the osmotic pressure as a function of the chain charge. Nevertheless, the existence of a universal critical chain charge for polyelectrolytes and the concept of counterion condensation have been very important in polyelectrolyte modelling to help develop approximate theories for polyelectrolyte properties. In the next subsection, we therefore discuss in some detail both the Oosawa two-state model and the full PB approximation for cylindrical double layers of strong polyelectrolytes without added electrolyte. Next, the case of added electrolyte is treated, again assuming strong polyelectrolytes. For weak polyelectrolytes, charges are determined by adsorption/dissociation equilibria. A final paragraph considers in detail the important case of proton exchange equilibria for linear polyelectrolytes. 2.2b Cylindrical electric double layers without added electrolyte To begin, we assume that the overall polyelectrolyte configuration is at least locally rod-like and study the cylindrical double layer in the PB approximation. While in practice there will always be some amount of electrolyte present, its presence may be neglected if the total number of co-ions is negligible as compared with the total number of counterions. Roughly speaking, this is the case when the concentration of added electrolyte is small as compared with the total polyelectrolyte monomer concentration,
11
N. Imai, T. Ohinshi, J. Chem. Phys. 30 (1959) 1115. F. Oosawa, Polyelectrolytes, Marcel Dekker (1971). 31 G. S. Manning, J. Chem. Phys. 51 (1969) 924. 21
POLYELECTROLYTES
2.7
a situation that is certainly not uncommon in practice. Polyelectrolyte electric double layers in the absence of added electrolyte are also important from an historical and theoretical point of view, since they have played an important role in the development of the theory of counterion condensation, one of the central concepts in polyelectrolyte modelling. A distinct difference with the case of excess added electrolyte is the absence of a well-defined bulk concentration of ions that can be used as a reference point (defining the zero of the electrostatic potential). Instead, models for electric double layers in the absence of added electrolyte depend on the definition of a certain volume to which the counterions are confined. That is, they are so-called cell models. In practice, the volume of the cell is related to the concentration of charged objects (colloids, polyelectrolytes). This reflects the physical effect that diluting the charged objects dilutes the counterions, and hence affects the electric double layers. Oosawa two-state model As mentioned, in the limit of vanishing concentration of both electrolyte and polyelectrolyte, the solution of the cylindrical PB equation exhibits a mathematical discontinuity at a critical chain charge. Simplified two-phase models for this limiting case, based on the PB equation, have been developed by Oosawa (loc. cit.) and Manning (loc. cit.). Their main virtue is that they clearly demonstrate the physical mechanism that leads to (electrostatic) counterion condensation. Below, we first discuss the Oosawa two-phase model. Following, we introduce the exact solution of the cylindrical PB equation in the absence of added electrolyte. Finally, we compare with experimental data for polyelectrolyte osmotic pressures. The Oosawa two-phase model provides a simple way to estimate the magnitude of the fraction /
of free counterions. The remaining fraction of counterions, ( 1 - / ) , is
assumed to be electrostatically captured or condensed into a cylindrical region close to the chain, of radius a (the polyelectrolyte chain itself is viewed as a line charge). The free counterions are assumed to move around within a cylindrical region a < r < R. The value of the radius R is set by the volume fraction (p of chains, which is approximated by (p = a2/R2
[2.2.1]
The geometry of the Oosawa two-phase model (which is a cell model) is illustrated in fig. 2.1. Without loss of generality, the strong polyelectrolyte is assumed to be negatively charged and to have a constant charge density of -v elementary charges e per unit length ( v is positive). Of course, both the counterion concentration and the electrostatic potential y/ are continuous functions of the distance r to the line charge at the centre of the cylinder. In the two-state model, the continuous distributions are approximated by bimodal distributions. The condensed counterions inside the cylinder
POLYELECTROLYTES
2.8
Figure 2.1. a) Geometry for the Oosawa two-phase model. Condensed counterions move in the inner cylinder of radius a. Free counterions move in the volume bounded by the inner and outer cylinders of radius R. b) Relation between the potential t//[r) arising from a line charge of strength -Jv and the potential difference Ayr in the Oosawa two-phase model. Note that the point of zero potential has not been defined in this graph. of radius a are assumed to have a constant concentration c cond , and the free counterions in the cylindrical region between a and R are assumed to have a constant concentration c free , given by the average values: C
free-TT-^T JVAv 7taz _
' 2 - 2 -21
1 (l-J)v
Like the counterion distribution, the electrostatic potential y/ is assumed to be bimodal: the electrostatic potential inside the cylinder of radius a is y/a and the potential in the cylindrical region between a and R is y/R . The distribution of the counterions over the two regions is then governed by a Boltzmann factor involving the difference in the electrostatic energy of the counterions: c
free cond
c
=
_J__ 1 ^ e xp(-eAiy/kT) ! " /
[2.2.4]
where Ay/ is the (average) potential difference between inner and outer cylinder (see [2.2.5]). To estimate this electrostatic potential difference, the inner cylinder is viewed as a line charge with a charge density of - Jv elementary charges e per unit length. This line charge gives rise to a continuous potential y/(r). From this potential, we estimate (as illustrated in fig. 2.1) Ayr= y/{R)-y/{a)
[2.2.5]
POLYELECTROLYTES
2.9
Standard electrostatics then gives ^ - = 2Jxln(R/a)
[2.2.6]
K1
where we have also introduced the (dimensionless) charge parameter x x = lBv
[2.2.7]
The Bjerrum length 1B used here, and throughout this chapter, is the same one that was introduced in the chapter on the adsorption of polymers and polyelectrolytes ([5.2.23]) I
B
e " 4ne0ekT
It is noted that (B is twice as large as rB , the corresponding length in low-M electrolyte solutions originally introduced by Bjerrum himself, see [1.5.2.30a]. For aqueous solutions at room temperature, ZB =0.7nm. Combining [2.2.4] and [2.2.6] gives an equation that can be solved for the fraction of free counterions J : -^— = (pfx-1
[2.2.8]
In the limit of infinite dilution,
r i x < \ f =\
[2.2.9] [l/x
x > l
As already alluded to, there appears to be a critical value for the charge parameter x c = 1, corresponding to a linear charge density of vc = 1 / lB. The above simple model predicts that, below this critical value (at infinite dilution), all counterions are completely free, and above it, any additional charge on the chain is compensated for by an electrostatically captured counterion. This special behaviour is the Manning limiting law mentioned previously. Solution to the full cylindrical PB equation Obviously, the above two-phase model is no more than a crude approximation. We next discuss the exact solution to the PB equation for the electrostatic potential y/{r) at a distance r from a negatively charged cylinder. As before, the cylinder radius is a, its charge density is -ve per unit length. Counterions are confined to the cylindrical volume for which a < r < R. For a negatively charged polyelectrolyte in the absence of added electrolyte, there are only positively charged counterions. These have a concentration profile c+ (r) . The Poisson equation reads
2.10
POLYELECTROLYTES
U - + ~ W ) = -(4;re/£0£)c+(r) u + ydr rdr)
[2.2.10]
and the Boltzmann distribution for the counterions is c+(r) = coexp(-e^(r)/fcT)
[2.2.11]
where c 0 is the counterion concentration at the point of zero potential {r = R) . The Poisson-Boltzmann equation for the dimensionless electrostatic potential y = ey//kT can now be written as ^ - + - ^ - y(r) = -A2exp[-y(r)] \dr rdr)
[2.2.12]
The screening parameter X is analogous to the Debye screening parameter K for the case of excess added electrolyte; A"1 is often called the Gouy-Chapman length. It is given by „ 4;rfRcn A2 = §_o. N
[2.2.13]
Av
The surface charge density at r = a is given by v/2na . This leads to a first boundary condition of d
l =2£ 3r r = a a
[2.2.14]
where * = vlB is again the dimensionless charge parameter. At the outer boundary of the cell, the electric field should vanish, which gives the second boundary condition: — 3r
=0
[2.2.15]
r=R
With respect to this equation, the point of zero potential, or equivalently, the reference concentration of counterions c 0 , can be chosen freely since adding a constant to the electrostatic potential does not affect its derivative, the electric field. A common choice in cell models is to set the potential to zero at the outer boundary of the cell. Then, c 0 equals the counterion concentration at this outer boundary. In the present case, it is more convenient to instead choose c 0 to be the average concentration in the cylindrical cell volume. For a section of length L, the number of counterions released by the charged cylinder is vL . In the limit a << R considered here, the volume in which the counterions are dispersed may be approximated by nR2L . This leads to: Av
A2=~ R
[2.2.16]
POLYELECTROLYTES
2.11
The analytical solution to the above PB equation and boundary conditions was first applied to polyelectrolytes by Fuoss, Katchalsky and Lifson1' and by Alfrey, Berg and Morawetz21. The solutions are, for a << R, \ix r2 y{r) = In — — s i n h 2 (Bin br)
[B2 R2
[2.2.17]
J
when the integration constant B is real (at low charge densities), and
[
n
2
1
.£__!_. sin 2 (|B| In br)
\B\ R
J
[2.2.18]
when the integration constant B is imaginary (at high charge density). The values of the integration constants B and b are determined by the boundary conditions at r = a and at r = R. WhenB is real, x = ( l - B 2 ) / [ l + Bcoth(Bln(R/a))] [2.2.19] 1
Blnb = - B l n R - t a n h " B and when B is imaginary, x = (l + |B| 2 ) / /[l + |B|cot(|B|ln(R/a))] [2.2.20] 1
|B|lnb = -|B|lnK-tan- |B| In the limit of vanishing polyelectrolyte concentrations
x <1
\B\ -> 0
x >1
[2.2.21]
What does this imply for experimental observables, such as the counterion osmotic pressure? The counterion osmotic pressure is usually given in terms of the osmotic coefficient >, n/RT
=
[2.2.22]
According to Marcus31, in the PB approximation, the osmotic pressure equals the ideal pressure resulting from the small ions at the positions where the electric field vanishes. In our case, it therefore equals the ideal ion pressure due to the counterions 11 21 31
R. Fuoss, A. Katchalsky, and S. Lifson, Proc. Natl. Acad. Sci. U.S.A. 3 7 (1951) 579. T. Alfrey, P. Berg, and H. Morawetz, J. Polym. Sci. 7 (1951) 543. R.A. Marcus, J . Chem. Phys. 2 3 (1955) 1057.
2.12
POLYELECTROLYTES
at r = R, whence 0 = exp(-y(R))
[2.2.23]
Using the limiting behaviour of the integration constant B, the osmotic coefficient is found to be (l-x/2 (p = \
x<\
[Il2x
x>\
[2.2.24]
This replaces the prediction for the fraction a of 'free' counterions of the much more approximate Oosawa two-phase model discussed earlier. The osmotic pressure itself is: [2x(l - x)
x<\
[l
x>l
n/nUm=\
[2.2.25]
The limiting value for the osmotic pressure is —M- = RT
1 NAv2xlBR2
[2.2.26]
Plots of the limiting behaviour of the counterion osmotic pressure and the osmotic coefficient are given in fig. 2.2. Note that despite the 'break' in the value of the integration constant B, the osmotic pressures does not exhibit the break predicted by the simpler Oosawa two-phase model and the Manning limiting law.
Figure 2.2. Counterion osmotic coefficient
Experimental values for osmotic coefficients of highly charged polyelectrolytes without added electrolyte are usually of the order q>= 0.1...0.3 for monovalent counterions, and agree reasonably well with the predictions of the full cell model. A general trend appears to be that the cell model overestimates the osmotic coefficients. Also, of course, it cannot explain the significant variation with counterion
POLYELECTROLYTES
2.13
type. For an example, see fig. 2.3, which shows recent data of Deserno et al. 1 ' for a rodlike polyelectrolyte, poly-(p-phenylene) with positively charged side chains. Presumably, the chlorine counterions have a stronger affinity for the polyelectrolyte chain than the iodine counterions, an effect that could be modeled in terms of a Stern model with specific affinities of the ions for the polyelectrolyte chain. Figure 2.3. Osmotic coefficient 0 as a function of monomer concentration c m for the rod-like polyelectrolyte poly-(p-phenylene) with positively charged side chains. Full symbols: iodine counterions, open symbols: chorine counterions. Full curve: PB cell model, dashed horizontal line: Manning limiting law. (Redrawn from Deserno et al. (loc. cit.)) 2.2c Cylindrical electric double layers with added electrolyte For the case of a charged cylinder with added monovalent electrolyte, the PB equation is written as [II.3.5.58] ^ - + - ^ - y ( r ) = x-2sinh y(r) ydr rdr)
[2.2.27]
First, consider the special case of a line charge of linear charge density v in the DebyeHiickel approximation. In terms of the charge parameter x = vlB , y(r) = 2XK0{KT)
[2.2.28]
More generally, in the PB approximation, the outer field ( KT » 1) of any cylinder is of the form y(r) = 2xeffK0(xr)
[2.2.29]
The parameter xeff is a function of both x and KCL . The interpretation of this parameter is that it is the charge parameter of an equivalent line charge that generates the same far field as the charged cylinder. The concept of the equivalent line charge will turn out to be very convenient when we discuss interactions of charged cylinders with weakly overlapping electric double layers. We should emphasize that the equivalent line charge parameter xeff plays an identical role as the diffuse double layer potential y d for colloids. We prefer to use xeff here since we use constant charge boundary conditions rather than constant potential or charge-regulating boundary conditions.
11
M. Deserno, C. Holm, J. Blaul, M. Ballauff, and M. Rehahn, Eur. Phys. J. E5 (2001) 97.
2.14
POLYELECTROLYTES
For a charged cylinder of radius a, the solution of the Debye-Hiickel equation is well known, and was discussed in II.3.5f. For this case, the charge parameter of the equivalent line charge is: x „= eff
[2.2.30] KaKx(Ka)
For highly charged cylinders, there is no known analytical solution, although an accurate analytical approximation has been developed by Philip and Wooding11, based on the idea that in the non-linear part of the cylindrical double layer, where y(r) » 1, the concentration of co-ions is negligible. Values for xeff (or equivalents thereof) obtained from numerical solutions of the cylindrical Poisson-Boltzmann equation have been tabulated by a number of authors, see for example Fixman and Skolnick21. Alternatively, accurate values for xeff can be obtained from the analytical approximations of Philip and Wooding. For the limit of infinite dilution and no added salt, the (approximate) Manning limiting law predicts a break at x = 1, \x
x <1
[1
x>\
*eff=
[2.2.31]
As Fixman and Skolnick observe, the limiting behaviour is only reached logarithmically upon decreasing the electrolyte concentration. Indeed, a comparison with the numerical values of Fixman (fig. 2.4) indicates that only at very small values of tea is Manning's limiting law a reasonable approximation. Figure 2.4. Effective charge parameter X = sraKj (/ra)xegas a function of bare charge parameter x. Symbols are values tabulated by Fixman (loc. cit), computed from numerical solutions of the full PB equation for various values of jra as indicated. Full curves are guide to the eyes. The lower dotted line is the Manning limiting law for tea —» 0 , the upper dotted line is the Debye-Huckel approximation.
For larger values of KCL , we slowly approach the regime where the cylindrical double layer becomes more and more similar to a flat Gouy-Chapman diffuse double layer, as mentioned in the beginning of this section, and also in II.3.5f. 11 21
J. R. Philip, R. A. Wooding, J. Chem. Phys. 52 (1970) 953. M. Fixman, J. Skolnick, Macromolecules 11 (1978) 863.
POLYELECTROLYTES
2.15
Next we turn to interactions between charged cylinders. The discussion closely follows Brenner and Parsegian11 and Fixman and Skolnick. For cylinders separated by more than a few Debye lengths, the electric double layers are only weakly overlapping. Far away from both cylinders, the electrostatic potential may then be approximated by the sum of the potentials of the isolated cylinders; this is nothing else than the linear superposition approximation (LSA) discussed, e.g., in chapter IV.3. The interaction of the cylinders depends only on the outer field of an isolated cylinder, and is conveniently calculated using the concept of the equivalent line charge.
Figure 2.5. Geometry of skewed line charges for evaluating the electrostatic interaction energy.
First consider parallel cylinders at a distance h. The interaction energy V{h) (per unit length) equals the electrostatic potential of one of the equivalent line charges, times the effective charge density of the other one: V(h)/kT = 2(B1x2ffK0(x-h)
[2.2.32]
To obtain the interaction energy V{h, y) of cylinders at a distance h and skewed at an angle y, we compute the electrostatic potential at some position Sj along one of the line charges by adding up the contributions from all positions s 2 along the other line charge, as suggested by Fixman and Skolnick, and illustrated in fig. 2.5; V(h,y) f r - j - ^ - = Jds 1 jds 2 u(s 1 ,s 2 ) u( S l , s 2 ) = I g l x ^
exof-x-ls - s i ) ^ ~ ^ ri
[2.2.33]
S
2
This gives the simple final result V(h,y) _27txlii fcT rig
11
exp(-rh) sin y
S.L. Brenner, V.A. Parsegian, Biophys. J. 14 (1974) 327.
|2 2 34]
2.16
POLYELECTROLYTES
which can be used down to angles of | y\ ~ a/L , where L is the length of the rods. For even smaller angles, it is more accurate to use the expression for parallel rods.
2.2d Proton exchange equilibria of polyelectrolytes So far it was assumed that some fixed number of polyelectrolyte groups was charged. In reality, of course, the charge is determined by adsorption/desorption equilibria, as explained in sec. II.3.6e. Here, the special, but important case, of proton exchange equilibria is treated in somewhat more detail. Proton exchange reactions can be symbolically represented by X H ^ X " + H+
[2.2.35a]
X H + ^ X + H+
[2.2.35b]
or
For simple monovalent acids or bases, the proton exchange reactions obey classical laws of chemical equilibrium, with a deprotonation constant K defined by: K=
x C
H
[2.2.36]
HX
where we employ the usual definitions pH = -loga H and pK = -logK . In these commonly accepted expressions, activities are defined with respect to an ideal solution of 1 M. To avoid taking logarithms of dimension having quantities, an alternative definition in terms of mole fractions was introduced in II.3.6e. [II.3.6.42] This latter definition differs by - log V m (- log 55.5 for aqueous solutions), where Vm is the molar volume of the solvent. Note that c x denotes both the species X" that occurs in reaction [2.2.35a] and the species X in reaction [2.2.35b]; similarly, c ^ stands both for HX and for HX+ . Also, note that we represent all proton transfer equilibria by their deprotonation constants pK, rather than using profanation
constants, which
would be just given by pKw - pK , where pKw is the dissociation constant of water; at T = 298 K, pKw =14 . We shall denote the fraction of groups (sites) that become deprotonated as a , and the fraction of charged sites as 9. If the charge is formed by deprotonation (as in case (a) above), a= 9; in the other case (b), 9=1-a.
Using this
definition, we have K=
ac a H V(1 - a)c
[2.2.37]
Hence, pH = pK + l o g ( - ^ J
[2.2.38]
As long as each macromolecule in a solution has no more than one single charge, similar proton exchange equilibrium will apply with an intrinsic equilibrium constant pK. However, polyelectrolytes usually carry many charges and these interact by
POLYELECTROLYTES
2.17
Coulomb forces. If a charge z.e on a chain in a medium of dielectric constant e is surrounded by JV other charges z^e at distances r^ , it has an extra electrostatic energy Uel given by Coulomb's law: Uel = Y - J - i —
[2.2.39]
Dissociation will cost this extra energy above the intrinsic (often termed 'chemical') work of dissociation for an isolated group. For a homogeneous polyelectrolyte, which has acquired many charges, L/ej is positive, implying that further charging becomes more difficult. In other words, each charged group must overcome the total potential energy due to all other charges, many of which are on the chain. As a result, a larger change in pH is needed to reach a certain a, i.e. the titration curve extends over a larger pH range and has a shape different from that of monovalent acids or bases. The proton binding behaviour of polyprotic systems has been extensively discussed in a review by Borkovec, Jonsson and Koper11. A general approach to a theory of titration (proton binding) isotherms requires calculation of the partition function for the macromolecule as a function of its degree of protonation, including the coupling between charge (distribution) and shape of the molecule. This is a very demanding task, often only tractable by means of molecular simulations. Here, we first neglect the coupling between the charge distribution and the shape of the macromolecule. We return to this issue at the end of this section when we discuss computer simulations of polyelectrolyte titrations. Two limiting cases are of special interest, and can be dealt with using simple approximations. If the interaction between the charged groups on the polyelectrolyte is weak, we can use a mean-field approximation. This is the limit of weak coupling. It usually applies when distances between charged groups on the polyelectrolyte are not too small. For polyelectrolytes with very small distances between charged groups, we are in the limit of strong coupling. In this limit, a charged group mainly interacts with its nearest neighbours. The problem of describing the titration of polyelectrolytes with only nearest-neighbour interactions is equivalent to the so-called Ising model, for which analytical solutions are available. First consider the limit of weak coupling. The Bragg-Williams approximation consists of assuming that the way the charged groups are distributed along the chain is not affected by the interactions. In other words, the configurational entropy is the same as that for a system of non-interacting sites, and given by s
conf =fc{alna+(l-a)ln(l-a)}
[2.2.40]
11 M. Borkovec, B. Jonsson, and G.J.M Koper, in Surface and Colloid Science 16, E. Matijevic, Ed., 99-339, Kluwcr Academic/Plenum Press (2001).
2.18
POLYELECTROLYTES
The deprotonation isotherm for this case now satisfies pH =P K - e ^ ( e ) + l O g p U V P fcTlnlO S U - « J
[2.2.41]
where if/>{0) is the potential on the chain's backbone at a particular degree of charging 0 (9= a° /o^ax )• It is quite customary to express experimental data on proton binding in terms of pKeff as a function of a or (1 - a). pK eff =pK + ApK
[2.2.42]
In the above mean-field approximation, ApK =
_ J ^ L JcTlnlO
[2.2.43,
In the Debye-Hiickel limit of low charge densities, the potential t//° is proportional to the charge density <j°, hence to 8. The shift is therefore linear in the dissociation. Hence, ApK = ~e6, with ~e defined in terms of the charge parameter x, as follows:
Figure 2 . 6 . pK e f f a s a function of 8 a t v a r i o u s c o n c e n t r a t i o n s of NaNO 3 , a s indicated in t h e figure; (a) for t h e p o l y b a s e poly (dimethylaminoethyl methacrylate) (0=1- a) a n d (b) for t h e polyacid poly (acrylic acid) [6= a).
POLYELECTROLYTES
2.19
x Ko0ra) ~ rfglnlO KjOra)
L
where Ko and K{ are the zeroth- and first-order Bessel functions of the second kind, respectively. An example is shown in fig. 2.6 for the protonation of polydimethylaminoethyl methacrylate at various concentrations of added salt. Since the above expression is based on various assumptions, such as the absence of correlation effects (mean-field assumption) and a constant double layer capacitance (Debye-Hiickel limit), deviations from linearity are likely to occur as soon as these assumptions do not hold any longer. Indeed, non-linear ApK plots are often observed in experiments, and can be represented in the form of a series expansion: ApK = W+e'02 + ...
[2.2.44]
Examples of the latter are given in fig. 2.7a and b, where the full (non-linear) PB equation was used to fit the data. At this point, we should also mention the Henderson-Hasselbalch (HH) equation [1.5.2.34], an empirical equation introduced as a generalization of [2.2.37]:
pH = pK + nlogf-^-J
[2.2.45]
Figure 2.7. a. Degree of protonation as a function of pH for hyaluronic acid at various concentrations of monovalent salt, as indicated. Experimental results are denoted by symbols. Solid curves are fits to the PB equation for a cylinder (radius 1 nm, charge spacing 1 nm and intrinsic pK = 2.75). b. Titration curves plotted as versus 1 — 0 (degree of protonation) for polylDL-glutamic acid) at two concentrations of monovalent salt. Experimental results arc denoted by symbols. The dashed and solid curves are for a PB model with cylinder radius 0.5 nm; dashed curve: flexible chain model (charge spacing 0.7 nm): solid curve: rigid rod model (charge spacing 0.5 nm). (Redrawn from (a) R.L. Cleland, J.L. Wang and D.M. Detiler, Macromolecules 15 (1982) 386; (b) D.S. Olander, A. Holtzer, J. Am. Chem. Soc. 90 (1968) 4549, with fits from M. Ullner, B. Jonsson, Macromolecules 29 (1996) 6645.
2.20
POLYELECTROLYTES
where pKgff and n are an effective deprotonation constant and a parameter accounting for the polyelectrolyte effect, respectively. This equation appears to work quite well for many cases as is best checked by making a 'Henderson-Hasselbalch plot' (pH versus log(cr/l-«)), which should be linear. Comparison with [2.2.35] suggests that if the HH equation applies, the change in pK must scale as (rt -l)log(a/l- a) , which is not justified by theoretical arguments. An illustration follows in fig. 2.12. An alternative method sometimes introduced in the literature is to use the Donnan capacitance. The idea is that the space in the neighbourhood of the chain has a welldefined 'inner' potential given by the Donnan model, and that it is this potential that modifies the proton binding and conversely. This approach may be adequate for strongly branched molecules, gel-like polymer particles or brushes, which have an inner region that is large compared with the Debye length. We shall use that approach in sec. 2.3e. For linear chains such a region does not really exist, and the Donnan approach is less appropriate; indeed, the linear dependence of the capacitance on ionic strength predicted by the Donnan model is usually not found in experiments with linear chains. When the distance between proton binding sites on the chain becomes smaller, the interaction between them becomes stronger, and at some point (when the interaction energy becomes of order kT or more) we can no longer assume weak coupling. That is, the distribution of charges along the chain is no longer entirely random. For this case, rather than separating the Helmholtz energy in an ideal entropy term and an averaged energy term, we must weigh all possible configurations of charges along the chain with their respective interaction energies. A useful approximation is to rewrite the Helmholtz energy F({Sj}) of a molecule with a specified distribution of protonated and deprotonated sites {Sj} in terms of a cluster expansion: F
(<si>) = - Z / ' i s 1 + X V i s J i=l
i<j
+
X Lijksisjsk + -
[2.2.46]
i<j
where the terms represent individual sites, pairs of neighbouring sites (doublets), triplets, etc., respectively. If the dominant contribution to the interaction energy comes from the nearest neighbours, it is reasonable to attempt a truncation after the second term:
^^H-X^i+Iv^ i-l
l2 2 471
--
i<j
This is called a (linear) Ising model11. The parameter fit = 2.303/cT(pK-pH) specifies the chemical potential of the proton, with respect to the reference state of a solution of non-interacting sites with a =0.5. The quantity u{. is the interaction energy between two neighbouring sites: it is zero unless both sites are charged, in which case it is given "After E. Ising, Z. Physlk31 (1925) 253.
POLYELECTROLYTES
2.21
as a first approximation by the screened Coulomb energy of two elementary charges at a distance C: u
= 1J
^^expHrn Ajt££Qt
(1 + ra)
The Ising model was treated in sec 1.3.8a in the context of adsorption of gas molecules (at pressure p) on a line of adsorption sites; there, the adsorption constant was taken to be KL and the interaction energy between two filled neighbouring sites was w^ j . Here, we use it for the desorption of protons from a linear polyacid, but otherwise it is entirely equivalent. Note that in sec. 1.3.8a the treatment was based on the quasichemical (Bethe-Guggenheim) approximation. It is possible to compute the partition function exactly using the method of transfer matrices, but the result is the same as that of the quasi-chemical approach for this one-dimensional case. We can therefore use the solution [1.3.8.4] and just make the substitutions: KL —> K^1
p -> a H+
6 -» a
wx j -> u
This immediately leads to the adsorption isotherm:
pH = pK-0.4343u + log l £ ± i z ^ £ ]
[2.2.49]
with p=[l- Aa(\ - a)(\ - e-ulkT)}112
.
[2.2.50]
Note that when u becomes very small, exp(-u/kT) and fi approach unity, and we recover [2.2.38] At large u, however, the charged sites tend to avoid each other. The result is that the effective pK changes strongly in a narrow window around a = 0.5 , but is changing relatively little outside that window. In other words, the titration occurs in two 'waves' as if the polyelectrolyte behaves like a divalent acid or base. Creating charges in an alternating arrangement on the chain requires no extra energy and this, therefore, happens first; in order to make all sites participate in the second 'wave,' the equilibrium constant has to change substantially. Examples of polymers with the latter kind of behaviour are poly(ethylene imine) (PEI), where the sites are nitrogen atoms (secondary amines) in the backbone chain, separated by two carbon atoms, i.e. only about 0.3 nm, and poly(maleic acid) and poly(fumaric acid), which have carboxyl groups attached to neighbouring carbon atoms. Branched variants of (PEI), like the regularly branched imine dendrimers, show these effects even more dramatically. Figure 2.8 shows data for linear PEI, both as the degree of protonation versus pH, and as effective pK as a function of degree of protonation. The pattern of two waves can be clearly discerned.
2.22
POLYELECTROLYTES
Figure 2.8. Titrations of linear poly (ethylene imine) at two different ionic strengths (indicated) plotted as degree of charging versus pH (a) and as effective pK versus 6 (b). Experimental data are indicated by points. Solid curves are calculated according to the Ising model. The parameters are: for 0.1 M, pK = 8.44, u = 0.0107, and for 1 M, pK = 8.63, u = 0.0178. (Redrawn from R.G. Smits, G.J.M. Koper and M. Mandel, J. Phys. Chem. 97 (1993) 5745.)
Figure 2.9. Titration of polymaleic acid, plotted as degree of protonation ( 1 — 0 ) versus pH, in different background electrolytes: LiCl, NaCl, and TMAC1. Solid curves are fits to a version of the linear Ising model in which chain tacticity is explicitly accounted for. (Redrawn from J. de Groot, G.J.M. Koper, M. Borkovec and J. De Bleijser, Macromolecules 31 (1998) 4182.)
Another example, for poly(maleic)acid, is given in fig. 2.9. Again, two waves can be seen in accordance with the explanations above. In this example, titrations were carried out with different monovalent counterions (lithium, sodium, and tetramethylammonium, TMA). Clearly, the differences are small in the low pH range (where less than 50% of the groups are charged, but they become very pronounced for the second 'wave': the midpoint of that wave for Li+ is almost 2.5 pH units lower than that for TMA+ . The interaction energy between charged sites is thus much larger for TMA+ , which suggests that this voluminous ion is much less strongly bound on charged sites than either Li + or Na + . The series corresponds with the specific binding sequence for alkali ions to carboxyl groups, see for instance sec. III.3.8b. As we saw above, for weak polyelectrolytes the charge density on the chain responds to the local environment, leading to a salt-dependent average charge density. As the ionic strength decreases, the charge density decreases too. This is not the only effect, though. When the size of the coil becomes comparable with the Debye length, its ends experience less interaction from neighbouring charges than middle segments. As
POLYELECTROLYTES
2.23
a result, the charges tend to accumulate at the ends, particularly when the fraction of charged monomers is on the order of a few percent. The charge density at the ends may then become 2-3 times that in the middle. This implies that when a polyelectrolyte solution is titrated, the charge 'grows in' from the ends1 .
Figure 2.10. Monte Carlo simulations of titration on flexible chains. The effective pK is plotted versus degree of protonation (1 - 0) . The chains consisted of 320 subunits connected by fully flexible bonds of 0.6 nm length. The dashed curves are calculated for a rigid chain; the points (connected by solid lines) are taken from Monte Carlo simulations.
So far, it has been assumed that coil swelling effects could be neglected, because the distance between groups separated by a long stretch of chain would be too large for any significant interaction. This turns out to be not entirely true. Monte Carlo simulations have shown beyond any doubt 21 that chain swelling does affect the effective pK. Comparison with experiments, e.g., on poly (D,L glutamic acid), confirmed that taking swelling into account improves the agreement between experiment and theory. Swelling effects are particularly dramatic for polyelectrolytes with a modestly or poorly soluble backbone. A well-known example is poly(methacrylic acid) (PMA). In its neutral state, this polymer has a certain internal cohesion, which prohibits it from forming a dilute coil. As a consequence, the distances between sites are initially almost fixed so that the pK must increase. Quite likely, the distribution of charges is very inhomogeneous at this stage with many more charges sitting at the periphery of the coil than inside. At a sufficiently high charge density ( ai.0.2 ), the electrostatic forces become strong enough to start breaking the intramolecular cohesion, and the molecule reaches a state where its pK is determined by the balance of the electrostatic and cohesive forces. Between a = 0.2 and a = 0.4, the molecule then swells substantially at a virtually constant pK. As soon as the entire molecule has reached the dilute coil state, pK may increase further because of nearest neighbour interactions. The typical behaviour of polymethacrylic acid can be shown either by means of a pKeff(a) plot (fig. 2.11) or as an HH plot (fig. 2.12). The latter plot is not simply linear as for a polyelectrolyte in a good solvent, but it has two linear sections of different slopes, and a non-linear section around a = 0.5 , which connects the two linear parts.
11 21
M. Castclnovo, Eur. Phys. J. El (2000) 115-125. See, for instance, M. Ullner, B. Jonsson, Macromolecules 29 (1996) 6645.
2.24
POLYELECTROLYTES
Figure 2.11. Effective pK as a function of degree of protonation (1 — 0) for polylmethacrylic acid) at two ionic strengths. Experimental data are indicated by points. The solid curve was calculated with a PB model for a charged cylinder (a = 0.55 nm, charge spacing 0.55 nm), but clearly docs not fit the experimental data. (Redrawn from M. Nagasawa, T. Murase, and K. Kondo, J. Phys. Chem. 69 (1965) 4005.)
Figure 2.12 illustrates how HH plots can give a quick view of conformational transitions in the chain. PAA does not exhibit such a transition; the plot is linear over the entire pH range studied. However, with the inclusion of a methyl group in the main chain (PMA) or by partial esterification of carboxyl groups (pe), the transition does occur as evidenced by the breaks. The slopes are measures of the polyelectrolyte effect in terms of n; the higher c s a j t , the lower n is. The compact state of PMA (at a < 0.2 ) also leads to an elevated n. At a = 0.5 , log(a/l - a) = 0 , pH = pK e f f , which can therefore be read from the figure.
Figure 2.12. Hcnderson-Hasselbalch (HH) plots of various polyfacrylic acid) (PAA) and polylmethacrylic acid) (PMA) either, or not, partially esterified (pe). Panel (a), •, PAA-pe + 0.01 M NaCl; x, ibid., without NaCl; A PMA, without NaCl. Panel (b), o,» A PMA-pe + 0.02 M NaCl; • , • , ibid., 0.2 M NaCl. Polyelectrolyte concentration 1000 ppm (courtesy J.T.C. Bohm).
POLYELECTROLYTES
2.25
The most pronounced swelling effect Is, of course, observed whenever the neutral chain is insoluble, so that the charging process becomes coupled to a phase transition. 2.3 Polyelectrolyte configurations
2.3a Dilute solutions No added electrolyte Kuhn et al.'' were the first to show that in the absence of screening by added electrolyte, polyelectrolytes adopt an essentially linear configuration. Consider a long, freely jointed polymer chain of contour length L, which consists of JVK Kuhn segments of length lK . In the Gaussian approximation, the Helmholtz energy of extending the chain to an end-to-end distance r is: fj^sL^J— kT LlK
[2.3.1]
Now, suppose that the chain has a linear charge density of v elementary charges e per unit length, corresponding to a charge parameter of x = iB v. Neglecting the influence of the counterions, the repulsive electrostatic interaction energy of the chains is estimated as Coulomb _ 2 ^ kT (Br
[2 3 21
where ! B is the Bjerrum length. Minimizing the total Helmholtz energy gives the scaling relation r =U2'3(!K/lB)'/3
[2.3.3]
For highly charged polyelectrolytes, it is no longer reasonable to assume that all counterions are dispersed to such an extent that their influence can be neglected. The scaling argument still applies, however, if e is interpreted as the number of 'free' counterions released into the solution per unit length of polyelectrolyte. The concentration range, where the rod-like polyelectrolytes do not overlap, is called the dilute regime. The boundary of the dilute regime is the overlap concentration c* of the rod-like polyelectrolytes, 1 c* =
NK „ ^ " ^ K
[2.3.4]
Note that in this context 'dilute' strictly means non-overlapping, which for salt-free polyelectrolytes does not necessarily mean that they do not interact. For sufficiently high polyelectrolyte molecular weights, the Afj^2 dependence implies that only solutions
11
W. Kuhn, O. Kiinzle, and A. Katchalsky, Helv. Chlm. Ada 31 (1948) 1994.
2.26
POLYELECTROLYTES
of extremely low concentration are dilute if no electrolyte is added. A further practical limit is, of course, that ionic strengths below, say, 10~5 M are difficult to obtain. Nevertheless, careful experiments have been performed in the dilute regime. A general feature of polyelectrolyte solutions without added electrolyte is that they exhibit a peak in their structure factor as measured by light-, X-ray and neutron scattering. The peak position q max is related to a characteristic distance l / q m a x • The fact that a peak is visible is thanks to two effects: 1) the strong electrostatic repulsion gives rise to rather well-defined distances between interacting polyelectrolytes and 2) the strong electrostatic repulsion gives rise to a very low compressibility and, hence, very low scattering at low q. For the dilute regime, the characteristic distance is simply the typical distance between the rod-like polyelectrolytes, q^ax ~ c • Indeed, in the dilute regime the position q max of the peak scales as q max <*= c 1 / 3 . Added electrolyte Eventually, at very high concentrations of added electrolyte, all electrostatic interactions are completely screened. Then, if the polyelectrolyte backbone is both flexible and water-soluble, its behaviour should revert to that of an uncharged, flexible, watersoluble polymer. For dilute solutions, we therefore expect a gradual transition from rod-like configurations in the absence of added electrolyte, to coiled configurations at high concentrations of added electrolyte. Quantifying this transition has been a major aim of many polyelectrolyte theories. Solving this problem has been achieved in part by mapping the problem onto the analogous problem of an uncharged polymer chain. Recall that standard theory for uncharged flexible polymer chains characterizes polymer segments in terms of two parameters: their length IK ("Kuhn length') and their second virial coefficient, or excluded volume [5 (introduced in volume II, sec. 5.2 as u = /?/fj| ). Chain expansion due to excluded volume is often characterized in terms of a linear expansion coefficient a , a or = —=a g,o where a
[2.3.5]
is the radius of gyration of the expanded chain, and a g 0 is the radius of
gyration of the equivalent chain without excluded volume interactions, a g 2 0 =ljV K !2
[2.3.6]
Chain expansion is determined solely by the dimensionless excluded volume parameter z z=
( 3 V / 2 Nl'2p —
- \ ! -
[2.3.7]
The expression for the linear expansion coefficient a that will be used here is the
POLYELECTROLYTES
2.27
Yamakawa-Tanaka11 expression: a2 =0.541 +0.459(1+ 6.04Z)046
[2.3.8]
which correctly goes to 1, as z —> 0 , and does not deviate too much from the correct asymptotic behaviour a °= z 2 / 5 at large z (see also chapter II.5). For polyelectrolytes, both the segment length (K and the excluded volume j3 can be considered as effective parameters that depend on the electrostatic interactions and, hence, on the electrolyte concentration. First consider the effective segment length. At issue is the 'stiffening' of polyelectrolyte chains due to electrostatic interactions. As we have seen, in the absence of added electrolyte, electrostatic stiffening is so strong that the polyelectrolytes have essentially rod-like conformations. Initially it was assumed that in electrolyte solutions, pieces of polyelectrolyte chains up to a Debye length long would be essentially rod-like. At larger length scales, the chain was assumed to be flexible. In other words, the 'effective' segment length was assumed to be proportional to K'X . Then, in 1977, Odijk21, and Skolnick and Fixman31 showed that, in fact, the electrostatic stiffening effect is much stronger than this, and gives rise to an effective segment length proportional to K~2 . Odijk41 (together with Houwaart) and Fixman and Skolnick51 were also the first to work out the theory for the electrostatic contribution to the segment excluded volume P, thus completing the mapping onto the uncharged polymer problem. Below, we first derive the expression for the effective segment length, and then that for the effective segment excluded volume. The theory of electrostatic stiffening was briefly introduced in II.5.14. It relies on describing the polyelectrolyte chain as 'wormlike'61 with persistence length q . Note that the Kuhn segment length equals twice the persistence length, lK = 2q . Consider a wormlike chain bent into a circle of radius R. The elastic energy of this circular wormlike chain is, per unit length :dG3asL= I J L [2.3.9] kT 2R2 For a circular polyelectrolyte, the persistence length q has both intrinsic contributions and contributions due to electrostatic interactions of the charges along the chain: q = qo+qe
11
[2.3.10]
H. Yamakawa, Modern Theory of Polymer Solutions, Harper & Row (1971). T. Odijk. J. Polym. Set 15 (1977) 477. 31 J. Skolnick, M. Fixman. Macromolecules 10 (1977) 944. 41 T. Odijk, A.C. Houwaart. J. Polym. Set Polym. Phys. 16 (1978) 627. 51 M. Fixman, J. Skolnick. Macromolecules 11 (1978) 863. 61 Such a model pictures the chain as an elastic cylinder. The persistence length q is the ratio of the bending elastic constant of the cylinder over the thermal energy kT (see also II.5.2f). 21
2.28
POLYELECTROLYTES
If the electrostatic interaction energy (per unit length) of the circular polyelectrolyte is / e l (R), the corresponding electrostatic contribution to the persistence length can be calculated from
q =_^llk He
R ^
[2.3.11]
kT dR
A serious problem in computing qe is how to deal with the electric double layers for curved polyelectrolytes of high linear charge density. This problem has not been solved satisfactorily. The pragmatic approach of Odijk, Skolnick and Fixman (and most authors after them) has been to model the polyelectrolyte as a line charge in the DebyeHuckel approximation. High linear charge densities can then only be accounted for in a very approximate way, by using an effective charge density. In the Debye-Hiickel approximation, the electrostatic interaction energy of a circular line charge, of charge density v (elementary charges e per unit length) is simply the local electrostatic potential times the local charge density (divided by two to avoid double counting of interactions). Per unit length of polyelectrolyte chain: ^ - =l v \ f d s e X p ( - y r ( S ) ) kT 2 B J r(s)
[2.3.12]
Due to the curvature, the distance r(s) between charges is smaller than the contour length s separating them (see fig. 2.13).
Figure 2.13. Circular line charge. The spatial distance r between charges is shorter than the contour distance s, which gives rise to additional electrostatic repulsion in curved configurations and, hence, to electrostatic stiffening.
This is the physical origin of the stiffening effect: curvature brings the charges closer to each other, which increases the electrostatic repulsion. The derivative of the electrostatic energy with respect to curvature, is
The derivative of the distance r(s) with respect to curvature R can be evaluated using r(s) = s ( l - ^ ( s / R ) 2 + ...)
[2.3.14]
POLYELECTROLYTES
2.29
In the other terms we may now set r{s) = s . Also, in the limit R —> °=, the integration limits may be extended to infinity. This finally gives v2in r n qe= 2_jdsexp(-x-s)(s + x-s2)
(2.3.15]
o In terms of the dimensionless charge density x = iB v, the result is q e
= ^ -
12.3.16,
As mentioned, an approximate expression valid for arbitrary charge densities is obtained by substituting an effective charge density xeff for x > 1,
%-^kr
[2.3.17]
A crude, but commonly used approximation, is to set xeff = 1 for x > 1 . The treatments of Odijk and Skolnick and Fixman (OSF for short) are much more complete than the derivation presented here, and properly take into account the fluctuating geometry of the polyelectrolyte chain. Various authors have tested the validity of [2.3.17] against numerical solutions of the full Poisson-Boltzmann equation for curved, highly charged cylinders11. For a highly charged cylinder of radius a, it turns out that [2.3.17], with xeff = 1, as compared with numerical solutions of the exact PB equation, is reasonably accurate provided KCL «1. At higher concentrations of added electrolyte, for tea = O(l) or larger, there are substantial deviations. Strictly speaking, the derivation of the expression for the electrostatic persistence length requires that the radius of curvature of the polyelectrolyte chain must always be large with respect to the Debye length K~1 . Provided the intrinsic persistence length is large enough, q 0 > x""1, i.e. for intrinsically stiff polyelectrolytes, this will certainly be the case. Furthermore, for these polyelectrolytes, excluded volume interactions are often negligible, except at extremely high molecular weights. Therefore, the above theory for electrostatic stiffening can immediately be applied to experimental data on intrinsically stiff polyelectrolytes. Indeed, shortly after the publication of the papers of Odijk and Skolnick and Fixman, it was demonstrated21 using magnetic birefringence that [2.3.17] with xeff = 1 almost quantitatively accounts for the observed electrostatic stiffening of long DNA molecules at low concentrations of added electrolyte (DNA has q0 ~ 50 nm). Its validity for intrinsically stiff polyelectrolytes, such as DNA, has since then been confirmed in many other experiments (see also fig. II.5.5) For intrinsically flexible polyelectrolytes, there are two further problems; first, the
11 21
M. Le Bret, J. Chem. Phys. 76 (1982) 6243, M. Fixman, J. Chem. Phys. 76 (1982) 6346. G. Maret, G. Wcill, Biopolymers 22 (1983) 2727.
2.30
POLYELECTROLYTES
validity of the K~2 scaling for intrinsically flexible polyelectrolytes ( q0 < K^1 ) and second, the problem of the electrostatic contribution to the excluded volume. Regarding the former problem, while some authors have reasoned that the tc~2 scaling is altogether invalid for intrinsically flexible polyelectrolytes, most authors argued that for flexible polyelectrolytes it remains valid, but only at length scales larger than a Debye length. Indeed, more recent detailed calculations have shown11 that at length scales larger than a Debye length, the K~2 scaling remains valid, even for flexible polyelectrolytes. The calculations also show that electrostatic stiffening at length scales below a Debye length is much less pronounced, which implies increased fluctuations at these short length scales. These can be corrected for by introducing an effective contour length, which is somewhat smaller than the full contour length of the polyelectrolyte chain due to fluctuations at short length scales. Next, consider the electrostatic contribution to the segment excluded volume2'31. Under conditions of strong electrostatic swelling of polyelectrolyte coils, the electrostatic contribution to the excluded volume completely overwhelms the non-electrostatic contributions; hence we can concentrate on the former. The effective segment length is
[2-3.18]
with qe given by [2.3.17] The simplest approach is that of Odijk who approximates the segments by rigid cylinders of length lK and effective diameter dgff . The effective diameter must account for the influence of the electrostatic repulsion of the segments. Assuming that a typical distance of closest approach is the distance at which the repulsive interaction energy equals the thermal energy kT, it was estimated that d e f f =2x- 1
[2.3.19]
The segment excluded volume can then be estimated using the well-known expression for the excluded volume of long slender cylinders,
£ = j£ d eff
[2.3.20]
Skolnick and Fixman instead use the expression [2.2.33] for the interaction energy V(h,y) of charged cylinders, and evaluate the full cluster integral for the excluded volume of two line charges of length lK :
P = jdr(l-exp{-V[h,y))
[2.3.21]
where r is the difference between the centre-of-mass coordinates of the rods, and the
11
H. Li, T.A. Witten. Macromolecules 28 (1995) 5921. Odijk and Houwaart, (1978) (loc. cit). 31 Fixman and Skolnick, (1978) (loc. cit.) 21
POLYELECTROLYTES
2.31
angle denotes an average over the internal coordinates of the rods (i.e. their orientations). It is convenient to define an effective diameter such that the excluded volume equals that of long slender cylinders, according to [2.3.20] An asymptotic expansion of the cluster integral results in the simple approximation deff =K- 1 (-lnvI B +21n£ eff +2.61)
[2.3.22]
This expression will be used below in the comparison with experimental data. Briefly, the procedure to calculate the polyelectrolyte radius of gyration is to first calculate the effective segment length from [2.3.17] and [2.3.18] For a polyelectrolyte of contour length L, the number of effective segments is then given by JVK = L/lK . The radius of gyration a 0 (unperturbed by excluded volume interactions) can now be calculated from [2.3.6] After calculating the electrostatic excluded volume from [2.3.20] and [2.3.22], the expansion factor a is calculated from [2.3.7] and [2.3.8] Finally, the radius of gyration a is calculated from [2.3.5] To illustrate the application of this theory, consider the data of De Nooy et al.1' on the flexible polysaccharide pullulan. By itself this polysaccharide is uncharged, but it can be given a variable linear charge density (up to about le/nm) by oxidation. As shown in fig. 2.14, the radius of gyration varies by only a factor of two when varying the salt concentration over two orders of magnitude. Adjustable parameters in comparing the experimental data with the theory are the intrinsic persistence length q0 of the polysaccharide, and a small contribution d0 to the effective diameter to account for steric repulsion at high salt concentration. As shown in fig. 2.14, the theory consistently fits the data for very reasonable values of the parameters. A scaling argument2 explains why the dependence of the coil size on the ionic strength is, in fact, rather weak. The dependencies of the electrostatic contributions to the Kuhn length and the excluded volume are:
Figure 2.14. Radius of gyration of oxidized pullulan (contour length L = 1100 nm) as a function of electrolyte (NaCl) concentration. Linear charge densities: triangles, x = 0.94 : circles, x = 0.73 ; squares, x = 0.49 . Lines: theory as described in the text, with q0 = 3 nm and d 0 = 0.3 nm. (Redrawn from de Nooy et al., loc. cit.)
1!
A.E.J. de Nooy, A.C. Bcscmcr, H. van Bekkum, J.A.P.P. van Dijk, and J.A.M. Smit. Macromolecules 29 (1996) 6541. 21 T. Odijk, Macromolecules 12 (1979) 688.
2.32
POLYELECTROLYTES
lK~K~2
[3.2.23]
deff-x- 1
13.2.24]
P~lldeK~K~5
[3.2.25]
In the limit of large excluded volume (excluded volume parameter z > > 1), a g - L3>5p'%l/S
~ x- 3 / 5 ~ ns-°-3
[3.2.26]
The terms involving the excluded volume and the Kuhn length exhibit opposite dependencies on ionic strength, causing a weak overall dependence on ionic strength. Whereas for chains without excluded volume the radius of gyration increases with increasing segment length (K (at constant contour length L), the dominant effect for chains with strong excluded volume is a decrease of excluded volume interactions, with increasing lK (see [2.3.7] and [2.3.8]). The scaling dependence ai ~ n~0& is close to the typical experimental observation that a2 ~ nj° 5 . Finally, note that Monte Carlo computer simulations11 have only recently been able to reach system sizes, required to critically test the OSF theory of polyelectrolyte coil sizes using simple bead-string polymer models, having effective Debye-Hiickel interactions between beads. These simulations confirm that the mapping to the uncharged polymer problem gives a good first approximation for the coil size, and that fluctuations at length scales below a Debye length do not significantly change the K~2 scaling of the electrostatic contribution to the persistence length. 2.3b Semidilute solutions No added electrolyte Salt-free polyelectrolytes remain one of the least understood systems in polymer physics, and this holds especially for semidilute (salt-free) polyelectrolytes. From a theoretical point of view, the main problems are (i) the long-range character of the electrostatic interactions in the absence of screening electrolyte and (ii) the existence of multiple, concentration-dependent length scales. This leads to a large number of distinct regimes, as we explain below. Here we only sketch some of the features that are more or less accepted, and do not attempt to compare theory with experimental data. Note that in the present context 'salt-free' is always defined with respect to the concentration of ions released by the polyelectrolytes. This is especially important for the more dense semidilute solutions: in a polyelectrolyte solution with a monomer concentration of, say, 0.1 M, the effects of adding small amounts of electrolyte (appreciably less than the monomer concentration) are negligible. The sequence of events upon increasing the polyelectrolyte concentration in a salt11
R. Everaers, A. Milchev, V. Yamakaov. Eur. Phys. J. E8 (2002) 3; T.T. Nguyen, B.I. Shklovskii, Phys. Rev. E 66 (2002) 021801.
POLYELECTROLYTES
2.33
free solution is believed to be the following. In the dilute regime, the polyelectrolyte configurations are essentially rod-like. As will be argued below, they remain rod-like as the rods start overlapping. If they would remain globally rod-like even at still higher concentrations, we would expect, at some critical concentration, a transition to an orientationally ordered, nematic phase. No experimental evidence has been found, however, for orientational order in salt-free, semidilute, polyelectrolyte solutions (with flexible backbones). Apparently, at concentrations where anisotropic phases would be expected, the increased concentration of counterions has already screened the electrostatic interactions to such an extent that the molecules have become too flexible to form an anisotropic phase. For concentrations above the overlap concentration, scaling arguments indicate that there are multiple semidilute regimes. These are illustrated in fig. 2.15. (b, c and d). For the purpose of the scaling arguments below, consider a highly charged polyelectrolyte of contour length L, with 'monomers' of length (B and volume (g , each of which carrying one elementary charge e. We work with the volume fraction of monomers rather than with the molar concentration c. By definition, the density of counterions equals the density of monomers. Assuming that all counterions contribute to the screening of the electrostatic interactions, the relevant screening length A"1 follows from X2 = (pr2
[2.3.27]
For a rough estimate of the effective segment length lK , we use the OSF expression [2.3.17], with xeff = 1, and with the Debye length /r"1 replaced by the screening length
Z K =-S
[2.3.28]
A s s u m i n g rod-like configurations, t h e semidilute regime s t a r t s at a volume fraction
At the overlap volume fraction
2.34
POLYELECTROLYTES
Figure 2.15. Solution regimes for salt-free polyelectrolytes. a) dilute rods, c < c 1 ; b) semidilute rods, c < c ' < c " ; c) semidilute and semiflexible: c"
[2.3.29]
Beyond this volume fraction, global configurations become coil-like, but the local configuration is still rod-like, as illustrated in fig. 2.15c Further increasing the monomer volume fraction decreases both the segment length and the correlation length, but the former decreases faster. At a further critical monomer volume fraction
Beyond this volume fraction, the screening by counterions is nearly complete and the behaviour of the semidilute solution becomes increasingly similar to that of uncharged flexible polymers (fig. 2.15d). Finally, consider the scaling of the position of the scattering maximum, as observed in light-, X-ray and neutron scattering experiments. As mentioned, in the dilute regime
1)
This is the characteristic length scale for density fluctuations (see also II.5.2d).
POLYELECTROLYTES
qmax~
2.35
[2.3.30]
For semidllute solutions with q>* < q><
[2.3.31]
as is indeed observed experimentally. For cp> q>** , the electrostatic interactions are screened to such an extent that the scattering maximum is no longer visible. Ultimately, in this regime, the scaling of the correlation length should tend to that for an uncharged flexible polymer (see eq. II.5.2.18), i.e. £~
[2.3.32]
Added electrolyte For the case of added electrolyte, the mapping to the uncharged polymer problem can again be used to develop approximations for the semidilute regime. In the discussion of polyelectrolytes below, following Odijk ' ', we use scaling theory for semidilute solutions of uncharged flexible polymers 21 . Some aspects of the scaling predictions are compared with the quasi-elastic light scattering data of Koene and Mandel for the highly charged, flexible polyelectrolyte sodium poly(styrene sulphonate) (NaPSS). As mentioned, for a highly charged, intrinsically flexible polyelectrolyte, the effective segment length and the segment excluded volume have the following dependencies on the Debye length: l
K~K~2
P=
K-h\~K-
[2.3.33] 5
For flexible, uncharged polymers in a good solvent, the monomer volume fraction at which coils start overlapping is (p*
^ ^ - 4 / 5 ^ - 3 / 5 ,-2/5 a
[2.3.34]
g
Substituting the expressions for the effective quantities for polyelectrolytes, it is found that
11 21
[2.3.35]
T. Odijk, Macromolecules 12 (1979) 688. P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell University Press, (1971).
POLYELECTROLYTES
2.36
Figure 2.16. Overlap concentration c* for NaPSS with M w = 6.5xl0 5 g/mol, estimated from the quasi-elastic light scattering data of Koene and Mandel, versus the concentration of added electrolyte (NaCl). The full line is the scaling prediction c* ~ c ~ 0 9 .
As shown in fig. 2.16, this compares favourably with the salt-dependence of the overlap concentration as deduced from the data of Koene and Mandel for the highly charged and intrinsically flexible polyelectrolyte NaPSS: the overlap concentration varies by about one order of magnitude when varying the salt concentration by one order of magnitude. The scaling prediction for the correlation length £, (the typical mesh size of semidilute solutions, see chapter II.5) of uncharged flexible polymers is: ^ag{
[2.3.36]
Again, substituting the expressions of the effective quantities for highly charged, intrinsically flexible polyelectrolytes, one finds,
£~x- 3 / 4 ~c°- 3 7 5
[2.3.37]
Note that the radius of gyration a g (in the dilute regime) and the correlation length
and the concentration factor {(pi (p*T^^ . Although
the radius of gyration decreases with increasing salt concentration, this decrease is overwhelmed by a stronger increase of the concentration factor (which essentially measures the distance to the dilute regime).
Figure 2.17. Effective diffusion constant of dilute (open symbols, less than 0.25 gf ) and semidilute (10g/( , closed symbols) NaPSS as a function of concentration of added electrolyte. Full lines are only guides to the eye.
2.37
POLYELECTROLYTES
The opposite dependencies of the radius of gyration and the correlation length on the salt concentration are nicely illustrated by the quasi-elastic light scattering data of Koene and Mandel for NaPSS, shown in fig. 2.17. This technique measures an effective diffusion constant Deff that is essentially inversely proportion to the radius of gyration in the dilute regime, and inversely proportional to the correlation length in the semidilute regime. 2.3c Grafted polyelectrolytes A brush is a dense layer of end-attached polymers on a surface. Brushes have rather special properties due to the fact that they can only swell in one direction, namely normal to the surface. A sketch of a polyelectrolyte brush is presented in fig. 2.18.
Figure 2.18. A polyelec-
trolyte brush.
Brushes consisting of uncharged chains were already discussed in chapter 1; here, we briefly recall the main results. The average extension of a grafted chain results from a compromise between the osmotic swelling pressure and the elastic force associated with the chain entropy. This can be cast into a simple scaling argument. Consider a flexible (Gaussian) polymer chain consisting of AT segments in a brush having a chains per unit area. If that chain is stretched to an extension L with L> rQ = NC2 (where r0 is the unperturbed end-to-end distance), the force /
between its ends equals
J = kT-^j and the associated pressure to keep the ends at this distance is Jo.
[2.3.38] This must be
balanced with the osmotic pressure 17 in the brush given by 77 = kT(fl/ 2)c2 , where c is the average segment (number) concentration equal to c ={<jN)/L and /? is the excluded volume of a segment: /3 = (1 - 2^)f3 = vC3 . Balancing these pressures immediately shows that the brush extension L scales as L/C = N{af:2v)i/3
[2.3.39]
2.38
POLYELECTROLYTES
The same scaling was found in chapter 1 (see [1.11.19]; (there is a prefactor 6~' /3 which has been omitted here). For the charged brush (with a fraction a of the segments carrying a charge) the osmotic pressure in the brush is different due to the charges. As is well known, a Donnan equilibrium exists between a solution of macro-ions and a pure salt solution. Since the brush is rather dense and since its thickness exceeds the Debye length, it can be well described as a macroscopic phase with an internal potential. For a strong polyelectrolyte, this suggests that there are several relevant cases. (1) At low salt concentrations, typically well below the concentration ac of charged monomers in the brush, the salt is almost entirely excluded from the interior of the brush, and the osmotic pressure difference is approximately given by the concentration of the counterions in the brush: 77 = kTac . This situation is called the osmotic brush. Balancing the osmotic pressure with the stretching force as given in [2.3.37], one gets the simple result: L/( = Na1/2
[2.3.40]
Note that there is no dependence on the grafting density a in this case! (2) At higher salt concentrations, the Donnan effect is that of a solution of screened cylinders, the radius of which is given by the Debye length tc~l (pertaining to the salt concentration outside the brush). This case, referred to as the salted brush, has an osmotic pressure n = kTlac\
IK-2 = kTc2 — —
[2.3.41]
for monovalent salt. This gives a brush height very similar to that of the neutral brush: L/£ = N(veo-e2f/3
[2.3.42]
but with an effective electrostatic excluded volume
V =
[2 3 431
* WT 1
l
--
BCs
(3) Finally, when the charges in the brush are completely screened (at ve « 1), the brush is expected to behave as a neutral brush with a constant 'bare' excluded volume
P=v(3 . All of this suggests a more general expression for the osmotic pressure covering all cases (1) - (3): n = kTclv.fAc) = kTcl\—
P
P[c s +orc p
+ v\
[2.3.44]
J
For a weak polyelectrolyte brush (annealed charge) the behaviour is more complicated
POLYELECTROLYTES
2.39
as the Donnan effect also acts on the protons, thereby affecting the degree of charging a . Qualitatively, the Donnan potential y/D is high at low salt concentrations; counterions are then strongly attracted and co-ions repelled, which implies that the effective
pK of the groups on the polyelectrolyte is shifted
by an amount
~ y D = {eif/D)/kT in the direction away from neutrality. At fixed pH, this implies that the fraction of charged sites decreases with increasing \yD\. Since y D increases with ac , but is also fixed by the pH, this means that dense brushes at low ionic strength have lower charge densities than dilute brushes or free chains. Added salt now has two effects: (i) it lowers y D and, hence, increases the charge density and the osmotic pressure inside the brush and (ii) it screens the charges. The net effect is that the brush height varies non-monotonically with ionic strength I: at low I (in the osmotic regime), it first shows an anomalous swelling with increasing I, and at higher I (in the salted regime) the brush shrinks. The anomalous swelling at low ionic strength can be understood more quantitatively as follows. The osmotic pressure at low I is dominated by the counterions for which we can write, for monovalent salt: 1/9
77 = kTc s exp(y D ) = A:T(e2yDj
[2.3.45]
The Donnan potential can be estimated from the electroneutrality condition: acp = 2c s sinhy D = cseyD
[2.3.46]
where a is now the fraction of dissociated monomers in the brush. This fraction responds to the external pH as follows: fy — u. —
i + gey
D
~ ~
gey
ro o A 71 [z..o.ti j
D
In this latter expression, the factor Q parametrizes the degree of dissociation in the bulk: g> = 1O(PH-PK| =
§-
[2 3 48]
«B
Eliminating a , one finds that e 2 y /7 =
fc c V / 2
= c / Qc& , and thus an osmotic pressure fcrU-^
[2.3.49]
Note that it is the geometric mean of the salt and monomer concentrations that gives the osmotic pressure here. Balancing again with the stretching pressure finally yields
2.40
POLYELECTROLYTES
K0 I / 3 L/f = JVM-
[2.3.50]
Thus, the brush height increases with increasing ionic strength and decreases with increasing grafting density. This latter effect is just the opposite of what happens with the neutral brush. Should 'bare' excluded volume be non-negligible, one might use: 77 = kTPcl
°KS „ + v\
[2.3.51]
Figure 2.19. (a) Thickness (in nm) of a layer of grafted polylacrylic acid) chains as a function of ionic strength / (on a logarithmic scale) for three pH values. The grafting density is 0.26 nm" 2 and the length of the chains is 368 monomer units, (b) Sketches of the structure of the polyelectrolyte brush in the various regimes of ionic strength, (c) Thickness of a weak polyacid brush as a function of ionic strength, calculated numerically for various grafting densities, as indicated; parameters: pH = pK + 2, JV = 20, v = 0 (courtesy of P.M. Bicsheuvel).
POLYELECTROLYTES
2.41
Figure 2.19a shows an experimental example of a weak polyelectrolyte brush, namely PAA grafted on a polystyrene film and immersed in a sodium chloride solution". Figure 19c gives a theoretical example. The height of the brush was measured as a function of salt concentration. The anomalous swelling at low salt concentration followed by a collapse at high concentration is clearly seen. 2.3d Polyelectrolyte gels By cross-linking polyelectrolyte solutions, one obtains polyelectrolyte gels or networks. The osmotic pressure of the small ions confined to the gel volume to preserve charge neutrality and give rise to the distinguishing feature of polyelectrolyte gels, their enormous swelling capacity. Not surprisingly, therefore, one of their most important applications is as a superabsorbent material. A further unique feature of polyelectrolyte gels is that they frequently exhibit first-order swelling transitions, in which the swelling of the networks changes in a jump-like fashion upon changing the solvent quality or the degree of ionization of the polyelectrolyte. The interest in polyelectrolyte networks dates back to Flory21 and Katchalsky31. Firstorder swelling transitions for polymer gels were first predicted theoretically by Dusek and Patterson41. Experimentally, they were first investigated for polyelectrolyte gels by Tanaka51 who also later discovered systems with multiple swelling transitions61 Below we give the simplest theory that captures the essential ingredients, and gives a qualitatively correct description of the swelling behaviour of polyelectrolyte gels, including the swelling transitions. After that we briefly discuss the current status of the theory of polyelectrolyte gels. We consider a polyelectrolyte gel of volume V' immersed in solvent bath of volume V » V'. The solvent is a monovalent electrolyte solution of concentration cs . Inside the polyelectrolyte gel, the electrolyte concentration is c's . The number of salt ions inside the gel volume is denoted by N's = c'sNAvV , the number of salt ions outside the gel is Ns = c s iV Av (V-V) . The polyelectrolyte gel consists of JV segments, each of which releases v counterions into the gel volume. The monomer concentration in the gel volume is c m = N / NAvV . The branch points of the gel have functionality / (number of polymer 'arms' extending from each branch point), and the number of polyelectrolyte segments between branch points is m . The so-called Flory-Rehner approximation (loc. cit.) consists in assuming that the canonical partition function can be factorized into three independent contributions. 11
E.P.K Currie, A.B. Sieval, G.J. Fleer, and M.A. Cohen Stuart, Langmuir 16 (2000) 8324. P.J. Flory, J. Rehner, J. Chem. Phys. 11 (1943) 521; P.J. Flory, Principles of Polymer Chemistry, Cornell University Press, (1953). 31 A. Katchalsky, J. Polym. Sci. 7 (1950) 393; A. Katchalsky, I. Michaeli, J. Polym. Sci. 15 (1955) 69. 41 K. Dusek, D. Patterson, J. Polym. Sci. A 2 6 (1968) 1209. 51 T. Tanaka, Phys. Rev. Lett. 40 (1978) 820. 61 M. Annaka, T. Tanaka, Nature 355 (1992) 430. 21
2.42
POLYELECTROLYTES
Then, the Helmholtz energy of the gel sample can be written as F
= F mix + F elast + Fions
I2-3-52!
where F mix is the contribution from polymer/solvent mixing, F elast is the elastic contribution of deforming the network with respect to some reference state, and F l o n s is the contribution due to the presence of mobile and bound ions. This factorization is the basis of the classic theory of polyelectrolyte gels ' ' . One of the issues of current theoretical interest is the extent to which this decoupling is a good approximation, and how to include corrections in case of its failure. This work has not yet led to a definitive conclusion, and here we proceed assuming that the factorization is valid. Historically, the mixing contribution is approximated using Flory-Huggins theory. For polyelectrolyte gels, which are usually quite dilute, we can also use a virial expansion. The advantage of the virial expansion over the Flory-Huggins theory is that the virial coefficients have a direct molecular interpretation in terms of configurational integrals (see sec. 1.3.9c), whereas the Flory % parameters do not. Following Khokhlov et al. , we therefore use a third order virial expansion: F
mix kT
= JV(Bc
+Cc2)
[2 3 53]
mm
where B and C are the second and third virial coefficients of the polyelectrolyte segments, respectively. The inclusion of terms up to the third order is neccesary to be able to account for the collapse of the network that occurs when the second virial coefficient turns negative. The elastic energy is approximated using the classic expression from the theory of rubber elasticity (see e.g. Flory and Rehner, loc. cit.):
(
9
9
9
\
at: + afr +af o where ax , a and az are deformation ratios of the network along the x, y and z directions with respect to a reference state. We consider free swelling and use ax= a = az= a. The use of this expression, valid for Gaussian chains, is, of course, questionable for highly extended polyelectrolyte chains, but it turns out that the qualitative results on gel swelling and volume phase transitions are insensitive to the specific expression used for the elastic energy. The choice of the reference state has been an issue of debate, but again, it turns out that the qualitative results are rather insensitive to this choice. Following Khokhlov et al. and de Gennes31, we assume here
11 J. Ricka, T. Tanaka, Macromolecules 17 (1984) 2916: H.H. Hooper, J.P. Baker, H.W. Blanch, and J.M. Prausnitz, Macromolecules 23 (1990) 1096. 21 A.R. Khokhlov, S.G. Starodubzev, and V. Vasilevskaya, Adv. Polym. Sci. 109 (1993) 123. 31 P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell University Press (1979).
POLYELECTROLYTES
2.43
that the reference state is close to the state of network preparation. The latter is approximated by a collection of Gaussian coils of length m that just start to overlap (the so-called c * hypothesis). This gives a reference monomer concentration of C
m.O
rn
_
1
[2 3 551
where a is the segment length. Next consider the polyelectrolyte contributions that are absent for ordinary polymer gels. The dominant contribution, which is the only one that we consider here, is the entropy of the small ions: -!5Si. = (N's + vN)ln(c's +cm) + N's Ync's
[2.3.56]
All other electrostatic contributions are neglected. This also implies that the virial coefficients in the mixing contribution only include non-electrostatic contributions. For the outside solution, the only relevant contribution to the Helmholtz energy is the entropy of the salt ions, pout
= 2i\Llnc, s
kT
[2.3.57]
s
The equation describing the Donnan equilibrium of the salt ions is the equality of the chemical potential of the salt ions in- and outside the gel volume: dF 3F out _Zf_ = ££
[2.3.58]
This gives 2 \l/2
fL= 1 +
fas-l
UcJJ
--^
[2.3.59]
2cs
At low external ionic strength, c s « vcm , almost all of the salt ions are expelled from the gel volume, whereas at high external ionic strength, c s » vcm , the concentrations of salt ions in- and outside the gel volume are nearly equal. In the latter limit, the ionic contribution to the swelling pressure becomes very small, and the polyelectrolyte network behaves similar to a neutral polymer network. At the swelling equilibrium, the total swelling pressure vanishes: '
- =0
[2.3.60]
da Even in this simple theory, this equation cannot be solved analytically. Numerical solutions are shown in fig. 2.20 for a polyelectrolyte with a backbone that is only
2.44
POLYELECTROLYTES
marginally soluble in the aqueous solvent (slightly negative value of B). First consider the limiting case of no salt in the external solution, c s = 0. Upon increasing the number of counterions v released into the gel volume, a discontinuous volume transition is found, in qualitative agreement with experiments. Then, increasing the salt concentration, the transition becomes less and less sharp, and finally disappears, again in qualitative agreement with experiments. In most experiments, the control parameter is taken to be the temperature, which basically tunes the second virial coefficient, B. At a certain temperature, corresponding Figure 2.20. Volume phase transition of polyelectrolyte gels. Deformation parameter a as a function of the number of counterions per polyelectrolyte segment v released into the gel volume. Parameter values: / = 4 , m = 100, Bcm0 m = 3 , Cc^Q m = l . Salt concentrations increase from top to bottom. For a polyelectrolyte segment length of a = 1 nm , the estimated reference monomer concentration is c m 0 = 0.17 M. For that case, from top to bottom: c s = 0 M , 1.7xlO"5M and 6.8xlO" 5 M.
Figure 2.21. Volume phase transition of mixed sodium acrylate/N-isopropylacrylamidc gels in the absence of added electrolyte. Total monomer concentration is 700mM; the concentration of ionic monomer sodium acrylate is indicated. The gels have been cross-linked with 8.6 mM JV.iVmethylenebisacrylamide. Plotted is the temperature versus the degree of swelling VI Vo , where V is the actual volume of the gel and VQ is its original volume. The full curves arc fits to an extended version of the classic theory of polymer gels, as discussed in this section. (Taken from Hirotsu et al. (loc. clt.).)
POLYELECTROLYTES
2.45
to a slightly negative value of B, there is a volume transition that may be either continuous or discontinuous. An example due to Hirotsu et al.11 is given in fig. 2.21, which shows volume phase transitions for polyelectrolyte gels of varying v in the absence of added electrolyte as a function of the temperature. The experimental data were compared with an extension of the theory as discussed here, which correctly predicted that upon increasing the number of counterions released by the polyelectrolyte segments, the transition changes from continuous to discontinuous. Clearly, a number of very drastic approximations have been made in deriving the theoretical results. Recent analytical work21 and computer simulations31 show that there can indeed be no strict decoupling of the elastic and ionic contributions in the partition function since the electrostatic interactions affect the polyelectrolyte configurations and conversely. However, it is not yet clear what consequences the breakdown of this coupling has for the dependence of measurable quantities, such as gel swelling and elastic moduli, on the degree of ionization and the ionic strength41. A further issue is that cross-links or branch points tend to be distributed rather inhomogeneously throughout the network volume in a manner that is highly sensitive to the conditions of network preparation. The consequences of this quenched disorder are being investigated theoretically, especially for neutral polymer gels51. A consequence of quenched disorder, specific to polyelectrolyte gels is that it may affect the degree of counterion binding61. 2.4 Polyelectrolyte viscosity 2.4a General features The viscosity of polyelectrolyte solutions contains a number of interesting trends as a function of the molecular mass M , the polyelectrolyte concentration c , the concentration of added low M electrolyte cs , if any, the charge on the chain, the extent of its screening and the nature of the polyelectrolyte and counterion. Generally speaking, there is no theory covering all these phenomena, but models have been elaborated for limiting situations. Lacking such general theory, it is expedient to consider well-defined experiments to assist in establishing trends. By 'well-defined' it is meant that the polyelectrolyte should be homodisperse if the effects of M , c , cs , line charge, etc. are to be investigated. These parameters should be varied over a sufficiently long range because polyelectrolyte and salt both contribute to the screening. The influence of c can most appropriately be studied by diluting the polyelectrolyte with increasing con 11
S. Hirotsu, Y. Hirokawa, T. Tanaka, J. Chem. Phys. 87 (1987) 1392. T.A. Vilgis, J. Wilder, Comp. Theor. Polym. Set 8 (1998) 61; J. Wilder, T.A. Vilgis, Phys. Rev. E57, (1998) 6865; H. Furusawa, R. Hayakawa, Phys. Rev. E58 (1998) 6145. 31 S. Schneider, P. Linse, Eur. Phys. J. 8 (2002) 457. 41 F. Schosseler, F. Ilmain, and S.J. Candau, Macromolecules 24 (1991) 225; R. Skouri, F. Schosseler, J.P. Munch, and S.J. Candau, Macromolecules 28 (1995) 197. 51 S. Panyukov, Y. Rabin, Y. Phys. Rep. 269 (1996) 1. 61 K.B. Zeldovich, A.R. Khokhlov, Macromolecules 32 (1999) 3488. 2)
2.46
POLYELECTROLYTES
centrations c s in such a way as to keep the ionic strength constant11. The greatest theoretical problem is accounting for the combined influences of electrostatic and hydrodynamic interactions. These effects are not additive since external mechanical forces can change the conformation of polyelectrolytes (sec. 2.3.) and the conformation determines the hydrodynamics. Obviously, electrolytes play a crucial additional role. Only under limiting conditions does the issue become simpler. For instance, weakly charged, dilute, polyelectrolytes behave as swollen coils (Einstein region with high volume fraction (p), whereas strongly charged polyelectrolytes in the absence of added electrolyte are extended, also if subjected to moderate shear forces. An ensuing problem is that of identifying the overlap concentration c* separating the dilute and semidilute regimes. Unlike the situation for uncharged polymers where this parameter can be simply related to M and the radius of gyration a and unlike &
polyelectrolytes at rest (fig. 2.15, [2.3.29 and 34]) 'dilute' now means that the polyelectrolyte molecules do not interact hydrodynamically. Generally, hydrodynamic interactions between particles have a much longer range than the electrostatic ones. So, for viscosity interpretations, c* may turn out lower than its static counterpart, the difference again being dependent on the line charge ve and c s . All told, rheological measurements do contain much information, but it is not yet possible to offer a general theory to rationalize the date. Accordingly, we shall in this section approach the issue empirically. To keep the scope within reasonable bounds, we shall limit the discussion to Newton fluids in the dilute and semidilute ranges. Under these conditions the measurements are straightforward (sees. IV 6.6 and 6.7), provided the system is well-defined. To describe the concentration dependence of the viscosity r]{c ) in table IV.6.1, some viscosity-quantities were defined, of which we shall mainly use the reduced viscosity2) niac=!LJ!sL
[2-4.1]
and the intrinsic viscosity li] =clim >W
[2-4.2]
In the mentioned table impenetrable particles were also considered, in which case the concentration can be well represented by the volume fraction q>. However, for polyelectrolytes, which can swell or compress depending on pH, cs and other variables, the (hydrodynamic) (p is not easy to define. Therefore, c is the preferred variable in [2.4.1 and 2]. If c is given in mol dm" 3 , rjiac and [rj] have the dimensions of a hydrodynamic molar volume, dm 3 mol"1 .
11 D.T.F. Pals, J.J. Hermans, J. Polym. Set 5 (1950) 773; Rec. Trav. Chim. Pays Bas 71 (1952) 458. These are IUPAC recommendations. Other authors sometimes use different names.
2.47
POLYELECTROLYTES
Figure 2.22 gives a sketch of often observed dependencies of the viscosity increment caused by polyelectrolytes. To the left side, the dilute regime prevails, c < c* and the relationship is linear. In this range, extrapolation to c = 0 is possible to yield [77]. At larger concentrations the polyelectrolyte itself starts to screen the charges on the other molecules, which gives rise to a reduction. The initial slope, the position of the maximum and the rate of the decrease beyond the maximum will depend strongly on cs and M (and, for heterodisperse polyelectrolytes, on its M- distribution), on the line charge ve and the countercharge distribution, in particular on the mobile fraction of the countercharge / , see [2.2.2-4]. Recall that highly charged polyelectrolytes [a = 1) at low c s are rod-like, whereas those at low a and/or high cs are coil-like; the viscosity differs substantially between these conformations. At very high c an increase is sometimes observed, probably resulting from alternative rheological behaviour (reptation ...), which is not necessarily entirely viscous. These phenomena will not be discussed here.
Figure 2.22 Schematic representation of the concentration dependence of the viscosity of polyelectrolyte solutions. Discussion in the text.
2.4b Lateral information Accepting that polyelectrolytes have their own idiosyncrasies, one can in some cases compare observations with those for more simple systems to which the behaviour has to reduce under limiting conditions. We call this 'lateral information;' it includes the behaviour of - uncharged polymers (limit for low v and/of high cs ) - low M electrolytes (limit for degree of polymerization—> 1) - particles (limit for impenetrable spheres of rods) Much of this information can be found in previous volumes of FICS, from which we shall now repeat some aspects that may turn out useful. For the range just above c* , particle sols and uncharged polymers, the Huggins equation11 sometimes works well. From [IV.6.11.9], we may write ' W = ['7] + *H['/] 2 c p +o(c p f 11
+
...
[2.4.3]
After M.L. Huggins, J. Phys. Chem. 46 (1942) 151; Ann. NY. Acad. Sci. 43 (1942) 1; J. Am. Chem. Soc. 64 (1942) 1712.
2.48
POLYELECTROLYTES
where fcH is the Huggins constant. Comparison with fig. 2.22 shows that such a simple series development is unlikely to apply for polyelectrolytes; beyond the maximum, kH should be negative; it is not a constant but depends strongly on cs , and it is questionable whether [T]]2 remains as a significant parameter. When screening causes the descending part in fig. 2.22, one should in the absence of electrolyte expect a c p 1 / 2 dependence. In fact, Fuoss en Strauss1'21 proposed a semi-empirical relationship of the type
^ = 17^2
^
where A and B are constants. In swamping electrolyte, c s » c (with c expressed as monomer concentrations), Bcp1/2 » 1, ^C=|SI/2
I2'4-5'
In passing we note that the viscosity of low M electrolyte solutions has a c p / 2 term (recall the Jones-Dole equation [1.5.3.5]), but this has a different origin. For intermediate c s I c ratios, more complicated equations are appropriate, of which illustrations will be given below. For uncharged polymers, a value for the overlap concentrations c* can be obtained from space filling (scaling) arguments (sec. IV.6.11)3).
As argued before, for polyelectrolytes c* is much lower, not only because the radius of gyration a is larger, but also because the range of hydrodynamic interaction is longer. From the particle sol domain, rather concrete information is available on the influence of the surface charge on the sol viscosity in terms of the three electroviscous effects that we shall now briefly review from chapter IV. 6. (i) The primary electroviscous effect is the influence of the electric double layer on the viscosity in the Einstein region. The ionic distribution around the polyelectrolyte chains will be polarized by the flow and this polarization will, in turn, affect the viscous energy dissipation. As interpolyelectrolyte interaction does not play a role in this phenomenon, it finds its way in [2.4.3] via modification of [77] at fixed kH . The phenomenon being determined by the mobile part of the double layer makes [77] dependent on Jve. Generally, [/;1 depends on cs , z, pH (for weak polyelectrolytes) and M . Recall
11 21 31
R.M. Fuoss, U.P. Strauss, J. Polym. Sci. 3 (1948) 602, 603. R.M. Fuoss, Dicuss. Faraday Soc. 11 (1951) 125. P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell University Press (1979) ch. 3.
POLYELECTROLYTES
2.49
that for uncharged polymers, see [IV.6.11.7], /,2\3
['] = * o ^ - = ^
/ a
a3.
*o^
I2-4-7"
where @0 is the Flory-Fox constant, under 0-conditions equal to = 2.5xlO 2 3 for [rj\ in m 3 kg~ 1 , r in m and M in kg . This equation applies to coils. The constant in it depends slightly on the solvent quality and, therefore, will be different for polyelectrolytes. (ii) The secondary electrouiscous effect accounts for the influence of pair interactions. This phenomenon controls fcH; on top of the value it has for uncharged coils there comes a second one, say kHel2- We refer to [IV.6.9.18], in which for solid spheres,
rin(«/lna)] 4/5 fcHei2~
, ,5 {icaf
J
e^eCa3 {Ka)e2lca a = -^-± -—2kT
[2.4.S]
[2.4.9]
indicating that for hard spheres the dependence on size (a) and electrokinetic potential C, is already not simple. When C, is not known, it can be converted to the mobile (diffuse) charge density o 4 usinga variant of [II.3.5.62] 11 ^ ^ K g O r o ) EQEK
[ 2 4 1 0 ]
KJ ( M I )
where KQ {Ka) and Kj {Ka) are the zeroth and first order Bessel functions of the second kind, respectively. Graphical illustrations of [2.4.10] can be found in figures II.3.14 and 15. However, this conversion is a minor problem in comparison to the issues of: (a) The rather narrow c -window where pair interactions suffice; as soon as these become measurable, multipair interactions also occur, which signals our departure from the dilute regime; (b) The objects under discussion being not hard impenetrable spheres, but more or less flexible charged chains; (iii) The tertiary electroviscous effect, which is a collective noun for the influences of conformational changes on the (reduced) viscosity. Here, we can think of increasing the hydrodynamic permeability of coils by swelling, alignment of chains in the flow field and consequences the variations of the persistence length may have on the flow behaviour. This tertiary effect is at the same time in the very heart of polyelectrolyte science but most difficult to describe, let alone generalize. It is even problematic to establish which
In this equation in Volume II, K in the denominator of the r.h.s. has to be deleted.
POLYELECTROLYTES
2.50
of these phenomena enter [2.4.3] through kH , and which through [77]; in fact, as many of these phenomena are coupled the distinction becomes academic. We shall therefore approach the issue from the empirical side by examining some literature illustrations.
Figure 2.23. Intrinsic viscosities of potassium poly (styrene sulphonate). Sample parameter as in the key. — rod limit. Drawn: Yamakawa model. (Redrawn from Davis and Russel, loc.cit.) Sample Mxicr3 % sulfonation A
1170.0
86
2.44
B
109.5
92
2.61
C
41.5
91
2.58
2.4c The intrinsic viscosity Figure 2.23 gives intrinsic viscosities for three aqueous samples of potassium poly(styrene sulphonate) solutions as a function of the reciprocal Debye length. The data are taken from a systematic and experimental study by Davis and Russel11; the Yamakawa model can be found in ref. 2) . The chain charge is expressed as the ratio lB/lc of the Bjerrum length over the distance between adjacent charges. As this ratio > 1 , no condensation is expected. Note the logarithmic scale: [//] varies by a factor of 10 3 . The [TJ] values are easily a factor of 10 2 higher than those for the corresponding uncharged polymers because of the chain expansion. As expected, [rj] increases with M and with K~1 : the lower the electrolyte concentration, the more extended the chain is. In the low salt limit, the values for extended rods are approached. To the left, at high c s a l t , the polyelectrolyte effect is suppressed and (9 conditions are approached. Here, the polyelectrolyte behaves as a non-draining Gauss-type coil. The drawn curves are based
11 21
R.M. Davis, W.B. Russel, Macromolecules 20 (1987) 518. H. Yamakawa, M. Fujii, Macromolecules 7 (1974) 128.
POLYELECTROLYTES
2.51
on the hydrodynamic theory by Yamakawa and Fuji11, combined with earlier work for such chains by Odijk21 and Fixman and Skolnick31. For theoretical background, see sec. 2.3c. We note that this reasonably accounts for the experimental trends. Key factors are chain stiffness (or for that matter, persistence length) and the excluded volume. As a trend, thermodynamic quantities are more easily accounted for than hydrodynamic ones. More to the quantitative side, it appears that [77] is, as a first approximation, proportional to M 1/2 . For a variety of systems this has been observed, for instance for poly( a -L-glutamic acid) in NaCl solution and various values of cNaC14), for poly(mono methyl itaconate) in organic solvents at various degrees of neutralization51, for sodium poly(acrylate) in aqueous NaBr solutions and at different degrees of dissociation6 and for oligo- and poly(methylmethacrylates) in acetonitrile, n-butylchloride and benzene71, where a lower slope than 1/2 was found for the low M samples. This proportionality is theoretically predicted by the familiar Stockmayer-Fixman equation81 [ti]/M1/2 = Ko+O.5l0oBM1/2
[2.4.11]
Here, B is a second virial coefficient, &0 is the Flory (or Flory-Fox) constant introduced in [2.4.7] and Ko is also a constant (independent of M ), which equals (r2\3/2
K =0
° °^r~
l2 4 121
--
where (r^ is the r.m.s. end-to-end distance under 0 conditions. Equation [2.4.11] works well if the chain is not too expanded. A plot of \rj\lM1^2 as a function of M 1 / 2 (a Stockmayer-Fixman plot) helps to assess the (hydrodynamic) virial coefficient. Intrinsic viscosities of polyelectrolytes typically depend on the electrolyte concentration c s , on the line charge and on the way this charge is distributed. As a first approximation for c s » c , one may expect [rf\(c ) to scale as c~ 1 / 2 . This scaling will hold when the polyelectrolyte behaves as a coil, expanded by intramolecular repulsion. A quick scan of the formulas for electric interaction Gibbs energies shows these to scale with c~1/2 provided / or ff11 are constants. Deviations from the c~ law are A expected for electrostatic ( \j/ and/or cfi are not constant but will regulate) and H. Yamakawa, M. Fujii, loc. cit. T. Odijk, J. Polym. Sci. Polymer Phys. Ed. 15 (1977) 477; Polymers 19 (1978) 989. 31 J. Skolnick, M. Fixman, Macromolecules 10 (1977) 944; M. Fixman, J. Skolnick, ibid. 11 (1978) 863; M. Fixman, J.Chem.Phys. 76 (1982) 6346. 41 M. Satoh, J. Komiyama, and T.Iima, Coll. Polym. Sci. 258 (1980) 136. 51 L. Gargallo, D. Radic, M. Yazdani-Pedram, and A. Horta, Eur. Polym.J. 25 (1989) 1059: also see Eur. Polym. J. 29 (1993) 609 for other solvents. 61 1. Noda, T. Tsuge, and M. Nagasawa, J. Phys.Chem. 74 (1970) 710. 71 Y. Fujii, Y. Tamai, T. Konishi, and H. Yamakawa, Macromolecules 24 (1991) 1608. 81 W.H. Stockmayer, M. Fixman, J. Polym.Sci part Cl (1963) 137. 21
2.52
POLYELECTROLYTES
conformational reasons (for high chain charges, the molecule assumes a rather rod-type conformation, see sec. 2.3). By way of illustration we give a few illustrations, emphasizing the c j 1 / 2 part and accompanying deviations.
Figure 2.24 Electrolyte concentration dependence of the intrinsic viscosity for NaPAA in NaBr solutions of concentration cs (moldm"3) . Given is the parameter B in [2.4.11]. The degree of dissociation is indicated. (Redrawn from Noda et al. loc. cit.) Figure 2.24 illustrates the expected trends for a weak polyelectrolyte. The data are plotted in terms of the parameter B in the Stockmayer-Fixman equation. This parameter is independent of M and has the dimensions of intrinsic viscosity, i.e. of a reciprocal weight concentration (in these experiments, decilitres per gram). One may expect B to consist of an electric and a non-electric part, which according to the figure, are linearly additive: B
= B non-el +B el
B e l = const •f{a)cl1/2
I2-4-13! [2.4.14]
where /(a) is a function of the degree of ionization. It appears that Bnon.ei is independent of c s and a ; this must be an intrinsic macromolecular parameter. The function J(a) is essentially a measure of the relation c^ia0) . As is always found, upon increasing a° further increases of a become gradually less active in increasing cfi (or the 'active' fraction). The trends of fig. 2.24 appear to be rather representative; they have also been reported (in less detail) by Satoh et al. (toe. cit.) for poly(a-L-glutamic acid) in NaCl.
POLYELECTROLYTES
2.53
Figure 2.25. Ion specificity in the intrinsic viscosity of Na poly(styrene sulphonatc) as a function of the electrolyte concentration. Measurements at non-zero shear (/ = 1000 s"1) . Redrawn from Jiang etal. 11 . Figures 2.25a and b are meant to illustrate the lyotropic effects for a strong polyelectrolyte solution. These measurements have been carried out as a function of the shear rate y . The intrinsic viscosity increases somewhat with decreasing y but maintains the same ionic specificity, of which the trends are typical. For instance, these have also been observed by Cohen and Priel 2 ', though at lower c s . The differences between the different counterions reflect their specific binding to the chain; increased binding leaves less charge in the diffuse {i.e. interaction-active) part. So, it is seen that the binding increases from z + = 1 to z + = 2 and for fixed z + , according to Li + < Na + < K+ and Mg 2+ < Ca 2+ < Ba 2 + . This is the familiar sequence for sulphate groups; it is a.o. reflected in the c.m.c. of dodecylsulphate micelles31 and in the surface pressure of Gibbs monolayers of the corresponding surfactants 41 . 2Ad The dilute and semidilute range Now we are considering intermolecular hydrodynamic interactions. For a variety of polyelectrolytes, the reduced viscosity as a function of c passes through a maximum, as sketched in fig. 2.22. The higher the c s , the more suppressed this maximum is. Examples include Na-pectinate in NaCl51, poly(n-butyl-4-vinylpyridinium) bromide in
11
L. Jiang, D.H. Yang, and S.B. Chen, Macromolecules 34 (2001) 3730. J. Cohen, Z. Priel, Macromolecules 22 (1989) 2356. P. Mukerjce, K.J. Mysels, Critical Micelle Concentrations of Aqueous Surfactant Systems, U.S. Natl. Bureau Standards, 36 (1971). 41 See the references in sec. III.3.10b. 51 D.T.F. Pals, J.J. Hermans, Rec. Trau. Chim. 71 (1952) 433. 21
2.54
POLYELECTROLYTES
NaCl, poly(2-vlnyl pyridine) in HCl-solutions in ethyleneglycol and triethylamine11, Napoly(vinyl sulphate) probably in NaCl or NaBr2), and Na-poly(styrenesulphonate) in saltfree31 and salt-containing media41. Figure 2.26 gives an illustration for the last-mentioned system by Cohen and Priel. In fig. 2.26a the position of the maximum is independent of M~5xlO~ 6 g ml"1 , which more or less corresponds to the overlap concentration c* . Added electrolytes lower the maximum, make it less distinct and move it to higher c . This last trend is about linear, with c (max)/ c s =4±0.5 and 77inc (max) also linear with M , and the steeper the lower c s is. The theoretical interpretation, offered by the authors, centres around this equation n J2r2 ^ i n c — ^
12-4.15]
Figure 2.26. Reduced viscosity as a function of polymer concentration for Na-neutralized poly (styrene sulphonate). (a) Influence of M at fixed, low electrolyte concentration (4xlO~ 6 M); (b) Influence of c s at fixed M = 16,000 . In fig. (a) the drawn curves are meant to guide the eye; in fig. (b) the dashed curves are theoretical. Redrawn from Cohen and Priel, loc. cit.
11 21 31 41
H. Eisenbcrg, J . Pouyct, J Polym. Sci. 13 (1954) 85. D.F. Hodgson, E.J. Amis, J. Chem. Phys. 9 1 (1989) 2635. H. Vink, Polym. 3 3 (1992) 3711. N.Imai, K. Gekko, Biophys. Chem. 4 1 (1991) 31.
POLYELECTROLYTES
2.55
with the screening now determined by K2 =4;rt R (c n + 2c.) 0
\ P
s
[2.4.16]
/
because the polyelectrolyte and the indifferent electrolyte both contribute; a h is the hydrodynamic radius. This equation was derived for extended chains, but Hess and Klein" obtained a very similar result for spherical coils. Hence, it is apparently not discriminative. In fact, apart from a numerical factor this equation contained the effective counterion charge z 4 , where z is defined as the number of counterions compensating the line charge, i.e. it is a measure of the diffuse, or mobile part of the countercharge. The factor z 4 also occurs in a paper by Antonietti et al.2), to be discussed below. That this factor does not occur in [2.4.16] stems from the formalism, according to which the structure factor S(q) is related to an 'exclusion radius' determined by K . It is also noted that the dependence o n e is a function of the cs I c ratio. Equation [2.4.15] can at least semi-quantitatively account for the trends in fig. 2.26: 77inc as a function of c passes through a maximum; this maximum increases with M via a h (for an extended chain it is linear) and gets lower with increasing c s . Even the slope cs(max)/ c s = 4 agrees with [2.4.15 and 16]. Equation [2.4.15] is, of course, more specific than the virial expansion [2.4.3]; in fact, this contribution has to be added to [2.4.12], hence the A in the l.h.s. of [2.4.15]. It is, on the one hand, gratifying that such a simple equation represents the data well over the given c range. On the other hand, one cannot infer too much from it about specific polyelectrolyte properties, such as the persistence length and the (counter) charge distribution.
Figure 2.27. Reduced viscosity increment for poly(Na-styrene sulphonatc) microgels in salt-free solutions. The molecular weight of the parental solution increases in the direction of the arrow. Degree of sulphonation 70%. Drawn curves, theory. (Redrawn from Antonietti c.s., loc.cit.}
11 21
W. Hess, R. Klein, Adv. Phys. 32 (1983) 173. M. Antonietti, A. Briel and S. Forster, Macromolecules 30 (1997) 2700.
2.56
POLYELECTROLYTES
Antonietti et a l . ' ' have presented another illustration of the same nature. These authors studied the rheology of branched polyelectrolytes, viz. cross-linked poly(styrene sulphonates) in electrolyte and salt-free solutions. So, the polyelectrolyte molecules behave as microgel particles. Below c*p there is no M -effect anymore, as would be the case for impenetrable spheres; but above c* , r][ac decreases with c just as in the r.h.s. of fig. 2.26. For this part of the curve, the authors derived an equation, which just as with the previous example consisted of the series expansion [2.4.3] plus an additional term, somewhat more elaborate than [2.4.15], but with the factor z 4 made explicit. Figure 2.27 gives an illustration. This double logarithmic plot shows that the theory can account satisfactorily for the results. Data below the predicted maximum are not available, but in analogy to fig. 2.26 a decline with decreasing c must be anticipated. For a gel the molecular mass is not so easily definable, the more so as the parental polymers (from which the gel particles were synthesized) were polydisperse. However, the trend is indicated by the arrow. This trend is reflected in the intrinsic viscosity, which in this direction increases from 0.044 via 0.052 to 0.090 dm 3 g"1 for the parental microgels, whereas the fitted values for the intrinsic viscosity and Huggins constant ku are, in this order, 0.139, 0.143 and 0.182 d m 3 g - 1 and 0.2, 0.6 and 0.9 , respectively. So, one accounts semiquantitatively for the trends.
Figure 2.28. Huggins constant for the same solution of K poly(styrene sulphonate) as in fig. 2.23.
Finally, fig. 2.28 gives an illustration of the salt effect on the Huggins constant, which displays a striking minimum between high values at low and high ionic strengths. To the left, 0 -conditions are approached. These minima result from an interplay of direct intermolecular contributions to the stress, the ensuing coil compression and the attractions appearing near 0-conditions. The authors discuss this in more detail, but 11
M. Antonictti, A. Briel and S. Forster, loc. cit.
POLYELECTROLYTES
2.57
the final quantitative answer still has to come. For further interpretations, see". Viscometry has also been invoked to monitor conformational changes and counterion specificity features. By way of illustration, we refer to the frequently discussed21 phase transition, which occurs as a function of pH in solutions of poly(methacrylic acid), but not with poly(acrylic acid). Figure 2.12 already gave an illustration. Sugai et al.3> confirmed these transitions for poly(ethacrylic acid) in NaCl and at various temperatures, and Klooster et al. 4) studied these for poly(acrylic acid) in methanol (MeOH) and found the occurrence of the transition to be ion-specific: for or > 0.1, it does not occur in LiOCH3 (as in aqueous solution of the Na-salt), whereas it is found in NaOCH3 . There appears to be more room for viscometric studies of ion specificities. In particular, the effects of higher valence cations like La3+ and Th 4+ deserve attention since it is known that, at the proper pHs these ionic species hydrolyze and give rise to superequivalent adsorption (sec. IV.3.9J) For polyelectrolytes, the phenomenon has been reported51. 2.5 Polyelectrolytes in electric fields Just as with low M electrolytes and particulate colloids, polyelectrolyte solutions can be subjected to an electric field, the response of the system acting as a tool to extract information. Three familiar techniques are electrokinetics (electrophoresis in particular), dielectric studies and conductometry. Experiments can be carried out in DC or AC. Electrophoresis is usually done in DC, dielectric investigations mostly in AC, with the aim of scanning the various relaxation ranges (dielectric spectroscopy). Conductometry is preferentially carried out in DC, but AC data are automatically produced as a side effect of dielectric spectroscopy: the (measurable) complex impedance consists of a real (resistive) part yielding the conductivity spectrum K(co) and an imaginary (capacitive) part giving a>£0£{a>), see [II.4.5.13] and sec. II.4.8a. These two quantities are related to each other through the Kramers-Kronig relations. The interpretation of the obtained electrometric data is cumbersome because a number of physical phenomena participate and interfere; for many of these no sufficient theory is available. For instance, the conduction mechanism differs substantially among the dilute, semi-dilute and concentrated regimes and may depend in a complex way on the presence of added electrolyte. In this section, some attention will be paid to the issue
11
J-L Barrat, J.F. Joanny, Adv. Chem. Phys. XCIV(1996) 1, I. Prigogine, S.A. Rice. Eds. Already in 1974 H. Okamoto, and Y. Wada cited 30 references, sec J. Polym. Sci. Polym. Phys. 12 (1974) 2413. 31 S. Sugai, K. Nitta, N. Ohno, and H. Nakano, Colloid Polym. Sci. 261 (1983) 159. 41 N.T.M. Klooster, F. van der Touw. and M. Mandel, Macromolecules 17 (1984) 2070. See for instance M. Drifford, J.P. Dalbiez, M. Delsanti, and L. Belloni, Ber. Bunsengesellsch. Phys. Chem. 100 (1996) 829. 21
2.58
POLYELECTROLYTES
of electrometrically bound counterions {= counterions that in electrometry appear not to move freely). Apart from the question as to how these are bound (electrostatically condensed, chemically or still otherwise associated), one of the issues is whether different electrometric techniques yield identical answers (are electrophoretically and conductometrically bound counterions identical?). Accepting that there is no established overall theoretical picture, it is sometimes helpful to compare electrometric properties of polyelectrolytes with those of low M electrolytes, on the one hand, and with particulate colloids on the other. One typical illustration is that the conductivity of a polyelectrolyte of JV charged monomers and its compensating counterions is always lower than that of the JV monomers free in solution. Apparently this is the result of counterion immobilization by the chain. Another line is the interpretation of the slip plane for solid particles and the application to polyelectrolytes. We shall address this issue first. 2.5a The issue of electrokinetic binding Although polyelectrolytes, together with their countercharge, are electroneutral, they move in an electric field, i.e. they exhibit electrophoresis. The occurrence of this phenomenon implies charge separation: upon moving, the polyelectrolyte may entrain part of the countercharge, but not all of it. Which part? The issue has a micromechanical origin. Upon tangential movement of a liquid (especially water) with respect to a solid, a thin, adjacent layer of this water is stagnant, meaning that it does not move with respect to the solid; in electrophoresis the stagnant water moves with the solid. In streaming potential experiments, it remains stationary with the particles constituting the porous plug. Countercharges captured in stagnant layers behave as if they are electrokinetically bound. The phenomenon is widespread; it is observed for inorganic solids, polymer latices, hydrophobic and hydrophilic surfaces, surfactant micelles, ... Hence, it presumably also occurs with polyelectrolytes, either with coils as a whole or with individual chains. Questions coming to mind include: (i) what is the origin? (ii) how can the phenomenon be experimentally detected? and (iii) can we state something about its magnitude? (i) According to present day insight, the origin has to be sought in the water structure adjacent to hard walls1'21. Mostly as a result of the repulsive part of the intermolecular interactions, this leads to the familiar stacking: oscillations of the molecular density pN{z) that peter out over a very few molecular diameters. This picture is in line with the generality of the phenomenon; it should apply to liquids adjacent to any surface that is hard (impenetrable) on the molecular scale. It is even observed for some L-L interfaces31, should not occur for clean water-air phase boundaries, but must be expected for fluids
11
J. Lyklema, S. Rovillard, and J. de Coninck, Langmuir 14 (1998) 5659. J. Lyklema, Oil Gas Set Technol. 56 (2001) 41. 31 A.M. Djerdjev, J.K. Bcatty and R.J. Hunter J. Colloid Inter/act Sci. 265 (2003) 56. 21
POLYELECTROLYTES
2.59
adjacent to polyelectrolytes. The structure formation leads to an increase of the tangential viscosity in the adjacent liquid that, because of the strongly cooperative nature of viscous flow, becomes so strong that phenomenologically speaking the layer behaves as if it were stagnant.
Figure 2.29. Relation between the electrokinetic and surface (- or line) charge, both expressed in the same units Cm" 2 or Cm- 1 . Tan a = / e k .
(ii) For particulate colloids, micelles, etc., the electrokinetically bound charge can be determined by simultaneously measuring the surface charge a° and the electrokinetically free charge crek . The latter can be inferred from £ potentials, using GouyChapman theory, of which the application is safe because the electrokinetically free part of the countercharge more or less coincides with the diffuse part of the double layer. Measurement and interpretation of f -potentials have been described in detail in chapter II.4. Figure 2.29 sketches the relationship between the two types of charges as it is usually found. At very low surface or line charge this charge equals the electrokinetic one (negligible fraction of countercharge in the stagnant layer), but upon increasing the surface or line charge, <7ek lays behind, to eventually level off in a plateau when any additional charge on the surface or the polyelectrolyte is compensated by a counterion in the stagnant layer. Phenomenologically, the trend is identical to that in fig. 2.4, but the interpretations are very different. (iii) Magnitudes can be expressed in terms of the ratio k
_ electrokinetically active charge surface- or line charge
which may, or may not, be identical to / in [2.2.2]. This ratio starts with unity, to decrease with increasing line charge, see fig. 2.29. For hydrophobic (Agl) or hydrophilic ( Fe 2 O 3 , TiO 2 ) macrosurfaces, / e k can be read from figures like II.4.13 and II.3.63; at high a° , / e k can be as low as 0.2. Such low values are also observed for ionic micelles and it will not come as a surprise if such values are also observed for polyelectrolytes. Application of these insights to polyelectrolytes remains a challenge for further study. Little is known of the micro-mechanics of water flow tangential to chains with internal degrees of freedom. The experience with macrosurfaces can perhaps best be translated to polyelectrolytes in the limiting cases of isolated chains (~ cylinders) and coil-like,
2.60
POLYELECTROLYTES
weakly charged polyelectrolytes, to which the more general and more intriguing, intermediate situations reduce under extreme conditions. In view of the developments in terms of two-state models (Oosawa, Manning, ..., see sees. 2.2b and c), there are arguments also to simplify electrokinetic binding in terms of only one parameter ( / e k ) demarcating electrokinetically, fully stagnant and fully free (as in the bulk solvent) double layer parts. This is an abstraction from reality because the tangential viscosity does, with increasing z, not drop discretely from infinite to its bulk value, but rather decreases gradually, though very steeply. However, in the case of macrosurfaces such idealizations are common and appear to serve well in practice; In electrokinetics it Is the familiar slip plane, which separates the fully stagnant layer from the fully mobile bulk, and in electrostatics this role is played by the outer Helmholtz plane, separating the Stern from the diffuse layer (see sec. II.3.6c). Actually, experience has shown that in many interpretations of data, one can get away with the identification of the potential at the slip plane (£) with that at the outer Helmholtz plane {y/1). We shall not repeat this discussion here. 2.5b Conductivity, backgrounds In principle, conductivity studies provide information on the numbers of ionic species in a solution and their mobilities. As association of ions lowers the number of charge carriers, the conductivity of a solution also depends on the degree of association. In the polyelectrolyte case it depends on the extent of counterlon association, but only if the association leads to electroneutral pairs that do not move in an electric field. On the basis of conductivities alone, one cannot obtain simultaneous information on numbers and mobilities. To that end, additional information is required, for example via measurement of transference numbers. Alternatively, one could analyze the conductivity as a function of adjustable variables, such as the degree of association of the polyelectrolyte or the concentration of added low M electrolyte. In sec. 1.6.6 the basic principles for the conductivity of simple electrolyte solutions have been discussed. Conductivity measurements are not particularly difficult provided the usual precautions are taken (avoidance of electrode polarization, appropriate thermos tatting, ...). Transference numbers can be extracted from electrolyte excesses near electrodes. The primary piece of information is the conductivity K (in Sm" 1 or Cy~ 1 s~ 1 m~ 1 J11, which follows simply from Ohm's law from the current density j (in A = Cs ) generated by an applied field E (Vm"1] j = KE
[2.5.2]
For an ideal, low M electrolyte solution, K consists of additive contributions of the ionic species i, so that it is possible to define specific ionic conductivities K{ with j^KjB
Previously called specific conductivity.
[2.5.3]
POLYELECTROLYTES
K = ZiKi
2.61
[2.5.4]
For non-ideal solutions, electric and hydrodynamic interactions have to be accounted for, but the additivity rule can be maintained (sec. 1.6.6b). It is a logical step to separate in K{ the number of charge carriers, determined by the ionic concentration c{ and the ionic velocity expressed as its mobility u; (in m 2 V*1 s^1) K^jZilFCjU;
[2.5.5]
and K = FIi\zi\ciui
[2.5.6]
Introducing for each charged species the ionic or molar conductivity Xi (in Sm~2mol~') as AJ^ZJIFUJ
[2.5.7]
[2.5.6] can be written as K = ZiciAi
[2.5.8]
For low M electrolyte solutions additivity persists, but Xi now becomes dependent on ci. The first term of the development A^cJ scales as c\12 (sec. 1.6.6b). Now, consider the application to polyelectrolytes, counting ionic species, distinguishing a polyelectrolyte, (monomer-) concentration c , a cation (c+) and anion (c ). Without loss of generality, we assume the polyelectrolyte to be negatively charged; hence cations are the counterions and anions the co-ions. K = F[| Z + |c + U + +|z_|c_u_+| Z p |c p u p ]
[2.5.9]
If c is counted per monomer, z is the number of charges per monomer. For polyelectrolytes in salt-free solution, the co-ion contribution drops out and [2.5.9] reduces to K = F[| Z + |c + u + | Z p |c p u p ]
[2.5.10]
Equations [2.5.9 and 10] are only formal in that they assume only one averaged concentration and one averaged mobility for each species. In reality, these parameters vary from place to place. Moreover, they depend on c p and cs (if electrolyte is added). Counterions have a very high concentration close to the chain, beyond which region c+ decays as in a diffuse double layer till its bulk value. For co-ions, if present, it is the other way around; their concentration is low near the chain (even zero if the line charge is high enough) and increases till their bulk value.
2.62
POLYELECTROLYTES
Here, we have to digress into the meaning of 'bulk electrolyte concentration.' For saltfree polyelectrolyte solutions, there is no bulk except for c —» 0 . Theoretically, in that case, c s = 0 , but in reality there is always some spurious electrolyte in the system (carbonate from CO 2 in the air, ions released from the vessel wall...), in addition to the H + and OH" ions stemming from the spontaneous dissociation of water. Actually, this reference state is not well-defined. In semidilute solutions the reference is the half-way cation concentration, but this reference obviously depends on c and a model is needed to establish it (compare sec. 2.2b). These ambiguities can be avoided by adding low M electrolyte to the system. However, a price has to be paid because in this situation negative adsorption of electrolyte takes place, leading to Donnan exclusion (sees. I.5.5f and 5.6a). Again, a model is needed to establish this effect because measuring c s in the dialysate would yield an overestimation. Only in swamping electrolyte (cs » c ) is the Donnan effect negligible. Intermediate situations cannot be dealt with without theory. Returning to the values of c and u in [2.5.9 and 10], there are arguments to use (again) a two-state approximation for u + , by distinguishing only bound and free counterions. The fraction n-_f c o n c M that is conductometrically bound has zero mobility. Counterions finding themselves close to the chain, but hopping along the chain, also belong to fraction (l - / c o n d J when for these ions c+A+ « c + (bulk)A + (bulk). For the free ions it is safe to presume u+ =u® , the mobility at infinite dilution, assuming c + (r) to be given by diffuse double layer theory, see [2.2.10 or 2.2.21]. The equivalent for flat plates has been developed by Bikerman 11 and gives rise to [II.4.3.60] for the excess surface conductivity in a diffuse double layer. The equivalent for cylindrical geometry has still to be elaborated. However, general experience with the interpretation of experimental surface excess conductivities in electrokinetics gives no indication that the Bikerman equation would not apply to a purely diffuse double layer. If problems arise, it is with the non-diffuse part (stagnant or Stern part), see chapter II.4. Translated into polyelectrolyte terms, it is an acceptable first approximation to work with immobile bound counterions and free counterions having bulk mobilities. With this picture in mind, setting in [2.5.9] for a symmetrical (z-z) electrolyte c+ = cs+ Jcond | z p | c p , c _ = cs and replacing the polyelectrolyte term by (1 - / c o n d ) | z p | C p U p to account for the conduction-active part, K = F[z(cs
+
/co«d|zp|cp)u0+Zcsu0+(l-/<=°»d)|Zp|Cplipj
[2.5.1!,
By the same token, for salt-free polyelectrolytes, K = F[z(cs+/c°»d|zp|cp)u0
+
(l-/»»d)|zp|cpUp]
[2.5.12]
These equations have a fair generality; variants can be found in the literature. Using 11
J.J. Bikerman, Z. Physik. Chem. A163 (1933) 378; Kolloid-Z 72 (1935) 100.
POLYELECTROLYTES
2.63
them at various c s and c , or as a function of the degree of dissociation, sometimes allows establishing parameters. However, this generality also implies the absence of any theoretical foundation in terms of polyelectrolyte configurations. In this respect, more research is wanting. For an attempt in this direction for semidilute polyelectrolytes in salt-free solutions, see11.
At this point it may be noted that so far a variety of bound (1 - / ) and free (/) fractions have been met. Different origins have been proposed (condensation, ion association, electrokinetic trapping...) and a variety of techniques have been used for measuring them (titration, osmotic pressure, electrokinetics, conductometry ...). We shall not discuss here the intriguing issue as to whether all these techniques measure the same / and, hence, whether all these fractions have the same origin. Given measurements of K and those of transference numbers, one can write equations combining the two. Generally, the transference number t of an ionic species in a mixture of species, is defined as the fraction of the current that species i transports. In terms of conductivities, K, K. 1 t=— — = —L 1 IiKi K
[2.5.13]
with 2^=1
[2.5.14]
Because of [2.5.14], one needs to measure the transport number of only one component in a mixture of two; for salt-free systems t
can be deduced if t+ is measured.
Finally, conductivities can also be represented in terms of molar conductivities A , which are the contributions per mole of electroneutral electrolyte. They are obtained by dividing K by the (molar) concentration of the electrolyte. Generally, for one electrolyte A = K/cs
(Sm 2 mor 1 )
[2.5.15]
In [1.6.6.13] it was shown that [2.5.15] also applies to asymmetric electrolytes, hence this equation may also be used for salt-free polyelectrolytes: Ap=K/cp
[2.5.16]
The problem is that in the more general situation of polyelectrolyte in low M electrolytes there are two electroneutral components, for which the molar conductivities are not additive, because of the Donnan exclusion. The usual procedure for obtaining a molar conductivity that is characteristic for the polyelectrolyte is by writing
11 R.H. Colby, D.C. Boris, W.E. Krause, and J.S. Tan, J. Polym. Sci B, Polym. Phys. 35 (1997) 2951.
2.64
POLYELECTROLYTES
/1_=^-^ P C P
[2.5.17]
This equation is simple but application is not obvious because Ks cannot be easily measured. Mostly for Ks the conductivity for c —> 0 is substituted, but this is only approximately true because of the Donnan effect. This issue comes back in establishing limiting values. For simple electrolytes, A{cs —> 0) can be found by extrapolation. If we do that for polyelectrolytes in the dilute range we obtain A (c -> 0), which is still a function of cs . In the semidilute range we can write11 A
P = AV(CP=O)
12 5 181
+ 0 C
' -
( P)
where A (c = 0) and <£> depend on cs . The overlap concentration c* also depends on c
s-
2.5c Conduction: illustrations We shall now present an anthology of experimental studies from the literature, showing the kind of results that can be obtained. The first illustration (fig. 2.30) shows the transference number of Na+ ions as the counterions of the weak poly(acrylic acid), PAA. The polyelectrolyte concentration is rather high and low M electrolytes are absent.
Figure 2.30. Transference number of Na+ (counterions) on PAA as a function of the degree of neutralization O, c — 1.5x10^^ monomol dm" 3 ; • c = 3.78 xlO~ 2 monomol dm" 3 . (Redrawn from Mandel, see ref. in sec. 2.8, who has based his figure on data by Wall et al. and Schmitt et al, respectively) Under these conditions, no c
d e p e n d e n c e is observed. For a—>0 , t P
r
. =0.5, Na+
meaning that 50% of the charge transport is accounted for by the counterion and 50% by the macro-ion. When a increases, t decreases because a fraction of the Na+ ions Na +
is bound to the chain, and therefore moves to the anode, together with the 11
H. Vink, Macromolec. Chem. 183 (1982) 2273.
POLYELECTROLYTES
2.65
polyelectrolyte. For larger a , t + even becomes negative, i.e. t > 1. This means that the polyelectrolyte does more than its share in the conduction; it even carries its counterparts in the wrong direction. The phenomenon is not unique; it is also observed for micelles and charged colloids in dilute electrolyte.
Figure 2.31 Concentration dependence of K - K{cp = 0) for poly(diallyl dimethyl ammonium chloride) (PDADMAC) over three c p ranges and for different molecular masses: O, Mn = 12,000; • 22,000; • 72,000; • 170,000 gmol" 1 . (Redrawn from Wandrey and Hunkclcr, toe. cit.)
Figures 2.31 and 32 stem from a review by Wandrey and Hunkeler11, containing much experimental information. In fig. 2.31 the average distance between the charges on the chain is 0.5nm , so the line charge ve = 3.2xlO~6 juC cm~' . As these charges reside at the periphery of the ally! groups (a five-member ring with an N(CH3);j; at its end), the backbone can be interpreted as well as a cylinder. If the radius a is ~ 0.6 nm C. Wandrey, D. Hunkeler, Study of Counterion Interactions by electrochemical Methods in Handbook of Polyelectrolytes and their Applications, S.K. Tripathy, J. Kumar, and H.S. Nalwa, Eds. Vol. 2 chapter 5. American Scientific Publ. (2002).
2.66
POLYELECTROLYTES
Figure 2.32. Molar conductivity A of Na poly(styrene sulphonate). Influences of molar mass and NaCl concentration: T = 20°C. Key: O, no added salt: V, 1CT6 (5xlCT 6 in (c)); • , 2xlO~ 6 ; D, 4xlO~ 6 ; • , 10~ 5 ; A, 2xlO~ 5 ; A, 5 x l 0 ~ 5 ; O, 10~ 4 M. The (number-average) molecular mass (in g/monomol ) is indicated. (Redrawn from C. Wandrey, loc.cit.; data points below 10 6 monomol"1 deleted.)
(depending on the radial positions of the allyl groups), this cylinder has a surface charge
C. Wandrey, Langmuir 15 (1999) 4069.
POLYELECTROLYTES
2.67
anionic polyelectrolyte Na poly(styrene sulphonate), (NaPSS); charge distance 0.5 nm ve = 4.8xl(r 6 uCcm""1; CT° = 17|iCcm-2 , also rather heterodisperse (Mw / MD <1.1). Results are collected in fig. 2.32. Curves are found with maxima that are suppressed by the addition of salt. The c range covers several decades and includes more than one concentration regime. To the right we are probably in the region (c) of fig. 2.15, where the characteristic length E, is independent of M . This regime is attained at lower c for higher M . Working our way to the left an increase of A is observed, which is more pronounced the lower c is. So, this is a typical polyelectrolyte effect. This upward trend has been observed more often. For example, Mandel in his review11 cites a number of illustrations. The occurence of a maximum is more rare; it was, for instance, reported for Na (carboxymethylhydroxyethylcellulose)21. Strictly, when in very dilute solutions the dilute regime would have been reached, A ought to be independent of c , see [2.5.18], but this is not observed. However, the caveat must be made that the measurement of A in such dilute solutions is tricky because the difference K-Ks in [2.5.17] is small; it requires a very accurate establishment of Ks and correction for salt exclusion. The positions of the maxima shift to the right with increasing cs . We note that this trend is also expected for the overlap concentration c*p ; this quantity increases because at high c the polyelectrolytes are screened better, reducing their excluded volumes.
Figure 2.33. Conductivity in dilute solutions of three carboxymcthyl-celluloses (from O via • to + with an increasing degree of substitution) and for poly(methacrylic acid) (A). (Redrawn from Vink. loc.cit.)
Another illustration was given by Vink and is reproduced in fig. 2.33. Here, K is plotted without attempting to subtract Ks . In this experiment extrapolation to c = 0 is linear and easy. The extrapolated conductivities are independent of the charge and nature of the
11
See sec. 2.8.
21
H. Vink, J. Chem. Soc. Faraday Trans. I 7 7 (1981) 2439.
31
H Vink, MacTomolec. Chem. 183 (1982) 2273.
2.68
POLYELECTROLYTES
polyelectrolyte and may, therefore, represent Ks . On the other hand, the slopes representing A are charge and nature specific, as expected. We note that the c range of this study (10~5-10~4 monomol dm" 3 ) is comparable with that in the previous study (fig. 2.32), so there are specific differences between polyelectrolytes of different natures. In the CMC series, the most highly charged samples have the lowest A values, apparently because the protons are more strongly bound. Ion specificities in ^ p (c p = 0) for polyfstyrene sulphonates) (PSS) and poly(vinylsulphates) (PVS) are collected in table 2.2; similar series were reported by Szymczak et al.1J and Kwak and Hayes21. Table 2.2. Limiting equivalent conductivities for PSS and PVS. Given is 104 A (c ->0) in S m 2 m o r 1 (fromVink, loc. cit.) Counterion
PSS
PVS
K+
46.1
64.5
38.1
55.3
112
162
Na H+
+
Analysis of such data, in terms of ionic concentrations and mobilities using [2.5.11 or 12], is ambiguous because there are two unknowns, Jcond and the mobilities of the bound counterions. For particulate colloids, there are safe ways for establishing _f c o n d , which incidentally correlates well with the free fraction inferred from particle interaction studies31. Hence, for such colloids the mobilities of bound ions can be obtained. All of this deserves to be elaborated for polyelectrolytes. Awaiting such further studies, it can be observed that A (c —> 0) in table 2.2 and in similar studies increases in the direction of Na+ < K+ < H+ , which agrees qualitatively with the increase of the ionic mobilities at infinite dilution (table I.6.5)4'. However, the increase is less quantitatively. The inference is that specific binding increases in the same direction, which is in line with the Gurney principle (sec. 1.5.4). The sequence depends on the nature of the polyelectrolyte5'. Regarding counterions of higher valencies, bivalent ions are more strongly (conductometrically) bound than monovalent ones. For example, Szymczak et al.6) report on this phenomenon for Mg2+ , Ca 2+ and Sr 2+ binding to the same polyelectrolyte. Counterions of still higher valencies like La3+ and Th 4+ can bind super11 J. Szymczak, P. Holyk, and P. Ander, J. Phys. Chem. 79 (1975) 269. Also sec P.R. Holyk, J. Szymczak and P. Andcr, J. Phys. Chem. 80 (1976) 1626. 21 J.C.T. Kwak, R.C. Hayes, J. Phys. Chem. 79 (1975) 265. 3 J. Lyklema, in [nterfacial Electrokinetics and Electrophoresis, A.V. Delgado, Ed. Marcel Dckkcr (2002) 87. See in this connection the IUPAC recommendation by van Lecuwen et al. in sec. 2.8a. 51 See for example H. Vink, J. Chem. Soc. Faraday Trans. (I) 85 (1989) 699. 61 J. Szymczak et. al, loc. cit.
POLYELECTROLYTES
2.69
equivalently, leading to overcharging11. Sometimes this phenomenon is attributed to ion correlations leading to a charge distribution, which differs from that obtained by a mean-field theory (like PB theory). However, in sec. IV.3.9j it was established that in all systems where multivalent ion binding was studied the origin was the formation and binding of hydrolyzed species. Only at rather low and rather high pH does no such charge reversal occur. To make sure that ion correlations are responsible for the overcharging, studies have to be carried out at such extreme pHs. As with viscometry, there remains much to be done in the field of polyelectrolyte conductometry, both experimentally and theoretically. 2.5d Dielectric properties We mention this theme for the sake of completeness, but refrain from a systematic discussion because the field is too complex and unsettled to review its fundamentals. For particulate colloids satisfactory techniques and elaborations are available, see sees. II.4.5e and 4.8. However, the numerous additional degrees of freedom, with the inherent number of conformational interactions for 'soft' particles, leads to almost unsurveyable complications. For an authoritative discussion of the state of affairs, a review by Mandel and Odijk is recommended21. 2.6 Electrostatically driven complexation and phase separation 2.6a Solubility of polyelectrolytes In this section we discuss the solubility of a polyelectrolyte, which carries a fixed number of charges (i.e. it has strong dissociating groups or the pH is fixed at such a value that a is unity, see [2.2.37]) and has a soluble backbone. In other words, we pose the question as to whether charges can reduce solubility. One may wonder whether this is a relevant discussion, because polyelectrolytes in water have been widely studied from the viewpoint of true solutions (i.e. one-phase systems) so that it is not immediately obvious why phase separation is an issue. Yet there is a general argument as to the reason we cannot ignore this point. This stems from the fact that in electrolyte solutions the ions are not distributed randomly. Due to electrostatic interaction, the ions tend to arrange in such a way that the distances between like (repelling) ions are larger and those between unlike (attracting) ions are smaller than the average distances. In other words, there are, even in dilute solutions, important ion-ion correlations. One of the simpler attempts to take such correlations into account in the theory of electrolyte solutions is the well-known Debye-Hiickel theory Sec, for example, M. Drifford, J-P. Dalbiez, M. Delsanti, and L. Belloni, Ber. Bunsenges. Phys. Chem. 100 (1996) 829. M. Mandel, T. Odijk, Dielectric Properties of Polyelectrolyte Solutions. Ann. Rev. Phys. Chem. 35 (1984) 75-108.
2.70
POLYELECTROLYTES
leading to the familiar limiting law for ionic activity. There is a delicate balance since both the electrostatic energy and the configurational entropy are lower than for the random case, but it is the entropy that must prevent the ions from clustering into a dense phase (see the discussion in chapter 1.5.2). An increase in the strength of the interaction as compared with the thermal energy may thus be expected to perturb that balance whereupon a dense phase appears. As is borne out by [2.2.39], such an increase can be brought about by an increase in the valencies zvz. of the ions or by a decrease in either the dielectric constant, the temperature, or the interionic distances. Experimentally, T, r, and e are not entirely independent; in fact, both r and e depend on T, but e may be controlled to some extent (e.g. in mixed solvents), and so are the valencies. As was shown by detailed experimental studies (see fig. 1.5.2) and later by Monte Carlo calculations, the correlation effects are relatively unimportant for monovalent ions in water at room temperature and moderate concentrations, so that it is justified for this case to use the Poisson-Boltzmann equation, which ignores such effects. However, correlations already become quite noticeable with divalent ions, to the extent that the Poisson-Boltzmann equation gives predictions qualitatively different from simulations in which non-smeared out charges occur. The energy/entropy balance mentioned above would, of course, shift in favour of insolubility when the valency of the ions increases or when they are tied to a macromolecule: charged groups, which are covalently linked to a polymer chain, have much less entropy than free ones. Qualitatively, this is one way of explaining why solutions containing polyions of opposite sign {polysalts) are often insoluble, whereas simple salts are soluble, and why DNA forms swollen coils when it has monovalent counterions, but collapses into a globular or toroidal state upon addition of the trivalent ion spermine or the tetravalent ion spermidine, or upon lowering the dielectric constant by, for example, adding some ethylene glycol. One theoretical approach sometimes used is to distinguish between free and bound counterions. The counterions are supposed to bind to charges on the chain, thus forming ion pairs; a polymer, which primarily carries such ion pairs, is called an ionomer. These ion pairs are strong dipoles and can again cluster with other ion pairs (into socalled multlplets). This effect, sometimes called the 'ionomer effect' produces an effective cohesion between monomers, which is opposed by the entropy associated with mixing polymer segments and solvent, and by the entropy due to the translation of the unbound ions. A feature of the theory is that counterion binding is described as a simple chemical equilibrium with a binding constant, which depends on the local dielectric constant. Kramarenko et al.11 have described models of this kind, which can account for phase transitions that are actually observed. An example is the titration of poly(acrylic acid) in
E.Yu. Kramarenko, I.Ya. Eruchimovich, and A.R. Khokhlov, Macromolecular Theory Simulation 11 (2002) 462
POLYELECTROLYTES
2.71
water/methanol mixtures with base. In their acid form, this macromolecule is soluble, but above a critical degree of neutralization the polysalts become insoluble. Another supporting observation is that polyacids in dimethyl sulfoxide (DMSO, relative dielectric constant 48.75) form very stable solutions, but as soon as a little neutral salt is added, they precipitate due to the formation of ion pairs. Hence, minor changes in the strength of Coulomb forces or in the effective entropy of the ions can tip the balance and render polyelectrolytes insoluble. 2.6b Polyelectrolyte complexes What is the physical reason for phase separation? As was said in sec. 2.6a, electrostatic forces always oppose the separation of opposite charges. The electrostatic energy of a system consisting of ions is lowest when the ions arrange in a dense phase, more or less in alternating order, as they would do in a salt crystal, with positive charges surrounded by negative ones and vice versa. To dissolve the complex, entropy is needed to overcome the attractive electrostatic energy; when the electrostatic forces are very strong, monophasic systems can only exist at extreme dilution. It should, therefore, be no surprise that stoichiometric mixtures of oppositely charged macro-ions will also phase-separate. The driving force comes from two effects: (i) the monovalent ions, which originally surrounded the macro-ion in the diffuse double layer, had a reduced entropy; when the complex forms, the diffuse double layers are stripped off the chains leading to an entropy increase; {ii) in the dense complex that forms, the average distance between positive and negative ions is smaller than that between the groups on the free macro-ion and their counterions, so that the electrostatic energy is lower. It also follows from this picture that the net change in Gibbs energy associated with the formation of the dense complex phase decreases with increasing ionic strength as the effects of (i) and {ii) both become smaller at higher ionic strength. Indeed, many complexes are redissolved at high ionic strength. A very similar case is that of colloidal spheres of opposite charge. Also here, phase separation may occur (heterocoagulation) around stoichiometric compositions. A difference is, however, that Van der Waals forces may be more important in lyophobic colloids. 2.6c Complex coacervation It has been known for a long time that mixtures of anionic and cationic polyelectrolytes phase-separate into a dense phase, containing most of the polymer, and a dilute phase essentially containing only small ions. The dense phase may sometimes have solid-like properties, in other cases it appears as a liquid. In the latter situation, the term complex coacervation is employed, and it seems likely that the phase-separated system is in thermodynamic equilibrium, so that one can meaningfully determine a phase diagram. When doing so, it is mandatory to consider the number of independent
2.72
POLYELECTROLYTES
components of such a system. The most general description calls for at least five components: solvent (water), polycationic salt, polyanionic salt, polysalt, and the simple salt consisting of the relevant counterions. Because of the Gibbs-Duhem relation, only four of these are independent. A simplified case is obtained when the simple salt is omitted and the polymers are stoichiometrically mixed in their pure acid or base form, so that only a polysalt is obtained. One could thus also consider simpler 3-component mixtures of polysalt, simple salt and water; off-stoichiometric mixtures will then require the two extra components, namely polyanionic or polycationic salt.
Figure 2.34. (a) Schematic phase diagram for complex coacervation in a mixture of a polyanion salt Pn"Cn+ and a polycation salt P n + C n ~, showing tie lines, (b) Experimental phase diagram for a mixture of gelatine and gum arabic at fixed pH (no added salt). Coexisting complex coacervate phases ( Cj , ... Cg ) and dilute electrolyte ( ej,... eg), obtained from the mixtures ( nij , ... m 5 ) are indicated. Critical points are indicated as 'Cr.p'. (Redrawn from H.G. Bungenberg dc Jong, in Colloid Science II, H.R. Kruyt, Ed., Elsevier (1949) ch. X, sec. 2g. In many cases, added salt narrows the two-phase region of the phase diagram and this often leads to complete dissolution. A schematic phase diagram is given in fig 2.34a. As can be seen, the tie-lines, corresponding to coexisting phases, are perpendicular to the polymer-polymer axis, which just means that the polymer-rich phase and the dilute phase have compositions that hardly differ when the overall composition is stoichiometric; further away from stoichiometry, the dilute phase becomes progressively rich in the excess component. For very asymmetric polyelectrolyte mixtures the system is monophasic. An example of an experimental phase diagram is shown in fig. 2.34b for the mixture of gum arabic and gelatine at pH = 5. Note that this kind of phase diagram, corresponding to associative phase separation (due to attraction between the polymers), is entirely different from the phase diagram of a mixture of two neutral, incompatible polymers in a common solvent, which usually splits into two separate polymer solutions due to repulsion {segregative phase separation). Compositions inside the closed two-
POLYELECTROLYTES
2.73
phase region will phase-separate into a dense phase (with approximately the overall composition) and a very dilute phase containing minute amounts of polymer. As is clear from this diagram, it is also possible that a concentrated phase appears with a composition, which is significantly different from the stoichiometric one (in terms of polymeric charges). Such a phase contains not only polyions of both kinds, but also necessarily a certain amount of small cations or anions to ensure overall electroneutrality. As the ionic strength increases, the composition range of the complex coacervate phase narrows, and above a critical ionic strength c sc complexation is entirely suppressed. We thus conclude that the insoluble complexes that form generally incorporate a certain amount of excess polymeric charge, which is zero in the strictly stoichiometric case and increases as the mixture becomes more asymmetric. For neutrality, a non-stoichiometric complex therefore has to take up small ions, which involves an entropy loss. Figure 2.35. (left) (a) Solubility diagrams (degree of coacervation as a function of composition) for a mixture of polyanion and polycation with added salt. Compositions within the closed area lead to phaseseparation, (below) (b) Solubility diagrams for mixtures of gum arabic and gelatine with added KC1 for various pH values as indicated. (Same reference as fig. 2.34, but sec. 2f.)
2.74
POLYELECTROLYTES
A useful representation of the phase behaviour is obtained at constant total polymer concentration, but varying polymer composition and ionic strength. Such a representation is schematically sketched in fig. 2.35a; an experimental example is presented in fig. 2.35b. One may wonder why asymmetric complexes are still insoluble. As in any electrolyte solution, the electrostatic energy is always lowered upon increasing the concentration because of increasing ion-ion correlations. For ordinary electrolytes consisting of small ions only, this attractive energy is counteracted by a significant (and salt concentrationdependent) entropy reduction, so that dissolution is possible although at the expense of reduced activity coefficients. For ions bound to a macromolecule, however, such an entropic effect is obviously much smaller. As a result, a stoichiometric mixture of macroions of opposite sign will easily become insoluble. A non-stoichiometric mixture will have to incorporate a certain number of small ions, which contribute an extra entropic term to the Gibbs energy. Hence, there is a balance between an attractive effect from the macromolecular ions and a repulsive effect from the small ions. This balance determines the shape of the phase diagram. One may also wonder what happens outside the two-phase region, but at salt concentrations below c sc . Under these conditions, the polyions still attract each other and form complexes, but there is a large mismatch in the number of available molecules, and the composition of the formed complexes will reflect that mismatch. Each molecule of the minority component will thus bind an excess of the minority component, so that complexes are formed with a high charge density. The small ions neutralizing that highly charged complex still have enough entropy to inhibit it from phase separating. Theoretical calculations of the shape of the phase diagram are complicated because the Gibbs energy of a mixture of polyions and small ions obviously depends strongly on the way in which the various ions arrange themselves. Not only will this affect the configurational entropy of the small ions, but it will also have a contribution from the entropy (conformational and translational) of the chains. To simplify matters, one could treat the electrostatics by the same model as that underlying the Debye-Hiickel theory, provided one takes into account the reduced entropy of ions on a chain. Needless to say, short range polymer-solvent and polymer-polymer interactions play a role as well; separating complexes consisting of polyelectrolytes with very hydrophobic backbones may be much more difficult than complexes, which are inherently hydrophilic. On top of this, specific ion effects will almost certainly lead to important differences in the binding energy of various kinds of ion pairs. A complete theory of complex coacervation is therefore a long way ahead. An early attempt to construct a theory for complex coacervation is due to Voorn and Overbeek11. These authors simply proposed two additive contributions to the Helmholtz 11 M.J. Voorn, Fortschr. Hochpolym. Forsch., 1 (1959) 192-233; J.T.G. Overbeek, M.J. Voorn, J. Cellular and Comparative Physiology 49 (1957) 7-26.
POLYELECTROLYTES
2.75
energy density: a polymer/solvent mixing entropy, given by the Flory-Huggins theory (see sec. V. 1.2.2) and an ionic contribution / i o n according to the Debye-Hiickel theory, given by 1 1 LJ™kT
= -K3(
3
/\2n:
[2.6.1]
Here, t is the size of a polymer segment and AT"1 is the Debye length given by the familiar expression [II.3.5.7] 2
e2 eeokT^
v
>'
In [II.3.5.7], Cj and zi are the number concentrations and number of charge per molecule of species i, e is the charge of the electron, and EQE is the dielectric permittivity of the medium. The number concentrations can be expressed in volume fractions ( CjJVj^/V =
where (B is the Bjerrum length. Note that at any combination of
A=^=i+vU^)+i kT
kT
N
2V \
3
C3
3+...
12.6.3]
J 2,
where a is the average number of charges per monomer unit (0 < a < 1) . This result is very similar to the osmotic pressure of an ordinary polymer solution [ 1.2.4], but with an effective excluded volume given by the term in braces with an electrostatic contribution to the excluded volume, which is negative and scales as f.|r"' . As expected, for low ionic
11
T.L. Hill, An Introduction to Statistical Thermodynamics, Addison Wesley (1960) ch. 18.
2.76
POLYELECTROLYTES
strength and vof order unity, this term is large and the excluded volume is negative. Hence, this polysalt will be insoluble. If simple salt is present, the osmotic pressure will have additional contributions from the salt. That due to the polymer can still be described by [2.6.3], but now the effective excluded volume is salt-dependent. With increasing salt concentration, it becomes less negative and eventually it may even switch from positive to negative; the polysalt becomes soluble in excess simple salt. Hence, the result is a phase diagram like that for a neutral polymer with an electrostatic term included in an effective solvency parameter. This result can be generalized to include polymer-solvent interaction by adding the usual Flory-Huggins interaction parameter x • as follows: ZeS=^-^°2+Z
= Zion+Z
[2.6.4]
where we defined the parameter j i o n = (2^/3)(^gX""1 /i3)a2 . The theory is, of course, approximate as it is based on the Debye-Hiickel theory for dilute electrolytes and low potentials, whereas complex coacervate phases may be rather concentrated. Moreover, it assumes that ions bound to a chain can distribute around a given ion in precisely the same way as free ions, which is probably only acceptable for ions separated by many Kuhn lengths, i.e. chains with low charge densities. Nevertheless, it seems to give qualitatively reasonable results. It is regrettable that the full set of equations has never been solved (which can only be done numerically), so that we do not know what the theory predicts for, for example, off-stoichiometry compositions. Finally, we should realize that the mean-field character of the treatment, which excludes fluctuations, cannot deal with a feature like the formation of soluble complexes. The next step was taken only much later by Castelnovo and Joanny11. These authors again considered mixtures of two polymers (A and B), both of which were considered to have a given linear charge density somewhere between zero and a high value. In addition, they took into account possible incompatibility effects (which lead to segregative phase separation for neutral polymers) by means of an adjustable Flory-Huggins parameter XAB for the interaction between different monomers. A contribution of the polymer's conformational entropy was included as well. The phase behaviour was analyzed in terms of solution stability, i.e. looking for spinodal lines. The theory predicts not only a region of associative phase separation due to electrostatic attraction, but also a region of miscibility and a region of segregative phase separation. As expected, associative phase separation occurs at high charge density and low ionic strength, whereas segregative phase separation occurs at high ionic strength and zero or low charge density. These two regions are separated by an intermediate one-phase region of complete miscibility. Unfortunately, results were again only given for stoichiometric compositions. We reproduce the diagram of states derived by Castelnovo and Joanny in fig. 2.36. 11
M. Castelnovo. J.F. Joanny, Eur. Phys.J. E 6 (2001) 377-386.
POLYELECTROLYTES
2.77
Figure 2.36. Theoretical phase diagram for stoichiometric mixtures of oppositely charged polyelectrolytes with added electrolyte. The parameter t characterizes the incompatibility strength as compared with the polyionic attraction, and s characterizes the salt-induced screening as compared with the polyionic attraction. Attraction is strongest for small values of s and t (lower left corner of the diagram) and decreases for points further away from this corner. (Redrawn from Castelnovo and Joanny, loc. cit.)
Complex coacervation is not restricted to pairs of flexible polyelectrolytes. Mixtures of one flexible polyelectrolyte and a globular protein also separate into dense, complex coacervate and dilute phases with qualitatively similar features. At the molecular level, the structure of such complexes has been investigated by means of various simulations11. 2.6d Polyampholytes Chain molecules with both negative and positive charges along the chain are called polyampholytes. Their behaviour in solution is strongly dependent on the ratio of positive to negative charges and on their distributions along the chain, much in the same way as the behaviour of polyelectrolyte mixtures depend on the stoichiometry. At about equal numbers of positive and negative charges, the attraction between these is the dominant feature, and the polymer collapses and forms an insoluble phase. When there is an excess of either positive or negative charges, chains may partially collapse, but the excess charge prohibits phase separation. Added salt will always counteract phase separation. The distribution of charges along the chain is also important. Strictly alternating polyampholytes (which are difficult to prepare!) are expected to dissolve more easily and 1
R. de Vries, F. Weinbreck, and C.G. de Kruif, J. Chem. Phys. 118 (2003) 4649; R. de Vries, J. Chem. Phys. 120 (2003) 3475.
2.78
POLYELECTROLYTES
be less prone to phase separation. Diblock polyampholytes, with one negative and one positive block, will show a strong tendency to form complexes, which are insoluble around the 1:1 charge ratio, but may form interesting micellar objects (mesophases) when one block is much longer than the other. Random copolymers are an intermediate case. If the dissociating groups are weak acids and bases, the various cases (excess positive, balanced, excess negative) will appear upon variation of the pH; the solubility minimum will be situated around the isoelectric point. 2.6e Polyelectrolyte multilayers The tendency of oppositely charged polyelectrolytes to form insoluble complexes also provides the mechanism underlying the fabrication of so-called polyelectrolyte multilayers. These layers are prepared by exposing a suitable (usually solid) substrate to solutions of polyanions and polycations in an alternating fashion. At each exposure, a finite amount of polyelectrolyte is deposited, and by repeating the alternating treatment one can build films up to tens or hundreds of nanometers in thickness. Many examples can be found in the literature 11 . That the alternating exposure method described above does often produces these multilayers may be surprising considering solubility diagrams like the one given in fig. 2.35. This diagram implies that, provided equilibrium can establish itself, a given quantity of complex coacervate (which forms at compositions within the two-phase region) must redissolve when a composition outside the two-phase region is imposed. Equilibration thus precludes multilayer formation. Indeed, cases have been observed where equilibration and the expected redissolution occur when a weak polyacid (polyacrylic acid, PAA) and a weak polybase polyfdimethylaminoethyl methacrylate), (PAMA) were used, and a sufficient amount of added phosphate buffer was present 21 . In this case, a rapid initial deposition was observed followed by a slow dissolution of material from the surface. An example is shown in fig. 2.37. When the ionic strength due to the buffer was lowered, the redissolution process was suppressed, and it was possible to build up a multilayer. These findings imply that polyelectrolyte complexes are either reversible, i.e. they can either equilibrate because the components are sufficiently mobile, or they are solid-like and, once formed, cannot redissolve upon changing the polyelectrolyte composition in the surrounding medium. Only in the latter situation is multilayering possible. Whether or not reversibility prevails depends on the chemical nature of the polyelectrolytes (e.g. the hydrophobicity of their backbone, the linear charge density, the nature of charged groups, etc.; typically, sulphonates form stronger complexes than carboxylates) and on the concentration and the type of added salt. Unfortunately, data underpinning such dependencies are rather scarce. A combination 11
G. Decher, Science 277 (1997) 1232; P.T. Hammond, Curr. Opin.Colloid Interface Sci. 4 (1999) 430; T.W. Graul, J.B. Schlenoff. Anal. Chem. 71 (1999) 4007. 21 D. Kovacevic, S. van der Burgh, A. de Keizer, and M.A. Cohen Stuart, Langmuir 18 (2002) 5607.
POLYELECTROLYTES
2.79
often used is poly(styrenesulphonate)(PSS)/poly(allylamine)(PAAm), which forms stable multilayers at concentrations of NaCl as high as 1 M. For the multilayering process to proceed successfully, an irreversible system is clearly required. The term 'multilayer' suggests that films prepared by the multilayering process have a well-defined periodic structure. However, neutron reflectivity studies 11 have shown that each deposited layer mixes largely with adjacent layers, so that no periodicity can be discerned at this level. In the cases studied, mixing did not extend to more than about two layers, as could be shown by reflectivity measurements on layers in which the contrast varied with a periodicity of several layers. Liquid-like complexes, in which individual polymers can move freely, would of course be expected to loose all spatial correlation over time.
Figure 2.37. Deposited mass of polymer as a function of time in an experiment where polycation (PAMA) and polyanion (PAA) are supplied at pH 6.8 to a silica substrate in alternating fashion. Curve A is for a low concentration of phosphate buffer (1 mM) and shows buildup of a stable multilayer. Curve B is for 5 mM added phosphate buffer; it is clearly seen that each deposition of PAMA is followed by a spontaneous redissolution process (the adsorbed mass decreases) and no stable multilayer is formed. (Redrawn from Kovacevic et al. toe. cit.)
2.6f Complex coacervate micelles Complex coacervation in mixtures of charged homopolymers leads to the formation of a macroscopic phase. However, if one of the homopolymers is replaced by a diblock copolymer, which has an extra neutral, water-soluble block, a macroscopic phase can be circumvented. Because incorporating the neutral block in the complex would substantially increase the Helmholtz energy, a complex coacervate particle cannot grow in three dimensions. Instead, objects of colloidal size are formed. These can be called complex coacervate core micelles (c.c.c.m.), as they are analogous to the association colloids formed by ordinary surfactants or by diblock copolymers in a selective solvent. J. Schmitt, T. Griinwald, G. Decher, P.S. Pershan, K. Kjaer. and M. Losche, Macromolecules 26 (1993) 7058.
2.80
POLYELECTROLYTES
Figure 2.38 Speciation diagram of particles occurring in mixtures of a polyelectrolyte and a charged/neutral diblock copolymer
Various systems have been reported in which stable micelles of this nature occur11. Evidence shows they are spherical, with a dense core consisting of the two kinds of charged chains in the form of a nearly neutral complex coacervate, surrounded by a more dilute corona of swollen neutral chains. The fact that two kinds of polymers are needed to form such objects implies that the sample composition plays an important role. Indeed, a pure micellar solution only forms at the 'preferred micellar composition' (p.m.c), which is somewhere close to a stoichiometric mixture in terms of charged polymer groups. Apparently, only very few small ions are incorporated. Away from the p.m.c, the number concentration of micelles decreases rapidly as they disintegrate into small soluble complexes carrying an excess positive or negative charge. This can be summarized in the following schematic speciation diagram (fig. 2.38), which presents the weight concentration of a given species as a function of the mole fraction x of either positive or negative monomers, based on the total concentration of charged monomers (i.e. we disregard the neutral monomers). We distinguish five kinds of species: free
11
A. Harada, K. Kataoka, J. Am. Chem. Soc. 121 (1999) 9241; A.V. Kabanov, T.K. Bronich, V.A. Kabanov, K.Yu, and A. Eisenberg, Macromolecules 29 (1996) 6797; S. van der Burgh, A. de Keizer, and M.A. Cohen Stuart, Langmuir 20 (2004) 1073.
POLYELECTROLYTES
2.81
polycations (p+), free polyanions (p-), positive soluble complexes (sc + ), negative soluble complexes (sc-), and complex coacervate core micelles (c.c.c.m). The composition axis can be split into four regions. In the two outer regions, 1 and IV, we have a p+ /sc+ mixture (I) or a p-/sc- mixture (IV), respectively. In the two inner regions, II and III, we have an sc+/ccm mixture (II) or an sc-/ccm mixture (III), respectively. The boundary between region II and III is at the p.m.c, where all polymer is just taken up in micelles. At the boundary between regions I and II, called 'critical excess cationic change' (c.e.c.c), we have a solution containing only sc+, and at the boundary between regions III and IV, denoted 'critical excess anionic charge'f c.e.a.c), we have just sc- in solution. Light scattering titrations have shown that the boundaries between the various regions are fairly sharp and marked by sharp changes in slope of the intensity vs /plots. An example is given in fig. 2.39. Figure 2.39. Light scattering intensity (arbitrary units) of a mixture of poly(methacrylic acid) (negatively charged, M = 113 kg/mol) and poly(dimethylaminoethyl methacrylate-co-poly glyceryl methacrylate) (positive PAMA35 and neutral PGMA100 block), as a function of mole fraction x of methacrylic acid monomers in the mixture, at starting pH 6.8 and 100 mM NaCl. The maximum intensity corresponds to the preferred micellar composition (p.m.c). (Redrawn from S. van der Burgh et al. (loc. cit.)
2.7 Applications of polyelectrolytes Polyelectrolytes are widely used to control the stability of colloidal dispersions. We discuss two specific applications, flocculation and enhanced stabilization. 2.7a Flocculation Many commercial flocculants are often cationic polyelectrolytes. In particular, high molar mass copolymers with a relatively low fraction of charged monomer units and the remaining units water-soluble but uncharged are very effective flocculants. The majority of practical systems where flocculation is required have negative surface charges (e.g. bacteria in active sludge of waste water treatment plants or anionic 'trash' in paper mill waste water); that is why cationic flocculants are preferred. Polyelectrolytes of low charge density are preferred because these do not reverse the surface charge and are thus less
2.82
POLYELECTROLYTES
prone to induce restabilization. Finally, a high molar mass is favourable because the number of bridges a given chain can form between two particles increases with increasing chain length. Some successful flocculants based on polyacrylamide copolymers can have molar masses of several thousand kg/mol. One example is the flocculation of montmorrilonite clay particles by copolymers of (uncharged) acrylamide and (cationic) N,N,N-trimethylaminoethyl chloride acrylate studied by Durand-Piana et al." The optimum flocculation occurred typically for polymers with 1% cationic units and molar mass M > 1000 kg/mol. The effect of molar mass in orthokinetic flocculation can be related to the time-dependent thickness of the adsorbed polyelectrolyte layer, in the same way as discussed in sec. 1.12.e; it is more complicated, though, because salt also influences the relaxation processes in the polymer layer21. 2.7b Stabilization enhancement Highly charged polyelectrolytes have the ability to adsorb strongly to charged surfaces such that the effective charge of the covered particle reverses. Moreover, the adsorption of weak polyelectrolytes (which is of course sensitive to pH) often shows considerable hysteresis upon cycling the ionic strength. For example, when 90% hydrolyzed polyacrylamide (effectively a copolymer of 90% acrylic acid and 10% acrylamide) adsorbs onto cationic latex from 0.5 M NaCl, the adsorbed amount is much higher than when adsorption takes place from 0.002 M NaCI. However, if adsorption is allowed to take place at 0.5 M salt, after which the ionic strength is lowered to 0.002 M, an intermediate adsorbed amount results, which is definitely higher than the one obtained upon adsorption at 0.002 M 31 . Meadows et al. called this 'enhanced adsorption,' and noticed that enhanced adsorption makes colloidal particles much more stable than particles prepared by the usual procedure (adsorption at low ionic strength). Similar enhancement effects occur upon cycling the pH41. The stabilizing effect of dense and strongly adsorbed polyelectrolyte layers is very effective indeed. Meadows found a tenfold increase in critical flocculation concentration for particles with enhanced adsorption. 2.8 General references 2.8a IUPAC recommendation Conductometric Analysis of Polyelectrolyte Solutions. Prepared for publication by H. van Leeuwen, R.F.M.J. Cleven and P. Valenta, Pure Appl. Chem. 63 (1991) 1251. 11
G. Durand-Piana, F. Lafuma, and R. Audcbert, J. Colloid Interface Sci. 119 (1987) 474; L. Eriksson, B. Aim, and P. Stenius, Colloids Surf. A70 (1993) 47. 21 Y. Adachi, T. Matsumoto, and M.A. Cohen Stuart, Colloids Surf. A207 (2002) 253. 31 J. Meadows, P.A. Williams, M.J. Garvey, and R. Harrop, J. Colloid Interface Sci. 139 (1990) 260. 4) J.G. Gobel, N.A.M. Besseling, M.A. Cohen Stuart, and C. Poncet, J. Colloid Interface Sci. 209 (1999) 129-135; C.W. Hoogendam, A. de Kcizer, M.A. Cohen Stuart, and B.H. Bijsterbosch, Langmuir 14 (1998) 3825.
POLYELECTROLYTES
2.83
2.8b Other references Aut. Div. International Symposium on Macromolecules, under auspices of IUPAC, Butterworth (1970). (Contains contributions on polyelectrolytes; old, but not dated, containing papers giving the then state of affairs.) J.-L. Barrat, J.-F. Joanny, Theory of Polyelectrolyte Solutions, in Adv. Chem. Phys. XCIV (1996) 1-66. (Review, 125 refs., emphasizing theory from a physical angle of incidence.) Polyelectrolytes in Solution and at Interfaces, J. Barthel, H. Dautzenberg, D. Horn and W. Oppermann, Eds., Ber. Bunsengesell. 100 Nr. 6 (1996). (Proceedings of the Symposium Polyelectrolytes Potsdam '93'. Containing 62 papers on polyelectrolytes in solution and at interfaces.) M. Borkovec, B. Jonsson and G.J.M. Koper, Ionization Processes and proton Binding in Polyprotic Systems: Small Molecules, Porteins, Interfaces and Polyelectrolytes, in Colloid and Surface Set 16 (2001), E. Matijevic, Ed., Kluwer/Plenum, 99-339. (Very detailed review on electric double layers in various systems, 390 refs.) H. Dautzenberg, W. Jaeger, J. Kotz, B. Philipp, C. Seidel and D. Shtcherbina, Polyelectrolytes; Formation, Characterization, Application, Carl Hanser, Miinchen (1994). S. Forster, M. Schmitz, Adv. Polymer Sci. 120 (1955) 51. M. Mandel, Polyelectrolytes, in Encyclopaedia of Polymer Science and Engineering, 2nd ed., H.F. Mark, N.M. Bikales, C.G. Overberger, and G. Mendes, Eds., Wiley, Vol. 11 (1988) 739. (Extensive review, 314 refs, covering the entire field. Although some features are nowadays better understood, this review still serves as an excellent introduction into the field.) M. Mandel, T. Odijk, Dielectric Properties of Polyelectrolyte Solutions, in Ann. Rev. Phys. Chem. 35 (1984) 75-108. (Review of the state of affairs indicating the many problems that have to be addressed.) T. Radeva, Ed., Physical Chemistry of Polyelectrolytes, Detcher (2001). K.S. Schmitz, Macroions in Solution and Colloidal Suspensions, VCH, New York (1993). Handbook of Polyelectrolytes and their Applications, S.K. Tripathy, J. Kumar and H.S. Nalwa, Eds., Am. Scientific Publishers 2002. (Three volumes: 1. Polyelectrolytebased Multilayers, Self-assemblies and Nanostructures; 2. Polyelectrolytes, their Characterization and Polyelectrolyte solutions; 3. Application of Polyelectrolytes and Theoretical Models.)
2.84
POLYELECTROLYTES
C. Wandrey, D. Hunkeler, Study of Polyion Counterion Interaction by electrochemical Methods, (2002).
3
ADSORPTION OF GLOBULAR PROTEINS
Willem Norde, Jos Buijs and Hans Lyklema 3.1
Introduction
3.2
Structure of globular proteins
3.3
3.2a
Conformational entropy
3.4
3.2b
Interactions that determine the 3D structure of proteins in aqueous solution
3.3
3.4
3.1
3.4
Adsorption of globular proteins from aqueous solution onto (solid) surfaces
3.10
3.3a
3.11
Adsorption kinetics
3.3b
Relaxation at interfaces
3.17
3.3c
Driving forces for protein adsorption
3.19
Adsorption-related structural changes in proteins
3.23
3.4a
How to measure structural properties of adsorbed proteins
3.24
3.4b
General trends
3.30
3.5
Adsorbed amount and adsorption reversibility
3.37
3.6
Influence of some system-variables on protein adsorption
3.40
3.7
3.6a
Protein and sorbent charge
3.40
3.6b
Hydrophobicity
3.41
3.6c
Protein structure stability
Adsorption at fluid interfaces
3.42 3.42
3.7a
Review of some general trends and techniques
3.43
3.7b
Some illustrations
3.45
3.8
Competitive protein adsorption and exchange between the adsorbed and dissolved states
3.52
3.9
Tuning protein adsorption for practical applications
3.55
3.10
General references
3.58
This Page is Intentionally Left Blank
3 ADSORPTION OF GLOBULAR PROTEINS WILLEM NORDE, JOS BUMS AND HANS LYKLEMA
3.1 Introduction This chapter may be considered as a sequel to chapter II.5. Chapter II.5 deals with the adsorption of relatively simple polymers and polyelectrolytes, whose molecules are built up of identical units. In solution, such molecules are rather featureless, having a flexible coily structure, and their adsorption behaviour has been well-modeled. In the present chapter we will discuss the more complicated biopolymers, which are often made up from a variety of monomeric units. For example, proteins are polymers of some twenty-two different amino acids. Because of the variation in physical-chemical properties-mainly in polarity and electrical charge, between the constituting amino acids, protein molecules are ampholytic, are more-or-less amphiphilic, and assume complex three-dimensional structures. As a result, the adsorption of biopolymers is more intricate than that of the simpler polymers treated In chapter II.5. Knowledge of the adsorption behaviour of biopolymers has progressed over recent decades but a unified predictive theory is still far away. However, the discussion of the principles of biopolymer adsorption may start from general trends observed for the adsorption of the simpler polymers, and, in particular, polyelectrolytes. Some main features of polyelectrolyte adsorption may be summarized as follows: (a) When a flexible polymer molecule adsorbs, its conformatlonal entropy is reduced. Hence, for adsorption to occur spontaneously (parts of) the polymer molecule should be attracted by the sorbent surface. Even if the attraction per adsorbing segment of the polymer is only weak the whole polymer molecule may adsorb tenaciously, because many segments adsorb. (b) A flexible, highly solvated polymer molecule typically adopts a train-loop-tall conformation in the adsorbed state, as depicted in fig. II.5.1. The extension and the density of the loops in the adsorbed layer are determined by the solubility of the polymer. A high loop-density is reached only with poor solvents. (c) Because of their polar ionic groups polyelectrolytes are usually well soluble in water. Also, mainly owing to intramolecular electrostatic repulsion, they adsorb with relatively little loop formation in a rather flat conformation.
Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
3.2
PROTEIN ADSORPTION
(d) More often than not, sorbent surfaces are electrically charged and they may electrostatically attract or repel polyelectrolytes. Even in the case of electrostatic repulsion the polyelectrolyte may still adsorb, namely if non-electrostatic attraction of segments outweighs the electrostatic repulsion. For example, polyelectrolytes containing hydrophobic apolar groups may adsorb readily on a (hydrophobic) surface under electrostatically unfavourable conditions. (e) Polyelectrolyte adsorption is strongly affected by the ionic strength. At elevated ionic strengths, charge-charge interaction may be effectively screened and the polyelectrolytes' behaviour approaches that of an uncharged polymer. The pH, which often controls the charge on the polyelectrolyte, and sometimes, that on the sorbent surface, influences polyelectrolytic adsorption in a similar way, at the pH where the polyelectrolyte attains a low charge density it adsorbs in a relatively thick, loopy layer, which is not sensitive to ionic strength variation. (f) In line with paragraphs c, d and e, the variation of the adsorbed amount of polyampholytes (macromolecules containing both anionic and cationic groups) as a function of pH, passes through a maximum at the isoelectric point. The pH-dependent adsorption profile flattens with increasing ionic strength. Polysaccharides, polynucleotides, and unfolded protein molecules that all attain extended structures in aqueous solution, adsorb according to the patterns mentioned above. However, globular protein molecules may deviate strongly from (some of) these principles. The polypeptide chains of globular proteins, to which enzymes, immunoproteins and most transport proteins belong, are folded into a denser structure and, as a rule, contain different structural elements such as a-helices and (3sheets as well as more unordered parts. The architecture of a globular protein is complex and heterogeneous, and highly specific for each type of protein. When interacting with an interface the protein structure may change. The kind of rearrangements depends in a complex way on the interaction with the sorbent surface. Because of the structure-specificity, the adsorption behaviour will differ between different protein species. Consequently, a general theory predicting globular protein adsorption is not to be expected but, based on the advances made during the last few decades, some general principles have been unraveled. Studies of the interaction between proteins and interfaces may help understanding of the mechanism that determines the 3D structure of protein molecules. In addition to this academic interest, protein adsorption has a high practical impact. Examples can be found in biomedical engineering, enzyme immobilization in bioreactors, drug targeting and -delivery, and its use in stabilizers in dispersions in foodstuffs, pharmaceuticals and cosmetics, etc. Conformational changes in protein molecules, which have consequences for their biological functioning, are the most intriguing and challenging aspects of protein adsorption, both from a theoretical and an applied point of view. This matter will be
PROTEIN ADSORPTION
3.3
discussed extensively in sec. 3.4. To understand the behaviour of proteins at interfaces we first have to discuss the main types of interaction that determine protein structure in (aqueous) solution. 3.2 Structure of globular proteins. Proteins are co-polymers built from some 20-22 L-a-amino acids linked together into a linear polypeptide chain. The polypeptide backbone consists of repeating identical peptide units. The structure of a peptide unit is presented schematically in fig. 3.1. Peptide units may interact through H- bonds (C=O ... H-N-) or with other adjoining hydrogenbonding entities, e.g., water molecules. Two of the three bonds in the peptide unit are free to rotate, whereas the C-N bond, shaded in fig. 3.1, is fixed because of its partial double-bond character represented by mesomerism. The side-groups, R, R', ... are specific for the type of amino acid. They vary in size, polarity, and charge. Depending on the distribution of the polar and apolar residues, the protein molecule is more or less amphiphilic and this is one of the reasons why they can be surface active. Some side groups contain cationic groups, others anionic groups. This makes the protein belong to the polyampholytes. Just as an essentially infinite number of words can be written by using the twenty six letters of the alphabet, an endless number of different polypeptides may be formed, characterized by a given sequence of amino acids. This sequence, the so-called primary structure of the protein molecule, rules the interactions in the molecule as well as those between the molecule and its environment. These, in turn, determine the spatial organization, i.e., the three-dimensional (3D) structure adopted by a protein molecule in a given environment. Within the 3D structure, different levels of organization may be distinguished, referred to as the secondary, tertiary, and quaternary structures. The secondary structure is the spatial arrangement of the polypeptide backbone, ignoring the side groups. The tertiary structure refers to the overall topology of the polypeptide chain, and the term, 'quarternary structure' is used for the non-covalent association of independent tertiary structures.
Figure 3.1. Structure of a peptide unit in a polypeptide chain. Two of the three bonds are free to rotate whereas the shaded bond is fixed.
3.4
PROTEIN ADSORPTION
In globular proteins the polypeptide chain folds into a dense structure in which the packing reaches volume fractions of 0.70-0.80. In such compactly folded conformations, the rotational freedom of the bonds in the polypeptide backbone (and side chains) is highly restricted. Such low-conformational-entropy structures are only thermodynamically stable if they are supported by interactions whose the net effect is sufficiently favourable to compensate for the low conformational entropy. Globular protein molecules in an aqueous environment have a few characteristic features: a. They are more or less spherical or ellipsoidal, with dimensions of the order of a few, to a few tens, of nanometers. b. Apolar amino acid residues tend to be accommodated in the interior of the molecule where they are shielded from contact with water. c. Almost all charged groups reside at the exterior of the protein molecule. Charges buried in the low dielectric interior occur as ion pairs. 3.2a Conformational entropy When apolar side groups are rejected from water to form the hydrophobic core of a protein molecule, part of the polar polypeptide backbone is pulled along into the interior. There, because of the absence of water molecules, the peptide units form hydrogen bonds among each other, which stabilizes the ordered structures of the polypeptide backbone. The best known secondary structures occurring in globular proteins include the a-helix and the fS-sheet. Schematic representations of these structures are shown in fig. 3.2. In distinction to a random polypeptide conformation, where two of the three bonds of a peptide unit are free to rotate, all three bonds in cc-helices and (3-sheets are blocked from rotation. This involves a reduction in conformational entropy of 2Rln2 = 11.53 J K"1 per mol of peptide units. Thus, if a polypeptide consisting of 100 amino acids (corresponding to a molar mass of ca. 10,000 Da) folds from an unordered (coily) conformation into a globular protein containing 50% of ordered structure, this goes at the expense of 580 J K^mol" 1 , which, at 300 K, yields an increase in Gibbs energy of 174kJmol~ 1 . Taking into account the restriction in the degrees of freedom of the side groups, the total increase in Gibbs energy upon folding of such a polypeptide may be in the range of a few hundreds of k j per mol. 3.2b Interactions that determine the 3D structure of proteins in aqueous solution Hydrophobic interaction Dehydration of apolar amino acid residues is the main driving force for polypeptide chains to fold into a compact globular conformation. To establish the contribution to the stabilization of a compact structure (taking the completely unfolded, hydrated conformation as the reference state) the distribution of all the amino acid residues and their individual hydrophobicities must be known. The distribution may be obtained
PROTEIN ADSORPTION
3.5
Figure 3.2. Ordered secondary structure elements in polypeptdide chains, (a) a- helix; (b) antiparallel |3- sheet; (c) parallel |3- sheet. from solving the 3D structure (by, e.g., X-ray diffraction and/or NMR) and the individual hydrophobicities can be expressed as the Gibbs energy of partitioning the amino acid between a non-polar medium and water. In this way, it has been inferred that the Gibbs energy of dehydration is about -9.2 kJnm~ 2 per mol 1 '. Based on an empirical relationship between the change in the water-accessible area upon folding and the molar mass of the polypeptide, and assuming that 60% of the interior of the folded protein molecules is made up of apolar amino acid residues21, the compact globular structure of a protein of molar mass 10,000 Da, is stabilized by a Gibbs energy of about SOOkJmor 1 at 25°C. Electrostatic interactions Proteins acquire their charge by (de)protonation of ionic groups in the side groups of the amino acid residues and the a -carboxyl group and a -amino group of both
11 21
F.M. Richards, Ann. Rev. Biophys. Bioeng. 6 (1977) 151. B. Lee, F.M. Richards, J. Mol. Biol. 5 5 (1971) 379.
3.6
PROTEIN ADSORPTION
terminal amino acids. Most of these charged groups are located at the aqueous periphery of the globular protein molecule, whereas in the unfolded conformation, depending on the primary structure, the charged groups are more or less homogeneously distributed over the fully hydrated expanded coil. The Gibbs energy of dissociation, AdissG , of a proton from any particular group can be split into a chemical ('chem') and an electrical ('el') term A
diss G = A diss G chem + A diss G el
I3-2- * 1
The chemical term contains the intrinsic contribution to the dissociation, and the electrical term the additional electrical work to remove the proton from the charged site to infinity (where the electric field strength is zero). In the simple case where all titratable groups belong to one and the same class Q0 = azFN
[3.2.2]
where Qo is the total charge of the sample, N the number of titratable groups, and a the degree of dissociation11. For the dissociation constant K diss the following expressions apply
p K diss=p H + i °g : 4r
l3-2 31
'
and AG° = -RTlnK d i s s
[3.2.4]
Combining [3.2.1], [3.2.3] and [3.2.4] yields pH = 0.434 AdissGc°hem + AdissGe°l _ j F RT
j B
^
[3 2 5 ]
a
It should be realized that A diss G° is a function of a ; at a high degree of dissociation, when the surface has attained a negative charge density and, consequently, a high negative potential, it requires more electrical work to remove a proton from the surface. The quantity AdissG°[ is often expressed as diss el =_2WE0_ RT F
=
_2WazN
[3.2.6]
in which W is the so-called electrostatic interaction factor. It depends primarily on the dielectric constant and the ionic strength of the medium. In differential form, [3.2.6] may be written as
The titration behaviour of polyelectrolytcs has been discussed in sec. 2.2d.
3.7
PROTEIN ADSORPTION
^ ^ - = -0.868 WzN + 0.434 da all-a)
[3.2.7]
M L °434 dg>0 zNFa(l-a)
[3.2.8]
or 0.868^ F
The quantity d p H / d Q 0 reaches a minimum value for a = 0.5 which is reflected by an inflection point in the titration curve for Qo versus pH. Proteins contain more than one class of titratable groups. Different classes j of groups are titrated in distinct pH regions. By way of example, the number of titratable groups, together with their intrinsic pK°-values for the protein ribonuclease (RNase), are given in table 3.1. For systems containing more than one class of titratable groups [3.2.8] has to be modified into *& = ^ dg>0 FljJVjZjajd-aj)
0.868 ?L F
[3.2.9]
and the differential titration curve, d p H / d g 0 versus QQ (or versus pH) displays more than one minimum, as is shown for RNase in fig. 3.3. Adsorption of ions other than protons can also lead to charging of the surface, provided that the adsorption is specific. The term, 'specific' implies that the adsorption
Table 3.1. Titratable groups and their intrinsic dissociation constants (pK°'s) in ribonuclease. Group
Number
pK°
Group
Number
pK°
a -carboxyl
1
3.75
e -amino
10
10.2
(5-, y -carboxyl
10
4.0-4.7
phenolic OH
3
10.0
imidazole
4
6.5
3
inaccessible
a -amino
1
7.8
4
> 12
guanidyl
Figure 3.3. Differential proton titration curve for bovine pancreas ribonuclease in 0.05 M aqueous solution. (Redrawn from W. Norde, J. Lyklema, J. Colloid Interface Sci. 66(1978) 266.)
3.8
PROTEIN ADSORPTION
Figure 3.4. Schematic picture of a globular protein molecule in solution. Shaded areas indicate hydrophobic regions. Charged groups originate from (de)protonation of amino acid residues (+/-) and from specific ion adsorption ( ©/© )• The dashed envelope iindicates the slip plane. Within this envelope the charge is Q e k . The compensating diffuse charge is not drawn.
Figure 3.5. Proton charge Qo and electrokinetic charge S e k of the same RNase as in fig. 3.3. The point of zero charge (p.z.c.) and isoelectric point (i.e.p.) are indicated. (Redrawn from W. Norde, J. Lyklema, J. Colloid Interface Sci. 66 (1978) 277.)).
forces are partly non-electric so that these ions can overcome the repelling electric potential at the surface and, by their adsorption, even increase the surface electric potential. In practice, often the electrokinetic charge @ek is measured, which results from both (de)protonation of amino acid residues and specific ion adsorption. This situation is depicted schematically in fig. 3.4. Figure 3.5 shows Q0(pH) and gek(pH) for RNase. The charge associated with specific adsorption of ions other than proteins is given by ( 9 e k - Q 0 ) . The difference between the point of zero charge (p.z.c.) and isoelectric point (i.e.p.) is a measure of the extent of specific adsorption. The protein charge is neutralized by countercharge. The electrical part of the Gibbs energy of a given charge-distribution can be evaluated as the reversible, isothermal, and isobaric work of charging the system. This is given by
Gel= J 9=0
y/'dQ'
[3.2.10]
PROTEIN ADSORPTION
3.9
where y/' is the electrostatic potential at the surface and Q' is the charge of the protein molecule during the charging process. To solve [3.2.10] a model for the charge distribution is required that relates y/ to Q . Some of such models are presented in chapter II.3. Away from the isoelectric point, where the protein surface contains a significant excess of negatively or positively charged groups, the charge may be thought of as being smeared out and the Gouy-Stern model may be applied. At, and near the isoelectric point, a model accounting for discrete charge effects is required to yield realistic results. Assuming a smeared-out distribution, and complete penetration without restraint of counterions in the expanded, unfolded polypeptide, the Donnan model may be a reasonable choice. The difference between the Gibbs energies for the compact and unfolded conformations, determines the stability of the one conformation relative to the other. It may be clear that, in the isoelectric region, a homogeneous charge distribution resists unfolding, whereas the opposite is true away from the isoelectric point. Values for the difference in Gibbs energies between the globular and the unfolded states are typically in the range of a few tens of k j per mole of protein, the effect becoming smaller at higher ionic strength. If ions reside in the interior of the protein they do so in pairs, and do not contribute significantly to the potential. In the unfolded state, such ion pairs are disrupted. Disruption of ion pairs is electrostatically unfavourable, but this effect is more or less compensated by hydration of the isolated ionic groups. For this reason, together with the fact that protein molecules usually contain no more than a few ion pairs, ion pair formation does not contribute much to the (de)stabilization of a compact protein structure. Lifshits-Van der Waals interaction Lifshits-Van der Waals interactions arise from forces between fixed and/or induced dipoles. For isolated molecules they decay steeply with the separation distance between the interacting dipoles, scaling as r"6 . Folding of the polypeptide chain into a compact protein involves the loss of Lifshits-Van der Waals interactions between units of the polypeptide and water molecules but, at the same time, such interactions are formed within the protein molecule and between water molecules. This is the Archimedes principle, discussed in sec. 1.4.6b. Quantitatively, the overall effect is not well understood for proteins. Because of the relatively high packing density in globular proteins, Lifshits-Van der Waals interactions are likely to promote a compact structure. The difference between the Hamaker constant of a protein and pure water points in the same direction. However, because of the opposing contributions of disruption and the creation of contacts, the net contributions of Lifshits-Van der Waals interaction to the stabilization of the protein structure is relatively small. Hydrogen bonds In the fully hydrated unfolded state, peptide units and some other moieties in the
3.10
PROTEIN ADSORPTION
amino acid side groups interact with water molecules through hydrogen bonds. When the protein folds into a compact structure, many of these bonds are broken and, instead, intramolecular hydrogen bonds are created in the protein molecule, and hydrogen bonds between water molecules. Peptide units dominate intramolecular hydrogen bonding, and determine the unique and fixed structures of helices and (3 sheets. Although hydrogen bonding is essential for maintaining ordered structures in the protein's Interior, its net contribution to the stabilization of the protein conformation is not particularly clear. Model studies1 2) suggest that peptide-water hydrogen bonds are more favourable than peptide-peptide and water-water hydrogen bonds. This would imply that hydrogen bonding (weakly) favours the unfolded conformation. However, hydrogen bonding between peptide units that, by the action of other factors (e.g., hydrophobic interaction) are forced into the non-aqueous inner parts of the protein, strongly stabilize ordered structures. Bond lengths and angles When the polypeptide chain folds into a tightly packed structure the bond lengths and angles may be more or less distorted. This could oppose stabilization of the globular protein structure by several kJ per mole31. In summary, the protein structure and structure stability is determined by various interactions inside the protein molecule, between the protein and water, and between the water molecules. These interactions compete with each other. As a result, the one structure is usually only some tens of kJ per mole more stable than the other. Hydrophobic interactions and the conformational freedom of the polypeptide chain play the leading parts. Both are mainly of entropic nature. Hence, the resulting protein structure is the outcome of an entropy battle between the polypeptide chain and water. Because of this marginal thermodynamic stability, none of the other factors influencing protein folding is unimportant. Consequently, even subtle changes in the environment of the protein, e.g. pH, ionic strength, ionic specificity, temperature, or additives, may cause structural transitions in the protein. In view of this, it is not surprising that most proteins, under most conditions, change their structure when adsorbing from solutions onto an interface. 3.3 Adsorption of globular proteins from aqueous solution onto (solid) surfaces The overall protein adsorption process is schematically depicted in fig. 3.6. Several steps or stages, indicated by the numbers in fig. 3.6, may be distinguished:
11
G.C. Kresheck, I.M. Klotz, Biochemistry 8 (1969) 8. W. Norde, Adv. Colloid Interface Sci. 25 (1986) 267. 31 M. Levitt, Biochemistry 17 (1978) 4277. 21
PROTEIN ADSORPTION
3.11
Figure 3.6. Schematic presentation of the protein adsorption process. Further explanation is given in the text. (1) transport from the bulk solution Into the subsurface region from where it is (2) attached at the surface. After initial adsorption the protein may (3 and 3*, etc.) relax at the sorbent surface. Relaxation may involve a fast surface-induced conformational change, but a slow spreading to optimize protein-sorbent interactions is also a very general phenomenon. Protein molecules may desorb from the sorbent surface (4, 4*, 4**). Obviously, the rate-constant of desorption is the smaller when the adsorbed protein molecule has reached a greater extent of relaxation. After desorption, the protein molecules (back in solution), may or may not (completely) regain their original native conformation. If not, the solution will eventually contain structurally perturbed protein molecules that may re-adsorb with an affinity different from the native ones. In this chapter we shall discuss all these aspects of the overall adsorption process in some detail, with special attention to adsorption-related conformational changes. In sec. 3.9 the chapter will conclude with some remarks on sorbent surface modification in order to adapt protein adsorption for practical applications. 3.3a Adsorption kinetics The rate of adsorption comprises two consecutive steps, (a) transport of the molecules to the interface and, (b), attachment at the sorbent surface. These steps will be considered separately and combined thereafter to derive an equation for the overall rate of adsorption. Transport towards the sorbent surface. The basic mechanisms of transport are diffusion and convection. The latter may be by laminar or turbulent flow. When the protein adsorbs, the subsurface region of the
3.12
PROTEIN ADSORPTION
solution becomes depleted and a protein concentration gradient is built up. This causes a flux J of protein molecules from the bulk solution towards the subsurface region. Under steady state conditions the flux is given as J = ktr(cb-ca)
[3.3.1]
where k^. is a transport rate constant, which depends on the transport mechanism and the dimensions and orientation of the protein molecules, and c b and cs are the protein concentrations in the bulk solution and the subsurface zone, respectively. In static systems, where the transport is by diffusion only, k^ may be approximated by [D/nt)l/2 in which D is the diffusion coefficient of the protein in bulk solution and t is the time1'. Clearly, the flux reaches a maximum value when cs = 0, that is, when attachment at the interface is much faster than transport to the subsurface region. This is the case when there is no activation energy for attachment, and in the initial stage of the adsorption process when there is no limitation of surface area available for adsorption.
Figure 3.7. Identification of ka and fcd .
Exchange at the interface. A simple situation is depicted in fig. 3.7. The protein molecules attach at, and detach from, the sorbent surfaces. This gives two fluxes from the subsurface region to the surface: one forward, dr/dt\+ , and one backward, dr/dt\_ , where the adsorbed amount r may be expressed in mass per unit of sorbent area. The net adsorption rate is given by, d/7dt = d/7dt| + -d/7dt|_
[3.3.21
The forward flux scales as cs and is proportional to the fraction (1 - 0] of sorbent surface area that is unoccupied with adsorbed protein. Hence, d/7dt| + =k a (l-0)c s
[3.3.3]
in which fca is the attachment rate constant. For the simple case that the adsorbing
11
For a general discussion and derivation, see sec. I.6.5e.
PROTEIN ADSORPTION
3.13
molecules do not change their sizes and shapes, as in fig. 3.7, the degree of surface occupancy 9 is defined as 77 .T 6 ^, where T"681 is the adsorbed amount when the sorbent surface is saturated with protein. For adsorption-induced conformational changes, as happens with most proteins and other (bio)polymers, the relationship between F and 6 is more complicated, and will be discussed later. The value of fca is lowered by any barrier for attachment, e.g., electrostatic repulsion, or hydration effects. A repulsive barrier may be created or reduced by the pre-adsorbed protein molecules and, hence, ka may vary with 0 and, therefore, with time. For the backward flux we can write
^j
= *d*
[3.3.4]
where kd is the detachment rate constant. In the course of the adsorption process, 6 increases and eventually a dynamic equilibrium (steady state) is reached, for which cb = cs = c
, the equilibrium concentration in solution. The equilibrium is character-
ized by fca(l-0)ceq =
fcd0
[3.3.5]
Assuming that d 77 d t_ is exclusively determined by kd and 6, the rate of adsorption off-equilibrium is still given by ka (1 - 0)c , so that ^
= ka(l-0)(cs-ceq)
With dT/dt
[3.3.6]
= J . Combining [3.3.1] and [3.3.6] gives
^ -
;b'Cef
dt
1
|
[3.3.7,
1
Polymers, including proteins, usually adsorb with a high affinity for the sorbent surface, characterized by an extremely high value of the adsorption equilibrium constant K (= ka/kd). For physical reasons, ka cannot attain extremely high values, and hence fcd must be very small. The reason for the almost infinitely low polymer desorption rate has been discussed in chapter II.5. High affinity adsorption isotherms (see fig. II.2.24) have the property that, below adsorption saturation, c is extremely low (usually below detectability). When reaching saturation, which is reflected by the (semi-)plateau in the isotherm, c (6) increases steeply and approaches c b . It follows that, upon approaching saturation, dF/dt drops strongly. This is reflected by a sharp transition in the curve for F(t) often observed for (bio)polymers. A typical example is given in fig. 3.8. Away from saturation, where c
= 0 , [3.3.7] may be rewritten as
3.14
PROTEIN ADSORPTION
Figure 3.8. Typical example of polymer adsorption kinetics.
c
b/rmax_
dO/dt
1 ,
ku
1
[3.3.8]
fca(l-0)
If fcjj. « ka(l-ff), dO/dt or, for that matter, dF/dt are determined by transport from the bulk to the subsurface region, independent of 0. This permits direct derivation of k^ from the initial linear part of the isotherm. If kb » ka (1 - 9), a condition that is probable when there is a barrier for attachment at the surface or when the surface is covered to some extent, dF/dt is determined by the rate of attachment, and hence, depends on 8. The value of ka(6) may be derived from F(t), using approximation [3.3.8]. So far, the discussion has been based on the oversimplified model shown in fig. 3.7. However, it is a rule rather than an exception that biopolymers, including proteins, adapt their conformations as a result of relaxation at the interface. This phenomenon is illustrated in fig. 3.9, where the native conformation is denoted, the 'N state' and the adapted, perturbed conformation the 'P state'. The rate-constant for relaxation is kT and those for desorption from the N state and P state, kd N and kd p . It follows that
d#M ^ f =
fc
a,Ncsa-8N-0P)-
kdN0N - kT9N
[3.3.9]
and d(9p
[3.3.10]
Figure 3.9. Adsorption and desorption of protein molecules followed by transition from the native (N) state to the perturbed (P) state. After desorption of a molecule in the P state the molecule may, or may not, return to the N state.
PROTEIN ADSORPTION
3.15
In the steady state, where dF/dt = 0 , k a V ^ N - ^ e q =fcd,N^N+
fc
d,P^P
I3-3-11'
In general, kdN *kdp. Now, contrary to the simple situation of fig. 3.7, the rate of desorption is not a unique function of the desorption rate constants but it also depends on 6P / 0^ . This ratio changes in the course of the adsorption process. As a result of surface relaxation it is expected that fcd p < kd N . The smaller the kd p / fcdN ratio, the larger the fraction of perturbed molecules in the steady state is. If kd p approaches zero, the adsorbed layer eventually consists of molecules that are all perturbed. It goes without saying, that when adsorption induces conformational changes in the protein, the expression for dF/dt is much more complicated than the one given in [3.3.7]. The adsorption kinetics are even more complicated when the increase in area per protein molecule ('footprint') due to spreading at the sorbent surface is taken into account. Let the molecular area of N state molecules be a N , and of P state molecules a p , with a p / aN = a r > 1. The maximum number of molecules in the N state per unit sorbent surface area is No , and in the P state it is JV. It follows that JV0 = aTN, and N o a N = 1. Further, 0N = nN / JV0 and 0V =np/N0, where n is the number of protein molecules in the adsorbed state. Hence, dft, -jf =
fc
a,Ncs(1-^N-aA)-'cd,N%x/(ar-%.ep)
I 3 - 3 - 12 !
in which /(a r ,# N ,i9 p ) takes into account the fact that the area-enlargement involved in the N —» P transition reduces the probability for other molecules to undergo the transition later. When a r = 1, _f (ar, #N, £?p) = 1 and it decreases for larger values of a r , the more so, the higher is the surface occupancy, {0N + 0p). Similarly, ^
= kr0Nx/(ar,0N,0p)-kdp0p
[3.3.13]
The steady state evolving from this model is given by k 0 a,NV- N
-aA)ceq
=k
d,fA
+fc
d,P0P
[3.3.14]
which is equivalent to [3.3.11] because (1-# N -ar8p) and (1-# N -&p) in the respective equations are the covered fractions of the sorbent surface. Both 0N and 6>p are expressed in number of adsorbed molecules per fixed number of sites per unit surface area. It means that (#N + 0p) is proportional to the adsorbed mass per unit surface area, F. This model explains the remarkable phenomenon of transient adsorption of polymers which has sometimes been reported. If fcdp<)c(iN, a r > 1 and the rates of attachment and spreading are of comparable magnitude, (#N + 0p) and, hence, F would pass through a maximum in the course of the adsorption process.
3.16
PROTEIN ADSORPTION
Figure 3.10. Sketch of excluded area for the deposition of molecules according to the random sequential mode.
Most theories for (blo)polymer adsorption, including proteins, start from the models presented above. Other models proposed to describe the kinetics of (bio)polymer adsorption include the random sequential adsorption (RSA) model11. In this model it is assumed that the molecules arrive randomly at the sorbent surface and that they stick where they hit. A subsequently arriving (spherical) molecule cannot be accommodated within the dashed areas around pre-adsorbed molecules, as illustrated in fig. 3.10. Clearly, the fraction 0 of the surface which is available for adsorption is a function of the degree of coverage, 0, of the surface by the adsorbate. For hard sphere molecules, 0=rmR2 (where n is the number of molecules per unit surface area, and R is the radius of the sphere. The RSA theory produces the following functionality
^=«p{ 2 -^ + i^ + i I n a - f l ) + -}
[3 3 151
--
which can be developed into 0(0) = l-46> + ^ 3 -6^+2.4243 ft + ....
[3.3.16]
K
Figure 3.11. Surface area available for the adsorption of spherical molecules in a random sequential mode. The drawn curve obeys [3.3.15], the other ones [3.3.16] with two, three or four terms of the expansion. 11
P. Schaaf, J.C. Voegel, and B. Senger, Ann. Phys. 23 (1998) 1.
PROTEIN ADSORPTION
3.17
The third term of the r.h.s. of [3.3.16] accounts for the overlap of the dashed areas in fig. 3.10, the fourth term for the double overlap, etc. Figure 3.11 shows curves for 0(9). In the kinetics according to the RSA model, the uncovered surface area (1 - 9) is replaced by
[3.3.17]
Curves for 9{t) are presented in fig. 3.12 for spheres and ellipses, and show that the saturated surface coverage decreases with increasing aspect ratio. Even with spheres it does not exceed fifty percent.
Figure 3.12. Adsorption kinetics according to the random sequential mode for spheres and ellipsoids of varying aspect radios, alb. (Redrawn from S.M. Ricci, J. Talbot, G. Tarjus, and P. Viot, J. Chem. Phys. 97 (1992) 5219.)
3.3b Relaxation at interfaces After attachment, the protein molecule will optimize its interaction with the sorbent surface, i.e., it relaxes. In the relaxation process the protein molecules expose an increasing number of segments to the surface, which usually leads to some degree of spreading. This involves rearrangements in the protein structure. The extent of spreading depends on the rate of spreading relative to the rate of attachment at the surface. More precisely, the extent of spreading is primarily determined by the ratio of the characteristic time of relaxation, rr , (which, in turn, is mainly governed by the internal cohesion in the protein molecule) and the characteristic time of filling the sorbent surface, rf (which is determined by the protein flux J towards the surface). If relaxation occurs much faster than filling the surface i.e., (rt /r f « 1) all molecules are allowed to relax completely after being attached, and the adsorbed amount under saturation conditions, P*81, is independent of J. Under most conditions, such behaviour is expected for polymers that relax quickly. However, in globular proteins the internal coherence is strong, and adsorption-induced relaxation relatively slow.
3.18
PROTEIN ADSORPTION
Then, the time of filling the surface, may approach the relaxation time (rt / rf > 1), and the degree of spreading will be affected by the protein flux, because a neighbouring site may already be occupied by a newly arriving molecule before the previously adsorbed molecule is given the time to relax. This results in less spreading, which is reflected in a higher value of r 6 ^ with increasing flux. The adsorbed layer may become heterogeneous with respect to the conformational states of the protein population. Molecules arriving at an early stage find sufficient area available for spreading, whereas this is much less the case for the molecules that arrive when the surface is already crowded.
Figure 3.13. Relaxation of protein molecules, adsorbed on silica. Surface area occupied per adsorbed IgG molecule as a function of the time required to fill the surface. (Based on data of M.G.E.G. Bremer, Immu.noglobu.lin Adsorption on Modified Surfaces, PhD thesis, Wageningen University, The Netherlands (2001).)
When the surface area occupied per adsorbed molecule (derived from / ^ a t ) is plotted vs. Tj (calculated from the flux), a curve is given, as shown in fig. 3.13. For Tt = 0, spreading is completely inhibited and the corresponding area may be compared with the dimensions of the native molecule. Extrapolation to constant r 8 ^ , where full relaxation is reached, allows estimation of the relaxation time. Thus, relaxation times in the range of tens- to thousands- of seconds are inferred for globular proteins. There is strong experimental evidence11 that at interfaces, as in solution, the conformational change of globular protein molecules is a distinctly co-operative, rather than a gradual process. It implies that the intermediate part of the curve in fig. 3.13 reflects the co-existence of native (N) and spread (P) molecules at the interface. Assuming a two state transition N —> P , the rate of spreading can be expressed as,
11
T. Zoungrana, G.H. Findenegg, and W. Norde, J. Colloid Interface Sci., 190 (1997) 437.
PROTEIN ADSORPTION ^
= -(cscN
3.19 [3.3.18]
where ks , the spreading rate constant, is determined by the activation Gibbs energy AG^ of the N —> P transition as
in which h is Planck's constant. Fitting the experimental curve to [3.3.18] yields estimates for ks and, hence, AG^ . For globular proteins, typical values for ks are in the range of lO^-lO" 1 s" 1 , and for AG+ some tens of kT . 3.3c Driving forces for protein adsorption In this section, the main contributions to protein adsorption on a smooth, rigid surface will be discussed. These contributions originate from (a) redistribution of charged groups (ions) when the electrical double layers around the protein molecules and the sorbent surface overlap, (b) dispersion forces between the protein and the sorbent material, (c) changes in the hydration of the sorbent surface and the protein, and (d) structural rearrangements in the protein molecules. Although these interactions are discussed separately, they are by no means independent of each other. Their actions could be synergistic or antagonistic. Redistribution of charged groups. As a rule the surfaces of the protein molecules and the sorbent are electrically charged, and surrounded by counterions and co-ions, together constituting the countercharge that neutralizes the surface charge. The surface charge and the countercharge together form an electrical double layer. Models for the electrical double layer are discussed in chapter II.3 and their interaction in chapter IV.3. There, it has been explained that the Gibbs energy Gcd required to invoke a charge distribution can be calculated as the isothermal, isobaric reversible work o° G c d = J y/°<&o°'
[3.3.20]
0
where y/°' and cf" are the variable surface potential and surface charge density, respectively, during the charging process. Equation [3.3.20] is the 2D equivalent of [3.2.10]; Gel in [3.2.10] is in J, whereas Gcd in [3.3.20] is in J m" 2 . As before, [3.3.20] can be integrated if a model for the double layer is available. To calculate the contribution from charge redistribution to the Gibbs energy of the adsorption process, AadsGcd , equation [3.3.20] has to be applied three times, i.e., Gcd for the bare surface and the dissolved protein has to be subtracted from Gcd for the protein-covered sorbent surface. For the sorbent surface, the Gouy-Stern model may be taken,
3.20
PROTEIN ADSORPTION
although for the protein molecule, a discrete charge model seems to be more appropriate. For the protein-covered sorbent charge-distribution models have been proposed by Norde and Lyklema11 and by Stahlberg et al.2). The protein molecules in the adsorbed layer retain a compact conformation (sec. 3.4). Under most ambient conditions of ionic strength the thickness of the adsorbed layer exceeds the Debye length, i.e., the distance over which electrostatic forces are effective. Byway of example, adsorbed protein layers usually reach a thickness of a few to a few tens of nm, whereas the Debye length in 0.01 M ionic strength is 3 nm, and in 0.1 M ionic strength it is only 1 nm. Therefore, the protein layer shields the protein-sorbent contact region from electrostatic interaction with the solution. To prevent an excessively high electric potential, the charge density in the non-aqueous, low dielectric, protein-sorbent contact region must be regulated to be essentially zero. Cross-differentiation allows calculation of the charge regulation from the dependency of the protein adsorption on the electrolyte concentration in solution. Charge regulation may occur through changes in the ionization of the protein and/or the sorbent surface, and also by the incorporation of indifferent ions from solution in the protein-sorbent contact region. As a consequence of the charge regulation, AadsGcd does not exceed a few tens of RT per mole of protein, and it is not very sensitive to the charge on the protein and the sorbent surface before adsorption. In addition to contributing to the charge regulation, ion transfer between the solution and the adsorbed layer includes a chemical effect. Compared to water, the low dielectric proteinaceous environment is a poorer 'solvent' for individual ions and, hence, the chemical effect of ion incorporation opposes protein adsorption. This explains why protein adsorption often reaches its maximum affinity when the charge density on the protein just matches that on the sorbent surface, so that no additional ions have to be incorporated to neutralize the protein-sorbent contact region. Dispersion interaction Dispersion interaction between macroscopic bodies may be computed using the Lifshits theory explained in sec. 1.4.7. There, the less accurate, but simpler microscopic treatment, known as the Hamaker-de Boer approximation, is also presented. The latter approach has the advantage of leading to relatively simple analytical expressions at the expense of less rigour. As is often the case, (adsorbed) protein molecules, and sometimes the sorbent material, are not precisely defined with respect to geometry and density, so application of the Hamaker-de Boer theory is acceptable. Then, for a sphere 1 (the protein molecule) interacting with a body 2 having a planar surface (the sorbent surface), across a medium 3, the contribution to the Gibbs energy of adsorption is given by, see [1.4.6.19 and 29] 11 21
W. Norde, J. Lyklema, J. Colloid Interface Set, 55 (1978) 285. J. Stahlberg, B. Jonsson, Anal. Chem. 68 (1996) 1536.
PROTEIN ADSORPTION
A
adsGdisp-
l 3)
6 [h
3.21
+
h + 2a + lnh + 2a\
l3 3
' -211
where A12(3) is the Hamaker constant for the system, a the radius of the sphere, and h the shortest separation distance between the sphere and the planar surface. The quantity A ads G di off
steeply with
increases with increasing dimensions of the sphere, and it drops increasing separation
distance:
A ads G di
reacnes
significant
magnitudes only for small values of h. For h « a , [3.3.21 ] reduces to I3322'
AadsGdiSp = ~ ~ ^ The Hamaker constant
A12(3) for the system can be derived from those of the
individual components, according to [1.4.6.29] A
12(3) = A2 ~ A 13 ~ A 23 + A33
[3.3.23]
When the Berthelot principle is valid, this may also be written as A -tAl/2 Al/2\jAl/2 A l/2\ 12(3)-i A ll ~ A 33 )yh.2 ~ A 33 j
A
,0 o 041 [J.J.^4]
In water, usually A,x > A 33 and A22 > A33 , and hence -A12(3) > 0 , so that AadsGdi < 0 , which implies attraction between 1 and 2. The Hamaker constant for interaction across water is ca. 6.5xlO~ 21 J for globular proteins, (l-3)xlO~ 19 J for metals and (4-12) xlCT21 J for synthetic polymers1'21. Based on these data, and applying [3.3.24] and [3.3.22], for a spherical protein molecule of radius 3 nm at a distance of 0.1 nm from the sorbent surface, AadsGdi at room temperature amounts to 6-1IRT per mol at a metal surface, and to 1-4 RT at a polymeric surface. Clearly, these estimates are semi-quantitative. Obtaining more accurate estimates requires more detailed knowledge of the system's parameters, such as Hamaker constants, sizes and shapes of the protein molecule and the sorbent material. Hydration changes Polar molecules attract water molecules, mainly through hydrogen bonding. They compete successfully with hydrogen bonds between the water molecules, so they are readily soluble in water. Apolar groups do not offer the possibility of a favourable interaction with water and therefore they are expelled from an aqueous environment. This is the hydrophobic effect. The water molecules at an interface of apolar material are strongly oriented so as to form as many hydrogen bonds as possible to other water molecules, as none can be formed with the apolar material. This reduces the entropy of the water adjacent to apolar compounds. 11
S. Nir, Progr. Surface Set 8 (1977) 1.
21
J . Visser, Adv. Colloid Interface Sci. 3 ( 1 9 7 2 ) 3 3 1 .
3.22
PROTEIN ADSORPTION
Protein molecules contain both polar and apolar groups. As discussed in sec. 3.2, for globular proteins in aqueous solution the apolar parts tend to be hidden inside the molecule, whereas the exterior is mainly polar. Other interactions and/or geometrical constraints interfere with these tendencies. Hence, although it is mainly apolar, the heart of the protein contains polar groups as well, and apolar groups occupy parts of the aqueous periphery of the molecule. However, water-soluble non-aggregating protein molecules do not have pronounced apolar patches at their surfaces. The surfaces of (model) synthetic sorbent materials are usually less complex than protein surfaces, but natural surfaces may also be rather heterogeneous. When the surfaces of the protein molecules and the sorbent are predominantly polar it is likely that some hydration water is retained between the sorbent surface and the adsorbed protein layer. However, if (one of) the surfaces are (is) apolar, dehydration would stimulate protein adsorption. The contribution from changes in hydration to the Gibbs energy of adsorption may be estimated from the partitioning of model compounds between water and a nonaqueous medium. It is thus estimated that dehydration of an apolar surface lowers the Gibbs energy by 10-20 mJ m~2 . In the case of a protein of 30,000 Da and an adsorbed mass of 1 mg m~2, this corresponds, at room temperature, to AadsGh d r ranging between -120 RT and -240 RT per mole of protein. It evidently indicates that apolar dehydration outweighs the contributions from overlapping electrical double layers and Van der Waals forces. Rearrangements of the protein structure When a protein molecule makes intimate contact with the sorbent surface, on one side of the molecule the aqueous environment is replaced by the sorbent material. As a consequence, intramolecular hydrophobic interaction becomes less important as a structure-stabilizing factor; i.e., apolar parts that were buried in the interior of the dissolved molecule may become exposed to the sorbent surface without making contact with water. Because hydrophobic interaction between amino acid side groups in the protein's interior support the formation of secondary structures, such as a-helices and |3 -sheets, a reduction of this interaction destabilizes such structures. A reduction in the a-helix and/or |J-sheet- content is, indeed, expected to occur only if peptide units released from these structures can form hydrogen-bonds with the sorbent surface, as is the case for oxides (e.g., glass, silica and metal oxides), or with surfaces retaining residual water molecules at the sorbent surface. Then, the decrease in secondary structure may lead to an increased conformational entropy of the protein. This may favour the adsorption process by tens of RT per mole of protein. If, in the non-aqueous protein-sorbent contact region, it is not possible for the peptide units to form hydrogen bonds with the sorbent surface, as is the case for hydrophobic surfaces, adsorption may induce the formation of extra intramolecular peptide-peptide hydrogen bonds, thereby promoting the formation of a-helices and/or p -sheets. Thus, whether adsorp-
PROTEIN ADSORPTION
3.23
tion on a hydrophobic surface results in an increased or decreased order in protein structure, depends on the subtle balance between energetic and entropic interactions. To summarize, it is concluded that protein-sorbent interaction is a complex phenomenon, comprising various subprocesses operating simultaneously. For example, hydrophobic interaction between the protein and the surface requires close contact between the two, which may be optimized by structural rearrangements in the protein. This may involve increased or decreased flexibility along the polypeptide chain. At hydrophilic surfaces, increased conformational freedom is expected in adsorbed protein molecules. Based on these considerations we arrive at the following 'rules of thumb'. Because of favourable dehydration, essentially all proteins adsorb on hydrophobic surfaces, even under electrostatically adverse conditions. With respect to hydrophilic surfaces, distinction may be made between proteins having a strong internal coherence ('hard' proteins) and those with a weak internal coherence ('soft' proteins). The hard proteins adsorb only if they are electrically attracted, whereas in the soft proteins the surfaceinduced conformational entropy gain is large enough to cause adsorption on a hydrophilic, electrostatically repelling surface. 3.4 Adsorption-related structural changes in proteins While searching for explanations of the clotting of blood, in the 1950s Leo Vroman discovered that water-repellent surfaces became water-wettable after contact with plasma. Later he discovered that some, but not all, of the proteins involved in blood clotting were adsorbed to hydrophobic surfaces. Together, these observations allowed him to propose that, 'globular proteins which can open easily will do so when they see a hydrophobic surface, and will turn themselves inside out to paste themselves with their fatty hearts onto that surface'11. About twenty years later Norde performed an extensive thermodynamic analysis of the protein adsorption process. This study confirmed that human plasma albumin changes its conformation when it adsorbs at polystyrene surfaces2'. This adsorption process was entropically driven, and part of the entropy gain originated from the conformational changes in the proteins31. Indications that proteins might change their conformation when they adsorb to air/ water and oil/water interfaces have been reported since the early part of last century. Information on the conformational state of proteins at the air/water interface was obtained by studying interfacial pressures and tensions using, for example, the Langmuir trough (III.3.3) and the Wilhelmy plate (III. 1.8a) techniques. Reviews cover-
11
L. Vroman, Blood, The Natural History Press (1968) 63. W. Norde, J. Lyklema, J. Colloid Interface Sci. 66 (1978) 257, 266. 31 W. Norde, J. Lyklema, J. Colloid Interface Sci. 71 (1979) 350. 21
3.24
PROTEIN ADSORPTION
ing this early work have been written by Chessman and Davies11, and Cumper and Alexander2'. For proteins adsorbed at the solid/liquid interface, information on the surface area per molecule can be derived from adsorption isotherms. However, without insight into the surface pressure or the geometry of the adsorbed protein layer, interpretation of plateau values in terms of conformational changes of adsorbed proteins becomes rather presumptuous. In one study, using infrared spectroscopy to analyze the bound fraction, the observed conformational changes of several proteins were so small that the authors felt it necessary to report that, 'caution must be exercised in interpreting the available surface area per adsorbed molecule as an indication of changes in conformation upon adsorption' 3). Generally, structural properties of proteins are either described by the energy that is required to unfold their natural state or by the relative location of all atoms in the folded structure. The folded structure is discussed in sec. 3.2. In the 1970s some experiments were performed to measure directly one of the structural properties of proteins in the adsorbed state. Structural elements of adsorbed proteins were investigated by using spectroscopic techniques, such as circular dichroism (CD), nuclear magnetic resonance (NMR), electron spin resonance (ESR), and fluorescence spectroscopy, while microcalorimetry was used to obtain thermodynamic information on the adsorption process. Reviews of this early work have often reported moderate changes in the protein structure, although the measurements were not sensitive enough to establish the type and extent of the adsorption-induced conformational changes4 5). From the late 1970's sophisticated experimental evidence became available that proved the extent of adsorption-induced structural changes to range from none to substantial denaturation of the proteins upon adsorption. A few decades of gathering experimental evidence has led to a number of general trends that relate the various physical and chemical properties of the protein/sorbent system with adsorption-induced structural changes. Nevertheless, a comprehensive theory with predictive value that describes the extent and rate of adsorption-induced structural changes in the proteins remains a challenge for the future. 3.4a How to measure structural properties of adsorbed proteins The development in experimental techniques has made it feasible to probe directly the conformational features of proteins in the adsorbed state. Not only have new and/or more sensitive instruments been developed, but more sorbent surfaces have also become available that interfere less with the measurement techniques. 11
D.F. Chessman, J.T. Davies, Adv. Protein Chem., 9 (1954) 439. C.W.N. Cumper, A.E. Alexander, Rev. Pure Appi Chem.. 1 (1951) 121. 31 B.W. Morrissey, R.T. Stromberg, J. Colloid Interface Set 46 (1974) 152. 41 W. Norde, Adhesion and Adsorption of Polymers, part B (1980) Plenum Publishing Corp. New York. 51 M.E. Soderquist, A.G. Walton, J. Colloid Interface Set. 75 (1980) 386. 21
PROTEIN ADSORPTION
3.25
Most of the techniques for measuring structural properties of proteins at interfaces are based on experimental methods that provide information on the structural properties of proteins in solution. For example, CD is a successful technique for measuring the secondary structure of proteins in aqueous surroundings. In 1974, McMillin and Walton constructed a cell that contained a stack of quartz disks to provide an optically transparent cell with enough protein layers to generate a reasonable signal11. However, difficulties with scattering and the alignment of the quartz plates made the researchers return to solution studies of desorbed proteins21. Finally, the technique regained much interest after it was shown that a suspension of small, nanometer sized colloidal particles covered with proteins generated high quality CD spectra3 . The most important approaches for determining structural properties of adsorbed proteins are based on spectroscopic methods that rely on the interaction of biopolymers with electromagnetic radiation (1.7.3), using techniques such as fluorescence-, NMR, infrared, and CD spectroscopy. Infrared and CD spectroscopy are mainly used to quantify the average amount of a -helical and (3 -sheet structures. As carbonyl groups on the protein backbone are involved In specific H-bonds in secondary structure elements, the wavelength at which infrared radiation is absorbed is slightly shifted. Circular dichroism occurs because secondary structure elements in proteins absorb left- and right- hand circularly polarized light slightly differently in the UV and far UV region. Infrared spectroscopy has a higher sensitivity for measuring changes In P -sheet structures, while CD has a higher sensitivity for quantifying a-helical structures. A major advantage of CD over infrared spectroscopy is that the spectroscopic signal is not affected by the presence of water. Fluorescence spectroscopy can generate information on the local surroundings and the solvent accessibility of fluorescent groups such as tryptophan residues or haem groups in proteins. Furthermore, fluorescence anisotropy and time-resolved fluorescence decay measurements provide information on the rotational mobility of adsorbed proteins, whereas fluorescence recovery after photo bleaching (FRAP) can be used to measure the lateral mobility. To date, applications of NMR in protein adsorption studies are rare51. Its use is limited to relatively small proteins for which complete 'H, 13C, and 15N assignments of the native structure are known. NMR spectroscopy can also be used in combination with hydrogen-deuterium exchange (HDX) experiments to probe the solvent accessibility of exchangeable hydrogens in the protein structure, because the NMR signal disappears when protein hydrogens are exchanged with deuterium atoms from solution. This type of HDX can be used to probe parts of the protein structure that are in contact with the sorbent
11
C.R. McMillin, A.G. Walton, J. Colloid Interface Set 48 (1974) 345. A.G. Walton, M.E. Soderquist, Croat. Chem. Acta 53 (1980) 363. 31 A. Kondo, S. Oku, and K. Higashitani, J. Colloid Interface Set 143 (1991) 214. 41 W. Norde, J.P. Favier, Colloids Surf. 64 (1992) 87. 51 K.Wuthrich, Acta Crystallogr. D51 (1995) 249. 21
3.26
PROTEIN ADSORPTION
surface and are thus less accessible for the deuterated solvent1'21. To characterize adsorbed proteins in terms of the energy or enthalpy involved in the stabilization of the protein structure, one can employ differential scanning calorimetry (DSC) and hydrogen-deuterium exchange in combination with mass spectrometry (HDX-MS). In DSC one measures the difference in specific heat as a function of the temperature between a sample with and without proteins. With such experiments one can derive the enthalpy and temperature of the transition between a folded and unfolded, or otherwise perturbed, conformation of proteins. In HDX-MS, one monitors the exchange between amide hydrogens on the protein backbone and protons in solution. When amide hydrogens are involved in hydrogen bonding in the secondary structure elements, their exchange rates decrease considerably. The time scale of such an exchange process is usually much larger than that used to study the solvent accessibility of exchangeable hydrogens located on the outside of the protein structure. By measuring the deuterium incorporation as a function of time, the protein is exposed to a deuterated solvent. With mass spectrometry, information is revealed on the folding/ unfolding equilibrium. This folding/unfolding equilibrium, in turn, is related to the Gibbs free energy that stabilizes the folded conformation of a protein. Compared to the spectroscopic techniques, DSC and HDX-MS monitor more general aspects of the conformational state of (adsorbed) biomolecules. Additionally, both techniques also provide information on the heterogeneity in the structural populations. Besides these techniques, most of the methods that can be used to characterize surfaces and adsorbed monolayers (III.3.7) also yield indirect information on the conformational state of proteins. For example, neutron reflectivity, ellipsometry, and
Figure 3.14. Atomic Force Microscopy images (0.75 um x 0.7 um ) of human plasma fibronectin measured in the intermittent-contact mode (Courtesy of M. Bergkvist). On the lefthand side, fibronectin is adsorbed at a hydrophilic mica surface, and on the right-hand side on a hydrophobic surface created by silanisation of silica with dichlorodimethylsilane11. 11 21
D.G. Gorcnstein, Z. Santago-Rivcra, and D.A. Keire, Polym. Mater. Sci. Eng. 71 (1994) 261. H. Nagadome, K. Kawano, and Y. Terada, FEES Letters 317 (1993) 128.
PROTEIN ADSORPTION
3.27
reflectometry can be used to study the average density and thickness of an adsorbed protein film, while imaging techniques such as scanning force microscopy (SFM) can be used to generate information on the dimensions that a protein assumes in the adsorbed state. By way of illustration, in fig. 3.14 an AFM image is given for the adsorption of fibronectin on two different surfaces. Fibronectin is a large flexible protein that is stabilized by intermolecular ionic interactions to form a compact structure. Upon altering the solution conditions the structure can revert to a more expanded state, thereby exposing previously hidden domains, for example cell-binding sites. Upon adsorption to hydrophilic surfaces, fibronectin often adapts an elongated structure, whereas on hydrophobic surfaces the compact structure predominates. The difference in morphology is explained by the interaction between fibronectin and the negatively charged groups on the hydrophilic surface that interferes with the stabilizing interaction between the protein domains. Finally, adsorption-induced conformational changes, as well as the influence of the sorbent surface on conformational transitions in the protein molecules, may be inferred from proton titration studies2'3'41. By way of example we present some results for the adsorption of a-lactalbumin (aLA ). To that end in fig. 3.15, proton titration curves for aLA in solution and adsorbed on negatively and positively charged polystyrene particles are presented. The curves for the dissolved and the adsorbed states differ over a wide pH-range. The shifts upon adsorption reflect the influence of the electric field caused by the charges on the polystyrene surface on the pK-values of the titratable groups of the aLA molecules. Figure 3.16 shows the titration behaviour in the pH-region where the carboxyl groups are protonated. The curve for aLA in solution comprises a region reflecting the translation from the holo- to the apo-state. The non-electric Gibbs energy, A holo ^ apo G°, associated with that transition can be
Figure 3.15. Proton titration curve for a -lactalbumin in solution ( ) and on negatively (A) and positively (o) charged polystyrene particles. T = 25°C; electrolyte, 0.05 M KC1. (Redrawn from W. Norde, F. Galisteo Gonzalez, and C.A. Haynes, Polym. Adv. Technol. 6 (1995) 518).
11
M. Bergkvist, J. Carlsson and S. Oscarsson, J. Biomed. Mater. Res. (2002). C.A. Haynes, E. Sliwinsky, and W. Norde, J. Colloid Interface Set 164(1994) 394. 31 F. Galisteo, W. Norde, Colloids Surfaces B4 (1995) 389. 4 ' W. Norde, F. Galsiteo Gonzalez, and C.A. Haynes, Polymers for Adv. Technol. 6 (1995) 518. 21
3.28
PROTEIN ADSORPTION
Figure 3.16. Titration of carboxyl groups in a -lactalbumin. The drawn curve refers to the protein in aqueous solution, the upper clashed curve to the protein, adsorbed on negatively charged polystyrene particles. T = 25°C; electrolyte, 0.05 M KC1. (Redrawn from W. Norde, F. Galisteo Gonzalez, and C.A. Haynes, loc. cit.)
derived according to l A
G
2 303
holo^apo ° = -
NmaxRT\[pK(a.po)~pKihol0))da
[3.4.1]
0
with pK = pH + log[ — - 1
[3.4.2]
where K is the dissociation constant of the carboxyl groups, iVmax the total number of carboxyl groups in the oeLA molecule, and a their degree of dissociation. Hence, Aj^^ a p o G° for aLA in solution may be evaluated from the shaded area in fig. 3.16. It is estimated to be ca. -3 RT per mole of aLA . Proton titration calorimetry data, presented in fig. 3.17, reveal an endothermic holo to apo-state transition of ca. + 18 RT per mole of protein. It follows that, at room temperature, the transition is
Figure 3.17. Isothermal proton titration enthalpy of a -lactalbumin in solution (•) and adsorbed on negatively charged polystyrene particles (A ). T= 25°C ; electrolyte, 0.05 M KC1. (Redrawn from W. Norde, F. Galisteo Gonzalez, and C.A. Haynes, loc. cit.)
PROTEIN ADSORPTION
3.29
driven by an entropy gain of about 0.17 kJ K"1 mol" 1 . Remarkably, the structural transition is virtually absent when aLA is adsorbed on the polystyrene surface. It demonstrates that the structural properties of aLA are very sensitive to the local environment. In this special case of aLA it implies that adsorption causes (a) spontaneous transition to the apo-state, (b), spontaneous transition to a new conformation from which the apo-state is no longer accessible, or, (c), stabilization of the native structure such that the apo-state is no longer favoured at low pH. All the above-mentioned techniques require specific characteristics of the type of interface that can be studied. For spectroscopic techniques, these solid surfaces should not interfere with the spectroscopic signal of interest. Furthermore, for adsorption studies a selection mechanism is required to obtain experimental signals that originate only from the adsorbed proteins. Besides this selectivity, the signal sensitivity can also be a problem, as going from proteins in solutions to proteins adsorbed at a twodimensional surface means that the possibilities for concentrating the sample are limited. Because of these requirements, studies of the conformational state of proteins are restricted to a limited number of available sorbent surfaces. In fluorescence and infrared spectroscopy one can selectively excite adsorbed proteins by using an evanescent wave that is created by total internal reflection of a lightbeam within an optically transparent material [1.7.10.a]. The evanescent wave penetrates the solution to a depth that is roughly half the wavelength of the radiation used. Two prominent techniques based on this principle are total internal reflection fluorescence (TIRF) (see sees. II.2.5c and III.3.7c, iv) and attenuated total reflection-fourier transform infrared (ATR-FTIR) spectroscopy (sec. II.2.5.C). Another advantage of using total reflection is that the polarization of the evanescent wave can be controlled easily, allowing investigation of the spatial orientation of adsorbed molecules. Techniques that utilize evanescent waves require optically inert and smooth surfaces that resemble Fresnel surfaces. For example, silica interfaces as present in oxidized silicon wafers, silicon crystals, polished glasses, and quartz in fluorescence or infrared studies. These silica surfaces can be modified easily by silanization. Self-assembled monolayers (SAM), or Langmuir-Blodgett films (sec. III.3.7a) can be used to obtain a wide range of surface characteristics. The smoothness of most of these above mentioned surfaces also makes them suitable for studies with microscopic techniques such as SFM. For infrared spectroscopy, adsorption can also be studied on polymeric material that is generally spin- or dip-coated on silicon, germanium, or zinc selenide crystals. Another way to apply common biochemical techniques which are used to characterize protein structures in solution, and are suitable for proteins in the adsorbed state is to utilize nanometer-sized colloidal particles. Because of the small size, the lightscattering is very low. At the same time, the small diameter of the particle provides a relative large surface area per sample volume, which benefits the sensitivity in measurements, using for example CD, DSC and NMR spectroscopy. Ultrafine particles of teflon, polystyrene, silica, and titanium oxide have been used successfully.
3.30
PROTEIN ADSORPTION
3.4b General trends To perform their biochemical functions, many proteins fold in specific well-defined and highly ordered structures. The folded conformations of globular proteins reach atomic packing densities, expressed in volume fractions, between 0.7 and 0.8. In such a compact structure the rotational mobility along the polypeptide chain is severely restricted, which implies that it has a low conformational entropy. This low entropy is balanced by hydrophobic and Coulomb interactions, the possibility of forming intramolecular hydrogen bonding in secondary structure elements, and interactions between fixed and induced dipoles. The net result is often a compact folded protein that is marginally stable, with stabilization Gibbs energies in the order of a few- to a few tens of kJ per molar quantity, see sec. 3.2. One should note that for a protein with a stabilization Gibbs energy of 15 kJ/mol, one out of 500 is unfolded, and that folding and unfolding processes often occur within fractions of seconds. Thus, although the structures of globular proteins are described as compact and well-defined, this structure is not static. The protein structure is readily disrupted if, for example, electrostatic interactions change by varying the pH in solution. In the same way, protein structures can be affected by adding organic solvents that affect the Van der Waals interactions and the solvation or by adding molecules that compete for the interactions, such as the chaotropics sodium dodecyl sulphate (SDS) and guanidine hydrochloride. Strength of Interaction: Hydrophobic interactions As mentioned before, the same type of interactions that drive the adsorption of proteins to interfaces are the ones determining the structure of proteins. Logically, a stronger interaction between a protein and a sorbent surface is expected to increase the probability and the extent of adsorption-induced structural changes. For example, when proteins adsorb to hydrophobic surfaces, dehydration of the hydrophobic surfaces is the main driving force and, at the same time, hydrophobic interaction is an important force that keeps the polypeptide chains tightly folded. In the adsorbed state the protein/sorbent interaction can be increased by having parts of the hydrophobic core of the protein exposed to the hydrophobic sorbent, leaving more hydrophilic parts of the protein in a more flexible, and thus entropically favourable conformation. Many studies have proved that structural changes are more substantial when globular proteins adsorb to hydrophobic surfaces compared to hydrophilic surfaces. For example, early TIRF studies demonstrated that the conformation of fibronectin is not affected when it is adsorbed onto hydrophilic silica but it is affected when the surface becomes more hydrophobic11.
11
J.D Andrade, V.L. Hlady, and R.A. van Wagencn, Pure Appl. Chem. 56 (1984) 1345.
PROTEIN ADSORPTION
3.31
Strength of interaction: Electrostatic interactions A tendency similar to that above has been observed with respect to electrostatic interactions, i.e., the stronger the attraction, the larger is the extent of structural changes1 . A similar observation was made by Larsericsdotter et al., who used DSC to show that, for the adsorption of lysozyme and ribonuclease-A at hydrophilic surfaces, strong electrostatic attraction leads to high affinities whereby the protein adsorption is accompanied by a reduction in the denaturation enthalpy. Shielding the electrostatic interaction by increasing the ionic strength reduces the affinity, and the adsorptioninduced reduction of the denaturation enthalpy is diminished . Another electrostatic contribution that can affect the protein structure in the adsorbed state is the net charge density in the interfacial layer (see also sec. 3.3c). In solution, most proteins unfold when the net charge density increases, for example at low or high pH values or at low ionic strengths. If proteins adsorb to charged surfaces the interplay between charged groups, including co-adsorbed low molecular weight ions, will affect the balance between the electrostatic interactions within the protein. It is generally observed that the maximum amount of protein is adsorbed when the solution pH is close to the isoelectric point of the protein. A study in which the amount of secondary structure of monoclonal antibodies adsorbed at silica surfaces was monitored while the electrostatic interactions involved were systematically varied, revealed a clear correlation between the reduction in adsorbed amounts when the pH is shifted away from the isoelectric point and the reduction in secondary structure3'. Structural stability of proteins Conformational changes in proteins not only result from the strong interaction between a protein and a surface but can also be seen as driving forces for adsorption41. This was inferred from experiments in which protein adsorption was monitored on hydrophilic surfaces under conditions where the proteins were electrostatically repelled by the sorbent. Under these conditions, the only driving force for adsorption is generated by a total increase of the entropy of the adsorbed proteins. It was also observed that only proteins having a relatively low structural stability adsorb under these otherwise adverse adsorption conditions. This led in sec. 3.3 to the introduction of the notions of 'soft' and 'hard' proteins. Proteins that adsorb to hydrophilic surfaces under electrostatic repulsive conditions are classified as soft proteins whereas proteins that do not adsorb under those conditions are considered hard proteins . Elegant studies on the conformation of mutants and the wild-types of T4 lysozyme'.6) 11
A. Kondo, F. Murakami, and K. Higashitani, Biotech. Bioeng. 40 (1992) 889. H. Larsericsdotter, S. Oscarsson, and J. Buijs, J. Colloid Interface Set 237 (2001) 98. 31 J. Buijs, J.W.Th Lichtenbelt and W. Norde, Langmuir 12 (1996) 1605. 4) W. Norde, J. Lyklema, J. Colloid Interface Sci. 71 (1979) 350. 51 W. Norde, J. Lyklema, J. Biomater. Sci. Polymer 2 (1991) 183. 61 P. Billsten, M. Wahlgren, T. Arnebrant, J. McGuire and H. Elwing, J. Colloid Interface Sci. 175 (1995) 77. 21
3.32
PROTEIN ADSORPTION
and carbonic anhydrase II11 adsorbed on nanometer-sized silica particles revealed that the extent of structural changes was inversely related to their stabilization Gibbs energies in solution. The nature of structural changes, however, was different between the two proteins. T4 lysozyme lost part of its secondary structure upon adsorption, while for carbonic anhydrase II the secondary structure is unaffected, but the protein adopts a more open conformation which is thermally destabilized compared to the protein in solution. The examples above prove that protein stability is an important factor that determines the extent of structural changes. At the same time, no clear relationship between the stability of a protein in solution and the extent of structural changes has been found. Apparently, the structural flexibility of a protein needs to be defined in more detail before such a relationship with predictive value can arise. It is repeated that one should distinguish between adsorption-induced changes that result from strong interactions between the protein and the sorbent surface, and structural changes that result from an adsorption process in which structural rearrangements are prerequisites for adsorption to take place. Surface coverage A number of CD studies, using nanometer-sized silica particles as the sorbent surface, demonstrated that the extent of structural changes was larger when the surface coverage was less 2 ' 3 '. Apparently, the adsorbed proteins need space for structural changes to take place. This phenomenon was also observed for protein adsorption at the air-water interface, where the biological activity of an adsorbed protein monolayer was stabilized by increasing the surface pressure 4 '. A theoretical approach to this phenomenon has been given in equation [3.3.12]. From this equation it is clear that the extent of adsorptioninduced conformational changes in proteins is a function of the adsorption kinetics, i.e., the faster a surface is covered with proteins, the lower is the probability that the proteins will undergo structural rearrangements. The strength of interaction also plays an important role, as it has been observed that the adsorbed amount F passes through a maximum in the course of adsorption, because adsorbed proteins spread at the expense of the total number of proteins attached to a surface. Time-dependency of the structure of adsorbed proteins In general protein adsorption is irreversible, implying that, as a rule, the interactions between a substrate and proteins increase with contact times. It is generally 11
P. Billsten, U. Carlsson, B.H. Jonsson, G. Olofsson, F. Hook and H. Elwing, Langmuir 15 (1999) 6395. 21 A. Kondo, S. Oku and K. Higashitani, J. Colloid Interface Set 143 (1991) 214. 31 W. Norde, J.P. Favier, Colloids Surf. 64 (1992) 87. 41 A. Tronin, T. Dubrovsky, S. Dubrovskaya, G. Radicchi and C. Nicolini, Langmuir 12 (1996) 3272.
PROTEIN ADSORPTION
3.33
accepted that when proteins adsorb they do so tenaciously, because of the large number of segment contacts that can be established between them and the surface. In turn, conformational changes in the protein structure are generally thought to be responsible for increasing the number of contact points, leading to protein adsorption process models that do not allow for conformationally changed proteins to desorb spontaneously1 2). These models are supported by the often observed reduction in protein activity at longer contact-times with the substrate, and by several studies that show that proteins slowly lose secondary structure in the adsorbed state. Unfortunately, little is known about the factors that determine the overall rate of conformational changes. Adsorption isotherms of IgG and fibronectin on hydrophilic and hydrophobic substrates, obtained either by direct addition or successive addition of proteins, clearly demonstrated that conformational changes occur more rapidly on hydrophobic surfaces31. Another observation is that structurally less stable and soft proteins not only have a higher tendency to undergo structural changes upon adsorption but the structural changes occur also more readily41. Moreover, structurally less stable proteins that adsorb with high affinity to surfaces show a higher irreversibility in adsorption and, for those proteins that do desorb, the extent of refolding is lower than observed for structurally stable proteins5'6'71. These observations indicate that rearrangements in the proteins' structure occur faster under adsorption conditions that lead to a large extent of structural changes, as intuitively expected. Besides the slow structural rearrangement, a rather rapid change in the structure is also observed when the first proteins adsorb . This effect can be explained by the lack of steric hindrance for spreading on the surface. See sec. 3.3 Recent stopped-flow fluorescence spectroscopy and anisotropy measurements9 revealed clearly distinct stages for protein structural rearrangements. Specifically, the conformational changes of a -lactalbumin adsorbed on polystyrene particles occur with rate constants of 50 s"1, 8 s' 1 and 0.001 s" 1 . Following the time-dependence of the (3 -sheet content of adsorbed antibodies to silica surfaces, using FTIR spectroscopy, clearly showed that both the relatively low amounts of structure for proteins that were adsorbed in the initial phase and the slow reduction in the [3 -sheet content after longer adsorption times (see fig. 3.18)10).
11
M.E. Soderquist, A.G. Walton, J. Colloid Interface Sci. 75 (1980) 386. I. Lundstrom, H. Elwing, J. Colloid Interface Sci. 136 (1990) 68. 31 U. Jonsson, I. Lundstrom and I. Ronnberg, J. Colloid Interface Sci. 117 (1987) 127. 41 B. Singla, V. Krisdhasima and J. McGuire, J. Colloid Interface Sci. 182 (1996) 292. 51 A. Kondo, J. Mihara, J. Colloid Interface Sci. 177 (1996) 214. 61 W. Norde, J.P Favier, Colloids Surf. 64 (1992) 87. 71 W. Norde, C.E. Giacomelli, Macromolecular Symposia 145 (1999) 125. 81 A. Tronin, T. Dubrovsky, S. Dubrovskaya, G. Radicchi and C. Nicolini, Langmuir 12 (1996) 3272. 91 M.F.M. Engel, C.P.M. van Mierlo, and A.J.W.G. Visser, J. Biol. chew... 277 (2002) 10922. 101 J. Buijs, J.W.Th. Lichtenbelt, and W. Norde, Langmuir 12 (1996) 1605. 21
3.34
PROTEIN ADSORPTION
Figure 3.18. The fraction of amino acid residues in adsorbed antibodies that adopt a (5 -sheet structure (closed circles), and the adsorbed amount (open circles), as a function of the adsorption time. Both the adsorbed amount and the percentage P -sheet structure were obtained with ATR-FTIR. The data presented were obtained when IgG molecules adsorb to a hydrophilic and negatively charged silica surface created by oxidation of a cylindrical crystal made of silicon. The crystal was surrounded by a solution containing 50 jig/ml of a monoclonal IgG with an isoelectric point of 5.8, and 5 mM acetate buffer at pH 5. In the native state, roughly 60% of the amino acids are involved in a p -sheet structure. (Redrawn from J. Buijs et ai., loc. cit.}
Structure of adsorbed globular proteins As might be apparent from the paragraphs above, the final structure of a protein in the adsorbed state is the result of the interplay between many factors. Even if all the interactions between a protein and a surface, including intramolecular interactions, that determine the protein stability, are fully characterized, the problem remains that the structure of the adsorbed protein will depend on the adsorption time and on the interactions with neighbouring proteins. Furthermore, structural alterations can appear on various structural levels. For example, in fig. 3.14 an example is shown where fibronectin is adsorbed in a more elongated structure when it is adsorbed on a negatively charged hydrophilic mica surface, whereas the structure is more compact when adsorbed on a hydrophobic surface. At the same time fig. 3.14 shows that fibronectin only changed its conformation when adsorbed to hydrophobic surfaces. Most probably, the contradiction between these data is based on the differences in structural properties that were probed using different measurement techniques. The first study utilized SFM, which can be used to monitor changes in the tertiary structure, while the second study was performed with TIRF, which is sensitive to local changes involving the relocation of a couple of amino acid residues. An interesting aspect of studies of adsorption-induced changes is that very few
PROTEIN ADSORPTION
3.35
investigators report a complete unfolding of adsorbed proteins. Cases of the absence of structural changes are reported, but they are definitely outnumbered by studies that indicate partial unfolding of globular proteins upon adsorption. At hydrophobic surfaces, some proteins even increase their content of secondary structure elements121 or one type of ordered structure may be converted into another. Both the conversion from a -helix to (3 -sheet structure31 and vice versa 451 have been observed. These results are noteworthy in view of the relatively low stabilization Gibbs energies of the protein structure in solution, where a small disruption in the structure often results in complete unfolding. One should note, however, that most techniques employed to study protein structures at interfaces only monitor average properties. This implies that the results obtained mean either that each protein is partially unfolded to some extent, or that a fraction of the adsorbed molecule is strongly unfolded while another fraction is in its native conformation. Some studies claim that the proteins adopt a well-defined structure similar to the intermediate structures created by local energy minima in the folding pathway of proteins in solution61. Nevertheless, measurements using techniques that provide information on the structural heterogeneity of the protein structure in the adsorbed state, such as DSC and HDX, generally show that structures of adsorbed protein molecules are perturbed differently. Also, with other methods, such as FRAP that is used to monitor the two-dimensional diffusion of adsorbed proteins, it has been observed that one population of BSA molecules, adsorbed to various polymeric materials, is mobile, while another population is immobile. The difference in mobility between populations was ascribed to the distribution of protein conformations, established in the early stage of adsorption71. Even for proteins that show no change in average structure upon adsorption, a more heterogeneous structure can be observed. For example, lysozyme adsorbed to silica surfaces has been subjected to a broad range of techniques to assess its structural properties, such as CD, DSC, TIRF, FTIR, and HDX-MS. These studies revealed that lysozyme does not undergo significant structural alterations when adsorbed to silica in compact monolayers. Nevertheless, the results from DSC, TIRF, and HDX studies indicate that the conformation of lysozyme is rather heterogeneous. This heterogeneity fits well into the models that allow structural rearrangements as long as space is available for the proteins to spread; whereas for proteins that adsorb at a later stage, the spreading is inhibited by spatial restrictions (sec. 3.3b). 11
E.J. Castillo, J.L. Koenig, J.M. Anderson, and J. Lo, Biomater. 5 (1984) 319. M.C.L. Maste, W. Norde, and A.J.W.G. Visser, J. Colloid Interface Set 196 (1997) 224. 31 C.E. Giacomelli, W. Norde, Biomacromolec. 4 (2003) 1719. 41 J. Buijs, W. Norde, and J.W.Th. Lichtenbclt, Langmuir 12 (1996) 1605. 51 A.W.P. Vermeer, C.E. Giacomelli and W. Norde, Biochim. Biophys. Ada 1526 (2001) 61. 61 P. Billsten, U. Carlsson, B.H. Jonsson, G. Olofsson, F. Hook, and H. Elwing, Langmuir 15 (1999) 6395. 71 R.D. Tilton, C.R. Robertson, and A.P. Gast, J. Colloid Interface Set 137 (1990) 192. 21
3.36
PROTEIN ADSORPTION
To illustrate the complexity of adsorption-induced changes in protein structures, an example will now be given of how the structural stability of various segments of myoglobin is affected by adsorption on silica particles. The results were obtained using HDX-MS, a technique that is sensitive to the stability of secondary structure elements in a protein, as the exchange-rate between amide hydrogens in the protein and deuterium atoms in solution is dramatically slowed if the amide hydrogens are involved in hydrogen bonding. In this study, the hydrogen/deuterium exchange process was followed as a function of the exchange time. After the exchange process, myoglobin was cleaved enzymatically, and the deuterium incorporation into four fragments, covering 90% of the sequence, was monitored using mass spectrometry11. Structurally, myoglobin is a 153 residue, tightly folded protein with a haem group that is bound noncovalently in a hydrophobic pocket. Myoglobin contains an extremely stable core, located at the C-terminus and linked to a part of the N terminus, while the centre part of the sequence, surrounding the haem group, is less stable. Upon adsorption to silica, the myoglobin segment located in the middle of the myoglobin sequence, and close to N terminal fragments are destabilized (see fig. 3.19). Although the structural stability of the segment around the haem group did not change upon adsorption, this structure became clearly more heterogeneous. Interestingly, for the N terminal fragment, comprising residues 1-29, two distinct and equally large conformational populations were
Figure 3.19. The structure of myoglobin in solution (left), and adsorbed at silica particles (right). The structural stability is indicated in terms of grey scales and is based on the average folding/unfolding equilibrium of four segments comprising residues 1-29, residues 30-69, residues 70-106, and residues 107-137. The structure is taken from S.R. Hubard, W.A. Hendrickson, D.G. Lambright and S.G. Boxer, J. Mol. Biol. 213 (1990) 215, and the grey scale ranges from 16 kJ/mol (light) to 32 kJ/mol (dark). Upon adsorption, only the structural stabilities of the C- and N-terminal parts are reduced whereas the a -helices close to the haem group are unaffected.
11
J. Buijs, M Ramstrom, M. Danfelter, H. Larsericsdotter, P. Hakansson and S, Oscarsson, J. Colloid Interface Set 263 (2003) 441
PROTEIN ADSORPTION
3.37
observed. One of these populations has a stability similar to that in solution (~ 23 kj/mol) whereas the other population is highly destabilized upon adsorption (—11 kj/mol). From these results it is clear that, even within one protein, the structures of parts are differentially affected by adsorption. Some parts of the protein and a part of the population are highly destabilized upon adsorption whereas other parts are only affected in their structural heterogeneity.
3.5 Adsorbed amount and adsorption reversibility Protein molecules may form numerous contacts with a sorbent surface (as synthetic polymers can). Although data are very limited, there is experimental evidence for such multi-contact adsorption1'. Multiple contacts probably lead to high-affinity adsorption of the whole protein molecule. Plateau-values of the adsorbed amount F corresponding to full coverage of the sorbent surface are reached at a very low protein concentration in solution c and the region in which F depends on c is limited to values very near the /"-axis. Dilution of the sorbate in the bulk phase creates a transient difference in the chemical potential of the sorbate at the interface and that in solution. This chemical potential difference is then eliminated by spontaneous desorption of the sorbate. For high-affinity adsorption, desorption is difficult, if discernible at all, because in the c regime where F is below its plateau value the protein concentration may well be below the detectable limit. Furthermore, when equilibrium is established at such low protein concentrations, where the protein concentration gradient between the subsurface region and the bulk solution is so small that, according to [3.3.1], the transport of protein molecules into the bulk solution is extremely slow. Occasionally, protein adsorption isotherms are reported whose initial rising part deviates from the /"-axis. However, upon dilution, F(c ) rarely (if ever), follows the same path backwards, thereby making the ascending and descending branches of the isotherm distinguishable. As a rule, it is found that dilution only leads to partial desorption, if at all, even when the observation time greatly exceeds the relaxation time of the protein at the sorbent surface2'31. Desorption upon dilution is minimal, especially at hydrophobic surfaces. Such hysteresis indicates a prohibitively high barrier for (further) desorption. The system has two (meta)stable states, one on the ascending branch and the other on the descending one, each being characterized by its local minimum in Gibbs energy. The fact that the adsorption and desorption isotherms represent different metastable states, implies that a physical change has occurred in the system between adsorption and desorption. 11
B.W. Morrissey, R.T. Strombcrg, J. Colloid Interface Set 46 (1974) 152. H.P. Jennissen, in Surface and Interfacial Aspects of Biomedical Polymers, Vol. 2 J.D. Andrade, Ed., Plenum (1985) 295. 31 W. Norde, C.A. Haynes, ACS. Symp. Series 602 (1995) 26.
3.38
PROTEIN ADSORPTION
Despite the irreversible nature generally observed for protein adsorption, many authors interpret their experimental data using theories that are based on reversible thermodynamics. The most common example is the calculation of the Gibbs energy of adsorption, AadsG , by fitting the (ascending) isotherm to the Langmuir equation or variations thereof121. It is, however, questionable whether the adsorption or desorption isotherm should be used for that purpose. Another approach sometimes encountered involves calculation of the Gibbs energy of adhesion, AadhG , using A
adh G = rsp - r™ - rpw
I3-5-1'
where y is the interfacial tension of the interface indicated by the superscripts. AadhG is the reversible work (at constant temperature and pressure) to form a protein (p)/ sorbent (s) interfaces at the expense of sorbent (s)/solution (ix>) and protein (p)/solution (u>) interfaces31. Equation [3.5.1] essentially corresponds to our [III.5.1.1] for the initial spreading tension. For further analysis, see sec. III.5.2; it is noted that the interfacial tensions may change with time. It is not clear how AadhG relates to AadsG, since the latter quantity is determined not only by creation and rupture of phase boundaries but also includes other contributions such as those resulting from electrostatic interaction and protein (and sorbent) structural rearrangements. The error involved in treating protein adsorption as a reversible process may be approximated by calculating the entropy production due to the irreversibility (reflected by the hysteresis) of the process. In a closed system, the entropy change associated with any process can be written as AS = ASe + ASj
[3.5.2]
where ASe is the reversible entropy exchange between the system and its surroundings, and ASj is the internally produced entropy. For a reversible process, AS{ = 0, but for an irreversible process ASj > 0 . According to Everett41, AadsSj can be calculated from the hysteresis loop, as the closed loop integral between the ascending and descending branches of the adsorption isotherm A
adsSi=RJ ^ f d l n c p
[3.5.3]
where r* is the adsorbed amount at the upper closure point of the hysteresis loop. The integral of [3.5.3] may be evaluated as the shaded area indicated in fig. 3.20. Regrettably F(c ) is not usually known in the very dilute region (i.e., at very negative
11
T. Mizutani, J.L. Brash, Chem. Phartn. Ball. 36 (1988) 2711. E.C. Moreno, M. Krcsak and D.I. Hay, Biofouling 4 (1991) 3. 31 C.J. van Oss, Biofouling 4 (1991) 25. 4) D.H. Everett, Trans. Faraday Soc. 50 (1954) 1077. 21
PROTEIN ADSORPTION
3.39
Figure 3.20. Adsorption ( —t) and desorption (<-) of BSA on/from silica particles. Temperature 25°C, 0.05 M phosphate buffer, pH 7. The shaded area repressents the entropy change associated with the hysteresis loop; it may be calculated using [3.5.3].
lnc values). Hence, only a minimum value of AadsSj can be estimated by letting the hysteresis loop close at the lowest experimentally detectable of c . Unfortunately, no consistent molecular picture has been put forward to explain the adsorption hysteresis. The most popular explanation is that, upon first attachment, the proteins form a relatively small number of contacts with the sorbent surface, and that they gradually relax to significantly enhance bonding. An alternative explanation for hysteresis is intermolecular aggregation between adsorbed protein molecules. Then desorption requires detachment of a cluster of protein molecules rather than single ones. As explained in chapter II.5, with flexible highly solvated polymers an increasing value for F is commonly found with increasing concentration in solution. It is explained by conformational adaptation of the adsorbed molecules, involving a decreasing number of attached segments per molecule. The adsorption isotherms of globular proteins, however, develop well-defined plateau-values. Usually, the plateau value is not too far from those corresponding to a close-packed monolayer of native molecules in a side-on or end-on orientation. Under some conditions, for example for highly charged and/or structurally very labile protein molecules, considerably lower plateau adsorptions are found. They could be explained by the formation of an incomplete monolayer and/or extensive unfolding of the protein molecule to attain a rather flat conformation of the polypeptide chain on the sorbent surface (which is typical for flexible polyelectrolytes). Sometimes, but not exceptionally, a 'step' or a 'kink' is observed in the isotherm at intermediate c . An example is shown in fig. 3.21. A reasonable explanation could be that the first plateau represents a layer with randomly oriented protein molecules. Then, at higher c , intermolecular nucleation occurs to form a, 'two-dimensional protein crystal', i.e., protein molecules align in orderly oriented arrays with an increasing amount of protein in a saturated layer. Unfortunately, the validity of this suggestion has not yet been verified experimentally. Alternative explanations for the
3.40
PROTEIN ADSORPTION
Figure 3.21. Adsorption isotherm with two plateaus, frequently observed for protein adsorption on solid surfaces.
appearance of a kink In the isotherm could involve tilting of the adsorbed molecules from a side-on to a more end-on orientation, or the formation of a second adsorbed layer on top of the first one. Of course, each of these possibilities may apply to different extents for different protein-sorbent systems. 3.6 Influence of some system variables on protein adsorption 3.6a Protein and sorbent charge A systematic approach to establish the influence of the electrical charge on protein adsorption is to vary the charge on the protein (by varying the pH In the system), while keeping the charge on the sorbent-surface constant. In fig. 3.22, plateau values for serum albumin are given as a function of the pH of adsorption at different sorbents of constant charge density. If electrostatic interactions between the protein and the sorbent surface were to dominate, the /"(pH) isotherm would be a monotonic function
Figure 3.22. Plateau adsorption of human serum albumin on negatively charged polystyrene (A), positively charged polystyrene (A), negatively charged silver iodide (x) and uncharged poly(methylene oxide) (•). T = 25°C; electrolyte, 0.01 M KNO3 . (Redrawn from W. Norde, Adv. Colloid Interface Set 25 (1986) 267.)
PROTEIN ADSORPTION
3.41
of pH. Instead, all of these curves show a maximum near the isoelectric point of the protein-sorbent complex. Such a bell-shaped /^'(pH) -functionality is a feature not only of albumin, but it is quite commonly observed for protein adsorption". Lateral electrostatic repulsion between adsorbed molecules could prevent the formation of a close-packed monolayer, and thus explain the reduction in /"P1 away from the isoelectric region. Another possible reason for the maximum in /T'(pH) is related to the degree of structural rearrangement of the adsorbed molecules. In the isoelectric region, the internal cohesion is expected to be at a maximum, and adsorption-induced conformation changes resulting in a larger molecular area at the sorbent surface would be minimal. Some proteins, in particular the hard ones, show a different .TP'tpH) pattern. Unlike the 'soft' proteins, for which conformational changes may be a major driving force for adsorption, the, 'hard' proteins adsorb only in case of hydrophobic interaction and/or electrostatic attraction. Hence, for the hard proteins /T'(pH)dependencies result like those shown for ribonuclease in fig. 3.23.
Figure 3.23. As fig. 3.22 but now for bovine ribonuclease on negative polystyrene (o), uncharged poly(methylene oxide) (•) and haematite (x). (Redrawn from W. Norde, loc. dt.) 3.6b Hydrophobicity
Protein surfaces contain both hydrophilic and hydrophobic residues. It is reasonable to assume that protein surface hydrophobicity determines, at least partly, the adsorption behaviour. Studies related to hydrophobic interaction chromatography have confirmed that the more hydrophobic proteins tend to adsorb more tenaciously to (solid) surfaces . In addition to the hydrophobic fraction of the outer surface, the over all hydrophobicity may influence the adsorption as well. The probability of structural 11 2)
C.A. Haynes, W. Norde, Colloids Surf. B2 (1994) 517. A.A. Gorbunov, J. Chromatography 365 (1986) 205.
3.42
PROTEIN ADSORPTION
rearrangements in the protein molecule, when it is transferred from an aqueous solution to a (partially) non-aqueous environment (as occurs in an adsorption process) increases with an increasing contribution from intramolecular hydrophobic interaction to the stabiliztion of globular structures in solution (see sec. 3.2b). Indeed, several studies1 2) report a strong correlation between the overall hydrophobicity of a protein and its tendency to change its structure upon adsorption. It is difficult to quantify the influence of the hydrophobicity on protein adsorption. We have to compare sorbent surfaces of different hydrophobicities, different chemical compositions, and possibly different electrical charges. Nevertheless, trends may be indicated. In general, because of the contribution of dehydration, the adsorption affinity increases with increasing hydrophobicity31. Furthermore, there are indications that conformational changes in the protein molecule are triggered more strongly with increasing hydrophobicity of the sorbent surface. 3.6c Protein structure stability Protein molecules may adopt a number of conformations that have only marginally different stabilities. Hence, interaction with the sorbent surface may trigger conformational changes. By using various experimental methods, adsorption-related conformational charges have been characterized (see sec. 3.4). For some proteins, a reduction in ordered (secondary) structure, involving an increased conformational entropy of the protein, could even be a dominating force for adsorption to occur spontaneously. Thus, unlike 'hard' proteins, 'soft' proteins do adsorb at hydrophilic, electrostatically repelling surfaces, thanks to a relatively large gain in conformational entropy. The propensity of structural alteration influences the adsorption affinity and the adsorbed amount. However, it is impossible to trace correlations when comparing different types of proteins, because of interference from other factors such as charge and hydrophobicity. Unambiguous data may be obtained from comparative studies using mutants of the same type of protein. For instance, Tian et al. have compared wild type lysozyme with different mutants4'. The role of protein structural changes on the reversibility of the adsorption process, i.e., the question whether or not desorbed molecules re-adopt their original, native structure will be discussed in sec. 3.8. 3.7 Adsorption at fluid interfaces Proteins at liquid-gas (LG) and liquid-liquid (LL) interfaces form a topic of their own. First, because of their practical interest, for example in relation to the creation and 1J
K.S. Birdi, J. Colloid Interface Set 43 (1973) 545. T. Aral, W. Norde, Colloids Surf. 51 (1990) 1. W. Norde, in Physical Chemistry of Biological Interfaces, A. Baszkin, W. Norde, Eds., Marcel Dekker (2000) ch. 4. 41 M. Tian, W. Lee, M.K. Bothwell, and J.M. McGuire, J. Colloid Interface Set 200 (1998) 146. 21
PROTEIN ADSORPTION
3.43
stabilization of protein-stabilized foams and emulsions, which are relevant for the food industry (ice cream, food emulsions), see chapters 7 and 8. respectively. Another field is that of enzymatic reactions in such interfaces, e.g., for lipases. Secondly, there is academic interest. The central questions are, "What happens to their conformations when proteins adsorb at fluid interfaces? Are these conformational changes similar to those at SL interfaces, which we have discussed extensively in preceding sections?" Insight into the properties of protein monolayers at LG and LL interfaces has made steady, but slow progress. This slowness is partly caused by experimental problems that may be of methodological nature (what techniques are available, and what pitfalls must be avoided in the interpretation of signals?), or inherent in the system (e.g., the adsorption of several globular proteins is largely irreversible, leading to history dependence). For example, it is possible that a given protein with a given surface concentration/', obtained by adsorption from solution, has a conformation which differs from that in the same final state, but is obtained from spreading at a low surfacepressure, n, followed by compression). Here, we can only review some relevant trends. For older, but not necessarily dated, reviews see those by Miller and Bach, by Wahlgren and Nylander and by Izmailova, mentioned in the General References, sec. 3.10. For older reviews see refs. 12) . MacRitchie has reviewed the adsorption-desorption reversibility31 and Dickinson, in two review papers 4 ' surveyed the relevance of protein adsorption at LL interfaces for the formation of (food) emulsions. 3.7a Review of some general trends and techniques A number of differences from adsorption at SL interfaces are readily noticed. These are partly of an academic, partly of a methodical nature; sometimes a mixture of both. l)We only consider macroscopic interfaces. The interfaces are smooth, and will usually remain so upon adsorption. Hence, there is no reason for worrying about surface roughness, surface heterogeneity, or ill-defined asperities, porosities etc. The flatness is a result of the contractile action of the interfacial tension y or the interfacial pressure, n. Problems arise only when the interface contains solid-like patches: or particles, that do not yield under the influence of n . Example: small proteinaceous gel particles. An advantage is that the interfacial area A is well known. 2) Assuming adsorbing ionic surfactants to be absent, the original interface is essentially uncharged, except for charge caused by the preferential adsorption of one of the ionic species of electrolyte that might have been added, including buffers. In sees 11
I.R. Miller, D. Bach, in Surface and Colloid Science, E. Matijevic, Ed., Vol. 6 Wiley-Interscience (1973) 185-260. 21 V.N. Izmailova, Progr. Surface Membr. Sci. 13 (1979) 141. F. MacRitchie, Proteins at Liquid Interfaces, in Physical Chemistry of Biological Interfaces. A. Baszkin, W. Norde, Eds., Marcel Dekker (2000) chapter 5. 41 E. Dickinson, J. Chem. Soc. Faraday Trans. 88 (1992) 2973; ibid., 94 (1998) 1657.
3.44
PROTEIN ADSORPTION
III.4.4b and 4c this phenomenon was analyzed and shown to be minor. Moreover, it is likely that these adsorbates are readily displaced by adsorbing proteins. The conclusion is that any influence of pH or c salt stems from the influence of these parameters on the protein in solution and in the adsorbate. 3) Air does not attract proteins (nor does it attract water). So, the driving force for adsorption at the air-water interface is expulsion from the bulk of the hydrophobic parts of the protein, which may be created by some reconformation. On the other hand, non-aqueous 'oil' phases may attract proteins, mainly through their hydrophobic parts. This difference may lead to conformational stability differences between proteins at these two types of interfaces. 4) As with SL interfaces, the adsorption is, to a large extent, irreversible. This extent depends strongly on the conformational stability. Desorption can, however, be achieved by displacers. Common surfactants can act as such, but sometimes spurious admixtures may do that as well. Purity of the protein sample is therefore an essential condition for reproducibility. 5) As the surface area-volume ratio is usually small compared with adsorption on dispersed particles, it is virtually impossible to determine accurately the surface concentration F by depletion of the solution. Of course, the adsorbing area can be greatly enlarged by foaming or emulsification. However the F[c) isotherms, obtained in this way are not representative for adsorption at quiescent interfaces because the interface has passed through a complex sequence of mechanical stretching, shearing and coalescence steps, which determine the adsorption of the proteins and emulsifiers, if any. 6) The determination of F from the surface pressure, K , is not a good alternative, either, because the 2D equation of state is only ideal under very restrictive (and not particularly interesting) conditions. In general, one may not expect even the relatively simple 2D equations that work reasonably well for uncharged polymers, such as [III.3.4.58] or [III.3.8.6], to apply to proteins. 7) When we look for viable alternatives to establish F and other properties of the adsorbate, the optical techniques discussed in sec.III.3.7b-c come to mind. The adjustment and implementation of these approaches to protein monolayers is not always straightforward, but useful developments have been made, with each technique having its advantages and limitations. For example, reflectometry is a powerful tool for studying the adsorption rate, especially in the earlier stages, but does not provide additional information on the adsorbate. Ellipsometry has now been adapted to protein adsorbates1'. Neutron reflectometry, in particular at small angles, helps to provide information on the layer thickness and averaged volume fraction profile2'31. This technique cannot pick up rapid adsorption or reconformation steps, but works both at LG
11
J.A. de Feijter, J. Benjamins and F. Veer, Biopolymers 17 (1978)1759. E. Dickinson, D.S. Home, J.S. Phipps and R.M. Richardson, Langmuir 9 (1993) 242. 31 A. Eaglesham, T.M. Herrington and J. Penfold, Colloids Surf. 65 (1992) 9. 2)
PROTEIN ADSORPTION
3.45
and LL interfaces. By TIRF (total internal reflection fluorescence) the translational and rotational diffusion of adsorbed proteins can be observed in detail11) and FRAP (fluorescence recovery after photobleaching) gives information on diffusion inside the adsorbate2'. 8) Of the non-optical methods for determining F, surface pressure and radioactivity measurements may be mentioned. The former technique, and its application to a variety of systems, not including proteins, has been discussed in detail in chapter III.3. As stated, it is not straightforward to relate n to F, but compression-expansion cycles in Langmuir troughs can be enlightening. Experiments with radio labeled proteins have the well-known advantage (high sensitivity) and proviso (that the tagging of proteins should not affect the conformation of the molecules). 9) A typical additional element for fluid interfaces is the option of studying interfacial rheology. A number of interfacial rheological characteristics are obtainable. For a review of the techniques, see sec. III.3.7e, in particular we refer to the interfacial shear (77^ ) and dilatlonal (77J) viscosities (Nm^1 s) and, for oscillatory measurements, the complex, interfacial dilational modulus: K° = K°'-iK°"
[3.7.1]
where Kg' is the dilational storage modulus and Kg" the corresponding loss modulus (all K's in Nm"1). The quantity Kg" = rf^a, where co is the frequency and ?/g the interfacial dilational viscosity; see table III.3.4. The values obtained for these characteristics usually depend on the method used, as a result of differences in the way in which forces are exerted on the monolayer. In this respect, the time-dependence and the history play an important, and sometimes decisive, role. The interpretation is not always unambiguous because the observed trends are usually caused by a non-additive combination of rates of supply from solution and rate of reconformation. Therefore, a review of a variety of experimental results for the same system can give some feeling for the basic processes involved, and for their complexity. Such information is much less accessible with SL interfaces. 3.7b Some illustrations The few examples from the literature to be discussed here mostly involve rheological measurements and/or comparisons of systems and/or conditions. For background information (for this section, and for the many studies not reported here), table 3.3 collects some relevant molecular properties of proteins in aqueous solution. As protein (re-)conformation is the complicated result of a range of subtle molecular transformations it is not feasible to characterize the conformational stability by using just one V. Hlady, R.A. van Wagenen, and J.D. Andrade, in Surface and Interfacial Aspects of Biomedical Polymers; J.D. Andrade, Ed., 2 (1995) 81. 21 M. Coke, P.J. Wilde, E.J. Russell, and D.C. Clark, J. Colloid Interface Scl. 138 (1990) 489.
3.46
PROTEIN ADSORPTION
Table 3.3. Molecular properties of some proteins . Ovalbumin
BSA
Myo-
Source
hen's eggs
bovine
sperm
serum
whale
M/kDa
45
69
17.8
i.e.p.
4.7
4.9
Molecular
compact
globular
structure
globular
Lysozyme
a-Casein
P-Casein
hen's eggs
cow's milk
cow's milk
14.6
23.5
24
10.7-11.1
globin
Shape
and spherical dimensions 2.9
7.8
5.1
5.3
compact rigid
random
random
globular globular
coil
coil
ellipsoidal
ellipse
prolate
prolate
14x3.8
4.5x3.5 4.5x3.0
ellipsoidal
ellipsoidal
ellipsoidal
(nm)
X2.5
X3.0
T.D.2) stability kJ mol" 1
54
60
% a-Helix
30
55
75
27
10
1-10
% P-Sheet
27
16
-
16
20
13-16
K -Casein
BLG
RNase
a-chymotrypsin
Cytochrome- C
Source
cow's milk
cow's milk
bovine
bovine pancreas
bovine heart
M/kDa
19
18
25.2
12.2
i.e.p.
4
5.2
9.4
8.1
10.0
Molecular
spherical
rigid
globular
globular
globular
ellipse
ellipsoidal
ellipse
pancreas
structure Shape
13.68
globular and
prolate ellipsoidal 3.6X7.0
dimensions (nm) T.D.21
3.8x2.8
2.8x3.0x3.4
X2.2 50
54
38
stability kJ mol" 1 % a-Helix
14
10
11
13
51
% P-Shect
31
50
33
25
-
Data obtained from several sources, collated by J. Benjamins and W. Norde. From differential scanning calorimetry; the Gibbs energy is given for the N —> D transition at pH of maximal stability.
PROTEIN ADSORPTION
3.47
Figure 3.24. Two-dimensional equations-of-state at the air-water interface for adsorbed monolayers of five proteins and PVA (M = 42.000 ). The concentrations, c, of the solutions from which the monolayers were made (given in g d m " ' ) are indicated. Discussion in the text (courtesy, J. Benjamins). number. The thermodynamic stability given refers to thermal denaturatlon, but conformatlonal alterations upon adsorption probably follow a different path, so the conformatlonal stability may also be different. Figure 3.24 contains n{F) curves for five different proteins, including the uncharged polymer poly(vlnyl alcohol) for the sake of comparison. In these experiments it was measured by the Wilhelmy plate technique, and r was obtained ellipsometrically on the same interface.
3.48
PROTEIN ADSORPTION
The first notable feature is that data points obtained at different values of c all collapse to one 'mastercurve'; surface pressures are fully determined by F, and not by the concentration in solution needed to obtain these surface concentrations. Stated differently, for these adsorbed monolayers the n{F) relationship is unique for each protein and so we may ask whether it makes sense to speak of a 2D equation of state. The second observation is that, for all five proteins, the first 0.7-1.1 mg m~2adsorbed do not contribute measurably to the surface pressure. This is a typical protein feature, which is not exhibited by the random and heterodisperse polymer PVA. For spread monolayers, the TI(F) curves are rather different (results not shown) but the horizontal initial parts persist. These parts coincide with the initial steep rise in the isotherm, F(c). In this part of the isotherm the adsorbing molecules have the entire surface at their disposal; they can complete any conformational adjustment during the time it takes to carry out an ellipsometry measurement (ca. 10 min.). If one is interested in the occurrence, and rates, of re-conformation on much shorter time scales, dynamic measurements have to be carried out. Indeed, such processes have been observed with relaxation times of a few seconds up to a few minutes. There is some correspondence between the length of the very low n range; for more rigid molecules, it is longer. Lysozyme being the top, lower for the albumins, still lower for the structurally relatively 'weak' caseins down to virtually zero for the fully structureless and heterodisperse PVA. Semiquantitative computations with the ideal 2D equation of state (n = RTF, with F in moles m" 2 ) show that, in these initial parts, K is indeed very low; the pressure starts to build up when this layer is completed and lateral interaction sets in. This lowpressure part was also studied optically for lysozyme by Ericson et al.1'. For proteins like the albumins the plateau adsorption passes through a maximum around the i.e.p., similar to that at SL interfaces (fig. 3.22), and for the same reason. The different rates of unfolding are also reflected in the dynamic surface tension. In fig. 3.25, this tension is given as a function of the relative expansion rate of a surface,
Figure 3.25. Dynamic surface tensions obtained by the overflowing cylinder technique for 0.25 gdm~ 3 solutions of p-casein, [3lactoglobulin, ovalbumin and lysozyme. Temperature 25°C. (Redrawn from Van Kalsbeek and Prins, loc. clt.).
11
J.S. Ericson, S. Sundaram, and K.J. Stebe, Langmuir 16 (2000) 5072.
PROTEIN ADSORPTION
3.49
Figure 3.26. Apparent interfacial shear viscosities at the water-n-tetradecane interface. A, myosin; B, lysozyme; C, K-casein; D, gelatin; E, Na-caseinate; F, a s -casein; G, (5casein; H, P-lactoglobulin and I, a-lactalbumin. (Redrawn from E. Dickinson, J. Chem. Soc. Faraday Trans. 94 (1998) 1657.)
as measured by the overflowing cylinder technique described in sec. III.3.7e and illustrated in fig. III.3.73. These data have been obtained by van Kalsbeek and Prins1'. With increasing expansion rate, all the curves tend towards the surface tension of water but the 'weaker' casein can unfold far more rapidly than lysozyme, and hence take segments to the interface at a rate that is more compatible with the rate of expansion. Hence, with this protein the dynamic tension remains below the static tension over the entire range studied. Long term interfacial viscosity changes are shown in fig. 3.26. Plotted are the apparent surface shear viscosities at the water-n-tetradecane interface. Adsorption takes place from the aqueous phase, which contains 10~3 weight % of the protein at neutral pH. The method for obtaining the apparent shear viscosity was not reported, but two conclusions can be drawn. First, 77 continues to increase over tens of hours, i.e., long after adsorption has been completed. This increase must be caused by structural changes at the interface, the lateral interaction between the adsorbed protein molecules being an important characteristic. Secondly, there are dramatic differences between the relatively rigid globular proteins and the rather random caseins, the former group displaying values that are higher by several orders of magnitude. One interpretation is that the rigid spheres can acquire a much higher packing density, perhaps with some cross-linking. An additional feature is that, in several cases, the proteins can form 2D gels, which may exhibit rupture: when that occurs, a kind of slip viscosity is measured. The corresponding interfacial dilational properties exhibit the same trend with respect to the protein specificity but the
H.K.A.I, van Kalsbeek, A. Prins, in Food Emulsions and Foams, Interfaces, Interactions and Stability. E. Dickinson and J.M Rodriquez Patino, Eds., Royal Soc. Chem (London ) (1999), 91.
3.50
PROTEIN ADSORPTION
difference between them is less1'21. The inference of these observations is that the adsorbates behave like a 2D gel, with their strength depending on the nature of the protein; several protein molecules unfold to a greater extent at oil-water interfaces than during heat-denaturation in aqueous solution3 . Our last illustration, fig. 3.27, combines at least three important features. It relates the interfacial rheology to the surface pressure, different non-aqueous phases are compared, and the experiments have been carried out conscientiously. The reported rheological characteristic is the interfacial dilational modulus, K£. This quantity can be obtained by several methods, either by an oscillatory technique, using [3.7.1] to obtain Kg', or by a monotonous expansion, yielding Kg directly. The data reported stem from experiments with the dynamic drop tensiometer, described in connection with fig. HI.3.72. They have been compared to those from the barrier-andplate method (essentially a longitudinal wave technique) and were found to be very similar at given n. (For the sake of simplicity we use the symbol Kg ). The matter of the identity of results from different techniques is a separate problem, and is not generally solved. The three different non-aqueous phases are air (as in figs. 3.24 and 3.25), n-tetradecane (as in fig. 3.26), and triacylglycerol, which is a sunflower-seed oil, an important fluid for the food industry, and has useful properties as a solvent. The purity of this oil was improved (e.g., by silica extraction) and verified (by surface tension measurements) to ensure the absence of surface-active admixtures. The monolayers are formed by adsorption. As a result, n and Kg depend on the adsorption time, but for the n{F) relationship there exists a time-independent mastercurve (fig. 3.24). As the moduli also collapse into a protein-specific mastercurve at different solution concentrations, it makes sense to plot K^ as a function of n. Thus, the adsorption time is eliminated as a variable. Another matter is the frequency a> of the measurement. The higher is co, the more elastic is the behaviour, i.e., the more the first term on the r.h.s. of [3.7.1] dominates. The figure only gives data at 0.1 Hz. One would intuitively expect Kg to increase with n, and for low n this is indeed observed. For an ideal monolayer the slope should be unity, but experimental moduli are higher than that. At higher n 's, Kg increases less than linearly, and even diminishes. (3 -Casein exhibits a more capricious behaviour than its more rigid globular counterparts. Generally, the behaviour in triacylglycerol (which may be a better solvent for hydrophobic and/or even hydrophilic parts of the protein molecules) is more unequivocal than for air and tetradecane, in that the curves consist of two linear parts, even for the unruly p-casein. A kind of scaling is suggested, with at least the solvent 11 F.J.G. Boerboom, A.E.A. de Groot-Mostcrt, A. Prins and T. van Vliet, JVeth. Milk Dairy J. 50 (1996) 183. 21 A. Williams, A. Prins, Colloids Surf.. A114 (1996) 267. 31 J. Lefebvre, P. Relkin, in Surface Activity of Proteins. S. Magdassi, Ed., Marcel Dekker (1996), p.181.
PROTEIN ADSORPTION
3.51
Figure 3.27. Modulus as a function of the interfacial pressure for the indicated proteins in three non-aqueous phases. Temperature 25°C, frequency of drop oscillation 0.1 Hz. BSA = bovine serum albumin, BLG = bovine lactoglobulin. (Same source as fig. 3.24.) quality of the protein parts for the oil as one of the parameters. The inference is that the descending branch is caused by reconformation or collapse; it is certainly not a result of desorption. Formation of a second adsorbed layer is also unlikely, because there is no reason why this would reduce Kg . Experiments like those above can suggest ideas for further theoretical and experimental studies. One could think of optical investigations in conjunction with the rheol-
3.52
PROTEIN ADSORPTION
ogical ones which, In turn, may be extended by looking in more detail at the influence of
OJ12).
For reasons other than merely academic, this type of study has enormous practical relevance. Several of the proteins discussed here are milk constituents and are relevant in the preparation of food products: See the examples in chapters 7 and 8, and the book by Dickinson and Rodriguez Patino, mentioned in the General References, sec. 3.10. Other interest stems from interfacial enzymatic reactions, whose time dependence can be followed from surface pressure and surface rheology studies. By way of illustration we refer to studies from the group of Nitsch on lipase31 and catalase41 an investigation by de Roos and Walstra51 on the loss of activity of the enzymes lysozyme and bovine chymosin owing to adsorption on emulsion droplets, and a review by Panaiotov and Verger61. There is no need to state that this is a challenging area where significant developments may be expected. 3.8 Competitive protein adsorption and exchange between the adsorbed and dissolved states Protein adsorption is generally highly irreversible with respect to variation of the concentration in solution. Nevertheless, proteins can easily be desorbed by surfaceactive molecules that compete for the interaction sites, such as surfactants or other protein species. By studying protein adsorption from plasma, Vroman and co-workers observed that the adsorbed protein layer was rather dynamic. They observed that more abundant and smaller proteins adsorb first, because the transport to the sorbent is quicker. Proteins that have a higher affinity for the sorbent surface then slowly replace these adsorbed proteins7'81. For plasma proteins adsorbing to hydrophilic surfaces, HSA adsorbs first and is then followed by IgG, fibrinogen, fibronectin, Hageman factor, and high-molecular weight kininogen. This cascade of competitive protein adsorption is generally called the Vroman effect. As surfaces with an abundance of albumin tend to be less thrombogenic than those coated with fibrinogen or IgG, controlling the protein Also see, J. Benjamins, E.H. Lucassen-Reijnders, Surface Dilational Rheology of Proteins Adsorbed at Air-Water and Oil-Water Interface in Proteins at Liquid Interfaces 7 (1998) 34184. 21 M.A. Bos, T. van Vliet, Interfacial Rheological Properties of Adsorbent Protein Layers and Surfactants: a Review,. Adv. Colloid Interface Sci 91 (2001) 43-71. 31
W. Nitsch, R. Maksymiw, and H. Erdmann, J. Colloid Interface Sci.. 141 (1991) 322 R. Maksymiw, W. Nitsch, J. Colloid Interface Sci. 147 (1991) 67. 51 A.L. de Roos, P. Walstra, Colloids Surf. B6 (1996) 201. 61 I. Panaiotov, R. Verger, Enzymatic Reactions at Interfaces; Interfacial and Temporal Organization of Enzymatic Lipolysis in Physical Chemistry of Biological Interfaces, A. Baszkin, W. Norde, Eds. (see sec 3.10) ch. 11. 71 W.G. Pitt, K. Park, and S.L. Cooper, J. Colloid Interface Sci. I l l (1986) 343. 81 L. Vroman, A.L. Adams, J. Colloid Interface Sci. I l l (1986) 3 9 1 . 41
PROTEIN ADSORPTION
3.53
adsorption cascade from blood proteins is a crucial aspect of producing nonthrombogenic biomaterials. Logically, most research dealing with competitive protein adsorption has long been focused on the above-mentioned proteins to a variety of surfaces 1 ' 2 ' 34 '. As a general trend, the displacement of proteins is strongly related to their affinity for the interface. For example, human serum albumin (HSA) adsorbed to hydrophilic silica can be partly replaced by fibrinogen, whereas HSA adsorbed at hydrophobic methylated silica surfaces prevent any displacement or further adsorption of fibrinogen or IgG. A similar result was obtained when trying to replace fi -casein and fragments thereof with (3-lactoglobulin. If (}-casein is adsorbed to a methylated silica surface, no further adsorption of, or replacement by, (5 -lactoglobulin occurs. However, when the hydrophobic C-terminal half of p -casein is cut off, (3 -lactoglobulin easily replaces the hydrophilic, N-terminal half of P -casein, with adsorption kinetics similar to those for (5-lactoglobulin adsorption to a bare methylated silica surface. Thus, strong hydrophobic interactions, originating from a hydrophobic substrate or a hydrophobic protein surface, diminish the displacement of adsorbed proteins by proteins from solution. As illustrated in fig. 3.6, a requirement of adsorption reversibility is that not only can the molecules attach to and detach from the surface, but detachment should lead to their original structure. The extent of structural perturbation of adsorbed protein molecules depends on various parameters such as the hydrophobicity of the sorbent surface, the structural stability of the protein, the charge contrast between the protein and the sorbent surface and, not least, on the degree of coverage of the sorbent by the protein (see sec. 3.4). Accordingly, the structure of the exchanged protein molecules and, hence, the reversibility of the adsorption process, may be affected by the same variables. Unfortunately, very little attention has been given to this issue so far. Only a few papers have considered protein refolding after exchange from a (solid) surface. Below, some trends emerging from the few systems studied, are presented. They deal with the hard protein, lysozyme and the soft protein, serum albumin, exchanged at hydrophobic and hydrophilic colloidal particles. Samples of exchanged protein were obtained by incubating an excess of protein in a colloidal dispersion for a period of about 16 hours. As reported for various systems, 56 ' 71 after 16 hours essentially all protein molecules have been on and off the sorbent surface. The supernatant solutions may contain different populations of protein molecules with respect to the number of 11
A. Baszkin, MM. Boissonnade, Am. Chem. Soc. Symp. Sen 602 (1995) 209. B.K. Lok, Y.L. Chang, and C.R. Robertson, J. Colloid Interface Sci. 91 (1983) 2104. 31 M. Malmsten, B. Lassen, Colloids Surf. B4 (1995) 173. 41 T. Nylander, N.M. Wahlgren, J. Colloid Interface Sci. 73 (1994) 151. 51 V. Ball, P. Schaaf, and J.C. Voegel, in Surfactant Science Series 75. M. Malmsten, Ed., Marcel Dekker (1998) p. 453. 61 J.L. Brash, Q.M. Samak, J. Colloid Interface Sci. 65 (1978) 495. 71 J.C. Voegel, N. de Baillon, and A. Schmitt, Colloids Surf. 16 (1985) 289. 21
3.54
PROTEIN ADSORPTION
Figure 3.28. CD spectra (left) and DSC thermograms (right) of native BSA (drawn curve) and BSA exchanged from polystyrene surfaces (dashed). Aqueous solution, 25°C, pH 7. (Redrawn from W. Norde, C.E. Giacommelli, J. Biotechn. 79 (2000) 259.) times they have been exchanged at the surface. As they could not be separated, the experimental data are averaged over these populations. Changes In structural properties of the exchanged protein are probed by monitoring the thermal stability, and the secondary structure. By way of example, fig. 3.28 shows CD spectra and DSC thermograms for BSA in solution, before adsorption and after release from polystyrene surfaces. The homomolecular exchange of BSA at polystyrene clearly provokes a change in the secondary structure. The shifts in the CD spectra indicate the formation of (3-sheet structure at the expense of an a-helix. The structural alteration is most pronounced when exchanged with a lower adsorbed amount, corresponding to a lower protein concentration in solution. As discussed in sec. 3.3b, adsorption from a low proteinconcentration solution allows more time for adsorbed molecules to relax at the sorbent surface before neighbouring patches are occupied by protein as well. Hence, the degree of spreading, i.e., the structural perturbation, is higher at lower protein concentration, and this is reflected in the exchanged molecules. The DSC data are in line with the CD results. The thermogram of the exchanged BSA molecules deviates from that of native BSA, the more so the smaller is the adsorbed amount. Exchanged BSA is more thermostable and the heat-induced transitions occur over a wider temperature range. A broader transition region could well be caused by a heterogeneous protein population with molecules of different thermostabilities. Similar results have been reported for BSA exchanged at other hydrophobic sur-
PROTEIN ADSORPTION
3.55
faces, i.e., those of silver iodide11, and Teflon21. Formation of (3-sheet structures is often associated with intermolecular aggregation. Aggregation between exchanged BSA molecules is indeed observed, and intermolecular hydrogen bonding invoking aggregation through pi -sheet formation3'4' may cause the increased thermostability. In contrast with the hydrophobic sorbent surfaces, mentioned above, homomolecular exchange at hydrophilic silica surfaces neither leads to a modification in the secondary structure of BSA, nor to a change in the thermal stability, at any degree of surface coverage. Still, in the adsorbed state the BSA molecules have different structural characteristics, but upon release from the hydrophilic surface the protein molecules fully regain their original native structures '. Lysozyme, a hard protein, recovers its native conformation after being exchanged, 2)
irrespective of the hydrophobicity of the sorbent surface . Thus, if from the limited number of experimental data, a trend may be inferred, it would be the following: a less stable protein conformation, a more hydrophobic sorbent surface, and adsorption from protein solutions of lower concentrations, promote conformational changes that, at least partly, persist after desorption by a homomolecular exchange process. Consequently, for a given protein, to avoid irreversibility and loss of biological activity associated with interfacial exchange, a hydrophilic sorbent and a high degree of surface coverage should be selected. 3.9 Tuning protein adsorption for practical applications In various applications it is desired that proteins should be in the adsorbed state. For example, in emulsions, foams and other dispersions, adsorbed proteins may be used to stabilize the dispersed particles against coalescence and, possibly, disproportionation. For more information, the reader is referred to chapter 8. In other applications, the adsorbed proteins have to be biologically active. This requirement holds, e.g., for immobilized enzymes in bioreactors and biosensors, immunoproteins in ELISA devices and other diagnostic test kits, and for proteinaceous farmacons in drug targeting systems. As a rule, in order to retain biological activity, the structural integrity of the protein should not be perturbed too much upon binding at the surface. In other cases, where surfaces are brought into contact with protein-containing fluids, the adsorption of proteins at these interfaces should be avoided as much as possible. Adsorption of proteins from (biological) fluids is generally considered to be the first event in the biofouling process. Subsequently, bacterial and/or other biological cells (e.g., blood platelets, erythrocytes) deposit on the adsorbed protein layer. Adsorption-induced conformational changes in the protein molecules usually enhance the 11
T. Vermonden, C.E. Giacomelli and W. Norde, Langmuir 17 (2001) 3734. W. Norde, C.E. Giacomelli, Macromol. Symp. 145 (1999) 125. 31 V.J.C. Lin, J.L. Koenig, Biopolymers 15 (1976) 203. 41 R.J. Jakobsen, F.M. Wasacz, Appl. Spectrosc. 44 (1990) 1478. 21
3.56
PROTEIN ADSORPTION
interaction with the cells. After deposition, microbial cells multiply, forming so-called biofilms. Adhesion of blood platelets at surfaces of cardiovascular implant materials may lead to thrombus formation. Biofouling causes great problems in areas as diverse as biomedicine (implants, catheters, artificial kidneys, contact lenses, teeth, and dental restoratives, etc.), food processing (heat exchangers, separation membranes, etc.) and the marine environment (ship hulls, desalination units, etc.). Knowledge of the mechanism of protein adsorption provides a few clues to control the process, for example by adapting the charge and hydrophobicity of the sorbent surface and the protein molecules, and by selecting environmental conditions such as pH, ionic strength, or temperature. However, practical applications often do not allow much freedom of choice. The composition of a natural fluid is a pre-set condition, and only the surface properties of a (synthetic) material contacting the fluid can be chosen to some extent. Mutatis mutandis, the same is true for immobilized enzymes and immunoproteins in bioreactors and bio- and immunosensors. Moreover, most natural fluids contain a mixture of proteins. Selection of a combination of surface properties with respect to, e.g., charge and hydrophobicity may involve a low adsorption affinity for one protein but a high adsorption affinity for another. A generic approach to influence the magnitude of the interaction between a protein molecule (and also other macromolecules and particles) and a sorbent surface is to manipulate both the long- and short-range interaction forces by grafting soluble polymers or oligomers onto the sorbent surface. Thus, the application of oligomers of ethylene oxide (EO) on polystyrene surfaces leads to the retention of the enzymatic activity of adsorbed a-chymotrypsin, whereas this activity is lost in the absence of such pre-adsorbed oligomers. The short EO chains prevent the enzymes' making intimate contact with the polystyrene surface, without hampering adsorption. As a result, the stress exerted by the sorbent surface on the protein molecule is less, so that the protein's structural integrity, and hence its biological functioning, is less perturbed, as shown in fig. 3.29. Another interesting example is the steering effect of pre-adsorbed polyethylene-oxide (PEO) molecules on the orientation of consecutively adsorbed IgG molecules. IgG molecules are anisotropic, and by realizing proper spacing between the adsorbed PEO molecules the IgG molecules can be forced into the desired orientation at the sorbent surface, i.e., with their antigen binding sites exposed to the solution in which antigens have to be detected. According to this sieving principle, the immunological activity per molecule of IgG adsorbed on a polystyrene surface could be doubled11. By far the greatest part of the recent research on modifying surfaces by grafting
" M . G . E . G . Bremer, Immunoglobulin Adsorption on Modified Surfaces, Wageningen University, the Netherlands, ch. 7 (2001).
PhD Thesis,
PROTEIN ADSORPTION
3.57
Figure 3.29. Temperature dependency of the specific activity of a-chymotrypsin in solution (o), adsorbed on polystyrene (A) and on polystyrene covered with octaethyloxide chains (•). pH = 7.1. (Redrawn from W. Norde and T. Zoungrana, Biotechnol. Appl. Biochem. 28 (1998) 133.)
soluble polymers aims at the prevention of protein adsorption and/or adhesion of biological cells". Because the natural habitat of proteins and biological cells is an aqueous medium the polymers used must be well-soluble in water. In most cases, PEO is used. Sometimes the use of polysaccharides, e.g., dextrans, is reported. The efficacy of the grafted polymers in reducing protein adsorption depends primarily on two characteristics of the polymer layer: (i) the grafting density and, (ii) the extension Into the solution. Despite controversy In the literature data, some trends emerge. As expected, protein adsorption decreases with increasing grafting density of the polymer. Also, as a rule, protein repellency increases with increasing length of the grafted polymer chains. At a grafting density where the separation distance between neighbouring polymer molecules is smaller than twice the radius (of gyration) of the polymer molecule, the grafted polymer chains have to stretch out into the solution. The polymer layer is said to attain a 'brush' conformation. The brush conformation determines the efficacy of protein repulsion. More specifically a few more trends may be mentioned. In the case of a not-too-thick polymer layer and relatively large protein molecules, long range dispersion forces may cause accumulation of protein molecules at the outer edge of the polymer brush. Furthermore, a higher brush density is required to resist adsorption of
11 E.P.K. Currie, W. Norde, and M.A. Cohen Stuart, Adv. Colloid Interface Sci. 100-102 (2003) 205.
3.58
PROTEIN ADSORPTION
Figure 3.30. The influence of the adsorption of BSA by the PEO grafting density and chain length: 700, 445 and - - 148 EO monomers. (Redrawn from ".)
smaller protein molecules. Currie et al.11 reported an even more complex interaction between proteins, i.e., bovine serum albumin (BSA), and PEO brushes. The results of that study are shown in fig. 3.30. At relatively low grafting densities long PEO chains in a brush stimulate BSA adsorption. However, with increasing grafting densities, the adsorption gradually decreases. This result can only be explained by assuming an attractive interaction between the PEO chains and the protein molecules which, in some unknown way, is determined by a combined effect of the length of the PEO chains and the grafting density. These results are supported by data on protein-polymer brush interaction forces, obtained by Norde and Gage21 and by Leckband's group3'41. They report an activation energy up to a few tens of kT for protein (i.e., streptavidin) molecules to penetrate into a PEO brush. However, when by applying a compressive load the activation energy barrier is surpassed, PEO-streptavidin contacts are stabilized by 1-2 kT. It seems that steric and osmotic forces are involved in rejecting proteins from polymer-brushed surfaces, but more subtle interactions such as the conformationdependent hydration of the EO units may be decisive with respect to whether the PEO's layer is protein resistant or not51. Theories describing these phenomena are currently under development. 3.10 General references Physical Chemistry of Biological Interfaces, A. Baszkin and W. Norde, Eds., Marcel Dekker (2000).
11 E.P.K. Currie, J. van der Gucht, O.V. Borisov, and M.A. Cohen Stuart, Pure Appl. Chem. 71 (1999) 1227. 21 W. Norde, D. Gage, Langmuir,20 (2004) 4162. 31 N.V. Efremova, S.R. Sheth, and D.E. Leckband, Langmuir 17 (2001) 7628. 41 S.R. Sheth, N.V. Efremova, and D.E. Leckband, J. Phys. Chem. B104 (2000) 7652. 51 M. Morra, Poly(ethylene-oxide) Coated Surfaces in Water in Biomaterials Surface Science, M. Morra, Ed., John Wiley (2001) ch. 12.
PROTEIN ADSORPTION
3.59
J.P. Brash, P.W. Wojciechowski, Interfacial Phenomena and Bioproducts, Marcel Dekker (1996). (Includes various applications of proteins at interfaces.) Colloidal Biomolecules, Biomaterials and Biomedical Applications, Surfactant Science Series 116, A. Elaissari, Ed., Marcel Dekker (2003). T.E. Creighton, Proteins: Structures and Molecular Properties, 2nd ed., W.H. Freeman (1993). (Textbook) Food Emulsions and Foams, Interfaces, Interactions and Stability, E. Dickinson and J.M. Rodriguez Patino, Eds., Roy. Soc. Chem. (London) (1999). (Role of proteins in stabilizing emulsions and foams, dynamics of proteins at fluid interfaces.) E. Dickinson, Adsorbed Protein Layers at Fluid Interfaces: Interactions, Structure and Surface Rheology, in Colloids Surf. B15 (1999) 161-176. (Review, includes rheology and competition with surfactants.) Food Colloids, E. Dickinson, Ed., Curr. Opin. Colloid Interface Sci. 8 (2003) 346421. (Update, contains various aspects of proteins at interfaces.) C.A. Haynes, W. Norde, Coilloids Surfaces B2 (1994) 517. (Review, emphasizing thermodynamics of protein adsorption.) V.N. Izmailova, G.P. Yampolskaya and Z.D. Tulovskaya, Development of Rebinder's Concept on the Structure-Mechanical Barrier in the Stability of Dispersions, Stabilized with Proteins, in Coll. Surf. A160 (1999) 89-106. (Paper with a review character; emphasis on the mechanical strength of proteinaceous adsorbates.) Y. Lvov, M. Mohwald, Protein Architecture: Interfacing, Molecular Assemblies and Immobilization Biotechnology, Marcel Dekker (1999). (State of the art on protein immobilization, emphasizing biotechnological and biomedical applications.) F. MacRitchie, Chemistry at Interfaces, Academic Press (1990). (Textbook) F. MacRitchie, Proteins at Interfaces in Adv. Protein Chem. 32, G.B. Anfinsen, J.T. Edsall and F.M. Richards, Eds., Acad. Press (1978) 283-326. (Protein adsorption and its effect on biological activity.) F. MacRitchie, Spread Monolayers of Proteins, Adv. Colloid Interface Sci. 25 (1986) 341. (Review, 122 references.) Biopolymers at Interfaces, 2nd ed., M. Malmsten, Ed., Surfactant Science Series 110 Marcel Dekker (2003). (Contains 30 chapters on principles, techniques and applications.) W. Norde, Adv. Colloid Interface Sci. 25 (1986) 267. (Review, 178 references.)
This Page is Intentionally Left Blank
4
ASSOCIATION COLLOIDS AND THEIR EQUILIBRIUM MODELLING
Frans Leermakers, Jan Christer Eriksson and Hans Lyklema 4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
Introduction
4.2
4.1a
Aqueous solution amphiphiles
4.2
4.1b
Hydrophile-lipophile balance (HLB)
4.6
4.1c
Critical micellization concentration (cm.c.)
4. Id
Surfactant packing parameter
4. le
Phase diagrams
4.7 4.14 4.16
Classical thermodynamics
4.18
4.2a
Thermodynamics of small systems
4.20
4.2b
The mass action model
4.22
4.2c
Implication for molecular modelling
4.24
4.2d
Fluctuations in micelle sizes
4.27
Molecular modelling
4.29
4.3a
Molecular simulations
4.30
4.3b
Self-consistent field (SCF) theory
4.31
4.3c
Quasi-macroscopic models
4.42
SCF for (spherical) non-ionic micelles
4.47
4.4a
4.49
Micelles at and above the c.m.c.
4.4b
Structure of C p E 5 micelles
4.53
4.4c
Trends for various micellar characteristics
4.56
4.4d
Quasi-macroscopic approaches to non-ionic micelles
4.59
4.4e
Pluronic micelles
4.61
SCF for (spherical) ionic micelles
4.64
4.5a
Micelles at the c.m.c. in various salt solutions
4.66
4.5b
Radial profiles
4.69
4.5c
Chain length dependence
4.72
4.5d
Specific ion effects
4.73
4.5e
Quasi-macroscopic approach to ionic micelles
4.75
Linear growth of micelles
4.76
4.6a
Phenomenological model for rod-shaped micelles
4.77
4.6b
SCF theory of infinitely long linear micelles
4.80
4.6c
The endcap energy
4.83
4.6d
Persistence length of wormlike micelles
4.85
4.6e
Second c.m.c. for ionic surfactants
4.86
Biaxial growth of micelles
4.88
4.7a
Thermodynamic stability of infinite bilayers
4.89
4.7b
Finite size disks
4.91
4.7c
Homogeneously curved surfactant bilayers
4.93
4.7d
On the thermodynamic stability of vesicles
4.96
Interactions between parallel lamellar surfactant layers
4.101
4.8a
Undulation forces between bilayers
4.102
4.8b
Intrinsic interactions between surfactant bilayers
4.104
4.8c 4.9
Liquid crystalline phases
Applications of the modelling
4.107 4.107
4.9a
Binary non-ionic ionic surfactant systems
4.108
4.9b
Solubilization of apolar compounds
4.113
4.10
Kinetic aspects of surfactant solutions near the c.m.c.
4.11
Outlook
4.118 4.120
4.12
General references
4.121
4 ASSOCIATION COLLOIDS AND THEIR EQUILIBRIUM MODELLING FRANS LEERMAKERS, JAN CHRISTER ERIKSSON AND HANS LYKLEMA
In this chapter we consider amphiphilic molecules in a solvent. Amphlphiles are (usually) short chain molecules that have two distinct sides. One moiety of the molecule dislikes the solvent, the other part favours it. When water is the solvent these moieties are called hydrophobic and hydrophllic, respectively. As a result of these interactions, the molecules associate and form objects of mesoscopic size, which are classically called association colloids. Without the process of self-assembly of amphiphiles into association colloids, interface and colloid science would not nearly be as challenging and significant as it is. The association of the hydrophobic (apolar) parts of the amphiphiles gives the driving force for the assembly (hydrophobic bonding). These fragments form the cores of such aggregates. The hydrophilic (polar) sides of the amphiphilic molecules accumulate at the water-core interface, forming the corona. The formation of the corona eventually counteracts the association process, and when finite objects are formed these are known as micelles. Micelles are prominent members of the colloid family and they typically introduce a large interfacial area to a system. Micelles are thermodynamically stable and it is possible to apply the powerful machinery of statistical thermodynamics to describe their self-assembly. The goal of such analyses is to understand such features as the micelle size, shape, the (in)stability against dilution and the influence of various physicochemical parameters in relation to the molecular architecture, as well as the solvent properties. In this chapter we will review the fundamental aspects of association colloids with a focus on the understanding of the underlying physicochemical phenomena. We cannot review all aspects of association colloids because the huge volume of experimental literature, the many implications to countless applications and the detailed modelling of generic and molecular-specific issues simply defy condensation into one coherent and concise text. Here we choose for an approach that largely relies on self-consistent field modelling. We like to show that, within one set of approximations, it is possible to cover a wide range of properties of these interesting systems. In particular, we will elaborate on the way in which modern computational methods can help to gain a better understanding of association colloids. Computer modelling cannot be regarded as the Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
4.2
ASSOCIATION COLLOIDS
end point of investigations. We therefore also pay close attention to various aspects of analytical quasi-macroscopic approaches. 4.1 Introduction Amphiphilic self-assembly introduces (liquid) hydrophobic domains in an aqueous solution. The sizes and shapes of the micelles depend strongly on the nature and concentration of the surfactant and other physicochemical conditions (pressure, temperature, ionic strength, additives) and determine in large the many applications. The use of amphiphiles in, for example, food processing or as a lubricant probably dates back to the beginning of mankind. The link between the physical properties and the molecular structure is much more recent. The term "micelle" dates back to McBain who referred to it in a Faraday Discussion remark" in a General Discussion on colloids and their viscosity (!). Early on the shape of small micelles was disputed, partly as a result of interpretational problems with X-ray data, but nowadays spherical micelles are considered as the prototypes for surfactant aggregates. Besides McBain, Hartley21, Debye31, and Stigter41, many others have in the early stages contributed to their understanding. In this connection it may also be mentioned that monolayers of surfactants at waterair interfaces have served as models for determining various parameters important for micellization. Recall Benjamin Franklin's seminal spreading experiments of oil on Clapham Lake (introduction of chapter III.3) and assessing molecular areas from j'(logc) curves for Gibbs monolayers (c.f. sec III.4.6d). 4.1a Aqueous solutions of amphiphiles Amphiphiles can hardly be discussed without explicit reference to the solvent in which they are dispersed. In order to have stable micellar structures, the solvent should be selective in the sense that it should be poor for one part of the molecule and good for the other. Self-association of surfactants in apolar media results in so-called reverse micelles with a polar core and apolar corona. These systems are close to water-in-oil microemulsions. More specifically, it is possible to tune the solvent selectivity, and thus the micellar structures, by using supercritical solvents. Space does not allow us to visit these systems (see e.g.51 for micellization in non-aqueous and mixed solvents). In this chapter the main focus is on aqueous systems. The hydrogen bonding capabilities of water makes it a highly associative liquid. 11
J. McBain, Trans. Faraday Soc IX (1913) 99. G.S. Hartley, Aqueous Solutions of Paraffin Chain Salts, Hermann (1936). 31 P. Debye, J. Phys. Coll. Chem. 53 (1949) 1; Ann N.Y. Acad. Set 51 (1949) 573. 41 D. Stigter, Rec. Trav. Chim. 73 (1954) 593; J.Th.G. Overbeek, D. Stigter, Rec. Trau. Chim. 75 (1956) 1263. 51 J. Eastoe, A. Dupont, and D.C. Steytler, Current Opin. Coll. Interf. Set 8 (2003) 267. 21
ASSOCIATION COLLOIDS
4.3
Molecules, or parts of them that are not able to take part in this hydrogen bonding network, are likely to be rejected by water. That is why oil and water do not mix. The hydrogen atoms in hydrocarbon molecules or tails are too strongly bonded to the carbon atoms to take part in hydrogen bonding. Hydrocarbon tails are therefore excellent entities to make the hydrophobic parts of surfactants. The longer the tails, the stronger the demixing with water and the stronger the driving force for micelle formation. In other words, the Helmholtz energy of mixing hydrocarbons in water is very unfavourable. It amounts to an unfavourable exchange energy of order kT per -CH 2 unit. There are many names associated with subsets of amphiphiles. Such names reflect some feature of the molecular structure or some physical property they represent or are coupled to a typical application. The name "surfactant" refers to the ability to adsorb strongly onto many surfaces while reducing the interfacial tension. Nowadays the keyword surfactant is used as a generic term. However, there are many other molecules that are surface active but which are not amphiphiles; homopolymers adsorb strongly onto almost any interface (see chapter II.5) and ions to (charged) mercury surfaces. If amphiphiles are used as a means to solubilize apolar compounds in an aqueous medium, they may be named emulsifier or they are called foaming agent if they stabilize thin liquid films. Surfactants are wetting agents when promoting, for example, the spreading of crop protection agents on leaves. If the surfactant is used to stabilize apolar particles in water (or a polymer matrix), the term compatibilizer is sometimes used. Amphiphiles are known to increase the solubility of other organic substances in water. Certain sub-classes of surfactants also have dedicated names. Substances that co-adsorb, or co-micellize, and in this way promote the activity of the surfactant, are called linkers, or, when in aqueous solution, hydrotopes. A class of its own is the set of lipid molecules, which are the main components of biological membranes. Molecules that have a hydrocarbon tail and a hydrophilic head group are the classical examples of a micelle-forming surfactant. When the head group is charged, we have the sub-class of ionic surfactants. Ionic surfactants are among the most prominent surfactants in aqueous media. We distinguish cationic and anionic surfactants. Sodium dodecyl sulfate (NaDS) or (sodium lauryl sulfate) is the best-known anionic surfactant. This surfactant is almost always polluted by dodecylalcohol, which is its hydrolysis product. Cetyltrimethylammonium bromide (CTAB) is a typical example of a cationic surfactant. We will see below that such micelles are strongly dependent on added indifferent electrolyte and that the kind of counterion is also relevant in these systems (see table 4.1). The phase diagrams of surfactants may differ significantly depending on the hydrophilicity or size of the counterion. Zwitterionic surfactants have both a negative and a positive charge in the head group. These are less sensitive to added salt. A snapshot of the structure of a small ionic micelle, as generated by molecular
4.4
ASSOCIATION COLLOIDS
Figure 4.1. a) Computer (MD) generated snapshots of an ionic micelle composed of cesium pentadecafluorooctanoate significantly deviating from the spherical structure". The dark spheres are the C-units (the fluor atoms are omitted, that is why one can "look through" the structure; the CH2 unit is somewhat more polar than the CF2 one), the light spheres are the oxygen of the carboxylic head group, the counterions (Cs) are dark gray spheres, b) MD snapshot of a spherical non-ionic micelle composed of C 12 E 6 surfactants21. The core is made up of densely packed alkyl tails (black, space-filling spheres). The lighter gray hairs are the EO fragments forming the corona. Water molecules are not shown, c) Schematic cross-section of a Na-dodecylsulphate micelle. The large spheres with a minus sign are the OSOg groups; the compensating countercharge is indicated by + and - signs, d) Schematic cross-section of a nonionic micelle, the core of densely packed alkyl chains are thin line parts, the ethylene oxides are the fat line parts. dynamics simulations, Is given in fig. 4.1a. In principle, we are interested in the average behaviour of surfactant assemblies and therefore it is always dangerous to discuss snapshots. Apart from this, we may be amazed at this stage about the complexity of the molecular assembly and the apparent importance of fluctuations (chain conformations, micelle shape, etc.). From the above, it follows that a CH2 group is a classical (but not the only) hydrophobic building unit. On the other hand, an oxygen atom in a larger molecule can 11 21
S. Balasubramanian, S. Pal, and B. Bagchi, Current Sci. 82 (2002) 8456. F. Sterpone, C. Pierleoni, G. Briganti, and M. Marchi, Langmuir 20 (2004) 4311.
ASSOCIATION COLLOIDS
4.5
accept a H-bond from water and is a typical hydrophilic unit. It is possible to combine these two in a regular fashion. The methyleneoxide (MO) chain -(CO-) n , where the Hatoms around the C are dropped for convenience, is the first member. Ethyleneoxide (EO) chain -(CCO-)n is the second; the propylene oxide (PO) chain -(C[CH3]CO-)n , where the CH3 branches off from the main chain, is the third compound. These homopolymers are unable to form micellar-like objects on their own because the amphiphilicity on the monomer length scale is insufficiently expressed. In this series, only the EO member mixes readily with water. The other two are not miscible in all proportions with water. A detailed understanding of this is still missing. Apparently the dimensions of the water H-bonding network "match up" with ethyleneoxide, but not sufficiently well with the other two. With this information, it is possible to design non-ionic surfactants. Combining sufficiently long aliphatic (hydrocarbon) tails and EO head groups leads to surfactants, which are nowadays available in high purity and with high homodispersity. The general constitution is C n E x , where E stands for EO. Results for these systems from before the 1970s should be regarded with some caution because at that time the products were rather polydisperse, especially with respect to the EO parts. When both n and x values are very small, we have weak surfactants. A typical example for a reasonably strong surfactant is for n = 12, x = 6 . I n this case, the lengths of the apolar and polar parts are about equal. A computer-generated example of the spatial structure of a spherical micelle composed of the C 12 E 6 surfactants is given in fig. 4.1b. Again, this example may serve to sharpen our intuition about these molecular aggregates. In this chapter our attention will be focused on strong surfactants, i.e. surfactants that form micelles at low surfactant concentrations. The schematic drawing of fig. 4. Id is derived from fig. 4.1b. Of course there are many polymeric amphiphiles, e.g. Pluronics, which have PPO as the apolar and EO as the polar moiety. The diblock copolymers are the polymeric analogues of the classical surfactants, but there are many other options. For example, one can have amphiphilic side chains also known as polysoaps. There exist protein molecules, such as caseins that have emulsifying properties. They have a rather apolar domain, but also a polar one, and further contain a conformationally disordered fragment that is highly solvated. This shows that amino acids can also serve as the surfactant building blocks. Polysaccharides are yet another class of water-soluble compounds. Linking aliphatic tails to a short string of these molecules leads to environmentally friendly, biocompatible amphiphiles ["sugar surfactants"). Surfactants based on n-alkyl chains generally form micelles for which the core is densely packed, c.f. fig 4.1b. In this case, the dimension of the apolar part of the surfactant molecule is a fundamental property. For this reason, it is evident that relatively small deviations from the linearity, which leads to more compact but less regularly packed tails, may have important consequences for self-assembly. We may consider a given polar head group connected to various isomers of a given apolar
4.6
ASSOCIATION COLLOIDS
constituent. The chain can be branched, it can be split into two or more subchains connected to the head, etc. The micelles composed of these isomers will differ noticeably and systematically. For example, when two ionic surfactants are coupled by a short bridge {spacer) at the position of the head groups, we have twin surfactants also called gemini. The name was coined by Menger et al.1'. In recent years the study of gemini surfactants has seen increasing interest, because it was shown that the micellization properties can be nicely varied by, for instance, controlling the length of the spacer. With the spacer, one can insert an extra steric contribution to the packing of head groups in the corona region. For overviews see refs. 23) . This again points to the fact that molecular structure is important for self-assembly. Alkyl chain lengths on the order of C16 or longer occur frequently in nature, as in lipid molecules. One then has to be aware of the fact that densely packed layers of these molecules will not necessarily be in a liquid-like state at room temperature. If linear, they can give rise to liquid-crystalline ordering. Unsaturated and branched segments in the chain will frustrate crystallization, and thus the degree of saturation is yet another parameter that should be taken into account, especially when fatty acid or mono or di-glyceride molecules are considered. An interesting, and in biological systems relevant example, is the possibility of semirigid molecules in which connected C5 and or C6 rings are grouped together, such as in cholesterol and steroids. Such molecules may have polar and apolar substituents. When the polar ones are predominantly on one face of the molecular plane and the apolar ones on the other side, we obtain facial amphiphiles that form aggregates of finite size that are distinctly different from the micelles that are discussed below. The above set of variables that are relevant for amphiphilic molecules is by no means complete, but we trust that the examples given suffice to indicate the vast scope. 4.1b Hydrophile-lipophile balance (HLB) From the above, it is clear that one can, at least in principle, generate series of surfactants in which the apolar part (usually called the tail(s)) and the polar part (usually called the head) systematically vary in size. This is particularly possible for the nonionic polymeric amphiphiles. Within a series, it is possible to rank them according to the polar/apolar ratio or the hydrophile/lipophile balance (HLB value). The empirical HLB notion was developed by William C. Griffin41 in 1949 who proposed to compute the HLB value as the molecular weight percent of the water-loving portion of the surfactant divided by 5. Experience showed that if a surfactant has an HLB = 1, it is very oil-soluble, however a surfactant with an HLB = 15 is water-soluble. Surfactants
11
F.M. Menger, C.A. Littau, J. Am. Chem. Soc. 113(1991) 1451. R. Zana, Adv. Colloid Interface Set 97 (2001) 205. 3 R. Zana, in Novel Surfactants: Preparation. Applications and Biodegradability, K. Holmberg, Ed., Marcel Dekkcr (1998) 241. 41 W.C. Griffin J. Soc. Cosmet. Chem. 1 (1949) 311. 21
ASSOCIATION COLLOIDS
4.7
with an HLB = 1-3 may be used to mix unlike oils, that water-in-oil emulsions can form with surfactants that have an HLB = 4-6, that one can compatibilize small particles in oil using surfactants with HLB = 7-9, that surfactants with HLB = 7-10 may be used to form self-emulsifying oils, that blends of surfactants in the range HLB = 8-16 can be used to make oil-in-water emulsions, that detergent solutions require surfactants with HLB = 13-15 and that surfactant blends with HLB = 13-18 may be used to solubilize oils (and form microemulsions) in water. It is also very useful to know that for making oil-in-water emulsions, the combination of the oil and the suitable HLB value changes from HLB = 6 for vegetable oils, HLB = 8-12 for silicone oils, HLB = 10 for petroleum oils and HLB = 14-15 for fatty acids and alcohols. In conclusion, the HLB notion has been, and still is11, quite useful for formulation purposes. However, the HLB value does not directly relate to the efficiency to form micelles, and for modelling purposes the HLB value is not a prominent quantity.
4.1c Critical micellization concentration (c.m.c.) Experience has shown that micelles only form above a threshold concentration, often referred to as the critical micellization concentration (c.m.c), which is for any surfactant system the most characteristic physical quantity. The lower the c.m.c, the more efficient the surfactant is to adsorb onto various surfaces. For applications in which the presence of micelles is needed, one obviously should choose a concentration above the c.m.c. The issue of sharpness of the c.m.c. will receive detailed attention below. The appearance of a c.m.c. is easily explained borrowing the ideas of macroscopic phase separation of, for example, oil and water, which also gives a rough estimate of the c.m.c. To first order, the surfactants are at low concentrations kept in solution as freely dispersed molecules because the translational entropy can (over)compensate the unfavourable local interactions the tail has with the solvent. With increasing concentration, the translational entropy per chain diminishes and, at a (fairly) well-defined point, it becomes more favourable to form a dense phase of surfactants. Obviously, the fact that objects of finite size are formed proves that such a phase separation model is flawed and more sophisticated models have to be introduced. Anticipating these extensions we nevertheless note that one can easily show, for instance, using the Flory-Huggins (FH) theory, that the maximum solubility (binodal) of oils in water (strong segregation) decreases exponentially with the chain length of the oil IV in line with experimental findings21. Below (see table 4.1) we will see that the c.m.c. indeed has this trend, proving the fact that the demixing of tails and water is crucial for determining the c.m.c. More importantly, we may use the FH theory to find our first estimate of the c.m.c.
11
X.F. Li, H. Kunieda, Curr. Opin. Coll. Interf. Sci. 8 (2003) 327. P.J. Flory, Principles of Polymer Chemistry, Cornell University Press (1953); C. Tanford, The Hydrophobic Effect. Formation of Micelles and Biological Membranes, 2nd ed. Wiley and Sons Inc. (1980). 21
4.8
ASSOCIATION COLLOIDS
The shortcomings of the simplistic phase separation model should not only be attributed to the fact that in the FH theory the solubility characteristics of the head groups are not included. Indeed, it is possible to set up a more elaborate FH-like theory for copolymers and also introduce interaction parameters for the head groupwater contacts, as well as for the head group-tail contacts. Such extended FH theory will also predict some trend of the c.m.c. as a function of the head group properties. Although such an approach is useful because the c.m.c. is reasonably accurately predicted, the theory remains fundamentally wrong. In such a simplistic phase separation picture (oil-water demixing), the chemical potential of the oil is an increasing function of the oil concentration only in the onephase regions. In the two-phase region the chemical potentials are fixed to the binodal values. Above the c.m.c. (corresponding to the two-phase region in oil-water systems), association colloids of finite sizes are formed, which has many important consequences. One of them is that the chemical potential of the surfactant is no longer fixed, but continues to increase with the micelle concentration. The key property that fundamentally changes the above picture is that the aggregates that are formed do not grow to macroscopic size but remain mesoscopic. There are stopping mechanisms, which limit the growth to be discussed, in more detail below. As finite size effects turn the step-wise phase transition into an apparent first-order transition, it is quite obvious that the sharpness of the micelle formation depends on the number of monomers in the micelle and the distribution over micelle sizes. One other consequence is that classical phenomenological thermodynamical arguments are not enough to understand micelle formation. This means that we should carefully consider how to properly use classical thermodynamic arguments in relation to the special model notions invoked. (i) Determination of the c.m.c. As c.m.c.s are characteristic properties of surfactant solutions, which indicate when micelles start to form, it is important to measure them. Such measurements can only be carried out with satisfactory precision when the c.m.c. is sufficiently defined, i.e. provided that a well-defined surfactant concentration exists above which micelles suddenly appear when the concentration is increased. When there is such a sharp c.m.c, a variety of physical properties, measured as a function of increasing surfactant concentration, show a break at the c.m.c, these breaks often coincide within experimental error. Figure 4.2 gives an illustration. We shall now briefly discuss these methods. One of the most obvious ways to probe micelle formation is to shine light on them because the scattering increases strongly with the volume of the scatterer (chapter 1.7). The shorter the wavelength of the light, the more details one can see. That is why Xrays or neutrons are better equipped to obtain insight into the internal structure of the micelles than visual light. Dynamic light scattering may be used to find the hydrodynamic or diffusion radius of the micelles. When the form factor and the structure factor are both playing a role in scattering experiments, one would prefer to construct a
4.9
ASSOCIATION COLLOIDS
Figure 4.2. Schematic figure for the dependence of some physical measurables on the surfactant concentration near the c.m.c. On a log
Zimm-plot (see fig 1.7.12). In principle this would give a direct measure of the overall mass of the micelle, its aggregation number and the second virial coefficient; the latter is a measure of the strength of interaction between the micelles. To make a Zimm plot, however, one has to change the concentration of micelles, e.g. by diluting the micellar system. However, one should be aware of the fact that in general micelles respond by making changes in the aggregation number when they are diluted. For this reason, one should be careful and restrict the dilution to regimes where size changes may be neglected. In any case, static light scattering can be of use to detect the c.m.c. because it reacts to the volume of the particles. Neutron scattering, as well as X-ray scattering, are excellent tools to study the sizes and shapes of micelles in terms of the structure factor. Extensive studies on well-defined surfactants with partially deuterated samples (for neutron scattering) have revealed much insight into the structure of micelles11. Measuring the interfacial tension as a function of the total surfactant concentration provides another frequently used method. When the interfacial tension of an air-water interface to which the surfactants adsorb is plotted as a function of the logarithm of the total surfactant concentration, and not as a function of the chemical potential (logarithm of the monomer concentration), one finds a break in the curve. This "break" is positioned where the overall surfactant concentration starts to deviate from the monomer concentration; in practice it will coincide with the c.m.c. One may use Gibbs' law to extract from the slope of the surface tension as a function of the surfactant chemical potential the excess adsorbed amount of surfactants at the interface, which is the inverse of the area per surfactant molecule a m . This is an essential parameter to make a rough estimate for the surfactant packing parameter P as given by [4.1.4]. The method has also been used to determine salt concentration effects on the c.m.c. Moreover, it is well known that 3y/31nc surf gives am I p , where p ranges from unity (high ionic strength) to 2 (no added salt). See sec. III.4.6d. There are also some practical problems. When, for example, an anionic surfactant is used one should be 11
J.R. Lu, R.K. Thomas, and J. Pcnfold, Adv. Colloid Interface Sci. 84 (2000) 143.
4.10
ASSOCIATION COLLOIDS
aware of traces of multivalent counterions in the system. These multivalent ions may coadsorb with the surfactant onto the interface, which affects the value of a m . Another familiar problem is that some minor admixtures may give rise to minima around the c.m.c. Space does not allow us to go into all the details here. Colligative properties react on the number of kinetic units and hence are sensitive to micelle formation. If a large number of surfactants group into one micelle, only the micelle (and the mobile fraction of the counterions) has full translational entropy. If one normalizes the osmotic pressure with the overall concentration of surfactant, we expect (for ideal systems) to find a constant value. However, as soon as micelles form, the osmotic pressure does not nearly go up as fast as expected for monomers and the normalized osmotic pressure goes to small values. However, osmotic measurements are tedious, therefore making this method not particularly attractive. Micelles may be used as carriers for hydrophobic molecules. We can detect the existence of micelles by measuring the capabilities of a surfactant solution to solubilize these apolar entities. Obviously, in this case one will not measure the c.m.c. of a pure surfactant system, but the c.m.c. of a mixed system. We will study solubilization in sec. 4.9b. For charged surfactants, one can use conductivity measurements to determine the c.m.c, at least when not too much extra salt is added. At low c s u r f , the molar conductivity is ideally concentration independent (in practice activity corrections cause them to be proportional to l/^c s u r f ). However, at the c.m.c. micelles appear and these micelles contribute differently to the conductivity and thus the molar conductance differs below and above the c.m.c. There are two reasons for this. First, one may expect that a micelle in which there are g surfactants will carry the charge more efficiently than g individual surfactants simply because the drag normalized per segment is less. On the other hand, counterion binding to a surfactant in a micelle is much stronger than that to isolated surfactant ions, therefore it turns out that a substantial fraction of the counterions are rather closely bound to the micelle so that the effective charge that is carried is much reduced. The overall result is a jumplike change in the slope of the molar conductivity as a function of
ASSOCIATION COLLOIDS
4.11
the overall system. For these techniques, the extent of counterion binding is relevant. These techniques thus tend to indicate the condition where the concentration of freely dispersed surfactants is approximately equal to that of micelles. Practice has shown that for pure surfactants in aqueous solutions, c.m.c.s, obtained by different techniques are within experimental error, sufficiently close to speak of THE c.m.c, which can therefore be analyzed in terms of physical principles. Small systematic differences may have a methodical origin because the micellar size distribution has in principle a certain width: it would be entropically very unfavourable to force all micelles to adopt the same aggregation number g. Different techniques may well react differently on micelles of differing sizes and therefore give a slightly different c.m.c. When experiments yield aggregation numbers, basically an average is obtained, which is different for different techniques. Colligative experiments, such as the osmotic pressure, give a number-averaged aggregation number; light scattering, on the other hand, gives a weight average. Rheology gives yet another average (see [IV.6.11.9]). The sharper the distribution is, the closer the aggregation numbers obtained by different techniques. (il) Windows for micelle formation The temperature sensitivity of micellization, both for ionic as well as non-ionic systems, is of great relevance for many applications. The c.m.c.s of ionic surfactants depend weakly on temperature and the obvious control parameter in these systems is the ionic strength. However, non-ionic surfactants, particularly those of the C n E x type, are much less sensitive to the ionic strength, but are much more temperature sensitive. In contrast to the common trend that molecules become more soluble with increasing temperature, the EO-based surfactants show the opposite trend. As a result, here the temperature is the main control parameter. Typically, such non-ionic micellar solutions rather abruptly become very turbid at a given temperature, known as the cloud point (c.p.), indicating the formation of large aggregates in the solution. We will pay attention to this phenomenon in sec. 4.4. The cloud point depends on the nature and concentration of the surfactant. We will return to this when we discuss phase diagrams (cf. sec. 4. le). For various (typically non-ionic) surfactants, there is yet another relevant temperature known as the critical micellization
temperature
(c.m.t.). When we follow
surfactant solutions at sufficiently high concentration, we may find micelles only above some rather well-defined temperature. In particular the temperature-sensitive surfactants of the Pluronic type show this phenomenon. Upon decreasing the temperature, one may lose the presence of micelles rather abruptly. Apparently, the solubility of the surfactant as a whole becomes too high to allow the formation of stable micelles in solution. The location of the c.m.t. depends on the surfactant concentration. We will discuss the c.m.t. when we consider non-ionic surfactants in more detail in sec. 4.4.
4.12
ASSOCIATION COLLOIDS
The c.p.t. and c.m.t. usually bracket the temperature window in which the non-ionic surfactant system may be used. For charged surfactants, the ionic strength is the main control parameter. In principle one can expect that upon varying the ionic strength there exists a window in which the micelles are colloidally stable. Indeed, in combination with temperature variations there can exist special conditions where this stability is lost. In this context, Table 4.1. Selected c.m.c. values1'21. Surfactant
Formula
Temp.°C
c.m.c./M
ref.
anionic sodium octyl sulphate sodium decyl sulphate sodium dodecyl sulphate sodium tetradecyl sulphate lithium dodecyl sulphate
C8SO4Na+ C10SO4Na+ C12SO4Na+ C14SO4Na+ C^SOjLi"1"
sodium dodecyl sulphate
C^SO^Na"1"
potassium dodecyl sulphate
C12SOjK+
rubidium dodecyl sulphate
C12SO4Rb+
cesium dodecyl sulphate
C^SO^Cs*
25 25 25 25 25 33 40 25 33 40 55 33 40 33 40 33 40
1.33 xlO" 1 3.33xlO- 2 8.2-8.3xlO" 3 2.1xlO" 3 8.68 xlO" 3 4.46xlO- 3 8.25X10- 3 8.2-8.3X10" 3 8.10xl0" 3 8.05X10-3 9.85xlO- 3 6.71x10-3 6.90x10-3 5.90x10-3 6.01xl0- 3 5.90X10-3 6.10X10-3
1 1 1 1 2 2 2 1 2 2 1 2 2 2 2 2 2
C12TMA+Br-
25
1.25xlO- 2
1
C,8TMA+Br-
25
3xlO" 4
1
C12Py+Br"
30
1.18x10-
1
65
1.63x10"
1
25 25 25 25 25 25 25 25 25 25
0.33 xlO~4 0.55xl0" 4 0.39 xlO" 4 0.64 xlO" 4 0.8 xlO" 4 l.OxlO- 4 9.0xl0- 3 0.87 xlO" 4 lxlO- 5 1.25xlO-4
2 2 2 2 2 2 2 2 2 2
cationic dodecyl trimethyl ammonium bromide octadecyl trimethyl ammonium bromide dododecyl pyridinium bromide linear non-ionics
C
12E2 C 12E3 C 12E4 C 12E5 C 12E6 < -'12E8 C 10E9 C 12E9
Ci 4 E 9 C
11 21
12 E 12
Mukerjee-Myscls compilation, see sec 4.12c. Taken from compilations in FICS III, tables 4.4. 4.5 and 4.6 (sec 4.6).
ASSOCIATION COLLOIDS
4.13
the Krqfft point should be mentioned. The Krafft temperature is the temperature below which the surfactants enter a two-phase region: a concentrated surfactant phase (often a lamellar phase) coexists with an aqueous solution dilute in surfactants. (ili) Experimental trends for the c.m.c. In table 4.1, a number of c.m.c.s are collected. Notwithstanding the fact that different authors (and/or) different methods) may find (or lead to) slightly different values, the list is sufficiently established to deduce the following convincing trends. (i) The c.m.c. decreases rapidly with the increasing length of the hydrocarbon chain. This is a general feature applying to all classes of surfactants; obviously it reflects the decreasing monomer solubility in this direction. Experience has shown that for each surfactant logc.m.c. = A-Bt
[4.1.1]
if t is the length of the hydrocarbon chain. Here A is a quantity of dimensions loglCg]"1, where c 0 is the extrapolated reciprocal c.m.c. for t —> 0 . This quantity has of course no physical meaning, although it is certain that A depends on the nature of the head groups only; it is more negative for non-ionics than for ionic surfactants. The constant B is about 0.3 for ionics and 0.5-0.6 for non-ionics11. It means that adding a CH2 group to the hydrocarbon chain reduces the c.m.c. by a factor of about two or three for ionics and non-ionics, respectively. The inference is that the Gibbs energy of transporting a CH2 -group from a chain in solution to that in a micelle depends on the charge of the head group. Below we will show that this effect can be explained taking into account that the ionic strength is not kept the same for all the systems considered. (ii) (Not shown in the table). The constant B also depends slightly on the nature of the head group. Branching of the chain increases the c.m.c, apparently because branching frustrates the packing, and one CH2 is sacrificed whereas one CH is added and the number of CH3 end groups increases. (iii) For ionics and non-ionics of the same tail length t, the c.m.c. is much lower for the latter. On the one hand, the ionic double layer around the micelles opposes micellization, whereas in non-ionics the repulsion between the EO fragments plays a much weaker role. The c.m.c.s reported in table 4.1 are without added salt. This means that the surfactants themselves provide the ionic strength in solution. The higher the c.m.c, the higher the ionic strength at the c.m.c. The opposing electrostatic force depends on the ionic strength. (iv) In line with the previous item, the addition of electrolytes reduces the c.m.c. because it decreases the double layer Gibbs energy. Empirically21,
11
C. Tanford, The Hydrophobic Effect. Formation of Micelles and Biological Membranes, 2nd ed. Wiley (1980). 21 H.F. Huisman, Proc. Koninkl. Nederl. Akad. Wetenschap B67 (1964) 367.
4.14 logc.m.c. = C-Dlogc s t o t
ASSOCIATION COLLOIDS [4.1.2]
where c s tot is the ionic strength, i.e. it contains the contribution of the surfactant c surf ~ c m c - a n f i t n a f of the added salt c s logc.m.c. = C-Dlog(c.m.c. + cs)
[4.1.3]
Here, C depends on t as in [4.1.1]. (v) For non-ionics, there is little influence of the EO length on the c.m.c. (vi) The counterion effect (lyotropic or Hofmelster series) is less than that of the presence of charges, but the trend is clear. On the sulphate group, specific binding increases from Li+ to Cs+ . This sequence is the same as that found for poly(styrene sulphonate)-latices (sec. IV 3.13c) and for Gibbs monolayers of NaDS (sec. III. 4.6d) and is in line with Pearson's rule (sec. IV.3.9i). In the same vein, for cationics with dodecyl trimethyl and dodecylpyridinium groups the binding also increases with the radius of the counterion, i.e. from Cl~ to I~ . Many Illustrations can be found in the Mukerjee-Mysels (toe. clt.) compilation. (vii) The Influence of the temperature is small except for non-ionics over large temperature ranges. There are indications that the c.m.c. passes through a minimum; when this minimum is close to the ambient temperature, as for many ionics, the effect is expected to be small. A minimum is in line with hydrophobic bonding as the driving mechanism. (viii) The combined effect of the nature of the head group and the affinity of the counterion is clearly visible; from table 4.1 it follows that the c.m.c. of C12TMA+Br" exceeds that of C,2SO^Na+ by a factor of about 1.5. The latter micelles form more easily. In conclusion, the tabulated data already give a first exploration of the properties of micelles. Obviously, a satisfactory theoretical treatment should account quantitatively for these observations. 4.Id. Surfactant packing parameter We mentioned above that the hydrocarbon-water system is in the strong segregation limit. This means that in the micelle the hydrocarbon tails are necessarily densely packed. Ninham and Israelachvili11 were the first to realize that for systems in which packing effects dominate one can predict to first order the physical chemical properties from a simple geometric formula featuring the (surfactant) packing parameter P also sloppily called the "surfactant parameter". We will introduce this parameter by assuming an idealized picture of a micelle. Most of the modelling afterwards will provide a justification of this picture. The first basic ingredient is that the core is a densely packed set of surfactant tails. The characteristic size of the core is given by the length
11 J.N. Israelachvili, D.J. Mitchell, and B.W. Ninham, J. Chem. Soc Faraday Trans II, 72 (1976) 1525.
ASSOCIATION COLLOIDS
4.15
Figure 4.3. Pictorial representation of various micellar topologies: the spherical micelle (left top), cylindrical micelle (left bottom), lamellar topology (middle), the cubic phase (right top) and inverted micelles (right bottom), redrawn from ref. .
of the surfactant tall I. The second Ingredient is that the volume v occupied by the tall(s) of the surfactant in the core is identical to the volume that a corresponding alkane occupies in the neat alkane phase. To first order, this volume Is proportional to the number of CH2 and CH3 groups in the surfactant tail. The third ingredient Is that the head groups cover the core such that the area per molecule am is similar to that found by the surface tension measurements at the c.m.c. (see: tables III.4.5 and 6). The combination of these quantities defines the dimensionless surfactant packing parameter [4.1.4] When P ~ 1/3 , the preferred micelle size is spherical. Cylindrical micelles are expected when P = 1/2 . Flat bilayers will dominate for P ~ 1 and reversed micelles may be found when P > 1. Here the values of 1/3, 1/2 and 1 are found from the combination V/(RA) where V is the volume, A the area and R the characteristic length in a homogeneously curved object. The surfactant packing parameter features two reasonably well-defined quantities, viz. the molecular constant / and a value v that is expected to be constant. The third parameter a m is not a constant because it depends in general on the physical chemical parameters in the system, such as the ionic strength. Moreover, this quantity is expected to vary weakly with curvature. The weakness of this surfactant packing parameter concept is thus that formally P is not a constant. 11 D.F. Evans, H. Wennerstrom, The Colloidal Domain where Physics, Chemistry, Biology and Technology Meet (1994).
4.16
ASSOCIATION COLLOIDS
We note, however, that the surfactant packing parameter incorporates various assumptions about the properties of surfactant micelles. Most of these will be subjected to molecular modelling: (i) Fluctuations of the positions of the head groups are small. To avoid a density drop in the centre of the micelle, at least one of the dimensions of the micelle should be proportional to (not exceeding) the length of the surfactant (. (ii) The density of the core is homogeneous and similar to that for the corresponding alkane phase. This also implies that little water will disperse in the core, (iil) The area per surfactant molecule is a well-defined quantity that is not a strong function of the curvature of the core. In other words, the stopping mechanism responsible for the finite aggregation number in micelles is similar to the stopping mechanism for accumulation of surfactants at the air-water interface. The fact that in the micelle there is no air-water interface introduces only second order effects. Most of these aspects will be considered in more detail below, and we will see that the fluctuations of the head groups are usually not small, that the area per molecule varies systematically with the geometry of the micelle and that the radius of the spherical micelle is larger than the cross-section radius of the cylindrical micelle, which in turn is larger than half the bilayer thickness. In fact, this implies that the apparent success of the packing parameter is not trivial. However, this concept remains rather powerful as a tool to estimate the first-order effects. Using this simple concept, one predicts that the best way to vary the surfactant parameter, and hence the structure of micelles formed by the surfactant, is mostly not to increase the tail length at given head group properties (as v <*• I, and a m does not depend much on the tail length, we expect that P varies only marginally), but rather, one should modify the area per molecule, i.e. by changing the degree of polymerization of the EO moieties in non-ionic surfactants or the ionic strength and nature of the counterion (c.f. table 4.1) in the case of ionics. Another very effective option is to change the architecture of the hydrophobic group. Splitting up a single chain into two will keep v constant but reduces I by a factor of two, which increases P by a factor of two. Indeed, this explains why single chain surfactants tend to form spherical micelles whereas double chain surfactants are more likely to form lamellar bilayers. Chain branching is also very effective in changing the packing parameter. 4.1e Phase diagrams We have stated that micelle formation is roughly related to the first-order phase separation transition between two immiscible liquids (oil and water). However, the finite sizes of the micelles give the system very distinct and unique characteristics. In sec. 4.2b we will prove that upon increasing the overall surfactant concentration above the c.m.c. the chemical potential of the surfactant does not remain fixed but continues to be a very weakly increasing function of the surfactant concentration. As a result, it is expected that upon increasing the overall surfactant concentration there are variations in the micellar size and shape. One can probe these changes and construct a phase
ASSOCIATION COLLOIDS
4.17
diagram. A plethora of such phase diagrams can be found in the literature ". By way of illustration, in fig. 4.4 we present a schematic phase diagram representing the non-ionic surfactant C 12 E 5 . On the abscissae axis we have the temperature, which is the major control parameter for this surfactant, whereas on the ordinate axis the surfactant concentration is plotted in roughly logarithmic co-ordinates. It is important to note that the c.m.c. is at rather low concentrations and various noteworthy events in the phase diagram only occur when the surfactant concentration is very high. There are many generic effects in surfactant phase diagrams. Space does not allow us to go into all the details. We just briefly comment on some of the features. Indeed, in fig. 4.4 there is a wide two-phase region at elevated temperature (the large gray region (s)) at a fixed concentration when spherical micelles exist (label Lj) upon a sudden increase in temperature of the two-phase region leads to the formation of a concentrated surfactant phase out of the small micelles (clouding; for ionic surfactants one would find the Krafft point(s)). The critical point of this phase separation region is indicated by a dot. In the phase diagram, the c.m.t. line is the first feature when coming from dilute solutions. We also have dashed the concentration above which the micelles
Figure 4.4. Generic phase diagram inspired from data of the non-ionic C 12 E 5 surfactant. The c.m.t. line is indicated by the dashed line at low surfactant concentration. The L phases are isotropic. The liquid crystalline Hj (hexagonal) and Vj (cubic) phases are found at high concentration and low temperatures. At high temperature, a large two-phase region is found (shaded parts in the diagram). The L is the micellar phase. The dashed line with labels marks the transition from spherical to rodlike micelles (second c.m.c). The La phase has a lamellar topology. The L3 (sponge) phase and the solid phase S are also indicated. Pictorial representations of various aggregate types are given in fig. 4.3. (Redrawn from R. Strey, R. Schomacker, D. Roux, F. Nallct, and U. Olsson, J. Chem. Soc. Faraday Trans. 86 (1990) 2253.)
R.G. Laughlin, The Aqueous Phase Behavior of Surfactants, Academic Press N.Y. (1994).
4.18
ASSOCIATION COLLOIDS
change from spherical (L^) to cylindrical ( L\). In many cases, this transition is rather abrupt and in that case one refers to this point as the second c.m.c. We will discuss this transition in sec. 4.6. The La phase consists of lamellae (fig. 4.3). For this particular example, these lamellae are only found at high surfactant concentrations. At high temperature the La phase can swell enormously, which is believed to be caused by undulation forces. Some aspects related to the equilibrium separation of surfactant bilayers are discussed in sec. 4.8. The L2 phase consists of inverse micelles. In this case, the hydrated EO heads form the cores of the objects. There can be coexistence between various phases. For example, the La phase can coexist with (worm-like) micelles. The gaps in the phase diagram are filled in gray. The L3 phase, also known as the sponge phase, has a local lamellar topology, but the lamellae are interconnected by bridges or so-called handles. The membranes divide the space into an inner and an outer half-space. In the region of high surfactant concentration, the transition to liquid crystalline phases is found. At sufficiently low temperature there may be a hexagonal phase (Hx) where rod-like micelles are ordered into a hexagonal lattice. There also may exist a cubic phase V composed of a complex but very regular network with zero mean curvature. Such a phase is also called the plumber's nightmare. Finally, at very high surfactant concentration we may find a solid phase of surfactant, which is labeled S. Phase diagrams can be more complicated. For surfactants with P ~ 1/3 , there may not be a strong tendency to form elongated or lamellar objects and as a result one may find a liquid crystalline phase of ordered spherical micelles. Such an ordered phase is often called a cubic phase (not indicated in the phase diagram). At present there is at least qualitative insight into most of the phenomena determining the surfactant phase diagram. We know that in phase diagrams there are many generic features, i.e. features occurring for almost any surfactant system. Although very useful, the presentation of phase diagrams for particular surfactant systems will not be the endpoint of experimental investigations because they are merely a catalogue, leaving many fundamental questions open. The challenge for the near future is to explain the molecular specific issues. Much more work is needed to explain why the phase diagrams of Cj E 5 and C 14 E 4 differ so much, although the difference between the two surfactants is just one O atom! Indeed, this simple observation calls for molecular modelling and this will be a main topic in the remainder of this chapter. 4.2 Classical thermodynamics Classical thermodynamics specifies the number of independent variables for any macroscopically homogeneous one-phase system. For each of these independent variables, there is one contribution to the change in the internal energy dU c
dL/ = TdS-pdV + ^// i dJV i i=l
14.2.1]
ASSOCIATION COLLOIDS
4.19
where the symbols have their usual meaning. Index i refers to all (c ) molecular (i.e. uncharged) components, including the solvent. Of the c + 2 variables { p , T and the c components), only c+1 are independent because of the Gibbs-Duhem relation. The most simple surfactant system has two components, a surfactant and water. For an open system of this kind, we can write for the change of the Gibbs energy dG = -SdT + Vdp + MSUT{dNSUT[ + // water dJV water
[4.2.2]
At constant p , T there is just one degree of freedom, the surfactant concentration. Classical thermodynamics provides the rules that should be obeyed when a macroscopic system goes through a phase transition and can be used to help understand those cases in which there is a macroscopic interface between two coexisting surfactant phases, e.g. when the system is above the cloud point or when a surfactant-rich lamellar phase coexists with a dilute surfactant solution which contains micelles. At issue is the question whether thermodynamics remains applicable on the mesoscopic level. In particular, the problem arises on how to use thermodynamic arguments for micelle formation. From a macroscopic point of view, the system remains completely homogeneous even when micelles are present. Thermodynamically speaking, nothing prevents or promotes molecules to choose to interact with each other and nothing prevents them from making molecular clusters or micelles still retaining [4.2.1]. Such interactions are determined by the values of the intensive variables. The extent of (non-ideal) interactions will then be a function of the chemical potentials, the pressure and the temperature. In an open system, one fixes the chemical potentials of the molecular components by letting the system equilibrate with a large reservoir. Below the c.m.c. there is no problem doing this, but as soon as the chemical potentials are increased to values consistent with the presence of micelles, the number of kinetic units of surfactants in the system suddenly becomes a very strong function of the chemical potential of the surfactants. Therefore, the closed system is preferred to examine micelle formation in the same manner as for any chemical equilibrium reaction. Macroscopic thermodynamics can actually provide further guidance to understand micellar systems. We can only make significant progress, however, upon pre-assuming that micelles exist. Formally, we then go somewhat beyond the realm of classical macroscopic thermodynamics. There exists a powerful framework, which is an extension of the classical thermodynamics, to deal with microscopically inhomogeneous systems. This approach is known as the thermodynamics of small systems. It was developed by Hill11 in the 1960s and elaborated by Hall and Pethica21 for micelle formation.
11 T.L. Hill, Thermodynamics of Small Systems. Part 1 and Part 2. Dover Publications (1991) and (1992). 21 D.G. Hall, B.A. Pethica Non-ionic Surfactants, chapter 16. Marcel Dckker (1976).
4.20
ASSOCIATION COLLOIDS
4.2a Thermodynamics of small systems In [4.2.1 ] the full set of independent variables of a homogeneous bulk composed of c types of molecules is given. From experiments, however, we know that surfactant systems organize themselves on a mesoscopic length scale. It is quite attractive to try to introduce this information into the thermodynamic analysis. However, in general one should be cautious when ad hoc extra degrees of freedom are added. Nevertheless, this is what we are going to do but we shall give an a postiori justification. Following Hill, we postulate that there is some hidden variable, which is most generally called the number of subdivisions. Without much ado, we directly identify this number as the number of micelles In the system Af = Xj-VJ, where Mi is the number of micelles of distinguishable state i. As all micelles are in mutual equilibrium, it suffices to use just one A/" • Conjugated to this hidden extensive variable, we introduce the intensive variable e. In effect, e is the energy needed to generate one extra micelle in the distribution { A/'i} at fixed entropy, volume and number of molecules. Now [4.2.1] is modified to c
dU = TdS -pAV + ^
fiidNi + £&A/
[4.2.3]
i=l
In a closed (N,V,T ) system, the Helmholtz energy is the characteristic function. For aqueous solutions we may as well choose to fix the pressure and switch to the Gibbs energy. In incompressible systems, these two quantities differ just by a constant. For a change of the Gibbs energy, we can write c
dG = -SdT + Vdp + ]T /ZjdJVj + aLV
[4.2.4]
i=l
From general thermodynamical arguments, we know that at equilibrium
an
isothermal-isobarlc system will minimize its Gibbs energy. In the small system analysis, the hidden degree of freedom is used to optimize the Gibbs energy, and thus (!^|
=£ =0
[4.2.5]
The corresponding stability condition reads
\dAr
h.p,{Ni}
W>T,p.{Nl}
The implication of [4.2.5] is that, at equilibrium, the Gibbs energy continues to be simply given by c
G
= X'"iJVi i=l
[4 2 71
- '
ASSOCIATION COLLOIDS
4.21
There is no excess Gibbs energy associated with the formation of micelles. This result is completely in line with the macroscopic thermodynamics for homogeneous systems. The extra term introduced in [4.2.3] thus does not, and should not, influence the Gibbs energy in the system. We note that the small system approach does not presuppose the presence of micelles of a particular size or shape. It anticipates, however, and this should follow from any statistical mechanical model, that there is a distribution in size (and shape). For any of these aggregates it must be true that there is no excess Gibbs energy associated with its presence, i.e. e = 0 . It is known that the micellar size (distribution) depends on a variety of factors, such as the nature of the surfactant, the temperature and the presence of additives that may be absorbed in, or repelled by, the micelle. This means that generally e is a function of p , T and the chemical potentials. However, as at equilibrium e is always equal to zero, from the e d V term no information can be obtained regarding the dependence of micellar properties on ambient conditions. To that end, models have to be developed. We note that in general we should expect that the average micelle size will be inversely related to the number of micelles in the system. This means that the Gibbs energy should also have a minimum with respect to the average micelle size. It is possible to split the Gibbs energy into a bulk part G b , referring to the solution in which the micelles are embedded and a part that can be attributed to the micelles Ga, i.e. G = Ga + G b . The total excess number of surfactants associated with micelles may be found if the bulk concentration of surfactant (i.e. outside the micelles) can be determined because the excess is counted with respect to a reference system, which fills the entire volume by the (dilute) bulk solution containing only monomers. The same applies to the other components and material balance gives N{ = N? + JVb for each component.
Because the volume work
does not enter
in the excess
thermodynamic potentials, the difference between the Gibbs and Helmholtz cases vanishes. Starting once again from [4.2.3], we may write c
Fa = Ga = ^ /ijJVf + EAf
[4.2.8]
i=l
where we have retained deliberately the EA/~ contribution to unravel more about its properties. Here we notice that the molecules in the bulk do not contribute to e. The value of e is the Gibbs energy that is in excess due to the presence of micelles. Normalizing the excess Gibbs energy by the number of micelles gives ga =
=Y//inf +£
[4.2.9]
where n? = N?IA/~ is the average excess number of molecules associated with one micelle. We may also refer to this as the most probable size of the micelle. Now E can be identified as the (excess) grand potential
4.22
ASSOCIATION COLLOIDS c
£ = ga - ^T nffii
[4.2.10]
i=l
At constant pressure and temperature we can write c
c
c
de = dg* - £ ^dnf - £ nf d/i, = - £ nf d//t i=l
i=l
[4.2.11 ]
i=l
which is the Gibbs-Duhem equation for micellization. It resembles Gibbs' law for adsorption. When we have a two-component system where surfactant {surf) is dissolved in water, we can assume that the chemical potential of the solvent is nearly constant so that the Gibbs law becomes (9^-)
=-
[4.2.12]
In words, this equation says that the excess grand potential is a decreasing function of the chemical potential of the surfactant because necessarily the aggregation number n f u r f > 0 . Recalling the stability condition [4.2.6] we use [4.2.12] and find that 3// surf / 3yl/< 0. Again in words, this says that in a closed system the chemical potential of the surfactant must go down with an increasing number of micelles. Let us next switch our attention from the closed isobaric system at constant T , where the Gibbs energy is minimized, to the open system. The grand potential is the characteristic function of the system in which all chemical potentials are fixed. In such a system both the micelle concentration, i.e. the number of micelles Af per unit volume, and the aggregation numbers (including the distribution) are variables, which the system may use to minimize its grand potential. The minimum of this grand potential should have the value zero in order to be consistent with the analysis in the closed system. In general, the thermodynamics of small systems becomes especially useful when it is applied to the interpretation of experiments. It also turns out to be of utmost importance for analysis of results of simulations or computations on micelles. 4.2b The mass action model A customary way to describe micelle formation is to consider the equilibrium between freely dispersed surfactants, usually called monomers and referred to by Xj, with micelles composed of g surfactants, X . In this approach, one assigns a chemical potential to the g -micelle, and the total number of surfactants in the system is split up into a contribution of the monomers Nsm and a contribution present in the micelles g x JVs( ,, i.e. N surf = Ns(1) + g x JVs( j . This equation is easily generalized to a range of micelle sizes. Ignoring for the moment the size distribution, the equilibrium may be expressed by gXx^±Xg
[4.2.13]
ASSOCIATION COLLOIDS
The equilibrium constant K
4.23
relates the micelle concentration to the monomer
concentration
K = J ^ i =^ L 9
^
9
[4.2.14]
This equation is often used to illustrate the first order-like transition the system passes through near the c.m.c. For large values of g , say g ~ 100 , we see that near and above the c.m.c. an increase of the monomer concentration by 1% already leads to a threefold increase in the micelle concentration. Indeed, experiments show that this is a realistic aspect of micelle formation in aqueous solution. The monomer concentration above the c.m.c. increases only very slightly when the micelle concentration is increased, e.g. by increasing the overall concentration. The model based on just one "reaction" is, however, unrealistic; intuitively one must expect the micelle to increase in size (increase the aggregation number) when the chemical potential of the monomers is increased. We will show below that this is true, as the Gibbs energy gain of size fluctuations is small compared with kT . The way to correct for these circumstances is to consider a large set of reaction equilibria and, connected to this, a set of equilibrium constants K ,. Now micelle formation becomes more complex because each of these constants has to be known before one can proceed. We shall not elaborate this route further but mention that it was shown already by Hill that the multiple equilibrium approach to micelle formation is equivalent to small systems analysis. For the time being, we accept the limitation of the single equilibrium model and proceed by pointing to the fact that equilibrium implies that S^s(l) =Mm
[4.2.15]
where in this notation jus,^ is the chemical potential of the monomer and JUSIQ^ is the chemical potential of the g -micelle (corresponding to adding one more micelle to the solution). In words, this equation points to the fact that the chemical potential of the micelle is found from the sum of the chemical potentials of the g monomers from which it is composed. Again, in line with the small system approach where £ = 0 , there is no extra Gibbs energy on top of that stored in the micelle. Ignoring activity corrections, we can write for the chemical potential of the monomer <"s(i) = >"s(i) + ; c T m ^ s i i ) a n d f° r t ^ le micelle MS{g) = fsig) + ^ m ^s(g) • w n e r e t n e volume fraction
[4.2.16]
In passing, we note that K = expf-AG^/fcT), where G®n is the standard Gibbs energy of micelle formation, which may be related to the c.m.c. (e.g. using [4.2.14], but only
4.24
ASSOCIATION COLLOIDS
after an appropriate definition of the c.m.c], and in general it is composed of an enthalpic AHj^ and an entropic ASj^ contribution, which are measurable by calorimetric methods. In a more elaborate model with multiple equilibria, similar equations apply for each equilibrium constant. We note that thermodynamics neither gives a clue about the value of g nor the values of K . To make progress, we need a molecular model to compute //°( j for a series of g values. 4.2c Implication for molecular modelling It is obvious that molecular models are needed for further progress. Models should be consistent with thermodynamics. This means, for example, that in equilibrium the (excess) grand potential of each aggregate should be zero [4.2.5], or equivalently that the chemical potential of a micelle is found to be the sum of the chemical potentials of its constituents [4.2.15]. Both the mass action model and the thermodynamics of small systems approach already require some extra-thermodynamic assumptions. In the small system approach, the existence of micelles of limited size is assumed and in the mass action model it is taken that distinct micelles exist to which a micellar chemical potential can be assigned. More detailed models will need these assumptions as well. The intensive variable e, associated with the number of micelles, is a central quantity, which in equilibrium is zero and whose derivative with respect to the number of micelles must be positive for stability. This quantity was identified as the (excess) grand potential of the micelles. It is possible, by using a model, to unravel this grand potential into its various components. Here we will illustrate this by splitting it up into a term that contains the mixing entropy and the remainder which we will call em . Referring once more to the mass action model of [4.2.13] and the equilibrium condition [4.2.15], we notice that the micelles have translational (dispersion) entropy. Ignoring the interactions between micelles and the non-ideal part of the mixing, one typically uses (,«s(1) - jU®m)/kT = lnfZ>s(1) for the monomers and
for the micelles; these logarithmic terms may be identified as (minus) the dimensionless (partial) mixing entropy term associated to the monomers and the micelles, respectively. We see that the driving force for micellization must at least overcome a significant loss of translational entropy; per micelle g monomers cede their translational entropy, whereas only the corresponding entropy of the micelle as a whole is gained. Assuming that at the c.m.c. the monomer concentration and the micelle concentration are of the same order, we see that the loss of translational entropy amounts to about -k{g -\)ln
ASSOCIATION COLLOIDS
4.25
of surfactant molecules in the micelle. Next we may set up a regular solution model to elaborate the idea of micelles having mixing entropy. Let us consider a (spherical) micelle with micellar volume Vm . We now construct a three-dimensional lattice with L sites of volume Vm , i.e. V = LVm . On these sites we distribute Af micelles that have athermal interactions with each other. From this we obtain the total (ideal) mixing entropy in the system -TS mjx //cT =
(// - ju°) / kT = In
{fiM-fJ%/l)/kT = ln{l-
[4.2.18]
We will refer a bit loosely to em as the translationally-restricted grand potential of a micelle. Alternatively, in Hill's terminology, em is the standard state subdivision potential. In fact, it is useful to point out that em is the characteristic function coupled to the partition function cpm . Here we use the notation cpm for the total volume fraction of micelles and do not refer to the size of the micelles.
4.26
ASSOCIATION COLLOIDS
All truly molecular models that aim at computing the partition function have in common that they focus on one (average) micelle at rest in a small system. We will now argue that for these models em is of key importance to evaluate the total micelle concentration. The first step is to realize that the total volume in the overall system may be split up into Af small systems with volume v = VI AT . As in the molecular modeling calculations carried out on a small system, we will optimize the Gibbs energy (for a small system) not with respect to the number of micelles (because we consider just one average micelle), but with respect to the volume of the small system v and note that dv/dAf < 0. In this small system, the molecular composition is identical to that in the overall system, and hence ni = N{/ Af for all molecules i . We take it that there is one micelle at rest in the centre of this system. The intrinsic Gibbs energy for this small system gm = £ 4 nilui + £m a n ( i hence the grand potential is e m . The subindex m reminds us of the fact that the system is constrained because it features a micelle at rest. The goal is to estimate what the overall micelle concentration is in the corresponding (unconstrained) macroscopic system. To do this we need first to convince ourselves that the micelle is stable. It turns out that it is useful to consider the excess quantities and the corresponding bulk values for each molecule I, i.e. n( = n? + H?3 , where the excess of the surfactant may be quantified in terms of the aggregation number g . Following the same arguments as above, we also find a Gibbs relation for the constrained system
(^M
=-"sVf=-9
[4-2.191
As g > 0, we infer that em is a decreasing function of ,usurf. At the same time, g = rigurf must be an increasing function of // surf , and thus e must be a decreasing function of g = n surf , to have stable micelles. Hence, the stability condition in the constrained system is given by — <0 9
[4.2.20]
Now we fix the chemical potentials in the small system and vary its volume v until the small system corresponds to the equilibrium one. Let us estimate the volume of the micelle um by the molar volume of a surfactant u surf times the aggregation number, i.e. vm = n° nrf u s u r f . Using this, the mixing entropy that a micelle will gain when the positional constraint is released depends on the volume of the small system, i.e. -S m i x Ik = ln(Pm = lnu m / v . Using [4.2.18], we have the equilibrium condition
^ m = e x p (~Sr)
|4 2 211
'-
Knowing the volume fraction of micelles, it is possible to compute v and with that the
ASSOCIATION COLLOIDS
4.27
system is completely defined. For a micelle with given em > 0 and dem/dn°urt
<0 ,
there exists an overall volume fraction of surfactant given by Psurf=«£urf
+
I42'22!
Pm
or, in other words, there is a one-to-one relation between the overall surfactant concentration, the micelle size and the micelle concentration. One of the main advantages of molecular modelling is that, besides the most probable micelle size, information can often also be obtained on the fluctuation in micelle size, which is directly linked to the micelle size distribution. Quite generally, a measure of the size distribution is found from (see sec. 1.3.7) kTJ(9L d
= a2{g)2_{g2}
[4 2.23,
<"surf
where we may identify (g) by g , that is the aggregation number of the most frequent micelle. Using Gibbs' law or [4.2.12], we may write (de / d(g))(d{g) / d/i) = -(g) or equivalently
4.2d Fluctuations in micelle sizes In general we should expect a distribution of micelle sizes. It is necessary to understand the role of the entropy associated with size distribution. We may start with the mass action model. Let there be a set of micelles with aggregation numbers g = 9mjn>"i>Sflnax • For each of them, there exists the equilibrium of [4.2.13], gXj ^ Xg . At equilibrium, AGg = fig -g/^ = 0 . We may identify AGg as the excess grand potential for the g -micelle. To distinguish this quantity from the most likely grand potential as obtained in the SCF theory and considered so far, we will refer to it as £2(g). Similarly, as done with e, we may split Q up as £2(g) = Qra(g) + kTlnip where
m(9» = -f3m«9» +
fc 2 3(9-(9»
[4.2.25]
where i2m[{g)) is the quantity that is computed in the SCF calculations and to which we refer as em [g). The measure of the width of the distribution k
is related to the
4.28
ASSOCIATION COLLOIDS
variance <jl = kT/(2k ). The statistical weight of a micelle with grand potential &mig) follows from Boltzmann's law. The distribution is Gaussian (at least around the minimum). For sufficiently large value of k&, the Gauss distribution is so narrow that integration from the smallest g = gmiB to the largest g = £fmax micelle size is the same as integrating from
g =-°°
to g = +<*>, which leads to the micelle probability
distribution
P
[K~
'H^T-
where
eXP
(-kJg-(g))2)
[
|4 2 261
kT J
'-
Ik / nkT normalizes the probability distribution function to unity. On top of
the translational entropy of the micelles, we now also need a mixing entropy. It accounts for the number of ways one can distribute distinguishable micelles into the systems. This mixing entropy may also be calledjluctuatton entropy and is given by S
fluct/'c = - X P ( 9 ) l n P ( f i f )
[4.2.27]
9
The average grand potential per micelle now contains the entropies of the translational and mixing degrees of freedom, as well as the standard state contributions: 12 = kT^
P(g) In (P(g)
9
9
9
^
[4.2.28]
'
Again, in equilibrium Q = 0 . The sum over all micelles gives the overall micelle concentration q>m =^q(p
and the volume fraction of g -micelles is found as
P(g)
I kT). Using [4.2.25], we arrive at
\j kg
y
{
kT j -J kg
V
,4.2.29]
{ kT )
Comparison of [4.2.29] with the relation [4.2.21 ] shows that the entropy due to the size distribution leads to a fluctuation correction given by Jk / nkT = -JUncr . Below we will analyze the variance of the size distribution that results from a MC simulation, as well as predicted from the SCF theory. At micelle concentrations not too close to the c.m.c, the relative fluctuations do not depend much on the micelle concentration and we typically find a I g ~ 0.2 . This means that the fluctuation contribution V2/r
ASSOCIATION COLLOIDS
4.29
appearance of micelles, and also near the sphere-to-rod transitions as well as for worm-like micelles, the distribution is not narrow and [4.2.29] is incorrect. As a result, we should keep in mind that the conversion from em to the total surfactant concentration is typically approximative, especially when for the sake of simplicity the entropy in the size fluctuations is ignored. It should therefore be understood that the relation between micelle size and micelle concentration is only expected to be accurate to first order, i.e. one should mainly apply the predictions to understand trends. Below we will present various methods that are regularly used to model micelles and micellar solutions. In the SCF theory, it is only possible to compute so-called most probable micelles (it focuses one m ) and we will use [4.2.24] to obtain information on the micelle size distribution. In analytical models that make use of surface-thermodynamics, it often proves more convenient to fix the chemical potentials and compute the grand potential &mig) (or equivalent, the g -micelle chemical potential) as a function of the micelle size. In this case, one has access to the micelle size distribution more directly, and [4.2.29] should be used to find the overall micelle concentration (provided the distribution is sufficiently
Gaussian and narrow). In (atomistic)
simulations, the size distribution of micelles is directly available from averaging over the simulation trajectory (provided that at any time in the simulation box there is a sufficient number of micelles present). In this case, it is not necessary to apply any of the thermodynamic rules. One of the most characteristic aspects of micelle formation is the pronounced gap in the size distribution. There are freely dispersed monomers in solution. These monomers have some affinity for each other and we should thus expect a small fraction of dimers, and even smaller quantities of trimers, etc. in the system. Then, the probability of finding multimers of 4, 5, ••• monomers is about zero, until at significantly higher aggregation number (order 100), the probability of finding aggregates peaks again at the most probable micelle size. Finally, this probability drops to zero again for large aggregation numbers. Any realistic molecular model should account for this gap. In this context, it is relevant to recall that for small micelles g < gmin , when dem/dg > 0 , the micelles are unstable and the Gibbs energy is at a local maximum. This is the signature for the existence of a gap in the micelle size distribution in models that focus on the most likely micelles (i.e. the SCF model). This does not mean that micelles with aggregation number g < gmin
do not exist. They may be present in the
tail of the size distribution for systems where the most likely micelle is sufficient in size, i.e. g > gmin • Such micelles may well function as an "activated state" through which growing micelles have to pass on their way towards their most probable sizes. 4.3 Molecular modelling The machinery of statistical thermodynamics gives, for any molecular model, the recipe to come to measurable observables. The partition function is the central quantity
4.30
ASSOCIATION COLLOIDS
that should be computed. Knowing the partition function gives, via the characteristic function, access to the thermodynamic, mechanical, and structural characteristics of the system. Exact partition functions for systems in which surfactants occur with molecular detail are not available. There are two options to proceed, viz. via the simulation route or one can choose to use mean-field approximations and solve meanfield partition functions. Within the latter approach, one can try to formulate the problem in such a way that (approximate) analytical results can be obtained. However, for the more detailed models computer assistance is unavoidable. 4.3a Molecular simulations For surfactant systems, one can use molecularly realistic simulations such as molecular dynamics (MD) or Monte Carlo (MC). The simulation route is not focusing on the partition function. There are simply too many possible states to visit all of them. As a result, thermodynamic quantities such as the entropy or Helmholtz energy are very hard (but not impossible) to evaluate. Instead, the aim is to generate, with the use of a computer, a large set of relevant realizations of the system. In MD the trajectory along various realizations is along a dynamic path generated by solving Newton's second law F = ma for each degree of freedom. Both structural, some mechanical parameters as well as dynamic parameters are obtained with an accuracy that is determined by the quality of the input parameters and the computer time available. It is possible to take the same molecular system as studied in an MD simulation and subject it to an MC simulation. Instead of a dynamic path, there are random changes made in the system, but the Metropolis algorithm makes sure that only relevant states are probed. Along the MC trajectory, one can then collect information on structural and mechanical properties. There is no systematic information on the dynamics. For a particular model (i.e. the specified, usually simplified, interactions and the large set of estimated force field parameters that parameterize the interactions on the atomic level) simulations give (when sufficient CPU time is available) access to close-toexact results, that is at each point along the simulation trajectory all correlations between all molecules are exactly accounted for. The system size is usually limited to linear dimensions on the order of a few nm. In addition, in MD simulations there is accurate dynamical information, but it is hard to proceed to times longer than a few ns. Both limitations make the simulation route less applicable for (dilute) micellar solutions. There are, however, several problems for which the simulations give the most accurate results available to date. One such example that will be discussed below is the lipid bilayer membrane. In the simulation of such systems, one typically makes sure that one has a suitable initial guess so that one can start the simulation already close to some equilibrated structure. In modern packages, all parameters necessary to account for all the interactions in the system, the so-called force field parameters, have
ASSOCIATION COLLOIDS
4.31
reasonable values and after a relatively short equilibration run one can already collect relevant information on the equilibrium state and fluctuations around the equilibrium. The simulation route is not limited to the atomic level. One can invoke models in which the molecules are represented by groups of atoms (united atoms) or with strings of segments (blobs). Then much larger system sizes are accessible and, for MD, longer time scales can also be probed 1 '. Of course, the standard force field cannot be used any more and one has to come up with an alternative set of parameters that specifies how the united atoms or blobs interact with each other. Such a coarse-graining step is not unambiguous. Quantifying the zooming out from the molecular details to the mesoscopic length is seen as one of the important challenges in the near future. We will not further discuss the technicalities here.
4.3b Self-consistent field (SCF) theory The second approach is the mean-field partition function route. The mean-field method pragmatically deals with the counting of the binary (and ternary, etc.) interactions in the system, which depend on the detailed local environments (i.e. distances between the molecules in each individual realization of the system), by substituting the actual surroundings by an average one. In this way, all pair interactions, three-body interactions, etc., are replaced by those for a particle in an external field. If this external field is made adjustable such that it corresponds to the average distributions of the molecules in the system, one speaks about a self-consistent field (SCF) theory. Such SCF theory may be used to model micelles in equilibrium with freely dispersed monomers. We will see below that the SCF method can be implemented on a level, ranging from that of a statistical segment (rather coarse) to that of the atom. Typically, however, the molecules are parameterized with less detail than in the all-atom simulations. For self-assembly, the structure of the surfactant molecules should be accurately accounted for. This means that architectural elements of the molecule, (one or two tails, linear or branched chains, ionic, zwitterionic nature of the head groups, etc.) must be realistically implemented. Important for a realistic modelling is that one should keep as many fluctuations in the model as possible, i.e. to impose as few constraints as possible. One weakness of SCF modelling is that it is needed to a priori specify the geometry of the system. This means that, when a spherical co-ordinate system is used, one can only obtain information on spherical micelles or spherical vesicles. In a cylindrical geometry one can model rod-like micelles or tubular vesicles. A flat geometry is used to generate information on lamellar bilayers, etc. As for each of these geometries, the partition function can be found as a function of the system variables, such as the chemical potentials and the temperature, one can a posteriori point to the thermodynamically most favourable geometry for the system. We will show this below in sec.
1
M. Krancnburg, C. Laforgc, and B. Smit, J. Phys. Chem. Chem. Phys., 19 (2004) 4531.
4.32
ASSOCIATION COLLOIDS
4.6 when we consider the sphere-to-rod transition in micellar solutions. The caveat is that more complex geometries, e.g. donut micelles or perforated disks, which may be important in practice, may be overlooked. Computer-assisted modelling very often leads to discrete rather than continuous functions. This applies also to the SCF model, which makes use of a discrete set of co-ordinates (lattice) with layers (sets of lattice sites along which the mean field approximation is applied) that depend on the chosen geometry. In other words, when the geometry is specified, the planes of lattice sites are positioned in space accordingly. For example, for a spherical micelle one can expect that it is possible to use a spherical co-ordinate system comprising spherical shells. Within each shell of sites, the density fluctuations are so small that it is possible to ignore them. Exactly how the lattice geometry is implemented will be discussed in some more detail in 4.4 and in appendix 1. As long as one is interested in the properties over length scales that exceed the size of a lattice site, there are little or no adverse effects of, nor advantages in, using a lattice. However, when one is interested in details that are small compared with the lattice site one should be aware of methodical artifacts. Modelling of micelles in which there are strong gradients in density is on the edge of the capabilities of lattice models. The adverse effects of the mean-field approximation, however, are more difficult to judge. As a result, it is necessary to compare SCF predictions with simulation results for corresponding systems. We will show results of such comparison below. The main disadvantage of using computer simulations or solving a mean-field partition function with the aid of a computer is that the equations that are used to generate a trajectory (in simulations) or to find density profiles from the optimum partition function do not yet say much about the physical properties of the system that is modelled. In general, one needs to analyze the results and do systematic variations of parameters before one can compare the results with experiments, and hence come to deeper understanding of the system of interest. Computer modelling is therefore rarely the end-point of investigations. Ideally, one would like in addition to have a model that incorporates yet other idealizations of the system, which also captures the main physics of the system of interest and ignores the less important effects. Such an approach may give insight into the balances offerees, the importance of fluctuations, etc. Somewhat closer to the theorist's dream is the advancement of analytical models. For self-assembly one can put forward formulae that capture various aspects of the problem at hand. These equations are the result of detailed theoretical modelling of a partial problem, e.g. the grand potential of the electric double layer in and around the micelle, or result from phenomenological fits of experimental or simulation (computational) data. Typically the start of such an analysis is an idealized view of the structures at hand. For instance, a usual Ansatz is to treat aqueous micelles as apolar droplets, surrounded by an oil-water interface with an adsorbate in it, to which an interfacial tension can be assigned. This means that, similarly as in computer-aided models, significant information input is needed, such as micelle geometry, the role of various
ASSOCIATION COLLOIDS
4.33
fluctuations, etc. It is fair to say that such quasi-macroscopic (analytical) models cannot (yet) take all degrees of freedom of the surfactant molecules in and outside the micelles into consideration. One rather hopes that the most important ones are sufficiently accurately accounted for. In this context it is necessary to mention that in a successful analytical model of this kind one would like it to contain the experimentally known correct results. This is typically not yet the case. The state of the art of successful quasi-macroscopic models is that there is a balance of equations, which compensates for each other's shortcomings. This point illustrates that direct comparison with a limited amount of experiments is usually insufficient. One should also critically confront these models with computer-aided molecular modelling. In other words, successful quasimacroscopic models should be developed in close harmony with molecularly realistic computational methods on the one hand and experimental results on the other. This desirable process of integration undoubtedly deserves more attention. (i) MC and SCF of surfactant micelles There are relatively few MC studies on micelles. In this section we will discuss the results of Wijmans and Linse1' who published a comparison between MC and SCF results for a rather primitive model (which at least allows such a comparison). Unfortunately, in the original paper they overlooked the fact that for a successful comparison the two models need interaction parameters that differ systematically. In particular, the through-bond contacts are irrelevant in the MC method and therefore a segment in the chain on a cubic lattice can effectively only interact with four neighbours, whereas the x parameters used in the SCF model were rescaled to six possible contacts. They incorrectly found major differences between SCF calculations and MC simulations. We will now present an improved analysis in which the original MC results are compared with the corresponding SCF model with the correct interaction parameters. We arrive at the conclusion that the results of the two methods are in almost quantitative agreement. This particular MC model is lattice-based and features an incompressible uncharged system. There is a cubic lattice with dimensions LxLxL
(L = 44) sites with
periodic boundary conditions. Chain-like surfactants of the type AN B N
are intro-
duced, where JVA = 10 is the length of the A-block and JVB = 10 the length of the B fragment in a monomeric solvent S . One may interpret an A unit as a hydrocarbon segment. In the adopted MC procedure the chains are freely jointed and self-avoiding. This means that each lattice site is occupied just once and that two segments along the chain occupy neighbouring sites. In the Metropolis weighting, nearest-neighbour con tacts were accounted for via the contact energies u AB = u A S . Assuming six neighbours, these were translated into a Flory-Huggins parameter %AB = %AS = 2.7 and ^gg = 0 .
11
CM. Wijmans, P. Linsc, Langmuir 11 (1995) 3748.
4.34
ASSOCIATION COLLOIDS
The correct parameters are a factor 2/3 smaller (at least for the middle segments) because middle segments in the chain have not six, but just four, possible contacts. Below we specify the corresponding SCF model. The simulation length is in the order of 107 Monte Carlo Steps (MCS) (in one MCS, the system is changed as many times as there are degrees of freedom). Results were typically averaged over 104 different micelles. In total 200 surfactant molecules were included in the system. The remainder of the sites was filled by solvent molecules. In the incompressible system the reference of the interaction energies is chosen such that in SCF ^ ^ = ^ B B = / c c = 0 and MC U
AA = U B B = U CC = ° •
The corresponding SCF model uses a spherical coordinate system. The molecular structure is very similar; here we choose the end segment of the A block to differ from the middle ones, i.e. -AjAN _jBN • Again, a freely-jointed chain model is used, however the chains are not self-avoiding but modelled using a first-order Markov approximation. In a Markov chain, the segments can fold back on previously occupied sites already taken by the same chain. In the SCF method, such approximate handling of intramolecular excluded volume effects is in balance with the similar handling of the intermolecular excluded volume interactions. At the cost of significant computer time, it is possible to systematically improve on the intramolecular excluded-volume correlations. Here we choose not to do this. By doing so the intermolecular excludedvolume problems (one site can be occupied by segments of two different molecules) remain. Consistent with the simulations, we take for the apolar segment-solvent interactions XAB = X^s = 2 - 7 x ( 2 / 3 ' = 1-8 and %A,S = 2.25 . The interaction parameter for A' with B was set equal to 1.8 and ZA'A= ® • ^ o r further details about the SCF method, we refer to appendix 1. Results are presented in figs. 4.5 and 6.
Figure 4.5. a) The normalized size distribution as found in the MC simulation for a system composed of lattice chains AJQBJQ in a monomeric selective solvent (left ordinate), volume fraction of micelles with size g as computed from SCF calculations
ASSOCIATION COLLOIDS
4.35
Figure 4.6. The radial volume fraction profiles for tail and head group segments of AJQBJQ surfactants in the micelle as found by MC simulations (continuous lines) and corresponding SCF calculations (dashed lines).
Of course an MC simulation reveals many details. We select a few relevant aspects. One important point Is that in the simulations there are, at any point along the MC trajectory, just a few (order five) micelles in the box. It is important that this number exceeds unity. If only one micelle forms in the simulation box, one cannot know whether or not the size regulation allowed the micelle to obtain its equilibrium size. The reason for this worry is that (in contrast to SCF) the thermodynamic properties are not evaluated in an MC run and one relies on the Metropolis algorithm to do its task. It is also important that the micelles continuously change in size and shape. These shape fluctuations can be estimated from the analysis of the principal moments of inertia measurable for each micelle in the system. It was found that the ratio between the main and the minor principal moments of inertia was largest (approximately 2.2) for micelles with relatively high or very low aggregation number and was relatively small (order 1.5) for intermediate (most likely) micelles (not shown). The size distribution found in the MC simulations is given in fig. 4.5a. According to expectations, there is a gap in the size distribution. Micelles with size g ~ 10 have a low (but finite) probability. Monomers and micelles with sizes around g ~ 30 are most frequent. Closer inspection of the micelle distribution shows that the distribution is not exactly symmetric, but near the maximum it can rather accurately be represented by a Gaussian. The freely dispersed monomer appears to be the most abundant species. It was found that a fraction of 0.17 unimers (volume fraction 0.008) is left in the system, while that of dimers is 0.06 (volume fraction 0.006). It suffices to add up these two amounts to obtain a measure for the c.m.c. for which we find
4.36
ASSOCIATION COLLOIDS
the aggregation number of the most likely micelles (at least in the range 17 < g < 40 ). The c.m.c. as predicted by the SCF calculations is cp° = 0.0202 (see sec. 4.4 for information on how the c.m.c. is found exactly). This value is close to the estimate found in the MC simulations. For the same overall surfactant concentration as in the simulations, the volume fraction of surfactants that is available to form micelles
ASSOCIATION COLLOIDS
4.37
fluctuations. Nevertheless, the two values for the polydlspersity are of the same order of magnitude, which gives confidence to the accuracy of the modelling methods. We note that in experiments with short-chain surfactants the fluctuations in micelle size are expected to be smaller, i.e. 0.1 < a Ig < 0.2 . The relatively large fluctuations in this case study may be attributed to the primitive parameter settings that were chosen to allow comparison between simulation and theory rather than to mimic the exact experimental situation. The important conclusion is that the SCF theory accurately reproduces the micellization characteristics of this simple model. Considering the fact that the SCF procedure takes about 1 second CPU per micelle whereas the simulations are at least four to five orders of magnitude more expensive CPU-wise, this is an important result. Hence, it is practical to continue and elaborate by SCF methods more sophisticated models, e.g. with respect to the role of the solvent (water) and other structural details. (it) MD and SCF of lipid bilayers All-atom molecular dynamics simulations are more expensive (CPU-wise) than MC. Virtually all MD simulations available for surfactant self-assembly have focused on the lipid bilayer membrane. Due to the potential impact for the life sciences, there is actually a drive to generate all-atom simulation results. MD simulations have the promise to give close-to-exact results on the ns time scale and nm length scale and can therefore serve as a reference for more approximate models. Of course, it remains true that the force field parameters used to parameterize all the interactions in the system have a phenomenological origin and are subject to uncertainty. Moreover, interactions such as bond stretching and bond bending, are usually treated in the lowest order. For example, the bond stretching is typically computed by Ust[l) = Kst(l-l0)2 where I is the length of the bond, (0 the equilibrium length and Kst the "spring" constant. In reality, the deviations of this parabolic dependence may be significant, but this is not incorporated. Relevant at this stage is also that the water molecules surrounding the bilayer in the simulation box are mostly modelled in a relatively rough way and that electrostatic interactions are typically cut off after a few water diameters in order to save CPU time. To be effective, MD simulations need extremely accurate initial "guesses" of the structure of interest. This means that the experimentally determined area per lipid molecule and the lamellar topology are often used to give the simulation a smooth ride to the equilibrium structure. Accurate MD simulations should allow the area per molecule to adjust in order to keep the bilayer tensionless. This means that the simulation box, in which the simulation is performed, is allowed to change its shape (i.e., its aspect ratio) during the trajectory. As it is extremely expensive to take many degrees of freedom into account and one likes to include as many lipid molecules as possible, the water film in between the membranes (periodic boundary conditions) is typically chosen to be very small (just a few water layers). Therefore, one should
4.38
ASSOCIATION COLLOIDS
consider MD simulations to be better representative for a lamellar phase at low water content than for isolated bilayers. In sec. 4.8 we will study the interactions between lamellar bilayers and show that this is indeed a serious limitation. Most of the MD work is done on the DMPC (di-myristoyl-phosphatidylcholine) bilayer, where the zwitterionic phosphatidylcholine head group is coupled via the glycerol backbone to two saturated C 16 chains. The SOPC (l-stearoyl-2-oleoyl-sn glycero-3-phosphatidylcholine) system is chosen here because a recent detailed comparison is made to a detailed SCF model11. The comparison between the two gives insight once more into the weak and strong points of more approximate SCF analysis. The corresponding SCF analysis of the SOPC bilayers is one of the most detailed available in the literature. We cannot go into all the details here, but a brief survey is appropriate. The lipid molecules were represented on a united atom level with segments with volume v (matching a lattice site). Two C18 tails, for which the CH3 end group was given a double volume, are connected to the backbone through an ester group (intermediate polarity). The tail coupled to the snl position was fully saturated. The one connected to the glycerol on the sn2 position (central carbon) had a double bond in between positions 9 and 10. The PC head group (on the sn3 position of the glycerol) consisted of a negatively charged phosphate with volume 5 v and a positively charged choline group with a central nitrogen carrying a positive charge. In a rotational isomeric state (RIS) scheme, the chain flexibility is given by the energy difference between a gauche and a transconfiguration. The value of 0.8 kT is used, except for the double bond halfway on the sn2 chain where the gauche configuration was strongly biased by favouring the gauche configuration by 10 kT (ad hoc choice). The interaction parameters were chosen such that the c.m.c. of this lipid is of order (p° = 3 10~12 . This means that the solvent is virtually free of lipids. The solvent phase was composed of water and a 1:1 electrolyte solution with (ps = 0.01. The effect of the ionic strength on the PC head groups has been studied by Meijer et al.2). In the SCF analysis, a special compressible water model was introduced. To this end, there was one extra monomeric component in the system, which represented vacancies V (we trust that this notation will not be confused by the macroscopic volume V of the system). The concentration of vacancies was fixed in the bulk and the local amount of free volume (local compressibility) is regulated by FH interaction parameters between V and the other segments in the system. The water molecules were considered to be composed of clusters of size y: W , where the whole set was taken, i.e. y = l,---,°° . The reaction of the type W + Wl ^± W +] with fixed reaction constant Kw was implemented at each lattice layer in the system. The cluster size distribution that follows depends on the local water density. To cut a long story short, such a model can be 11
A.L. Rabinovich, P.O. Ripatti, N.K. Balabaev, and F.A.M. Lccrmakcrs, Phys. Rev. E67 (2003) 011909; F.A.M. Lcermakers, A.L. Rabinovich, and N.K. Balabaev, Phys. Rev. E67 (2003) 01 1910. 21 L.A. Mcijcr, F.A.M. Lcermakers, and A. Nelson, Langmuir 10 (1994) 1199.
ASSOCIATION COLLOIDS
4.39
used to more accurately account for water and partitioning in the core of the bilayer. The classical way to do the chain statistics in an SCF model is such that the orientation of the chain is not affected by the orientation of its neighbours. This is a poor assumption for highly ordered systems such as lipid layers. It is much better to implement a first-order correction, as used in the SOPC results, which gives a bias to the orientation of a test chain to be oriented in the same direction as the average chains in the surroundings. This has been implemented in the so-called self-consistent anisotropic field (SCAF) approach. In such a model, it is possible to find the orderdisorder phase transition in the lipid bilayer, which in the biological literature is called the gel-liquid phase transition
1(
.
It is obvious that the MD and SCF models both have a plethora of parameters. In MD these parameters are hidden in the force field. In the SCF model the number of parameters is much less. Here we choose not to mention the set, but rather refer to the
Figure 4.7. Comparison of overall structural data obtained for hydratcd SOPC lipid bilayers. In panels a and c, the MD data are plotted where the density in g/cm3 is given as a function of the distance z from the centre of the bilayer. In panels b and d, the corresponding SCF results are presented where the volume fraction is plotted as a function of the layer number z (the centre of the bilayer has z = 0 ). Profiles of the two tails, the glycerol backbone units, the head group units, the phosphate and the nitrogen (the charged members in the PC head group) and water are given for both approaches. In the SCF results, the profiles of the 1:1 electrolyte are also given (in panel d), and the profile of the free volume is indicated by the letter V. 11
F.A.M. Lcermakcrs, J.M.H.M. Schcutjcns, J. Chem. Phys. 89(1988) 6912.
4.40
ASSOCIATION COLLOIDS
original literature for a full discussion of this1'. The primary results of both methods are predictions of the structural characteristics of the bilayer system. In fig. 4.7 we present a selection. The first problem that has to be overcome when a detailed comparison is made is that the MD method is not lattice-based whereas the SCF model is. This means that the natural units in MD simulations are densities in mass per unit volume whereas the SCF model gives scaled densities in terms of volume fractions. The translation of true to scaled densities depends on the Van der Waals radii used in the simulations and in the simulation this parameter is different for each atom. Accepting this problem, we still can compare the results on a qualitative rather than a quantitative level. The spatial coordinate also differs between the two methods. In the MD simulations, true dimensions are found whereas in the SCF method all lengths are normalized to the characteristic length of a lattice site. In principle, the size assigned to a lattice site in the system contains some compromise. Water molecules will fit on a lattice site with size I ~ 0.3 nm. The C-C bond is significantly shorter. Therefore, an average of 0.2 nm was chosen as a reasonable compromise between these two lengths. We met this problem before (sec. 1.3.8b). For the segment potentials, both the short-range, nearest-neighbour contacts and the longer range electrostatic interactions are accounted for as discussed in sec. 4.5 as well as in appendix 1. In a way we can refer to this approach as a variant of the GouyChapman model in which the finite size of the ions and the chain statistics of the chainmolecules are accounted for. Referring to figs. 4.7a and b, we find many remarkable points of resemblance and interesting differences regarding the overall concentration profiles across the bilayer. One directly notices the fact that the MD data are noisy, which is due to the limited statistics in the sampling. As there was no averaging done over the two halves of the bilayer, one can judge the level of noise from the left-right asymmetry. The major important observation is that water apparently does not penetrate into the core. In both models the water penetrates up to and perhaps slightly beyond the glycerol units, but not much further. This is in line with general experience. The overall density in the tail region is very homogeneous throughout the core; there is no dip in density in the centre as sometimes reported in MD studies. We expect that in those investigations the membranes were not sufficiently equilibrated. The core density closely resembles the density of a hydrocarbon melt (of, for example, C lg chains). The overall profile of the head groups is decomposed in a P and N profile for the phosphate group and the choline group in panels 4.7c and d, respectively. It is seen from these profiles that in both models on average the head group is laying rather flat, i.e. parallel to the membrane surface. However, the N profile, which is positioned at the end of the head group, has a wider distribution. This has important consequences for this potential profile (not shown, but deductible from the ion A.L. Rabinovich, P.O. Ripatti, N.K. Balabaev, and F.A.M. Leermakers. loc. cit.
ASSOCIATION COLLOIDS
4.41
distributions). It turns out that the electrostatic potential is somewhat positive on the water side of the membrane and also positive at the inside of the bilayer, but negative at the positions of the phosphates. Outside the bilayer in the aqueous phase, there is the classical diffuse part of the electric double layer with a decay length given by the Debye length. Inside the bilayer, there is also a diffuse-like layer. Here the difference in the distributions of the positive and negative ions decays much slower than in the water phase. The fact that at the centre of the bilayer the densities of positively and negatively charged ions approach each other shows that at these ionic strength conditions the electric field in the core approaches zero. In the SCF calculations, there is a free-volume profile across the bilayer (c.f. panel 4.7d). The fraction of free volume in the bilayer is higher than that in the water phase. Indeed, the free volume is more expelled by water phase than by the alkyl chains. This is consistent with the fact that alkyl chains spontaneously adsorb from water onto the vapour-water interface. The free volume further tends to be slightly elevated in the interfacial regions, that is near and around the glycerol backbone, the reduced density of molecular components allows for a reduction of unfavourable interactions. As the presence of free volume is important, e.g. for the transport of gases across the bilayer (see also sec. 6.7a), there are attempts to come up with such numbers from MD simulations11. Further, we should note that without any free volume the membrane would be in the gel state (chains highly ordered in all-trans conformation). The unsaturation in the sn2
chain keeps the bilayer in the liquid state. Within the same
parameter setting, the DPPC membrane is in the gel-state (C 18 lipids). Such gel membranes have been studied both in MD21 as with the SCF model31. There is a large number of measurables in MD. The SCF computations are less detailed, but nevertheless one can find various detailed predictions of measurable quantities. The order parameter profile of both alkyl tails, the interdigitation of tails across the midplane of the bilayer, the average position of segments along the tails and their fluctuations and several other parameters of the bilayers were found to be very similar between the two approaches. Moreover, they appear to follow the experimental findings. For example, order parameters can be extracted from deuterium NMR41. Considering the efficiency with respect of the computation times needed between SCF and MD (at least an order of 105 in CPU time), it is obvious that only the SCF method can be used for a parameter study, e.g. on the effect of the interdigitation of the alkyl tails as a function of the solvation of the phosphate group. Here we refer to the original literature51. In conclusion, this comparison shows the potentialities of MD and SCF simulations. 11
S.J. Marrink, R.M. Sok, and H.J.C. Bcrcndscn, J. Chem. Phys. 104 (1996) 9090. R.M. Vcnable. B.R. Brooks, and R.W. Pastor, J. Chem. Phys. 112 (2000) 4822. 31 F.A.M. Lecrmakers, J.M.H.M. Scheutjcns, J. Chem. Phys. 89 (1988) 6912. 41 K. Gawrisch, N.V. Eldho, and I.V. Polozov, Chem. Phys. of Lipids 116 (2002) 135 5i F.A.M. Leermakers, et al.(2003) loc. cit. 21
4.42
ASSOCIATION COLLOIDS
4.3c Quasi-macroscopic models Another strategy to extend the theoretical modelling of association colloids is to construct (based on experimental evidence or insight from realistic molecular models) new theories in which some carefully selected degrees of freedom are deliberately Ignored In order to end up with a simpler, and, more importantly, more transparent description of the system. These models typically deal with micellar aggregates on the macroscopic level although it remains necessary to mix in some molecular features as well. As a result, these models typically pragmatically combine features of the macroscopic with that of the microscopic world and we may refer to the approach as being quasi-macroscopic. They may subsequently be used to consider problems that are beyond the reach of detailed simulations or numerical SCF methods. The target of this level of modelling is the Gibbs energy of a micelle (or the micellar solution), which is assumed to be composed of a set of separable contributions for which relatively simple expressions are formulated. Although the computer is often still needed to solve the equations, we may also refer to this approach as being analytical. Important contributions may be found in the literature1'. The mass action model is the starting point for many models of this kind. The important property is the difference in standard chemical potential of a g -aggregate and g monomeric surfactants in solution, //°( , - g/i0^ • This quantity is once more identified to be the average grand potential of a micelle In the micellar standard state £m identical to the translationally-restrlcted grand potential. Upon dividing by, g we obtain the difference in the standard chemical potentials between a surfactant molecule present in an aggregate of size g and a singly dispersed surfactant in water or, equlvalently, the surfactant contribution to em uslg)-9Msa) 9
[ 4 3 i ]
g
which is a function of the size of the aggregate (as well as the temperature and pressure). The target of analytical modelling (as with molecular modelling) is to come up with accurate formulas that account for the major contributions to nfl. Typically it is assumed that these are additive and that one can evaluate them individually. In general, there may be a relatively large number of contributions. We will introduce and subsequently discuss these in more detail 12° =U2°) tr+ U2O) def( +U2£) intf +U2°) s t e r i c + -
[4.3.2]
' R. Nagarajan, Theory of Micelle formation: Quantitative Appraoch to Predicting Micellar Properties Jrom Surfactant Molecular Structure in: Structure-Performance Relationships in Surfactants, K.Esumi and M. Ucno, Eds., Surf. Sci. Series 70 (1997) 1; D. Blankschtcin, A. Shiloach, N. Zocllcr, Current Opinion Coll. Int. Sci. 2 (1997) 294; J.C. Eriksson, S. Ljunggren. Langmuir. 6 (1990) 895; J.C. Eriksson, S. Ljunggren. and U. Henriksson J. Chem. Soc. Faraday Trans II. 81 (1985) 833.
ASSOCIATION COLLOIDS
4.43
In this equation, the meanings of the various contributions are the following. The term labeled tr accounts for the transfer of a hydrophobic tail from water to the micelle, deft expresses the deformation of the tail when it is densely packed and oriented (on average) in the normal direction in the micelle as compared with the bulk liquid state, intj
gives the work to create a core-water interface, and steric points to the
interactions between the (uncharged) head groups. When the surfactants are charged we need to add an extra term (•OSlei • w n l c n accounts for the electrostatic contribution. For non-ionic surfactants, it is necessary to account for the solvation (i2°)sol and stretching of the corona chains (i2° ) def . Let us illustrate the procedure by considering a primitive model for a spherical micelle composed of a linear alkyl-based surfactant with a small uncharged hydrophilic head group with area a h , postponing more detailed treatments to sees 4.4 and 4.5. Assume that the core has a density close to that of a liquid hydrocarbon. The transfer of Gibbs energy of a CH2 , ( G2 ) differs from that of a CH3 , ( G3 ) and may be estimated from solubility data for hydrocarbons in water. At room temperature one has for a tail with (t -1) CH2 s and one CH3 (the other end is connected to the surfactant head) a transfer-free energy per chain amounting to: — s - i L = G 2 (t-l) + G3 =1.47(£-l)-3.53-lnx s ( 1 )
[4.3.3]
The translational entropy of the monomer, -lnx^j, where x s ,j, is the monomer mole fraction, opposes micellization whereas the two non-logarithmic terms are the most important contribution to the driving force for self-assembly; both terms are negative. The values for G2 and G3 can be obtained from alkyl-water phase equilibria as reported by Tanford11. In the literature the temperature dependence of the transfer energy is also available21. ( &)ti
depends on the tail length t and not on the aggregation
number, and in fig. 4.8 (where the translational entropy contribution is dropped) it is just a horizontal line. One may argue, however, that in a spherical aggregate with g tails in the core, not all segments manage to find a place deep in the core and as a result some segments have to be interfacial. This would imply that there is a g dependence. This feature is incorporated in the intj term to be discussed below. The volume of the core, composed of g chains with tail length t, can be computed when the volumes occupied by a CH2 and CH3 are known31. The molecular volumes v2 for a CH2 and u3 for a CH3 are estimated from the density of aliphatic hydrocarbons. One chain occupies the volume ut
1 C. Tanford, The Hydrophobic Effect. Formation of Micelles and Biological Membranes, 2nd ed. Wiley and Sons, New York (1980). 21 R. Nagarajan, E. Ruckenstein, Langmuir 7 (1991) 2934. 31 R. Nagarajan, Theory of Micelle Formation: Quantitative Appraoch to Predicting Micellar Properties from Surfactant Molecular Structure in: Structure-Performance Relationships in Surfactants, K.Esumi and M. Ueno, Eds., Surf. Set Series 70 (1997) 1.
4.44
ASSOCIATION COLLOIDS
Figure 4.8. a) The (tr) transfer (excluding the l n x ^ , contribution), (def) tail deformation, (intf) = interfacial, and (int) head group interactions contributions to the micellar standard state grand potential per surfactant of a g -micelle, as well as the (tot) total standard chemical potential as a function of the aggregation number for a surfactant with ( = 12 and head group area of approximately a^ =; 0.5 nm^. An alkyl-water interfacial tension y = 50 raN/m is assumed, b) The micelle size distribution around the maximum on a logarithmic scale (arbitrary units). vt = u 2 ( t - l ) + u 3 = [0.0269(t-l) + 0.0546]10- 2 7
[m]3
[4.3.4]
The estimate for the tail length contributes to the length of a methylene group l2 and a methyl group I3 is It = l ^ t - 1 ) + I3 =[O.13(£-1) + O.28]1CT9
[m]3
[4.3.5]
For a spherical micelle, we can now directly compute the area per molecule a g (t) occupied by a g -micelle as a function of the tail length t
„ ( « = ( — ] (vt)213 g \ 9 )
[4.3.6]
The radius of the g -micelle with surfactants with length t is given by Kg(t) = H f - M g ^ 4/r )
[4.3.7]
where it is assumed that there is a sharp interface between core and aqueous phase. In reality the interface is somewhat diffuse; we argued above that it is a few water molecule diameters wide and for macroscopic interfaces, this is well established (chapter III.2). As a result, the radius requires an assumption as to how to define and locate the core-corona interface. In homogeneously curved assemblies, the surfactant tails are on average all oriented in the normal direction of the micelle surface and as one end of the surfactant tail is attached to the polar head, which is constrained at the core-water interface, we should expect that the chain conformation differs from that in the bulk. There are several ways to estimate this conformational contribution, which involves an
ASSOCIATION COLLOIDS
4.45
entropy loss. One way to do this is to enumerate in a dedicated mean-field theory the conformational entropy of tails that are grafted by one end on a curved interface such that the density of chains in the core has a preset value. The results for such singlechain partition function for various g and t values can be fitted to obtain interpolation formulas". Typically, for very short chains it is possible to account for all intramolecular excluded-volume effects, finite extensibility, etc. The disadvantage of such interpolation equations is that they do not give much physical insights. An alternative method was suggested by Semenov21 who launched an equation inspired by the entropy loss that is suffered by stretching an ideal or Gaussian chain. When a chain with contour length L = tl, where / is the length of a methylene unit, is forced into a micelle with radius R q (t), the entropy loss depends on the flexibility of the chain. The statistical segment size of an alkyl chain is often taken to be 3.6 methylene groups. This means that the statistical segment size (k = 0.46 nm and the cross-sectional area of an alkane molecule in the liquid state is (£ =0.21 nm 2 . The number of statistical segments in an alkyl chain of length t is thus given by rk = t/3.6 . The deformation contribution to the standard chemical potential is given by t^gW
Kit)
where for the spherical geometry Ks = 3/r2 / 80 . In cylindrical geometry with radius of the cylindrical crosssection R , the coefficient Ks is replaced by Kc = 5/3Ks. For lamellar geometry Kj = 2KC and R is replaced by the half-bilayer thickness. The deformation of the tails potentially contributes to a significant stopping force especially when the characteristic size of the micelle is on the order of the fully extended length of the tails. In the Gauss-type stretching model, the chain can extend unrealistically further than its contour length and typically the entropy loss of the fully stretched chain is underestimated. Therefore, in the Gauss-type model the chain stretching usually contributes to a minor extent to the stopping force. As can be seen in fig. 4.8a, the deformation term is positive and it only weakly depends on the aggregation number. When a micelle is formed, there is a core-water interface left because the head groups cannot cover the entire area. This means that we overestimated the transfer energy contribution. We have to implement a correction. Assuming a sharp core-water interface, a first estimate is -G 2 A(u2) 2/3 l a N'')~ a h'' w n e r e " R I ^ ' 2 3 i s *-ne a r e a e x ' posed to the water phase of a surface CH2 group and (a N (£)-a h ) is the area that is not shielded by the head group (the dimensions of the head group area is given by a h ). However, in this approach the finite width of the core-water interface is ignored. Therefore, one usually uses the interfacial tension approach 11 J.C. Eriksson, S. Ljunggren Langmuir 6 (1990) 895; J.C. Eriksson, S. Ljunggren, and U. Henriksson J. Chem. Soc. Faraday Trans 2, 81 (1985) 833. 21 A.N. Semenov, Soviet Phys. JETP, 61 (1985) 733.
4.46
^
ASSOCIATION COLLOIDS
L
=^ ( a g ( t ) -
V
[4.3.9,
where y^w is the macroscopic interfacial tension yhw = 50 niN/m of the interface between water and hydrocarbon (composed of chains with length t ). The interfacial term has the tendency to enlarge micelles as for given t the area is a decreasing function of the aggregation number and the free energy cost per molecule decreases as is also shown in fig. 4.8a. Hence, this term raises the cooperativity of the micellization process. An equation such as [4.3.9] gives a first-order approximation of a quantity that is more complex. The free energy associated with the interface between core and corona has non-trivial curvature corrections, non-trivial anisotropic chain orientational aspects, etc. We further stress that the interfacial tension as used in this equation is not thermodynamically defined because there is no way to measure unambiguously. We will see below that one can not extract this quantity from molecular modelling either. In this primitive model the most important stopping mechanism is the contribution due to repulsive interactions between the head groups. The simplest way to account for the head group repulsion is the van der Waals (regular solution) approach. Assuming head groups of size a h = f^ , the area fraction occupied by the heads is given by Ja-gah/{4:7rRg{t)2). The ideal pressure of a 2D lattice gas is given by (nah)lkT = -\n(\-(t>a)
or
tester ^ t e r =-ln(l-/a)
[4.3.10]
Of course, this ideal lattice gas pressure will not be appropriate at large values of Ja and [4.3.10] should be replaced by (i2°)ster /kT = ( l - / a ) " 2 , which follows the simulation results for hard disc pressures in 2D rather accurately even at relatively high values of Ja . We mention once more that we must include extra terms if the head group is charged (of electrostatic origin) or oligomeric (elastic-free energy of stretching of the head group). Such contributions will be discussed separately below. The micelle size distribution is, for given surfactant length t, given by
*>m(9)-exp U ^ ? -
[4.3.11]
The shallow minimum of the total i2° in fig 4.8a leads, even on a logarithmic scale, to a sharp maximum in
ASSOCIATION COLLOIDS
4.47
stretching of the tails may be significantly underestimated and we may consequently have underestimated the contribution of chain-stretching to the "stopping" force for micelle formation. Further, the absence of micelles with very small g is only qualitatively correct because the assumptions of sphericity of the structure of such small aggregates is likely to be far too restrictive. A better description should also feature curvature corrections for y hw . Evaluating the quasi-macroscopic model, we see that the interfacial tension yhc essentially increases the cooperativity for self-assembly whereas the head group repulsion, I.e. the interaction term, is the anti-cooperative term. The combination of the two gives rise to a minimum in the standard chemical potential of the micelle as a function of g and the corresponding maximum in the size distribution. State of the art quasi-macroscopic models are developed in rather large detail and are targeted to be molecularly specific. For example, such analytical models may be used to obtain insight into the selection of the geometry for a particular surfactant system. The temperature dependence of micellization may be investigated after the temperature dependencies of the parameters of the model are specified11. In the following sections, we will elaborate specific molecular models for the selfassembly of a specified set of surfactants. We base our analysis on the self-consistent field model because this method is best suited for this. Results will be complemented by those from the corresponding quasi-macroscopic models and checked against experiments where appropriate. 4.4 SCF for non-ionic (spherical) micelles In this section we will discuss two types of non-ionic surfactants, the C n E x and Pluronic surfactants, which have the EO moiety as the head(s) and the alkyl chain or PPO unit, respectively as the tails. Most of the discussion will be focused around SCF predictions. Quasi-macroscopic models for non-ionics will receive some attention at the end of this section. Water is a reasonably good solvent for polyethylene oxide at room temperature. On the E = C2O -level, the FH parameter with water (W) is at room temperature at approximately XWE ~ 0-4 • PEO has a lower critical solution temperature (LCT) meaning that there exists a temperature, higher than room temperature, such that / W E > 0.5 and, depending on the molecular weight, the system has a solubility gap. The surfactants that rely on the EO part of the molecule to have micelles that are soluble in water are correspondingly sensitive to the temperature. Likewise, the solubility of the PPO chain is temperature sensitive. As a result the most important parameter to control the micellization of EO-based surfactants is the temperature. The micellization for the
J)
R. Nagarajan, loc.cit.: D. Blankschtein, et ai. Loc. clt.: J.C. Eriksson, et at, Loc. cit.
4.48
ASSOCIATION COLLOIDS
family of non-ionic surfactants is a very weak function of the ionic strength. Only at relatively high ionic strength do electrolytes affect the solvent quality of the ethylene oxide part of the chain. In this chapter, we choose to construct an SCF model for which the united CH2 atoms determine the discretization length. The simple reason for this is that the alkyl chain may also be modelled using these entities. Using these united atoms allows one, in principle, to account accurately for the chain architecture. We use a model pioneered by Barneveld11 for EO-containing and PO-containing surfactants that has fewer parameters than the two-state model, but still captures the main temperature characteristics (c.m.t, c.p.t). The generic features of micellization can easily be illustrated by a very primitive model, cf sec. 4.3b ad (i). However, eventually one needs to be molecule specific. The challenge for the near future is, of course, to choose parameters and molecular architectures such that the results can be generalized. One of the problems for a realistic molecular model is dealing with the water phase. Water is an associative fluid that forms via relatively strong and many H-bonds small water clusters that dynamically break and form. This complexity is beyond reach for models that are targeted to describe association colloids. At present we can only advance via ad hoc water models. Below we will discuss SCF results with spatial resolution in two dimensions, and for such a coordinate system we need (for the time being) to be less ambitious. In short, we will model water to be composed of a cluster of W units that occupy five sites (one central surrounded by four others). In addition, we allow for discrete amounts of free volume. Free volume is implemented as units called V that occupy one lattice site each. For details about this model, we refer to appendix 1. The alkyl-EO non-ionics are modelled as(Cg)1[C]n_[[(O)1(C)2]x(O)j. Here we choose to introduce the united atom level of description with the C 3 pointing to the terminal methyl group and the O representing oxygen. The terminal O represents the alcohol group. We will model the methyl group to be a bit more hydrophobic than the methylene units and therefore set %C W= 1 • 1 a n d Xc w = 1 ^ • The overall solubility parameter for an ethyleneoxide unit should be such that it is soluble in water. Typically it turns out that one should choose a negative value and use the Ansatz that this parameter is inversely proportional to the temperature Xow = /£ow'300)-22fi where %ow(300) = -0.6 is the value at room temperature. In passing, we note that in the literature there exists a two-state model to deal with the temperature dependence of non-ionics2'. We will not discuss this approach here. An important parameter that drives the demixing of the heads and the tails is the repulsion between O and C for which we choose Xoc = %oc = ^ • The exact value of this parameter is uncertain; this value becomes important when the
1
' P.A. Barneveld, The Bending Elasticity of Surfactant Monolayers and Bilayers, PhD thesis Wageningcn University (1991). 21 P. Linse, M. Bjorling, Macromolecules 24 (1991) 6700.
ASSOCIATION COLLOIDS
4.49
Table 4.2. The Flory-Huggins interaction parameters used in the SCF analysis of the non-Ionic surfactant system at T = 300 K X
ca
c
O
W
V
ca c o w
0
0.5
2
1.5
1.5
0.5
0
2
1.1
2
2
2
0
-0.6
2.5
1.5
1.1
-0.6
0
2.5
2
2.5
2.5
0
V
1.5
bilayer is normally compressed. Bilayers under compression will be discussed in sec. 4.8. To make the set of FH interaction parameters complete, we mention that with respect to the Interaction with the free volume the O is treated as W , i.e. Xyo = 2.5 , and the mixing energy between methylene and methyl groups is set to j
cc
= 0.5 .
Below all lengths are made dlmenslonless by normalization to the characteristic size of the lattice cell for which a value of I = 0.2 nm Is used. The surface tension (grand potential per unit area) is given in dimenslonless units kT 112 . This implies that y = 1 corresponds to 100 mN/m. The tabulated set of FH interaction parameters is relatively large and admittedly to some extent ad hoc. We have used the comparison with the more detailed modelling to estimate most of the parameters, but they still should be regarded as adjustable. We may judge the quality of the set by their consistency with respect to the predictions that result from it and to the comparison of the results with experiments. One should realize, though, that the MD or MC simulations typically include even more parameters In their force fields. Similarly, in analytical models the number of parameters is not small either. We have to live with this. We note that the present free-volume model has approximately twice the number of input parameters as compared with the more primitive incompressible model of sec. 4.3b (i). However, as we will see, this set suffices to deal both with the whole set of non-Ionics, including the Pluronics and related compounds, as well as mixtures of these components. Finally, this parameter set is sufficient to predict the adsorption of the surfactant at the air-water interface without the need to introduce additional parameters (space does not allow us to illustrate this).
4.4a Micelles at and above the c.m.c. We will now discuss in some depth the state of the art of SCF modelling of non-ionic surfactants of the C n E x type, for which we choose three representatives (n,x) = (10,6), (12,5), and (14,4). Later we will also discuss other members of this surfactant family. This set of three surfactants has been selected because all three surfactants have the same number of carbons and as compared with the (12,5), the (10,6) has one more O
4.50
ASSOCIATION COLLOIDS
Figure 4.9. SCF results for three C n E x surfactants, (n,x ) = (10,6), (12,5) and (14,4). a) The translationally restricted (standard state) grand potential as a function of the aggregation number, b) The bulk concentration of the freely dispersed surfactants (monomers) as a function of the aggregation number of the micelles, c) The total volume fraction of surfactant in the system as a function of the monomer volume fraction. The arrows point to the theoretical c.m.c. of the surfactant system, d) The relative size distribution ag/g as a function of the average aggregation number g . The open circles in panels a and b represent the system, which coexists with infinitely long cylindrical micelles to be discussed in sec. 4.6b.
and the (14,4) has one less O unit. Hence, the overall molecular weights of the three components are similar. Nevertheless, the three surfactants are remarkably different in their association behaviour as is known experimentally11. We may refer to fig. 4.1b to a detailed MD-generated snapshot of a non-ionic micelle. Here we will be Interested not in snapshots but in the average micellar properties. Any SCF analysis of micellization starts by analyzing £m(g) as a function of the aggregation number. These functions are given in fig. 4.9a. Typically, £m(g) has a positive slope at very small values of g (dotted line), there exists a part with a negative slope (drawn line), and subsequently there may again be a part with a negative slope (again dotted). As explained, only the sections with negative slopes correspond to micelles that are macroscopically stable. The maxima in the curves correspond to the smallest micelles that are stable. We will refer to these micelles as the ones corresponding to the theoretical c.m.c. " R.G. Laughlin, The Aqueous Phase Behavior of Surfactants, Academic Press (1994).
ASSOCIATION COLLOIDS
4.51
For the surfactant with the largest head group (10,6), there is an aggregation number above which em is negative. We argued above that the physically realistic values for this quantity should be positive, but we have dotted the continuation of the curve towards the negative values. The two other curves do not reach negative values, but develop a minimum in £m{g) • This minimum signals the upper limit of the size of spherical micelles that is thermodynamically stable. This minimum also tells us that the micelle concentration cannot increase further. For these systems we need to consider alternative aggregate shapes. In fig 4.9b we present the volume fraction of the freely dispersed surfactants (monomers) that coexist with micelles with aggregation number g . We note that the chemical potential of the surfactant is (in first order) given by the logarithm of this volume fraction and hence one can interpret the figure also as the chemical potential of the surfactant versus g . It follows from [4.2.19] that there is a direct relation between em and the chemical potential. According to [4.2.19], when £m(g) decreases, ^ s u r f (g) increases and vice versa. Again, we have used a solid line for those parts that are physically significant and dotted the irrelevant parts. Consistent with the idea of a c.m.c, the smallest stable micelles are found at the minimum of the (monomer) chemical potential curve. We will see that other definitions of the c.m.c. may be given. To elaborate on an alternative measure of the c.m.c, we present in fig. 4.9c the relation between the overall surfactant concentration, computed by [4.2.22], i.e. (^ =
4.52
ASSOCIATION COLLOIDS
be sufficiently stable to allow for a high concentration of (spherical) micelles. Below we will argue that this system is close to a cloud point (at room temperature). Several alternative measures of the c.m.c. exist. A practical definition of the c.m.c. is the concentration where half of a differential surfactant addition goes into micelles, i.e. where S
Figure 4.10. Overall radial volume fraction profiles for spherical micelles of C12 E 5 . a) Near the theoretical c.m.c. for an aggregation number g = 45 . b) Near the experimentally accessible c.m.c. for an aggregation number g = 102 .
ASSOCIATION COLLOIDS
4.53
4.4b Structure of C 12 E 5 micelles Once the set of macroscopically stable micelle sizes is known, it is of interest to discuss their structural characteristics. In fig. 4.10 we present radial volume fraction profiles of C 12 E 5 surfactant micelles. In panel (a) we show the smallest stable spherical micelle at the theoretical c.m.c. In this case the aggregation number is just g = 45 and aggregates only exist at extremely low micelle concentrations. In panel (b) a significantly larger micelle is presented, which is more characteristic of a spherical micelle at such concentration as might be seen by light scattering (i.e. near the experimentally expected c.m.c). The micelle with the smaller aggregation number appeals to our intuition. In the core the tail density is rather homogeneous; the core is virtually free of water and the free volume has slightly accumulated in the tail region with respect to the water phase. All these properties are in broad agreement with expectations. Less obvious is the amount of overlap between tails and head groups. The ethyleneoxide unit is not completely insoluble in a hydrocarbon phase1' because there is just one O near a pair of C's. Hence, we should anticipate that the separation between heads and tails is not as dramatic as one would expect for a charged surfactant. Indeed, the transition region between core and corona is very wide and extends over much of the micelle structure. Such a rather diffuse core/corona interface poses a major problem for accurate analysis by quasi-macroscopic models where this interface is assumed to be (fairly) sharp. The micelle size is rather modest and the core radius is significantly less than the extended tail length. From the thermodynamic analysis of this surfactant system (fig. 4.9a), it was found that £mig) has a minimum. This minimum indicates that spherical micelles above a particular size are unstable. We wish to find some molecular interpretation of this instability. The radial volume fraction profiles presented in fig. 4.9b show a micelle that is close to this instability regime. In this case the aggregation number is approximately 100 at or near the "experimental" c.m.c. As compared with the small micelles, the aggregation number has doubled and the size of the core has extended significantly and is now on the order of the extended length of the tails. From this point of view, it is not surprising that growing the micelles still further becomes incompatible with the spherical shape. There are a number of significant changes in the radial volume fraction profiles as compared with the small micelles that are consistent with the incipient instability. The tail density in the core tends to drop to some lower value in the centre. The head groups now interpenetrate significantly into the core. The water phase, however, is still kept outside the micelle. Eventually, with a still larger aggregation number, the head groups start facilitating the uptake of water in the micelle and the micelle structure becomes unstable. In passing, we note that this scenario of micelle size instability, which presents itself in the SCF calculations, will remain unnoticed for all approaches that impose 11
S. Burauer, T. Sachcrt, T. Sottmann, and R. Strey, Phys. Chem. Chem. Phys 1 (1999) 4299.
4.54
ASSOCIATION COLLOIDS
conformational constraints on the surfactants. One may wonder how sensitive these instabilities are for the parameters in the model. Without going into all the details, it is necessary to mention that all presented results are seen in all reasonable parameter settings. Of course, one can suppress the instabilities by increasing ^ c o or by making Xyjo m o r e negative. These observations are relevant in the context of the temperature sensitivity of the micellar solutions, which will be discussed in the section where we consider the Pluronic surfactants. It Is of considerable interest to also present the radial distributions of individual segments along the surfactant. To this end, we evaluate the number of segments with ranking number s at coordinate r given by ns[r) which are related to the segment volume fraction profiles <jo(r,s) by ns{r) = L{r)(p{r,s), where L{r) is the number of sites in the spherical coordinate system at coordinate r (see appendix 1). We have selected the C 12 E 5 surfactant at the theoretical c.m.c. for this exercise and have numbered the terminal CH3 as segment s = 1 and the terminal oxygen as segment s = 28. One intuitively expects that each segment will have a distribution that is centered around a given most likely position. This average position is expected to be at a lower r for smaller values of the ranking number s . In other words, the tail end should be in the micelle core and the end of the hydrophilic head should mostly point towards the water phase. Moreover, the width of the distribution is expected to be wider for terminal segments than for middle segments. Inspection of fig. 4.11a,b, which provides the segment distribution functions, gives several rather surprising results. Referring to fig. 4.1 la, where some of the tail segments are presented, we notice that most of the tail segments have virtually the same distribution! All distributions peak near the edge of the core and the width of the distribution is only a very weak function of the ranking number. In fig. 4.11b, where the distributions of the carbon groups just next to the oxygen atoms are shown, the situation is more according to intuition. The distribution shifts to larger r and widens for increasing s . The fact that radial distributions of the tail segments are rank independent must be attributed to the spherical topology; only few segments are needed to populate the
Figure 4.11. Segmental distribution functions ns(r) for segments with ranking number s as indicated: a) for segments in the tail and b) for segment next to an oxygen unit in the head group for the C 1 2 E 5 surfactant at the theoretical c.m.c.
ASSOCIATION COLLOIDS
4.55
centre. Moreover, the observation that the core-corona interface is wide makes it is possible that almost every segment in the tail can visit the centre. As most of the volume of the core is at its periphery, this explains that the most likely position of each tail segment is at the core-corona interface. Most "artists' view" representations of spherical micelles are definitely misleading in this respect. From the appendix and specifically [A.5.2], we know that it is possible to define and compute a grand potential density a{r). It is possible to show that this quantity is equivalent to the difference between the normal and the tangential pressure a>(r) = lp N (r)- pT{r)] ( a o / u o ) , where ao/vQ <x \/l is the ratio of the area and the volume of a lattice site. Close inspection of [A.5.21 also reveals that there are various options for defining this difference p N - p T (i.e., the grand potential density). Indeed, this quantity cannot be unambiguously defined. The fundamental reason for this is that there are no strict rules on how to attribute the non-local interaction terms to the various contrib.utions to the local pressure tensor components. The choice made here to compute the grand potential density is that the binary interactions are assumed to be equally distributed over the two components that take part in each pair interaction. Other options are possible; these may lead to different grand potential densities. The caveat is that one choice is not necessarily better than the other. Hence, there should not be any physical consequence of how this bookkeeping is done. Having mentioned this ambiguity, we still may be interested in the question of how the radial distribution of the grand potential density develops in the micelle. In fig. 4.12 we present these distributions, made dimensionless with I3 IkT for the two micelles with g = 45 and g = 101 for which the overall distributions were shown in fig. 4.10. The proper volume integral over the grand potential density distribution is equal to £ m , which is different in the two cases ( em ~ 25 for g = 45 and £m - 1 3 for g = 101.). The grand potential density assumes both positive and negative values. In the head group region it tends to be positive, indicating repulsion between the head groups. The grand potential density drops sharply to negative values near the core-corona interface; the negative values keep the micelle together. In the centre of the micelle the pressure tends to grow again. For the large micelle the local value for p N - p T is clearly positive in the centre, while it remains negative for the smaller one.
Figure 4.12. The radial grand potential density profiles for two micelles of C 12 E 5 with number of surfactants g = 45 and g = 101 as indicated.
4.56
ASSOCIATION COLLOIDS
4.4c Trends for various micellar characteristics It is known experimentally and all models also predict that log c.m.c. drops linearly with the tail length, see [4.1.1 ]. In fig. 4.13a this linear dependence is illustrated for the non-ionic surfactant series with eight EO segments. In the calculations this linear dependence is predicted accurately. The slope is strongly dependent on the actual value used for the j c w parameter for which a value of Zcw = * •1 w a s chosen. One should realize that the conversion from volume fraction to molar concentrations involves a division by the chain length, which is JV = 25 +1. Implementing this conversion has minor effects on the slope. The present parameter settings give logic.m.c.) <* -0.351. Most likely this is a slight underestimation of the experimental slope. Theoretically one can easily increase the slope by increasing ^ c w . In fig. 4.13a it is shown
Figure 4.13. a) The concentration of surfactant at the c.m.c. (in volume fraction units) on a logarithmic scale as a function of the length of the surfactant tail for CtEg non-ionic surfactants. The solid line is the theoretical c.m.c. (i.e. first measure of the c.m.c. where the maximum occurs for e m (g)), the dotted one is the second measure for the c.m.c. (where the volume fraction of micelles equals that of the freely dispersed surfactants, i.e.
ASSOCIATION COLLOIDS
4.57
that two definitions of the c.m.c. ((i) theoretical c.m.c. defined by the appearance of the first stable micelles and (ii) an "experimental" c.m.c. where the micelle concentration equals that of freely dispersed ones) differ systematically in line with the fact that the c.m.c. is not infinitely sharp. However, the scaling of the c.m.c. with the tail length is not affected by this ambiguity in the definition of the c.m.c. Applying arguments based on the surfactant packing parameter, we argued above that the aggregation number in spherical micelles should grow more than linearly with the tail length. As shown in fig. 4.13b, we see that the aggregation number for both measures of the c.m.c. is almost linear with the tail length, albeit that the data are better fitted by a quadratic dependence. We notice however that near the theoretical c.m.c. virtually all surfactant systems form spherical micelles, but that especially as the surfactant packing parameter is not close to 1/3, one should anticipate that shape changes may take place, which cause the aggregation numbers to grow much more quickly than predicted in fig. 4.13b. The growth in aggregation number as a function of the tail length must imply the increase of the size of the core on the one hand and the micelle as a whole on the other 1 '. As shown above, we have information on the distribution function of the individual segments in the micelle. It is rather straightforward to obtain the first moment of such distribution functions to find the average location of a particular segment. We estimate the core size by the average position of segment s = t, which is the last apolar segment of the tail connected directly to the head group. The size of the whole micelle is estimated from the first moment analysis of the terminal oxygen segment s = N . These two size measures are presented in fig. 4.13c. Completely in line with expectations, we find that the core radius increases with the length of the surfactant tail. On top of the core there is a corona, the size of which being a good approximation, independent of the surfactant tail length. This can be deduced from the constant difference between the micelle size and that of the core. From this constant corona size, we may anticipate that the area per molecule am is a weak function of the tail length. We may compute the dimensionless area per molecule am straightforwardly as am = 4/rR20re Ig . The conversion to a real area per molecule is a technical issue. We suggest using the conversion factor 2!2 = 0.08 nm 2 . The factor 2 corrects for the anisotropic dimensions of a lattice site (which may be traced back to the value of Aj =1/3, cf. appendix 1). In fig. 4.13d, it is shown that the area per molecule is a slightly increasing function of the tail length. In real dimensions we predict the area to be on the order of 0.4-0.5 nm 2 per molecule. From experiments it is known that the c.m.c. is only a weak function of the head group size (see table 4.1). In fig. 4.14a we present the volume fraction of monomers at the theoretical and "experimental" c.m.c. for a series of non-ionic surfactants for which 11
C. Tanford, Y. Nozaki, and M.F. Rhode, J. Phys. Chem. 81 (1977) 1555.
4.58
ASSOCIATION COLLOIDS
Figure 4.14. a) The concentration of surfactant at the c.m.c. (in volume fraction units) on a logarithmic scale as a function of the length of the surfactant head x for C16EX non-ionic surfactants. The solid line is the theoretical c.m.c, the dotted one is the experimentally expected c.m.c. b) The corresponding aggregation number at the theoretical c.m.c. (closed symbols) and that at the experimental c.m.c. (open symbols) as a function of the tail length. The drawn curves are power-law fits with an exponent of approximately -0.6 . c) The size of the micelle core and the overall micelle in lattice units as a function of the head group size for the C 16 E X with £m = 10 kT . d) Corresponding dimensionless area per molecule as a function of head group size.
the tail length is fixed to 16 units and the head group size varies from 8 to 32. In line with expectations, the experimental c.m.c. shifts to a higher concentration in line with the theoretical one. The chain length JV, needed for the conversion from volume fractions to molar quantities, is a strong function of the head group size. In this case AT = 16 + 3x + l. With this type of correction, the c.m.c. is an even weaker (but still growing) function of the head group size. The corresponding aggregation numbers are presented in fig 4.14b. As expected, with increasing head group size the aggregation number decreases. The SCF predictions indicate that the aggregation number follows the scaling behaviour g « x~° 6 , both for the micelles near the theoretical c.m.c, as well as near the experimental one. Referring to fig 4.14c, we note that the overall size of the micelle is a linearly increasing function of the head group size, i.e. Rm = 10 + 0.09x . Indeed, the aggregation number goes down, and directly coupled to this the dimension of the core reduces, but this does not compensate for the growth of the corona layer. The thickness of the corona is
ASSOCIATION COLLOIDS
4.59
•^corona x x ° * a n d it exceeds the unperturbed radius of gyration of the head group moiety significantly. The area per molecule is an increasing function of the head group size as can be seen in fig 4.14d. To a good approximation we find that am =*= x02 . This scaling exponent is significantly smaller than 0.5, indicating again that the head group moiety is stretched normal to the core-corona interface.
4.4d
Quasi-macroscopic approaches to non-ionic micelles
From the above it is obvious that it is not trivial to make transparent and predictive (analytical) models for non-ionic surfactant systems. Nevertheless it is of interest to review briefly some of the attempts that are present in the literature. For this, we must return to sec. 4.3c and, in particular, to [4.3.2] where various contributions to the chemical potential of exchange of a surfactant from the solution to the micelle are summed up. To use such a model for the alkyl non-ionics, we need to add terms that capture the relevant aspects of forming a corona made up of the EO head groups. Basically we must account for the solvation effects (i7i ) m i x and the deformation of the head group U^) d e f h • We have seen, e.g. in fig. 4.10, that the corona chains do not have a trivial distribution and there are several ways to approximate what happens. As a first model, it is assumed that the head group density is uniform over some region D, which may be called the effective corona thickness. In such an approach the tail is non-uniformly stretched. In a second model, we assume that there is a radial distribution of segments and propose that the stretching of the corona chains is uniform. Although these models are not completely confirmed by accurate SCF predictions, both lead to relatively simple equations that may serve as first estimates. Regarding the mixing contribution in a corona of homogeneous density, it must be realized that the surfactant in solution has an EO fragment, which is viewed as an isolated free polymer coil swollen in water. In the corona, it is viewed as forming a solution denser in polymer segments compared with the isolated polymer coil. The differences in hydration between the two states contribute to the Gibbs energy of the aggregate. To estimate this, we need the volume of the corona Vs, which may be expressed in the volume of the core Vc , the radius of the core Rc and the corona thickness D VS=VC (1 + — ) 3 - l R
I
c
[4.4.1]
)
The average volume fraction is given by (pc = g(3x +1)/3 / Vc . Introducing the excluded volume parameter v = 1 - 2 j
E W
, where / E W is the overall FH interaction parameter
for an ethylene oxide segment with water, we write for the mixing energy
(i?U
i
4.60
ASSOCIATION COLLOIDS
When the density is allowed to be non-uniform and when it is assumed that the stretching is uniform, the segment density drops with 1/r2 . In this case the mixing term is given by (^W —
1 1 = -xmv
M/1Q]
4.4.3
c
kT 2 l + D/Rc where
[4.4.4,
More recently Persson et al.2) have proposed complementary equations derived from the conventional Flory-Huggins theory for a homogeneously distributed corona region composed of ethylene oxide layers, which account for the contact energy and the mixing energies of core-corona and corona-solvent contributions. When the volume fraction in the head group region is expected to drop with 1/r2 , we may use Flory's31 rubber elasticity theory to estimate the deformation free energy of a chain
kT
2[xl2
D
\
We are not going to review the performance of these phenomenological equations here. For this we refer to an extensive review given by Nagarajan41. In general it will be clear that the SCF results are significantly more in line with experimental findings than the analytical approximations. The fundamental reason for this is rather obvious: the EO head groups are typically too short to allow polymer theories to become accurate and the profiles of the corona cannot be accounted for with sufficient accuracy. Moreover, the overlap between core and corona is, at least for the short EO surfactants, important. The latter effects may be accounted for by a renormalization of the interfacial Gibbs energy y such that the finite fraction of EO groups in the water phase is accounted for.
11
A.N. Semenov, Soviet Phys. JETP 61 (1985) 733. CM. Persson, U.R.M. Kjellin, and J.C. Eriksson, Langmuir 19 (2003) 8152. 31 P.J. Flory, Principles of Polymer Chemistry, Cornell Univesity Press (1953). 4 R. Nagarajan, Theory of Micelle Formation: Quantitative Approach to Predicting Micellar Properties from Surfactant Molecular Structure in: Structure-Performance Relationships in Surfactants, K. Esumi and M. Ucno, Eds., Surf. Sci. Series 70 (1997) 1. 21
ASSOCIATION COLLOIDS
4.61
4.4e Pluronic micelles Not all surfactant systems are based on alkyl chains as the hydrophobic fragment. It is of considerable interest to briefly discuss surfactants that have the propylene oxide PO as the hydrophobic unit. PO is composed of two CH2 s one O and a side-group CH3 . Here we refrain from the side-chain complexity and model the PO unit as (CJgtCHgljfOlj. There is an interesting family of commercial surfactants, the so-called Pluronics, which are tri-block copolymers that have the PPO as a central block and two PEO blocks as the hydrophilic heads. Here we will present the SCF predictions for Pluronic type surfactants. It is of interest to mention that the parameter set used for the alkyl-based non-ionic surfactants (table 4.1) can be used directly to model the Pluronic systems. Below we consider P84, which is (EO)19 (PO)43 (EO)19 . For this surfactant, which according to the literature has an HLB = 14, a c.m.t. (of a 1 % solution) of 28.5°C and a c.p.t. of 74°C1), we will present structural information on the micelles and investigate what is expected upon changing the temperature. Little is known in the literature about the internal structure of Pluronic surfactant micelles. As the PO is not a pure alkyl moiety, the amount of solvent accumulated in the core is unknown. All-atom MD simulations on these polymeric micelles are rather expensive and we do not know of any attempts in this direction. However, SCF predictions for the overall segment volume fraction profiles can be made, and results are shown in fig. 4.15a. In this graph, we present a micelle at the theoretical c.m.c. for which the micelle has an aggregation number of approximately 50. Not unexpectedly the overall structure of the micelle is very similar to that of the shorter alkyl chain surfactants. Because of the polymeric nature of the surfactant, all length scales have increased somewhat. The core has a homogeneous density of approximately 0.8. This includes the CH2 and CH3 and the O units of the PO moiety. Again there is a wide
Figure 4.15. a) Radial volume fraction profiles for the hydrophobic (PPO) and hydrophilic (PEO) moieties of the P84 pluronic surfactant, the water and the free volume profiles, b) The radial grand potential density for the micelle at the theoretical c.m.c. ( g = 49 ). 11
P. Alexandridis, T.A. Hatton, Colloids Surf. A 96 (1995) 1; V.G. de Bruijn, L.J.P. van den Broeke, F.A.M. Leermakers, and J.T.F. Keurentjes Langmuir 18 (2002) 10467.
4.62
ASSOCIATION COLLOIDS
overlap region of core and corona and the head group profile has a characteristic shape very similar to that found for the short-chain non-ionics. The water penetrates significantly into the core. This is also not unexpected because of the O unit in the PO, which attracts water to some extent. One may argue that as the core is significantly larger, the grand potential density has developed such that one would be able to find some signature of an isotropic (Laplace) pressure. This is not the case as shown in fig. 4.12. It is of interest to see that the grand potential density is a scaled-up version of the grand potential density of the non-ionic micelle at the c.m.c. given in fig. 4.12. There is a positive value on the outside where the hydrated head groups feel a positive pressure. The grand potential density is negative inside the micelle. Although the segment density in the core is rather homogeneous, the chain conformations in the micelle are not isotropic, they are stretched because of the "anchoring" on the EO groups in the water phase. As a result, the local pressure difference p N — p T is not homogeneous in the core.
Figure 4.16. a) Grand potential as a function of the aggregation number for P84 Pluronic surfactants for various values of the Zow a s indicated. Note that the line parts with a positive slope correspond to unstable micelles, b) The ^QW " e ^ ordinate) and c.m.t. (K) as a function of the surfactant bulk volume fraction cp° for P84.
(i) C.m.t. andc.p.t As both PO and EO solubilities in water depend on the temperature, it is no surprise that the temperature is one of the main control parameters for non-ionics. De Bruijn et al.11 showed that the SCF model can be used to predict the c.p.t. and the c.m.t. for these surfactants in a very simple way. These authors found an excellent correlation for these quantities between experimental findings and the theoretical predictions, assuming that /ow is the only FH parameter that is temperature-dependent. Of course this is a simplification because all interaction parameters depend on the temperature. However, Zcw a P P e a r s to be a very weak function of the temperature, at
" V.G. dc Bruijn, L.J.P. van den Broeke, F.A.M. Lccrmakers, and J.T.F. Kcurcntjes Langmuir 18 (2002) 10467.
ASSOCIATION COLLOIDS
4.63
least around room temperature. We may also assume that this is the case for the CH3 water contacts. As it is not known how the various parameters that control the free volume depend on the temperature, we keep these parameters fixed here as well. Hence, we propose the following relation between XQSN a n c ' t n e temperature T *OW = " 0 . 6 ^
[4.4.6]
This equation shows that with increasing temperature J O w becomes less negative. This renders both PO and EO less water-soluble. In general, the temperature dependence of the FH parameter can be written as % = A + B/T,
where the A part
refers to the entropic and B to the enthalpic part. Using [4.4.6], therefore, means that we assume here that for all interactions, the FH parameter is temperature independent, i.e. that B = 0 except for Xow f° r which the value of A is zero. This is of course far from realistic. In fig. 4.16a we present predictions for the grand potential as a function of the aggregation number for a series of ^Ow
vames m
the range -0.4 to -1.0. There are
large effects for relatively small changes of the Xow parameter. This means that the micellization is a strong function of this parameter. For %ow > - 0 . 6 , the surfactant becomes gradually less soluble. The region of £m(g) with a negative slope becomes smaller and smaller and eventually for Xow > ~ 0 - 4 vanishes. The absence of the stable micellar region effectively means that the solution cannot maintain free floating spherical micelles. The situation is very similar to the situation discussed for C 14 E 4 , where the stability region was also much narrower than for surfactants with a larger head group. Here we will formally identify the cloud point by the temperature [ Xow^ for which no stable micelles are present. In this case the cloud point is close to %ow > ~0-4 • We realize, however, that already for more negative values of ;fOw • * n e maximum value of the micelle concentration is sufficiently low so that the micelles will remain experimentally undetected. In practice, the c.p.t. is therefore not as high as predicted here. De Bruijn et a!.1' examined a large set of Pluronic molecules and proved that there is a strong correlation between the c.p.t. identified by the loss of stability of the spherical micelles and the experimental data. Experimentally it is known that for non-ionics the c.m.c. is a strong function of the temperature 2 . In an experiment where the surfactant concentration is fixed and the temperature is changed, one may find a micellization transition at a temperature known as the c.m.t. In fig. 4.16b we present the predictions of the c.m.t. as a function of the given bulk concentration of the surfactant P84 using the ad hoc relation [4.4.6] between Xow
anc
'
tne
temperature T . In line with experimental observations, it is
found that that the c.m.t. is a strong function of the bulk concentration, albeit our 11 De Bruijn ct al., loc. cit. 2 M. Tanaka. S. Kaneshina, G. Sugihara, N. Nishikino, and Y. Murata', Solution Behaviour of Surfactants. Theoretical and Applied Aspects 1 (1982) 41.
4.64
ASSOCIATION COLLOIDS
temperature scale is obviously not accurate. Again, de Bruijn et al.1' proved a very good correlation between experimental data for the c.m.t. (for a 1% solution) and the theoretical predictions for a large class of Pluronic surfactants. In conclusion, we expect that the first-order effects of the temperature are reasonably well accounted for in this simple model. 4.5 SCF for (spherical) ionic micelles An important class of surfactants is that of the alkyl-based, single-chain, ionic amphiphiles. Spherical micelles composed of these surfactants are centrally placed in this section. Basically we will take the same approach as in the previous section. We will first discuss the SCF predictions and thereafter summarize the pertaining aspects of the semi-macroscopic approaches. SCF analyses of micelles composed of charged surfactants give a unique opportunity to include electric double layers. We not only study the diffuse part of the double layer, but also account for the mechanism by which the surface charge is generated (i.e. the self-assembly process) and for specific interactions with counterions. The surfactant charge and diffuse double layer in ionic micelles have many similarities with the electric double layer of ionic surfactants densely packed at the air-water interface [see III.fig. 3.17, p.3.52]. In the SCF model, the conformational statistics and the short-range interactions are already accounted for. The extension to account for the electrostatic interactions involves the integration of the formalism with the Poisson equation, which allows the evaluation of the electrostatic potentials resulting from given charge distributions. In order to compute the distribution of the (charged) molecules, one should generalize the segment potentials by including the appropriate electrostatic terms. We again refer to appendix 1 for the details and limit ourselves to the most important issues here. The idea is to borrow as much as we can in terms of molecular characteristics from the surfactant models discussed above. In fig. 4.17 we give a schematic representation of the molecules that we shall consider in the modelling of the ionic surfactant systems. We will use the same water model as for the non-ionics, including the free volume component. In addition, there are small cations and anions where it is understood that the theory can be developed in the absence and presence of extra electrolytes. We take the ionic surfactant as the alkyl chain with a charged head group. In fig. 4.17 we give the cationic surfactant CTA, which has a C16 tail and a positively charged group surrounded by three CH3 units and the anionic surfactant DS, which has a C12 with a sulphate group modelled here as a unit that occupies five sites all having a charge of v = -0.2. These molecules mimick the paradigms of surfactant science CTAB and NaDS. 11
De Bruijn et al., toe. cit.
ASSOCIATION COLLOIDS
4.65
Figure 4.17. Molecular building bricks of the molecules involved in the SCF modelling of ionic surfactants. CTA = cetyltrimetylammonium (cation). DS = dodecylsulphate (anion), a cation composed of 5 volume units, a similar anion and a water unit with the same size.
We evaluate the electrostatic potential in and around the charged micelles from the charge distribution using the Poisson equation. In order to accomplish this, the dielectric permittivity profile in the system must be known. In the core of the micelle the dielectric permittivity is low, but in the water phase the relative dielectric constant should be close to 80. To account for the gradients in the dielectric permittivity, we use the local density average, i.e. f £
o A^A' r '
£0£{r)=
[4.5.1]
A
where the index A runs over all segment types at position r, and £A is the relative dielectric constant of the phase composed of pure component A and eQ is the permittivity of vacuum. In the following calculations, we take all relative dielectric constants to be 80 except for the V units for which we choose unity (vacuum) and except for ec = £CH = 2 . It is typical for this kind of analysis that the concentrations of dissociating substances (NaBr, NaDS, CTAB) are counted as electroneutral volume fractions. This is in line with the thermodynamic tenet that the solution and a micelle (including the counter charge) are electroneutral. The conversion from volume fractions to concenTable 4.2. The Flory-Huggins interaction parameters used in the SCF analysis of the ionic surfactant systems at T = 300 K. S is a unit of the sulphate head group, N is the central unit in the trimethyl ammonium group.
c3
C
S
N
cation
anion
W
V
0
0.5
2
2
2
2
1.5
1.5
c s
0.5
0
2
2
2
2
1.1
2
2
2
0
0
0
0
0.5
2.5
N
2
2
0
0
0
0
0.5
2.5
cation
2
2
0
0
0
0
0
2.5
anion
2
2
0
0
0
0
0
2.5
W
1.5
1.1
0.5
0.5
0
0
0
2.5
V
1.5
2
2.5
2.5
2.5
2.5
2.5
0
X
4.66
ASSOCIATION COLLOIDS
trations depends on the molar volume. A conversion factor of 10 for the small ions is rather accurate, I.e. (ps = 1 corresponds to 10 Molar. As compared with the modelling of the non-ionic systems, we now have to consider several additional molecular species. Except where mentioned otherwise, the model encompasses the following features. The ions are modelled as the water molecules, but with a central unit that carries either a positive or a negative charge (c.f. fig. 4.17). The Flory-Huggins parameters for all molecular species are collected in table 4.2. Here we specifically have chosen to model the cationic and the ionic species to be structurally identical and completely symmetric in respect to the short-range interactions. So, in this approximation specific differences between cationics and (structurally identical) anionics are ignored. So, the main difference between the two surfactants is the chain length; secondary differences, such as those mentioned in sec. 4.1c (iii), trend (viii), are ignored, but differences in hydrophilicity of the head groups are accounted for. Typically we assigned a large repulsion to the interactions between the charged species with the hydrocarbon units C and CH3 , even more repulsion with the free volume V and athermal interactions with water. We know that ion-specific effects do matter in micellization and briefly touch upon these effects in sec. 4.5d. The short-range interactions felt by the head group units (sulphate and N in the trimethylammonium) are of course much less important than the effects of their charges. We give these groups a weak repulsion with water, an athermal interaction with the small ions and a significant repulsion with the hydrocarbon units. 4.5a Micelles at the c.m.c. in various salt solutions We will start our analysis by comparing the thermodynamic data for the two mentioned surfactant systems over a large range of ionic strengths. As before, we consider em . In fig. 4.18a we present a set of curves for the anionic NaDS and in fig. 4.18b similar data for the CTAB. With respect to the excluded volume characteristics, we have modelled both surfactants very similarly (c.f. fig. 4.17). The CTAB surfactant has three methyl groups surrounding the positive charge and is therefore more hydrophobic than the NaDS. We have modelled the sulphate by five identical units, all with a reasonable solvency and a charge of -0.2 elementary units. From fig. 4.18a, one can observe a gradual transition regarding the ionic strength contribution of the added electrolyte. For relatively low electrolyte concentration, i.e. cps < 0.002 , the micellar properties are independent of the added salt. At higher ionic strength, there is a significant effect on the micellization of the amount of added salt. This reflects the fact that the ionic strength in solution is for an increasing extent determined by the added electrolyte. When this concentration becomes high, increasing screening of the double layer allows the micelle to grow in size, in the example by up to a factor of two. For the more hydrophobic CTAB surfactant, the same is found at lower ionic strength. However, as shown in fig. 4.18b, it is predicted that the stability region for
ASSOCIATION COLLOIDS
4.67
Figure 4.18. Standard state grand potential e m as a function of the aggregation number for various values of the salt concentration of a 1:1 electrolyte for a) NaDS and b) CTAB surfactant system. To convert the volume fractions of ions to a molar concentration, one needs a conversion factor of 10, i.e. the molar concentration is 10 times the volume fraction of the ions.
micelles (i.e. dem{g)/dg <0) is gradually lost with increasing ionic strength. For tps > 0.002, there appears a minimum in £m{g) and the value of this minimum shifts to higher £m(g) with increasing ionic strength. At cps > 0.05 the stability is lost and the surfactant does not dissolve anymore (Krafft point). This behaviour mirrors the clouding of non-ionic surfactants. Referring to 14.2.24] we know that the slope of e(g) is related to the fluctuations in the aggregation number. The steeper this curve, the more narrow the relative size fluctuation. The width of the size distribution has many features in common with that of the non-ionic surfactants. Near the c.m.c, there are relatively large size fluctuations at surfactant concentrations near some experimental c.m.c, i.e. where the surfactant monomer volume fraction is approximately equal to the volume fraction of micelles. The fluctuations are relatively independent of the surfactant concentration and almost independent of the average aggregation number. These fluctuations are found smaller for the ionic micelles than for the non-ionics. Inspecting fig. 4.18a, we notice that for NaDS the fluctuation in aggregation number is a weak function of the ionic strength in the solution. In line with experimental findings, we predict a I g-0.2 for low salt and a /g ~ 0.15 for the high salt concentration. The same applies for CTAB in the regime of relatively low ionic strength. However, when the ionic strength is increased to values near the Krafft point, a strong increase in the size fluctuations is observed. The volume fraction of freely dispersed surfactants that coexist with the smallest stable micelles is the (theoretical) c.m.c. In fig. 4.19a we present the c.m.c.s for NaDS and CTAB as a function of the ionic strength. In line with the results discussed above, we see that at very low ionic strength the c.m.c. is a very weak function of the ionic strength. At relatively high ionic strength, there is a power law-like dependence on the ionic strength, i.e. c.m.c. »= [
4.68
ASSOCIATION COLLOIDS
Figure 4.19. a) The theoretical c.m.c. in volume fraction units as a function of the (bulk) volume fraction of 1:1 electrolyte for NaDS and CTAB surfactant systems, b) The aggregation number at the c.m.c. as a function of the volume fraction of 1:1 salt for NaDS and CTAB.
range solvatlon; Its size etc.) starts to determine the repulsion between head groups rather than the electrostatic repulsion. Such leveling off of the electrostatic contribution is not (yet) seen in the presented systems. The dependence of the c.m.c. on the ionic strength of NaDS is similar to that of CTAB except that for the latter spherical micelles cannot be found at salt concentrations in excess of cps = 0.05; the CTAB curve stops at this salt strength value. Apparently CTAB micelles need a sufficiently strong electrostatic repulsion between the head groups to have a sufficient stopping force for micellization; the three methyl groups in the head group give it a hydrophobic character. In this case, the scaling exponent of the ionic strength dependence of the c.m.c. is slightly higher than in the NaDS case. For CTAB we find an exponent near -0.8. In experiments similar power law values are found11. The aggregation number at the theoretical c.m.c. also depends on the salt concentration. In fig. 4.19b we present the corresponding aggregation numbers in the ionic strength regime of 10~3 <
C. Tanford, J. Phys. Chem. 78 (1974) 2461.
ASSOCIATION COLLOIDS
4.69
4.5b Radial profiles In fig. 4.20a we present overall radial density profiles for the spherical micelle at the c.m.c. of NaDS surfactants at the added salt volume fraction of tps = 0.01 (equivalent to 0.1 molar). Radial distributions are presented for the apolar tails, the polar sulphate group, the water and the free volume. The apolar core is virtually free of water and the free volume enriches the core slightly more than it does bulk water. The head group profile is situated on the aqueous side of the micelle. There is a significant overlap of the tail distribution with that of the head groups. The head groups are not positioned at an infinitely sharp plane. In fact, the distribution of the head groups is bell-shaped and extends over some five lattice layers. This means that the head group distribution is at least 1 nm wide, which is comparable with the radius of the core. The distribution of the ions is given in fig. 4.20b. As expected, the co-ions, labeled Cl, are expelled from the head group region, and even more so from the core. The counterion (Na) concentration is significantly increased in the head group region. Indeed, the counterion can easily penetrate between the head groups. The head group density in the corona remains very low. The latter is of course correlated to the relative large width of the head group distribution. In passing we note that the low concentrations of the co-ion (Cl) in and around the micelle imply the general phenomenon that micelles expel salt.
Figure 4.20. a) Radial volume fraction profiles for spherical NaDS micelles at the c.m.c. at volume fraction of (salt NaCl)
4.70
ASSOCIATION COLLOIDS
As a result of the penetration of ions into the head group region, only a fraction of the countercharge resides in the diffuse part of the double layer. In addition, no very high potentials are created. The radial charge density profile is presented in fig. 4.6c in combination with the radial electrostatic potential profile. There is an excess negative charge in the head group region and an excess of positive charge both in the core and in the bulk solution just outside the micelle. The electrostatic potential is negative with respect to the bulk solution; it is most negative inside the core and decreases exponentially outside the micelle, as expected for the diffuse part of the electric double layer. In fig. 4.20d we present yet one more example of the dimensionless radial grand potential density profile. We note that the non-local contributions to the grand potential density are equally distributed over the monomers to which the interactions refer. This choice is arbitrary and in line with the fact that the local pressure tensor components p N and p T in the micelle are not experimentally observable quantities. Nevertheless, we find a complex structure of the grand potential density, which we are not going to try to explain in any detail. Note that the local grand potential density in the core is negative. This means that the volume contribution to the formation of micelles is negative (i.e. favourable); clustering of tails is the driving force for micellization.
Figure 4.21. a) Radial volume fraction profiles for spherical CTAB micelles at the c.m.c. at volume fraction of salt q>s = 0.001 . Given are the tail, head, water and V components, b) The distribution of the anion and the cation, c) The radial electrostatic potential (left ordinatc) and charge density (right ordinatc) profiles. In the inset the electrostatic potential profile on the outside of the micelle is plotted on a logarithmic scale, d) The radial grand potential density profile.
ASSOCIATION COLLOIDS
4.71
In fig. 4.21 we present the radial profiles for the CTAB surfactant system at cps = 0.001. In first order all the results are very similar to those of NaDS, but there are significant, specific differences worth discussing. The spherical micelle at the c.m.c. CTAB is slightly larger than that of NaDS: the interface between core and corona is shifted slightly to larger r values. It is significant that the overlap of tails and head groups is larger for the CTAB micelles than for the NaDS and, therefore, the average head group position is approximately the same for both micelles. The overlap between heads and tails must be attributed to the more hydrophobic nature of the CTAB head group. This has corresponding consequences for the charge-potential relations. The electrostatic potential inside the micelle is significantly higher for the CTAB micelle (about 225 mV as compared with roughly -70 mV for the NaDS micelle). The main reason for this is that the ionic strength is ten times lower in the first case. The high electrostatic potentials in the corona of the CTAB micelles result from the fact that the counterions do not readily enter the interior of the micelle; there is only a modest increase of the counterions in the head group region in the CTAB micelles as compared with that in the NaDS ones. As a result, the excess charge in the head group region is significantly larger for the CTAB micelles (about 50%) as compared with that of the NaDS. We may compare the electrostatic potential just outside the micelle, i.e. at layer r = 10, and compare this with the £ -potential (potential at the shear plane). The absolute value of this potential is i// ~ 7 mV in 0.1 M salt (NaDS micelle) and a factor of 6.5 higher ( ^ = 45 mV) in 0.01 M salt (CTAB micelle). Ion specific effects can disturb this (idealized) prediction. Such effects are discussed below (sec. 4.5d) in more detail. The fact that not all counterions are in the diffuse part of the double layer, but that a significant fraction of the counterions can penetrate the corona and locally compensate the head group charge is familiar. This effect is not easily quantified from the SCF results. Before one can do this, one needs to define the head group region and specify some rule to judge those ions considered to be diffuse and those ions directly taking part in the charge compensation. To get a rough estimate, we evaluate the direct charge compensation at r = 6 at the maximum in the head group profile. At this coordinate, a fraction 0.46 of the charges of CTAB (0.01 M salt) is masked by counterions whereas this fraction grows to 0.67 for the NaDS case (0.1 M salt). Otherwise stated, the "free" fractions of the countercharges are / = 0.54 and / = 0.33 , in 0.01 and 0.1 M electrolyte, respectively. Again, these numbers are likely to increase when counterions are specifically bound to the head groups. In the MD-generated snapshot of an ionic micelle as given in fig. 4.1a, we can also see that many of the counterions have accumulated in the head group region. Quantifying this counterion binding via MD simulations is not trivial either; it is similarly difficult in SCF modelling. In passing, we point to the inset of fig. 4.20c where we present the logarithm of the electrostatic potential as a function of r in the diffuse part of the electric double layer. The linear slope proves that in this region the electrostatic potential profile drops
4.72
ASSOCIATION COLLOIDS
exponentially in full accordance with the Debye Hiickel prediction. The Debye length is approximately 1.6 nm (here we use the lattice site with length 0.3 nm). For reasons of completeness, we also present the grand potential density profile across the CTAB micelle. Without going into further details, we mention that the pressure profile in the head group region is very similar to that for NaDS micelles; there is a positive excess pressure in the head group region and a negative one in the tail region. 4.5c Chain length dependence Experimental evidence (see [4.1.1 ]) has shown that often log c.m.c. = A-Bt where the slope B is about 0.3 for ionics and 0.5-0.6 for non-ionics. SCF theory can account for this difference. To this end, in fig. 4.22 we present computed c.m.c.s for a series of alkyl trimethylammonium bromide surfactants for three values of background electrolyte concentration. From this figure it is seen that both at very high ionic strength, i.e.
Figure 4.22. The c.m.c. as a function of the tail length for trimethylammonium bromide surfactants for three values of the bulk volume fraction of added salt as indicated.
ASSOCIATION COLLOIDS
4.73
4.5d Specific ion effects Ion specificity can also be modelled and by way of illustration we shall now demonstrate how to account for variations in ionic solvation and radii. In practice these two trends are closely related, see 1.5.3d. In a model, one can separate these issues. (i) Counterion solvation We choose the CTAB system with (ps = 10~3 as our reference system. Referring to fig. 4.17, we have modelled the anion as a negatively charged unit surrounded by four units of W (mimicking water). The key idea here is to replace the W component by some new unit named Y and give the anion the chemical composition (Br)(Y) 4 . If we choose the interaction parameters %YA = _£WA for all segment types A, we return to the CTAB system. Let %Yc a n c ' %Y CH ' 3e *-ne on ty parameters that differ with respect to this set. These two parameters will control the change in Gibbs energy upon placing a unit Y near or in the hydrophobic core. We have seen above that the corona of the CTAB micelles is not too polar (the head groups are buried more or less in the interface between core and water) and in this way we expect to modifiy the efficiency by which the counterion can enter the core as well as the corona region. This, in turn, is expected to influence the micellization. As in the starting system, we set %Y CH - %YC = 0.4 for simplicity. This means that in this model the results are fully specified by the value of %xc • The thermodynamic data of the micellar systems collected for the CTA surfactant systems with modified counterion hydrophobicities are given in fig. 4.23a. All systems contain a volume fraction of a 1:1 electrolyte (with the same counterion and the default co-ion) of
Figure 4.23. a) The standard state grand potential £nl as a function of the aggregation number for cetyltrimcthylammonium surfactants with a counterion, of which the hydrophobicity varied as indicated. The volume fraction of the 1:1 electrolyte is tps = 10~3 and all other parameters arc the same as in the CTAB system, b) The c.m.c. in volume fraction units as a function of the hydrophobicity of the counterion.
4.74
ASSOCIATION COLLOIDS
more easily come near the core and partition into the corona. Consequently, the charge is more efficiently screened. It is therefore no surprise that the grand potential curves of the micelles develop similarly as with increasing ionic strength (cf. fig. 4.18b). The curve for %YC = 0.6 , is very near the Krafft point, and it resembles the starting system with the added salt volume fraction of cps = 0.05 . Apparently the overall hydrophilicity of the head group is the sum of that of the head and the corresponding counterion. With decreasing %YC, the hydrophilicity of the head group diminishes and the surfactant packing parameter P increases. As a result, the stability of the spherical micelles diminishes. In line with the discussion of these systems at elevated ionic strength, we anticipate that the more hydrophobic head groups will more likely promote the formation of cylindrical micelles. These effects are well known in the literature and we will return to these effects below. Above we have analyzed what happens when the counterion is made more hydrophobic. Similar physics may be observed when the counterion is explicitly attracted to the head group. This is another aspect of specific adsorption. Again, we should expect that the charges in the corona are more effectively screened and the effective hydrophilicity of the head group diminishes. (ii) Counterion size The second important ion-specific effect that can be sequestered in SCF modeling is the size of the counterion. We choose the NaDS system to investigate this. As mentioned above, the counterion in the original system occupies five sites and may be expressed as A4 = (Na)1(W)4 with the four W units surrounding the centrally charged unit. The key idea to investigate the effect of the size is to reduce the number of W units around the Na but retain the lattice size. Introducing A3 = (Na)j(W)3 , A2 = (Na)1(W)2 , Al = (Na)j(W)j and A0 = (Na)j, we obviously reduce the size of the counterion. We should realize from the start that by removing W units we are also going to change the overall hydrophilicity of the counterion. The smaller the ions, the more the ions will behave as point charges and the less they are influenced by the non-electrostatic, excluded volume effects. The standard state (translationally restricted) grand potential is plotted in fig. 4.24a as a function of the aggregation number g . The micelles formed in the presence of small counterions are typically larger than those in the presence of large counterions. Indeed, the smaller counterions can more easily penetrate into the head group region making the micellization more favourable. In fig. 4.24b it is shown that the c.m.c. decreases with the decreasing size of the counterion and that the aggregation numbers go up (as mentioned above). These changes are significant but modest. The results are in line with the experimentally observed dependencies of the c.m.c. for various counterions (see table 4.1). We add the caveat that in practical situations we must expect that by changing the counterion one invariably also changes more parameters at the same time and the experimental situation is obviously more complicated.
4.75
ASSOCIATION COLLOIDS
Figure 4.24. a) The standard state grand potential e m as a function of the aggregation number g or 5 different counterions A0, Al, A2, A3 and A4 as defined in the text (only the first and last one are indicated), b) The theoretical c.m.c. (in volume fraction units) (left ordinate) and the aggregation number at the c.m.c. (right ordinate) for various counterions A0, ..., A4 as indicated. The lines are to guide the eye. 4.5e Quasi-macroscopic approach to ionic micelles When anionic or cationic surfactants form micelles, one should account for the presence of the electric double layer(s) in the system. The analytical computation of these interactions is obviously complicated by a number of factors, such as discrete charge effects (which is beyond Poisson Boltzmann and SCF), specific ion effects (as discussed in sec. 4.5d), the exact size and shape of the surfactant head group and the dielectric permittivity in the corona. One invariably has to make drastic approximations. Typically one considers the head groups to lay in a plane with a Stern layer on top. Let the radius of the core be given by Rc and within a Stern approximation, let the positions of the first counterions be located a distance S from the core. The head group area a h is evaluated at the distance 8 from the core and will be expressed as as. There is an exact solution of the grand potential of the diffuse part of the flat electric double layer, see [II.3.5.20]. Evans and Ninham11 also give an accurate curvature correction so that the results can be applied to various geometries
^ ^ = 2[lnf- + Vl + (s/2) 2 ]--(Vl + (s/2}2-l) - ^ l n f I + iVl + (s/2)2] kT
[ (2
) s\
I
KS
(2
2
j
[4.5.2]
where s/2 is the dimensionless charge in the diffuse part of the double layer which in [II.3.5.14] was written as s/2 = - p o 4 , where s=
Ane2
[4.5.3]
eoeKaskT and K the reciprocal Debye length. The first two terms of [4.5.2] are appropriate for the flat electric double layer whereas the third term comprises the curvature correction. The parameter C must be chosen according to the geometry. It is C = 2/(Rc + S) 11
D.F. Evans. B.W. Ninham, J. Phys. Chem. 87 (1983) 50025.
4.76
ASSOCIATION COLLOIDS
for spherical geometry and C = 1/(RC + S] for cylindrically shaped surfaces. Note that [4.5.2] yields an electrostatic surface (pressure) contribution of approximately 2/cT/a charge (planar case)1'. We expect that the radial density profiles of the spherical micelles of the charged surfactant systems as given in figs 4.20 and 4.21 are realistic. However, using [4.5.2] to stand for the complete electrostatic contribution to i2£ implies a number of ad hoc simplifications. We have argued above that the success of any model lies in a matching balance of approximations. Taking the full charge density of the head groups sitting on the micelle surface obviously ignores the distribution of the head group charges, the local charge compensation of small ions that penetrate the corona region and effectively reduce the charge, the fact that the head group charges are in a relatively low dielectric permittivity (especially when the charge is very close to the core-water interface), etc. Moreover, it neglects the part of the electric double layer inside the core of the micelle. A micellar model, which makes use of [4.5.2], has 8 as an adjustable parameter, which may be used to change the surface charge density of the micelle to some extent. In such a way, one can "correct" for the fact that only a fraction / of the ions is in the diffuse part of the double layer and the fraction (1-/) is in between the head groups. All of this is somewhat artificial. For an illustration of a model in which Stern layers are explicitly accounted for, see ref. . 4.6 Linear growth of micelles With increasing overall surfactant concentration, the primary effect is that the number of micelles rises. A secondary effect is that the aggregation number increases. When the number of micelles per unit volume grows, the translational entropy assigned to each micelle invariably goes down. This means that em diminishes as well. Of course, in reality systems compensate for this by increasing the aggregation number. For sufficiently low values of em, it becomes relatively inexpensive (in terms of Gibbs energy) to allow for shape fluctuations. As a result, we might anticipate a transition from spherical geometry into elongated ones. When that occurs the strategy of the system has completely altered. The primary effect of raising surfactant concentration is now the growth of micelles rather than the increase in number of micelles in solution. It is often found that a system rather suddenly changes its mode of response and in experiments one can notice this change in the rather abrupt changes in measurable properties. The transition from spherical to elongated micelles takes place at a concentration that is known as the second c.m.c. Depending on the surfactant architecture and concentration, there are also limits for further expansion of the aggregation number and linear micelles will in turn give way to lamellar aggregates. We
11
V. Srinivasan, D. Blankschtein, Langmuir 19 (2003) 9932.
ASSOCIATION COLLOIDS
4.77
shall now discuss the basic elements of these transitions starting with the appearance of linear micelles.
4.6a Phenomenological model for rod-shaped micelles Consider a linear micelle with a cylindrical middle part of length Lc , which on both sides ends with a hemispherical cap. A central assumption in the modelling of linear micelles is that the excess grand potential of a surfactant in such a micelle can be expressed as a superposition of the excess grand potentials of the surfactant felt in the various microenvironments. A surfactant in the cylindrical part is in a different environment than those positioned in the ends and already from a chain-packing point of view we then expect that there are gradients in (Gibbs) free energy densities along the micelle. For linear micelles to compete favourably with spherical micelles, it is necessary that the excess grand potential per molecule for a cylindrical part i2j? j is lower than that of the molecules in the endcap .£§phere . The latter quantity was discussed in the previous sections. The cylindrical counterpart can be evaluated similarly. For example, the conformational entropy loss for the surfactant tail packed in a cylindrical geometry differs from that in a spherical geometry, and the Gibbs energy to develop an electrostatic double layer is a function of the geometry. At this point we assume that for a given chemical potential of the surfactant both £2^x, as well as •Gl?pilere . are known. We note that at equilibrium there are no gradients in the chemical potential of the surfactants. This means that the differences between Q®, and •fi°iiere will give rise to an excess Gibbs energy associated with the endcaps. For a spherocylindrical micelle containing g-m surfactant molecules in the cylindrical part and m 12 molecules in each hemispherical end, we have for the excess chemical potential per molecule in a spherocylindrical micelle with an aggregation number g, &!c{g) <&<9> = ^ y ^ y l ^ ^ p h e r e = < 1 + J(Sphere " < l )
I4-6-1!
for g > m . Note that for such micelles the value of the aggregation number g can go to infinity without a violation of the packing constraints. From [4.6.1] we can define the excess Gibbs energy associated with the presence of the two endcaps AGec= S p h e r e - < l )
I4'6'21
We may also consider the formation of rings. For a worm-like micelle longer than a few times its persistence length, it should be possible that the two ends of the same micelle merge forming a closed loop. The system will gain the endcap energy by doing so. However, the entropy loss associated with the formation of a ring is typically larger than this gain so that rings can safely be ignored. To get information on the length distribution of linear micelles, it is quite common
4.78
ASSOCIATION COLLOIDS
to set up a mass action model in which between all micelle sizes an equilibrium occurs where the aggregation number g grows by integer numbers11. It appears to us that in the first-order model without any loss of physical reality the "discretization length" can be chosen larger, i.e. the size of the spherical micelle. In such a coarse model, the linear micelles exist with discrete lengths L = 1,2
,°° , where the unit length Rs is the
size of a spherical micelle size that coexists with the linear ones. Let there be various equilibria between micelles of different sizes. In particular, we consider the equilibrium between micelles that differ by one length unit X
L+Xl^XL+l
[4.6.3]
for each value of L . In [4.6.3] the X refers to a species and the subindex refers to the length. When we consider the reaction constant for this reaction the true reason for the coarse way of discretization becomes clear. In this "reaction," the endcap energy of the Xj component is gained when the micelle grows from length L to L +1 . The reaction constant K is assumed to be independent of L and only a function of this endcap Gibbs energy K = exp
^
[4.6.4]
Now the volume fraction of a micelle with length L can be expressed in terms of that of the volume fraction of micelles with unit length. Realizing that the micelle with length L assumes L times the volume of that with L = 1 , we find the size distribution
^e-aL
[4.6.5]
where -a = lnfK^), in line with more detailed treatments11. This size distribution in terms of the length of the worm-like micelles may be converted into the corresponding distribution as a function of the aggregation number. In first order one assumes that the aggregation number is proportional to the length L . The proportionality constant is the number of surfactants per unit length gL . The size distribution is for very small L linearly increasing with L (and thus with g ) and for large L (large g ) exponentially decaying. Such an exponential size distribution is of course much wider than the size distribution found for spherical micelles (in first approximation Gaussian). The value I / a may be interpreted as the (dimensionless) most probable length (in units of size Rs ) of the micelle. At fixed value of K, we see that the most likely micelle size increases with an increasing concentration of the spherical micelles with which they coexist. Similarly, at a fixed concentration of spherical micelles the size of the wormlike micelles increases when the endcap energy increases. Note that in this model the product Kip^ will always be smaller than unity.
11
J.C. Eriksson, S. Ljunggren, Langmuir 6 (1990) 895.
ASSOCIATION COLLOIDS
4.79
The total volume fraction of micelles is found by
Pm = P l + 2 K t f + 3 K g g f + - =
( i
_^)2
[4.6.6]
This equation may be inverted to find the volume fraction of the micelles with unit length as a function of the overall volume fraction of micelles 1 l-j4K
[4.6.7]
It is easily shown that the average size is given by L (L) = L~l = ^ ^ 11=1 ^L ^ ^ l
[4.6.8]
In the limit of rather low values of the endcap energy or very low volume fractions of micelles with unit size (spherical micelles), we see that the average size approaches unity (meaning that only spherical micelles are present). However, when K
4.80
ASSOCIATION COLLOIDS
like micelles may be relatively unstable. We will use these results to argue that the second c.m.c. is a relatively sharp phenomenon. 4.6b SCF theory of infinitely long linear micelles The first step in the SCF analysis of worm-like micelles is to consider the limit of very long micelles, i.e. L -> <x>. In this case the ends can be ignored. In such a system it is natural to probe all micellar properties per unit length. The unit in which the lengths of the micelles are counted is here the discretization length in the calculations, which is length of a lattice site I . Hence, the task is to determine the structural and thermodynamic properties of these micelles per unit length. We position a micelle with its centre at the axis of a cylindrical co-ordinate system. By doing so, we ignore both the conformational degrees of freedom of the worm-like micelle as well as the translational ones. When the persistence length is sufficiently large (we are going to determine this in sec. 4.6d), it is not a serious approximation to ignore the flexibility of the worm-like micelle. In first order we may also ignore the translational entropy. The linear micelle as a whole may have a finite amount of translational entropy (typically of order 10 k), but per unit length this amount of entropy is negligible (especially considering the limit of L —> <» ). The only relevant (infinitely long) cylindrical micelle is the one with vanishing standard state (translationally restricted) grand potential per unit length, e^ = 0. The stability constraint de^/dg < 0 still applies. In these computations, the aggregation number g = gc corresponds to the number of surfactants per unit length in the cylindrical micelle. For the details related to the model, we refer once again to appendix 1. Let us revisit the C n E x surfactant system using the same parameter setting as in fig. 4.9. In fig. 4.25a ej^ is presented as a function of the aggregation number g per unit length. Physically the (excess) grand potential per unit length e^ is the work needed to
Figure 4.25. a) Standard state grand potential e m for a cylindrical micelle, normalized per unit length e^ as a function of the number of surfactants g per unit length, for three nonionic surfactants (n,x) = (10,6), (12.5) and (14,4) as indicated, b) The volume fraction of free monomers as a function of the number of surfactants per unit length for the non-ionic surfactants as indicated. The open circles represent the cylindrical micelles for which e m vanishes.
ASSOCIATION COLLOIDS
4.81
lengthen the micelle at fixed chemical potentials. Let us assume that we can impose a value of £^ by forcing some length, e.g. by holding it to its ends. A s^ > 0 implies that it is not favourable to have long micelles in the system; when the constraints on the end are released, the length will spontaneously decrease (positive tension). Inversely, when £ rt\ < 0 i* i s f a v o u r a ble for the system to have linear micelles and the length will increase upon the removal of the constraints (negative tension). Only when e^ = 0 does the length of the micelle not change upon the removal of constraints. We describe these (equilibrium) micelles as tensionless. It is seen that both the (10,6) and (12,5) surfactants can reach the tensionless state of the infinitely long cylindrical micelle. Such a tensionless state appears to be impossible for the (14,4) surfactant. The (14,4) surfactant is the one with the smallest head group and the largest surfactant packing parameter. Hence, for this surfactant e^ remains positive for all values of g . We do not expect cylindrical micelles for this surfactant. We also did not find significant spherical micelles for these surfactants either. Below we will show that for the (14,4) surfactant, the lamellar state is the stable one. In fig. 4.25b we present the logarithm of the volume fraction of the monomers that coexist with the cylindrical micelles with given aggregation number per unit length g. As in the spherical micellar systems, the chemical potential of the surfactant is an increasing function of the aggregation number. For the cylindrical micelles, however, there is just one relevant (equilibrium) point, viz., the cylindrical micelle must be in the tensionless state. We have marked these points as open circles on the (10,6) and (12,5) curves in fig. 4.25b. The natural question presents itself, whether tensionless cylindrical micelles can coexist with the spherical micelles. The necessary requirement for coexistence is that the two types of micelles coexist at the same chemical potential of the surfactant. The corresponding data for the spherical micelle have been shown and discussed above in fig 4.9b. It is easily seen that for the (10,6) surfactant the spherical micelle is more stable than the cylindrical ones for all the stable spherical micelle sizes, so (10,6) surfactants associate into spherical micelles up to very high surfactant concentration. The situation is different for the (12,5) surfactant. As indicated by the open circle in fig 4.9b, the spherical micelles with g = 125 exist at the same chemical potential as the tensionless cylindrical micelles. The Em of these spherical micelles is em =8.8 kT. The volume fraction of micelles that coexists with the cylindrical micelles is therefore (»m=exp(-8.8) = 10- 3 . We can now estimate the endcap energy of spherocylindrlcal micelles of C 12 E 5 surfactants. We have seen above that exp(AGec / kT)
4.82
ASSOCIATION COLLOIDS
Figure 4.26. a) The radial density profile of the C 12 E 5 surfactant in a tensionless cylindrical micelle, b) The radial density profile of a corresponding spherical micelle (e m = 8.8fcT ; aggregation number is g = 125 ) at the chemical potential of the surfactant ((£>surf = 0.00088609) as in (a). The overall volume fraction (head plus tails), the individual parts (head, tail), the water and the free volume profiles are given.
across the tensionless cylindrical micelle are given in fig. 4.26a. This micelle contains just over 6 surfactant molecules per unit length. The bulk concentration of surfactants is given by
ASSOCIATION COLLOIDS
4.83
4.6c The endcap energy The second step in the modelling of cylindrical micelles is to consider finite size micelles explicitly. This is possible in a two-gradient SCF analysis. Referring again to appendix 1, we can design a cylindrical lattice coordinate system in which gradients are allowed in the radial and longitudinal direction. We position the cylindrical micelle such that the central axis of the coordinate system coincides with the long axis of the micelle. The number of surfactant molecules determine the length of the central cylindrical micelle in the system. Observables are again the density profiles and the grand potential of such a micelle. As the micelle is fixed to "live" on the axis of the coordinate system, we once again ignore the translational and conformational degrees of freedom of the micelle.
Figure 4.27. A two-dimensional cross-section of the overall surfactant volume fraction through a spherocylindrical micelle composed of g = 575 C^Es molecules. Black is high density; white is low density. The bulk concentration of the surfactant is extremely close to that of the tensionless, infinitely long cylindrical micelles p s u r f = 0.00088609. The computation box is 150 lattice units long and 40 sites wide (one unit = 8C ). A typical example of a spherocylindrical micelle composed of C 1 2 E 5 surfactants is given in fig. 4.27. Here we present a cross-section through the rod-like micelle in such a way that the long axis is positioned in the horizontal direction and the radial direction is in the y -axis. The gray scale is chosen to be black for the highest density and white for the lowest one. The sum of the volume fractions of head groups and tail segments (i.e. the overall surfactant volume fraction profile) is given. As is seen from this example, the endcaps are larger in radius than the middle section. This observation is in line with expectation and was first predicted by Eriksson et al.11. We can make cross-sections through the middle part of the cylindrical micelle and find a result very similar to the profile given in fig. 4.26a. A cross-section in the part where the endcaps have the maximum size in the radial direction gives a volume fraction profile very similar to that in fig. 4.26b. It is further relevant that the volume fraction of surfactants in the bulk that is in equilibrium with the short rod of fig. 4.27 is extremely
11
J.C. Eriksson, S. Ljunggren Langmuir 6 (1990) 895; J.C. Erikson, S. Ljunggrcn, and U. Henriksson, J. Chem. Soc. Faraday Trans 2, 81 (1985) 833.
4.84
ASSOCIATION COLLOIDS
Figure 4.28. The standard state grand potential of the rod-like micelle e^ as a function of the aggregation number for the C[2E5 surfactant. The limiting value of e m f°r l ar 6 e values of g is the endcap energy. Considering that every six surfactants add one lattice site to its length allows conversion from the aggregation number to the length. close to that for infinitely long cylindrical micelles. In line with this, one can increase the length of the rod-like micelle by one lattice site for every 6 surfactant molecules added. This means that the two endcaps are sufficiently far remote from each other so that they do not "feel" each other and more. We conclude that the picture sketched above that the micelle might be considered as being composed of a central homogeneous middle part with two spherical caps at the ends is useful. There are, however, some intricacies. Upon very close inspection one may notice that the middle part does not smoothly increase its size up to the end caps. In between the cylindrical body and the end caps there is a neck, i.e. a region with a reduced thickness. This shape modulation results from the conformational properties and the packing constraints of the molecules in the local environments. The physical consequence of the neck is extracted from fig. 4.28 where we show the grand potential as a function of the aggregation number ej^fg). For very small, dumbbell-like micelles the two necks may overlap giving rise to a minimum for e^ [g]. This occurs near g ~ 300 . Micelles for which dem /dg>0 (for 200 < g < 300 ) are unstable. Physically this implies that the size distribution of the rod-like micelles is not smooth and that micelles in the range 200 < g < 300 are rare. This gives food to the abrupt nature of the second c.m.c. We note once again that for the stable micelles the slope -3eJ^(g)/3g is a measure of the polydispersity. Indeed, for sufficiently large g -values the slope becomes zero indicating a very large polydispersity. We saw above that the conservation of surfactants leads to an exponential size distribution. We may further extract the endcap energy from fig. 4.28. We notice that in the limit of very large g the standard state grand potential goes to a well-defined limit of approximately 10 kT . The fact that this value becomes independent of the aggregation number implies that the middle part of the micelle is tensionless and that the limiting value must be interpreted as the endcap energy. To elaborate on this, we may first assume that the middle part has some line tension r and the endcap energy is AGec . The grand potential of the micelle is then e^ = r L + AGec , where L is the length. For a fixed number of surfactants in the micelle, the grand potential can be minimized with respect to its length and d£^n/dL = r = 0 . This thus means that the micelle has optimized its length when the middle part is tensionless.
ASSOCIATION COLLOIDS
4.85
Above we have made a crude guess of the endcap energy based on the concentration of spherical micelles that coexist with the infinitely long linear micelles (8.8 kT ). The more accurate two-gradient analysis gives a result that is only 1.2 kT higher. This trend is found in other surfactant systems as well. In all the systems we have studied to date, we have seen that the "true" endcap energy is slightly higher than that found from the crude guess based on one-gradient SCF calculations. This means that the crude guess has a predictive power. We will use this observation in sec. 4.6e, as well as in the applications section. 4.6d Persistence
length of wormlike
micelles
The third step in the modelling of linear micelles is to consider the persistence length £, which is a measure for the stiffness of a wormlike micelle. It is expected that the stiffness of the micelles increases with the core size and is less sensitive to the dimension of the corona. The analysis of the persistence length starts with the observation that long linear micelles are tensionless, i.e. per unit length, e m = 0 . Above it was argued that in practical situations closed uniaxial micelles are not expected to occur frequently. Closed uniaxial micelles are however very useful theoretical devices for the analysis of the persistence length. The procedure to find the persistence length is to homogeneously curve a linear micelle to form a torus. A torus has no ends and therefore the value of em cannot contain the endcap energy. The value of e m will not be zero for homogeneously curved cylindrical micelles because of the curvature energy stored in the micelle. Let us analyze the curvature energy per unit length stored in a weakly curved cylindrical micelle. We denote the grand potential per unit length as r and let the radius of the torus be R . For large R , the curvature J = 1 / R is small and therefore we may use a Taylor series expansion of r in the small parameter J
Here r(0) = 0, a linear micelle that is not curved and is infinitely long with no excess Gibbs energy per unit length. We did not yet discuss the sign of J . From the point of view of a long linear micelle, one may call bending to the right as positive and bending to the left as negative. However, this difference is arbitrary and thus the bending energy should not depend on the sign of J . As a consequence, the linear term in J (and all odd terms of the expansion of [4.6.9]) should be zero. The first non-zero term is the quadratic term J 2 . The next higher order term will be proportional to 1 / J 4 , which we may safely ignore. Only for strongly curved micelles should we consider these higher order terms. Let us introduce the stiffness parameter, ks = O 2 z73J 2 ) J=o . Then we can write for sufficiently small curvatures r(J) = ^ksJ2 . Note that ks has the dimensions [kTl], i.e. energy times length, and thus ks /{kT) has the dimension of a length. In the
4.86
ASSOCIATION COLLOIDS
Figure 4.29. The bending coefficient ks as a function of the curvature J = 1 / R of the torus composed of C ^ E g surfactants. Each symbol represents an SCF result. Lattice artefacts cause the noise in the data.
literature this ratio is identified as the persistence length £ . To show that the persistence length is linked to ks /{kT), we take a linear fragment of the cylindrical micelle with length L , curve it into a torus with radius R = L /(2/r) and compute the curvature energy stored into this piece. We find that TL = nks I R = 2n2ks /L . The persistence length t, may be defined as the length of the cylinder L for which the Gibbs energy of bending into a torus is just given by 1 kT . This means that £ = 2n2ks /(kT). Hence, in this definition the persistence length is In2 larger than our previous identification. We will keep the convention from the literature, i.e. Z=ka/kT. In fig. 4.29 we show SCF results in which ks = em[R/ x) is plotted as a function of the curvature J = 1 / R of the torus. The radius is found from the length of the torus R = Ll(2n], where in turn the length is found from the excess number of surfactants in the torus divided by the number of surfactants in the cylindrical rod per unit length. In the limit of J —> 0 we expect that the expansion of the line tension t of [4.6.9] up to the term proportional to J 2 is exact, which means that ks will become independent of J . From this figure, we see that in the curvature range given this independence is nearly reached. The very weak increase with J is due to higher order terms, which are not completely negligible yet. The value of the persistence length is g=ks/ kT ~ 38, which corresponds to approximately £ = 10 nm in real dimensions. Lauw et al.11 have analyzed a series of non-ionic surfactants and found that £ is a strong function of the tail length £«=£/? where fi ~ 2.5 and is almost independent of the head group size. The latter finding was rationalized by the fact that with the increasing head group size the overall dimensions of the micelle do not change much as the core size diminishes with increasing size of the corona. 4.6e Second c.m.c.for ionic surfactants Above we have seen that the concentration of surfactants at which the cylindrical micelles becomes stable, i.e. the second c.m.c, is strongly correlated to the endcap 11
Y. Lauw, F.A.M. Lcermakers, and M.A. Cohen Stuart, J. Phys. Chem. 107 (2003) 10912.
ASSOCIATION COLLOIDS
4.87
energy of the wormlike micelles AGec . More directly it was found that the em value of the spherical micelle that coexists with the cylindrical micelles is systematically some 20% below the endcap energy. This correlation is useful because the SCF computations to calculate the true endcap energy requires relatively long rod-like micelles and are time consuming. However, the computation of the second c.m.c. is extremely inexpensive in terms of CPU. It may be necessary to repeat the theoretical procedure to find the second c.m.c. This starts with the evaluation of the chemical potential (bulk concentration of free surfactants) of the surfactant system in the case of tensionless, infinitely long linear micelles. The second step is to use this chemical potential (bulk volume fraction of free surfactants) in the spherical coordinate system and identify the spherical micelle that exists at this chemical potential. Let us name the standard state grand potential of this micelle as em. Now the second c.m.c,
r f
is the bulk volume fraction of free monomers that is in equilibrium with
these micelles. In fig. 4.30a we present a typical result for an ionic surfactant system. With increasing ionic strength, one expects the charged head groups to be increasingly screened. As a result the surfactant packing parameter P tends to become larger and consequently the tendency to find cylindrical micelles increases. This trend is indeed observed for the NaDS system11. The curve with the label A4 is the NaDS system with the default 1:1 electrolyte. We see that for this system e*m > 0 at relatively high ionic strength. This means that we expect that cylindrical micelles are only found at these high ionic strength cases. In first order e*m is a linear function of the logarithm of the ionic strength. In fig. 4.30b we present the volume fraction of micelles at the second c.m.c. This volume fraction strongly decreases with ionic strength, indicating that above a given ionic strength the first micelles that will be observed experimentally will be cylindrical. As can be seen in fig. 4.30b we find an extremely strong scaling of the volume fraction of micelles at the second c.m.c. with the ionic strength cp*m <^(
R.G. Laughlin, The Aqueous Phase Behavior of Surfactants, Academic Press (1994).
4.88
ASSOCIATION COLLOIDS
Figure 4.30. a) The standard state grand potential em for spherical NaDS micelles that coexist with infinitely long tensionless cylindrical micelles as a function of the volume fraction of added salt, b) The corresponding volume fraction of spherical micelles at the second c.m.c. as a function of the ionic strength. The effect of the size of the counterion is shown for the A0, Al, A2, A3, and A4 (default) counterions as indicated.
course in line with expectations. More surprising is the magnitude of the effect. Going from the default ion A4 to the bare ion A0 gives a shift on the ionic strength axis of almost one decade in ionic strength. Knowing this relatively large effect, It is understandable that the importance of the type of counterion on the micellization did not remain unnoticed in the experimental literature. 4.7 Biaxial growth of micelles In the sequence of micellar objects with gradually less curvature, we now consider twodimensional sheets of surfactants with (local) lamellar topology. Lamellar systems are important in, for example, detergency applications. In a quasi-phenomenological model one can once again consider the local environment for the surfactant in the lamellar geometry and elaborate on the various aspects of surfactant solutions in which lamellae occur11. We will not elaborate further on this approach and concentrate more on the SCF predictions. Bilayers must be expected to be stable when the surfactant packing parameter is of order unity. This means that one needs surfactants with a relatively small head group. For the alkylethyleneoxide family we may expect the C 14 E 4 surfactant to be a candidate. For this reason we will take this surfactant as our paradigm. As in the case of linear micelles, for ionic surfactants lamellar objects are expected for head groups that are sufficiently screened, i.e. at high ionic strength. So, such structures may result for high overall surfactant concentration and/or a large amount of added salt.
11
J.C. Eriksson, S. Ljunggren, Langmuir 6 (1990) 895.
ASSOCIATION COLLOIDS
4.89
Figure 4.31. a) The standard state grand potential per unit area in the flat geometry e^ as a function of the number of surfactants per unit area g (two sides of the bilayer) for C 1 4 E 4 surfactant molecules, b) The equilibrium concentration of surfactants in the bulk as a function of the number of surfactants per unit area in the corresponding bilayers. The open circle represents the tensionless bilayer for which
4.7a Thermodynamic stability of infinite bilayers The first step in the SCF modelling of bilayer membranes is to consider surfactant aggregation in a flat geometry. We will first concentrate on large flat bilayers for which the ends (rims) may be ignored. In this case we may study the bilayer properties per unit area. This of course implies that conformational fluctuations (undulations) are ignored. In an SCF calculation one can strategically position the bilayer with the centre at a preset coordinate. In this way, translational degrees of freedom are ignored. The grand potential per unit area in this system is denoted by efm, where the superindex f reminds us about the flat geometry. As this grand potential is normalized per unit area, one can interpret this as the surface tension of the bilayer. We present the surface tension of the bilayer as a function of the number of molecules per unit area (of both halves of the bilayer) in fig. 4.31a. Again, the part with the negative slope represents stable systems, i.e. systems for which the Gibbs energy has a local minimum. In principle, we should also consider the entropy associated with the translational and conformational degrees of freedom of the membrane. The membrane as a whole may have some (order 10 kT ) translational entropy, but this can be ignored for the grand potential per unit area (per lattice site). Below we will define and compute the membrane persistence length. On the length scale of the persistence length there is some conformational entropy. For sufficiently large values of the persistence length we may also ignore the conformational entropy of the membrane per unit area. As a result, the equilibrium membrane should essentially be tensionless. We may briefly digress on this. Let us for the sake of the argument suppose that we have the bilayer in a frame so that a finite value of the surface tension y can be imposed. If y> 0 the system is not happy with the unfavourable surface area in the system, and if we free the bilayer from the frame it will contract. In this process the number of surfactants per unit area g increases. As 3//3g<0, this implies that the
4.90
ASSOCIATION COLLOIDS
surface tension goes down. Correspondingly, we may try to impose y < 0 again using a frame. This is a favourable case for the system. Indeed, as soon as the bilayer is free from the frame, it will try to increase its area in order to minimize the Gibbs energy. However, by doing so the surface tension increases. So, an equilibrium situation arises when y = 0 . The bulk concentration of the freely dispersed surfactants molecules <^urf is a function of the number of molecules per unit area in the membrane. We present this function in fig. 4.31b. In line with all the results above, the bulk concentration of surfactant should be an increasing function of the number of molecules per unit area. The tensionless bilayer is indicated by the open circle and is found at a bulk volume fraction of (p^^f = 0.00015439 . Inspection of similar data for micelles with spherical symmetry (fig. 4.9b) indicates that the tensionless bilayer is indeed more stable than any of the spherical micelles. We showed above that tensionless cylinders could not form (c.f. fig. 4.25b), and thus we conclude that the tensionless bilayer is the most favourable structure for this surfactant system. In micellar systems, the slope demdg was found to be inversely proportional to the width of the micelle size distribution (c.f. [4.2.24]). For the bilayer system only the region very close to e^ = 0 , that is the tensionless membrane, is relevant. We may define the membrane compressibility Ka = dlnam/deim, where a m is the area per molecule in the surfactant bilayer. The compressibility has the inverse dimensions of efm , i.e. [ I2 /kT] and physically says how much the area of the membrane has to change in order to give it a surface tension of unity [kT/l2]. As am = 1/ g we have 31na m = -3 In g and the compressibility modulus is also given by Ka = -3 In g I de^ . The relevant compressibility is thus found from the slope defm/dg in fig. 4.31a at e^ = 0 . The steeper the slope, the smaller the compressibility. The slope de^/dg is rather constant for a range of g values as can be seen from fig. 4.31a, which means that the compressibility decreases with decreasing area per molecule. Numerically the compressibility for the tensionless bilayer is «•= 0.068 (2 /kT , which corresponds to approximately 1.5 m/N. Sometimes the inverse of this quantity, which is a twodimensional expansibility a?, is reported for bilayers in the literature. For lipid bilayers, area expansibilities of approximately 200 mN/m are reported. Here we find that the expansibility for the surfactant layers is somewhat larger than that of lipid bilayer membranes. Let us conclude this paragraph by presenting volume fraction profiles for the flat bilayer system. In fig. 4.32 results for a tensionless bilayer composed of C 14 E 4 surfactants are presented. In panel (a) the tail and head group profiles are shown. As compared with the corresponding data in cylindrical or spherical symmetry, the head group densities in the flat bilayer systems tend to be rather high. The overlap between heads and tails is of the same magnitude as in the other geometries and again we observe that some of the EO head groups can penetrate into the core. In line with all the data presented for similar surfactants in the curved geometries, we find that water
ASSOCIATION COLLOIDS
4.91
Figure 4.32. a) Volume fraction profiles across the tensionless surfactant bilayer composed of C 14 E 4 surfactants. The head and tail distributions are presented as well as the water and free volume profiles, b) The grand potential density across the corresponding bilayer membrane. The equilibrium concentration of the surfactants is ^ u r f = 0.00015439 and g = 0.382, corresponding to a m = 5.235(2 .
does not enter in significant amounts into the core and that the free volume is somewhat increased in the core compared with that in the bulk. In fig. 4.32b the dimensionless grand potential density across the bilayer is shown. Again, we note that this quantity is not uniquely defined and that we have assigned the pair interactions equally over the two segments involved. This "natural" choice is not the only one and alternative grand potential density profiles can be generated. In line with results discussed above, we have a positive excess pressure in the head group region and a negative one in most of the tail region. Exactly in the centre the pressure happens to become positive again. We do not attempt to rationalize these details further. The integral over the grand potential density is zero (tensionless bilayer). 4.7b Finite size disks The second step in the modelling of free-floating bilayer membranes is to also consider the extra Gibbs energy associated with the rim. This system is a two-dimensional analogue of a small oil droplet in water, which features Laplace pressure and a surface tension. Qualitatively we no longer expect that the flat bilayer remains tensionless because the rim will insert a compression force. Quite generally we may decompose the (excess) grand potential of the disk ej^ into the work term to have a flat bilayer yA and an extra contribution to the existence of a rim TL , i.e. £J^ = yA + rL , where the area A , the length of the rim L and the line tension r depend on the choice made for the radius R , whereas y can unambiguously be defined (as the grand potential density in the central part of the disk that exists far from the rim). Of course, fj^ should be independent on the choice made for the R . For a given disk (with fixed aggregation number), we see that T[R) = (e^ - yA)/ L . There exists an R for which [dr/dR] = 0 , where the change in fi is a notional change (to be distinguished from the increase of the disk that takes place upon increasing the
4.92
ASSOCIATION COLLOIDS
Figure 4.33. Cross-sectional volume fraction profiles for the overall surfactant for the disk composed of g = 2960 Cj^E^ surfactants in grey scale (black high density, white low density). The dimensions are 40 lattice sites in the normal and 65 lattice sites in the radial direction. From visual inspection, the radius is approximately 50 sites wide (one unit is 8C).
number of surfactants in the object). This particular radius may be called the radius of tension R*. For this radius we have the two-dimensional Laplace equation / = -r/R*. Apart from n(R* )2 we may estimate the area by A = gam / 2 , where am is the area occupied by the surfactants in the planar tensionless bilayer. The division by 2 is needed because g is the number of molecules for the bilayer (not for the monolayer). From this estimate, we directly find a value for the radius R . We expect that this radius is close to R*. In a third option, we may visually investigate the density profiles for a flat, finite-sized disk and estimate the radius from that. Let us start by presenting in fig. 4.33 a disk-shaped micelle composed of g = 2960 C 14 E 4 surfactant molecules in a gray-scale volume fraction profile plot. Here we present a cross-section from the centre of the disk to the bulk, both in the normal z direction as well as in the radial r -direction. From visual inspection, the disk has a radius of approximately 50 sites. The flat bilayer shown in fig. 4.33 has an area per molecule a m = 5.235 I2 and R = <Jg/2am/x ~ 49.66 . The latter result is in very good agreement with visual inspection of the profiles. It is possible to compute the surface tension of the bilayer in the central part of the disk for which indeed a negative value is found, i.e. y = -0.009984 [kT/l2]. This tension is constant over a large part of the central region of the disk (not shown). The standard state grand potential is also available from the SCF results, f^ = 77.800 kT. The radius of tension is found at R* = yj-E^/^7 = 49.8 I. From the latter value of the radius, we find the line tension associated with the rim of the disk as z=-Ry = 0A92 [kT/l]. In fig. 4.34 we present results for a wide range of disk sizes. The standard state grand potential is plotted in double logarithmic coordinates against the aggregation number. The fact that e^ « g0-5 shows that the thermodynamic analysis of the disk may be accurate for significantly smaller disk sizes as well. Using a fixed value of the area per molecule a m = 5.235 I2 , the radius was computed from R = jg12am I n and the line tension is thus estimated from r = eJ^A/rR). This estimate of the line tension is also shown in fig. 4.34 from which it is found that the line tension is approximately constant over the whole range of disk sizes. Close inspection shows that the line tension increases with decreasing disk size proving that there is a curvature correction to the line tension. We do not further analyze this and mention that this analysis of the
ASSOCIATION COLLOIDS
4.93
Figure 4.34. The grand potential of the disk as a function of the aggregation number g of Cj 4 E 4 surfactants in double logarithmic co-ordinates (left ordinate). The open symbol corresponds to the disk with g = 1200, ^ u r f = 0.0006075, and e^ = 50 kT . The corresponding line tension of the rim z as a function of the logarithm of the aggregation number (right ordinate).
line tension breaks down at very small disks (not indicated) for obvious reasons. This negative tension that occurs in the plane of the disk indicates that the system can gain Gibbs energy when it creates more membrane area. It can do so by buckling up. We thus conclude that a large disk is not stable and it should be possible to prove that at some disk size the closed vesicular state is thermodynamically more stable. To that end, we should analyze the curved bllayers in more detail. 4.7c Homogeneously curved surfactant bilayers The third step in the analysis of isolated bilayer membranes is to analyze homogeneously curved ones. Here we shall consider two homogeneously curved surfactant layers, viz. a spherically closed bilayer, known as vesicle or liposome, and a cylindrically shaped vesicle. This type is only homogeneously curved if the object is infinitely long so that end effects can be ignored. For such cylindrical or tubular vesicles, we will analyze the structural and thermodynamic parameters per unit length. For a curved interface one needs to specify two radii of curvature at each point on the surface. For homogeneously curved objects, the radii of curvature are constant along the surface. As a result, the shape of the object can be described by two radii of curvature Rj and R2 . Our first interest is in bilayers that are weakly curved. This means that the radii of curvatures are large. The curvature is small and therefore we have the two curvature parameters, the mean or total curvature J = l/Rj+1/R 2 a n d the Gauss or saddle splay curvature K = l/RjXl/R 2 , see [III. 1.1.4 and 5]. There are other possibilities to define two independent curvature parameters, but the present choice of J and K is very convenient because their values are insensitive to the exchange of Rj and R2 . We use a Taylor series expansion of the membrane surface tension in the small parameters J and K around the flat state
*/,« = *0,0,+ (f2:) J + i t e ) \6J)00
2{dJz ) Q 0
J>+(m
K + ...
[4.7.1,
\dK)00
In principle this expansion of the surface tension covers the possibility that the flat
4.94
ASSOCIATION COLLOIDS
bilayer state is not the state with the lowest tension. However, bilayers for which this is the case must be very rare. One would anticipate such a bilayer when it is composed of more than one type of surfactant and when it has a solubility gap. One way to prevent an unfavourable region between (phase) separated domains is to position one surfactant on the inner side and the other one on the outer side. Then there may be a spontaneous curvature. In biological systems spontaneously curved states may be maintained actively, but for in-vitro systems we expect such a scenario to be rare. In the typical case where the flat bilayer is the most stable state, we can simplify [4.7.1]. We argued above that the unsupported flat bilayer has no tension and hence ^(0,0) = 0 . We further notice that one can have positive and negative values for J and K . With respect to the flat bilayer, we can choose at will which direction of the radius of curvature to count positive or negative. As a result, in the expansion [4.7.1] all odd terms in the curvature should be zero as well. Defining the mean bending modulus Jcj = {d2y/dJ2)00
and the Gauss or saddle-splay bending modulus k2 =
0y/dK)oo
we may write for small curvatures y{J,K) = -klJ2
+ k2K
[4.7.2]
which was first introduced in the context of surfactant bilayers by Helfrich1'. We now may compute the total bending energy stored in a spherical vesicle with radius R or in a tubular vesicle per unit length I and radius of the tube R . As the vesicle is homogeneously curved, we can multiply y{J,K) by the surface area. The result can be identified by the standard state grand potential of the vesicle: (2 l "\ e^ = 4xR2y\ —,— = 4;r(2k1 + k2) 1 A k \ = K-±- = 7tk J (—,0 l
R
J
spherical vesicle
[J]
[4.7.3]
tubular vesicle
[Jm" 1 ]
[4.7.4]
R
The interesting result is that for the spherical vesicle, e^ is independent of the radius. This has important consequences for the thermodynamic stability of closed vesicles. We will return to this below (cf sec. 4.7d). In the previous section we have shown that the disk-shaped micelle has a grand potential, which is an increasing function of the aggregation number. The grand potential of the closed vesicle however is not a function of the aggregation number. As a consequence there must be a critical aggregation number for which the closed vesicle has a lower grand potential than the disk. In fig. 4.35a we present £^{g) • As expected from the Helfrich analysis the total curvature energy of the spherical vesicle converges to a constant value for sufficiently large values of g. The limiting value is
11
W. Helfrich, Z. Naturforsch. 28c (1973) 693. The constants fc, and k2 are identical to kc and k often used by Helfrich and others.
ASSOCIATION COLLOIDS
4.95
4/r(2kj +k2) ~ 50 kT . For very small values of the aggregation number g < 150 , there is just a spherical micelle. In the regime 150 < g < 2000 there is clearly a positive slope de^ /dg > 0 indicating that these intermediate vesicle sizes are not yet stable. For very large values of g the grand potential becomes essentially a constant. Again, as the slope does not become negative, the thermodynamic stability of these larger vesicles remains an issue and we return to this below (cf. sec. 4.7d). At this stage we notice that for g ~ 1200 the grand potential of the flat disk becomes equal to the grand potential of the closed spherical vesicle. The disk with g = 1200 is indicated by an open symbol In fig. 4.34. To have a transition from the disk to the vesicle, the chemical potential of the surfactant should be lower in the latter case. In fig. 4.35 we prove that this is the case. In this figure we compare the bulk volume fraction of free surfactants in equilibrium with closed vesicles and flat disks with aggregation number g . In the relevant range of sizes g > 1000 , the bulk volume fraction is lower for the vesicular system. The conclusion is that the transition from disks to vesicles is expected to take place at least for g > 1200 , but perhaps already at lower g . Although the chemical potentials signal that the vesicles are more stable than the disks, the grand potential of the latter is lower. The dynamics of the closing of the disk to form vesicles is therefore not a trivial matter.
Figure 4.35. a) The standard state grand potential of a spherical vesicle e^ composed of C-14E4 surfactants as a function of the aggregation number g. b) The volume fraction of free C 14 E 4 surfactants in the bulk in equilibrium with micelles with aggregation number g for the spherical vesicle (solid line) and the flat disk (dashed line), c) The radial density profile across a homogeneously curved vesicle composed of g = 1200 surfactants. The centre of the vesicle is at r = 0 . d) The grand potential density in units kT/fi across the corresponding bilayer.
4.96
ASSOCIATION COLLOIDS
In fig. 4.35c we give the radial volume fraction profile of a very small vesicle with g ~ 1200 surfactants. After closing the disk, this is the first vesicle that is expected to form. The radial density profile should be compared with the cross-section volume fraction profile shown in fig. 4.32a. As expected, the curvature of the bilayer has significant effects on the density profiles. The head group density is relatively high on the inside, and somewhat lower on the outside than it is in its flat counterpart. On the other hand, the tail density tends to be higher on the outer side. As a result the overlap between head and tail profiles is stronger on the outside. The dimensionless grand potential density profile across the vesicle is given in fig. 4.35d. Although the vesicle is not yet large enough to have a homogeneous water phase inside, it can already be seen that the grand potential density inside the vesicle approaches zero. This implies that the vesicle has no Laplace pressure inside. This observation is confirmed by inspection of the corresponding profile for much larger vesicles. From the fact that the inner water phase is thermodynamically in equilibrium with the same outer water phase, one does not expect to find such a pressure. On the other hand, one may argue that the vesicle bilayer has a surface tension y= e^/{AfiR2). The Laplace equation SP = 2y/R is "saved" by the observation that the surface of tension is positioned at R —> °° . The p N - p T profile across the curved bilayer can be compared with the flat one given in fig. 4.32b. Before we address the thermodynamic stability of the vesicles, it is of sufficient interest to evaluate the two bending moduli, kx and k2 , separately. From the spherical vesicle we only have access to the combination kb = 2fcj +k2 (see also HI page 1.81). However, the analysis of cylindrical (tubular) vesicles gives information on the mean bending modulus only. In fig. 2 we plot e^ as a function of the curvature J = 1 / JR . From the Helfrlch expansion [4.7.2] we know that the slope of this curve equals k^n. To a good approximation we find in this case fcj = 4/cT . Using this result in the curvature energy of the spherical vesicle, we extract k2 = -4fcT . 4.7d On the thermodynamic stability of vesicles Taking a membrane piece from the flat state (ignoring end effects) and curving it homogeneously into a spherical vesicle will cost an amount e^ of curvature energy.
Figure 4.36. The standard state grand potential per unit length of the tubular vesicle composed of Cj^E^ surfactants as a function of the curvature J = 1 / R . The dotted line is the extrapolation of the computed data to the limit of J —> 0 .
ASSOCIATION COLLOIDS
4.97
For sufficiently large vesicles, this value becomes independent of the radius. As a consequence, the process by which two vesicles merge into one larger one is favourable, as it will lower the GIbbs energy of the system by an amount of e^ . We will neither elaborate on the mechanism of vesicle fusion nor on the rate of this process. We expect, however, that vesicle fusion is a very slow process. One might therefore also consider the transport of individual surfactants from small vesicles to larger ones (analogous to Ostwald ripening). The driving force for this process is almost zero because taking one lipid out of a small one and inserting it into a larger one does not significantly reduce the Gibbs energy (the total number of vesicles remain the same). Following the argument, however, we expect that, given enough time, vesicle disproportionation will proceed up to the situation that the total volume of the solution is packed with large vesicles. Then one needs to account for the membrane-membrane interactions (cf. sec. 4.8), which will stop the process. However, there is experimental evidence that the sketched scenario of continuously growing vesicles is not the full story. Dilute vesicle solutions can be found routinely in experimental situations and such vesicle solutions seem to obey equilibrium characteristics11. We have made the explicit assumption above that all vesicles are homogeneously curved. We should, however, consider the consequences of random changes along the bilayer in, e.g., thickness, curvature, etc. These changes have both energetic and entropic consequences. It appears that the bending properties are of extreme importance for tensionless bilayers. Therefore, we need to consider the physics of curved bilayers in somewhat more detail. The Gauss bending modulus, k2 is of interest for the phase behaviour of the surfactant layer. If fc2 > 0 , the Gibbs energy of the interface can be lowered by forming saddle shapes. Saddles have radii of curvature of opposite sign and thus a positive value of K. It follows from the Gauss-Bonnet theorem21 that the integral of the Gaussian curvature over a surface is related to the number of handles (genus) p formed on a closed interface: JKdA = 4^(1 - p )
[4.7.5]
For a spherical interface, p = 0 and the integral equals 4TT . All surfaces that can be transformed into the sphere without cutting the surface are topologically equivalent to the sphere and have p = 0 . The genus is unity ( p = 1) for the torus and related shapes, and generally the genus is equal to the number of handles that characterize the closed surface. For example one can have a cubic phase (cf. 4.3 for which the genus is unity
11
M.M.A.E. Claessens, Size Regulation and Stability of Charged Lipid Vesicles, PhD thesis, Wageningcn University (2003). 21 S. Hyde, S. Andersson, K. Larsson, Z. Blum, T. Landh, S. Lidin, and B.W. Ninham, The Language of Shape: the Role of Curvature in Condensed Matter Physics, Chemistry and Biology, Elsevier (1997).
4.98
ASSOCIATION COLLOIDS
per unit cell. This means that the integral over the Gauss curvature does not contribute to the curvature energy in this case. When k2 > 0 the formation of each spherical vesicle has a Gibbs energy penalty of 4nk2 • On the other hand, when k2 < 0 there is a gain in Gibbs energy. We conclude that for the stability of vesicles, k2 < 0 is a prerequisite. The second criterion for intrinsic stability of the vesicle is that the standard state grand potential must be positive. As a result, stable vesicles are only expected when -2kj < k2 < 0 . In passing, we note that the vesicles composed of C 14 E 4 surfactants obey both criteria as shown above. An important consequence of [4.7.5] is that the value of k2 does not control the shape changes of the closed vesicle. Instead such fluctuations are regulated by the mean bending modulus fcj. Fluctuations of the membrane shape are known as undulations. Undulations develop on the tensionless bilayer due to thermal motion. We may define the persistence length £m of the membrane as the distance over which the normals become decorrelated. This means that the membrane is locally flat on length scales smaller than £m . With a length scale larger than £m , the membrane executes a random walk in space (similarly to a polymer chain). De Gennes and Taupin11 have shown that for surfactant bilayers the persistence length is an exponential function of the mean bending modulus
4,-'=P[^] where the value of a is still subject to debate. Simulations tend to be consistent with a value of a ~ 3 . The length I is a molecular length, for which we here take the crosssection of the tail (lattice site). The persistence length is important to understand the existence of vesicles. The key idea is that on length scales larger than the persistence length, a membrane-piece should be allowed to close and form the vesicle without a penalty to pay for any curvature energy. Apparently, there is an effective bending fcj modulus of the bilayer, which is a function of the length scale of the membrane patch L as
(L)=
ta
*' 4-^ (t)] We may use this equation to estimate the curvature energy of the undulating vesicle £^(R) = 4^r2fcj(R)+fc2l
[4.7.8]
This effective e ^ becomes a decreasing function of the aggregation number ( R °= g2 ), and we conclude that vesicles are made thermodynamically stable because of the
" P.G. dc Gcnncs, C. Taupin, J. Phys. Chem. 86 (1982) 2294.
ASSOCIATION COLLOIDS
4.99
Figure 4.37. a) The overall bending energy of a spherical CTAB vesicle (with radius R = 100 ) e^ as a function of the volume fraction of added salt in the solution (left ordinate; closed symbols) and the bending moduli fcj and fc2 (right ordinate; open symbols), b) The volume fraction of free surfactants as a function of the added volume fraction of salt, for CTAB micelles at the c.m.c. (dotted line), for micelles with £ m = 0 (dashed line, which discontinues at ^>s cs 2 x 10~3 ), and coexisting with spherical vesicles. undulation entropy. We may now estimate the vesicle size of the C 1 4 E 4 surfactant system of the previous section. Assuming that a renormalized grand potential of e^ ~ 10 kT can be compensated by translational entropy, we find using kc = 4 kT , fc = - 4 kT , and a = 3 , vesicles with a size of order R ~ 150 nm. It will be clear that for a surfactant system with a significantly higher mean bending modulus, the Gauss bending modulus must be sufficiently negative to keep the vesicles to a mesoscopic size.
(i) Charged bilayers Stable vesicles are expected when the mean bending modulus is of order kT and the Gauss bending modulus is sufficiently negative, or -2fcj < k2 2 0 . These mechanical parameters not only depend on the surfactant architecture, but also on the physicochemical conditions. We will illustrate this for the CTAB surfactant system. In fig. 4.37 we present the curvature energy of CTAB surfactant vesicles over one decade in ionic strength, i.e. 10"3 <
4.100
ASSOCIATION COLLOIDS
curve stops when it is about to cross the value found for the vesicles. For this ionic strength, the grand potential no longer reaches the value zero for spherical micelles. For low ionic strength cases, the dotted and dashed lines are below the vesicle line indicating that vesicles are not the preferred geometry. This observation coincides with the fact that the curvature energy of the vesicle becomes negative. This also signals instability of the vesicles. For the ionic strength, at which the total curvature energy of the spherical vesicle is positive, we still find that micelles at or near the c.m.c. have a lower chemical potential. This means that for these systems vesicles only form at sufficiently high surfactant concentration, and that we should expect the vesicles to coexist with small micelles. The total curvature energy contains two contributions kx and k2 for the mean and Gauss bending modulus, respectively. In fig. 4.37a we show that kj is a weakly decreasing function of the ionic strength. We find a power law with an exponent of - 0 . 1 . This dependence is very weak but still significant. For low ionic strength where the stability of the vesicles vanishes, the bending modulus is relatively high and the persistence length of the surfactant bilayer the largest. In first order we expect the vesicles to have a radius, which is proportional to the persistence length and, therefore, we expect the vesicles to increase in size with decreasing ionic strength. Claessens has studied this phenomenon in detail for lipid vesicles11. For these systems one can go to much lower ionic strengths and as kl is relatively large for lipids, it is possible to find (at low ionic strength conditions) very stable giant vesicles that exceed the //m scale. The Gauss bending modulus k2 is a stronger function of the ionic strength; it is strongly negative for low ionic strength, but approaches zero for high ones. Its low value at high ionic strength signals the instability of the bilayers at these physicochemical conditions. We will study the CTAB system at higher ionic strength when we consider the interaction between bilayers below in sec. 4.8.
Figure 4.38. The overall bending energy of a spherical vesicle (left ordinate) and the mean bending and saddle-splay modulus (right ordinate) for a C 12 E 5 surfactant as a function of the hydrophilicity of the head group.
11
M.M.A.E. Claessens, Size Regulation and Stability of Charged Lipid Vesicles. PhD thesis, Wageningen University (2003).
ASSOCIATION COLLOIDS
4.101
(ii) Non-ionic bilayers The effect of added salt in ionic systems is mirrored by an increase of the temperature in non-ionic systems. This is again shown in fig. 4.38 where we present the total bending energy of a spherical vesicle in combination with the bending moduli of the C 12 E 5 surfactant system as a function of ^QW (rented to the temperature as specified by [4.4.6]). In line with data presented for the ionic system, the overall bending energy increases with increasing Xow • m e m e a n bending modulus decreases when the head groups become less repulsive and the Gauss bending modulus increases. For the nonionic system, it appears that above %ow ~ -0.41 the bilayers loose their stability (because k2 > 0); instead, we expect that the system will generate saddle-shaped surfaces. In sec. 4.8 we will show that for such a low value of J O w ^ i s P oss ible that the intrinsic interaction between the bilayers is not necessarily repulsive. However, the decrease of the mean bending modulus kx becomes less with increasing temperature. This means that the membranes become more flexible. In sec. 4.8a we will also show that these flexible membranes tend to repel each other because of undulation forces. The fact that undulations become more important at higher temperatures is consistent with the experimental observation that the lamellar spacing in the La phase increases with increasing temperature. The contradictory trends that (i) the increases of the undulation repulsion and (ii) the intrinsic interaction becomes attractive, in combination with the appearance of a positive value for the Gauss bending modulus, correlate with the appearance of the L3 or sponge phase (c.f. fig. 4.4). We note that alternative explanations of the appearance of the L3 phase invoke the bending moduli and spontaneous curvature of individual monolayers that are part of the bilayer11. These mechanical parameters can be extracted from SCF computations as was recently shown by Kik and coworkers21. We cannot go into these details here. 4.8 Interactions between parallel lamellar surfactant layers In this chapter we have thus far assumed that the association colloids did not interact with each other. This is a good approximation for dilute micellar solutions sufficiently far from the c.p.t. or the Krafft temperatures (non-ionic and ionic surfactant systems, respectively). However, the complete phase diagram of any surfactant system becomes particularly rich at relatively high concentrations. Then the micelles are forced to interact and these interactions have many consequences. The topic of strongly interacting association colloids is considerably more complex than the corresponding problem with classical (rigid) colloids. Interactions may induce changes in the aggregation number or topological transitions, say, from spherical to cylindrical and even to lamellar geometry. Indeed, in the concentrated regions of many surfactant phase
11 21
U. Olsson, H. Wennerstrom, Adv. Colloid Interface Set, 113 (1994) 113. R.A. Kik, J.M. Kleijn, and F.A.M. Leermakers, Phys. Rev. Letters, submitted (2004).
4.102
ASSOCIATION COLLOIDS
diagrams a plethora of liquid crystalline phases occur. Various cubic hexagonal or lamellar phases are only stable when strong interactions occur. Concentrated surfactant phases also form when the association colloids become attractive, e.g. as expected above the c.p.t. or below the Krafft temperatures. In this case a phase rich in surfactants coexists with a dilute phase. To understand the surfactant-rich phase it is again necessary to study the interaction between micelles. For geometric considerations it is relatively simple to focus on the interaction between parallel lamellar surfactant bilayers. Even this problem is complex and challenging; we introduce it in the next subsection. 4.8a Undulation forces between bilayers There are two generic contributions to the pair interactions of bilayers. Both are due to fluctuations: (i) Fluctuations in the electronic densities give rise to the (mostly attractive) Lifshits-Van der Waals forces, dealt with in I. sees. 4.6 and 4.7. (ii) Fluctuations of the shape of the bilayers lead to the (repulsive) so-called undulation force. As the latter is less well-known, we will elaborate shortly on undulations. The undulation force is easily explained realizing that the tensionless bilayer features shape variations on various length scales. The entropy associated with these fluctuations contributes to the stability of the bilayers. The spectrum of shape changes that can develop depends on the distance between the bilayers. The closer they are packed the more these fluctuations are damped. As a result, the fluctuations will push bilayers apart (repulsion) (sec. III.2.9c). This can also be quantified. Let us consider a square piece of a surfactant bilayer with a surface area LxL = As oriented on average parallel to the x,y plane. The shape of the bilayer is at any moment characterized by the height u(x,y) of the bilayer, which can be positive or negative. The average position is such that (u) = 0 . It can be shown11 that the mean square amplitude scales as the size of the membrane squared ( u 2) = _WL(M a
[ 4 .8.1,
This result is as intuitively expected. The shape of the interface can be described by a set of sine and cosine waves, so called modes. The modes with the largest wavelengths have the larger displacements u from the x,y plane. Each mode represents a degree of freedom and contributes equally to the elastic Gibbs energy with -g- kT (equipartition theorem), and therefore the averaging of all the modes gives a root mean square amplitude, which scales with the size of the bilayer. This type of scaling is known as scale-invariant, and it is an important property of the system. From 14.8.1] it follows that the mean square amplitude is proportional to the temperature and inversely
' W. Helfrich, Elasticity and Thermal Undulations of Fluid Films of Amphiphiles, in Liquids at Interfaces, Les Houches, J. Charvolin. J.F. Joanny. and J.Zinn-Justin. Eds. (1988).
ASSOCIATION COLLOIDS
4.103
proportional to the mean bending modulus fcj. When undulating bilayers are in close proximity, the excursions of the bilayers are limited to the average spacing between the bilayers. In other words, some modes are suppressed from the spectrum of fluctuations and thus the membranes are subjected to an entropic repulsion between them. A simple argument to find the scaling law of this undulatory interaction can be obtained by an independent membrane-piece approximation. Let us consider a single membrane of negligible thickness between two parallel rigid walls of spacing 2h . The membrane is thought to be cut up into equal squares of area Ai. The area Ai is chosen such that the fluctuations can fit into the slit of 2h (u 2 )och 2
[4.8.2]
where the proportionality constant should be of order unity. Combining [4.8.1] and [4.8.2] gives A. oc 1
L
[4.8.3]
kT
It is now assumed that the pieces of size Ai behave as free particles in a onedimensional ideal gas. The number of these particles equals JV = As I Al. Therefore, the pressure (force per unit area) exerted by the membrane on either wall should obey
p u «_5L_ f c r = _ l _ f c T . e K ? £ u
Asx2h
Aix2h
kxh3
[4.8.4]
Integrating -J Jdh gives the undulation Gibbs energy per unit area
The numerical factor in [4.8.5] is to our knowledge still subject to debate and ranges from 6/32 to 3/128 . The difference in the numerical prefactor is not unimportant because the attractive van der Waals force between two bilayers with thickness b and separation h is of the same order of magnitude and has the same scaling with the distance according to [1.4.6.24]
g
™=-if^-^+i^W)
[4 8 61
--
where A is the Hamaker constant. The total interaction energy follows superposition and as a first approximation it can be assumed that the two are linearly additive. We do not expect that this is very accurate. Nevertheless, we believe that typically the undulation repulsion and the van der Waals attraction are compensating each other at least to some degree, and the
4.104
ASSOCIATION COLLOIDS
binding (effectively they are attracted to each other) or unbinding of the bilayers (i.e. when they are repulsive) is a subtle issue, which will be dominated by the surfactantspecific interactions, e.g. the electrostatic or steric interactions. For these specific interactions, once more we need a molecular realistic model to account for the relevant features. We will turn once again to SCF modelling. 4.8b Intrinsic interactions between surfactant
bilayers
At constant p and T the Gibbs energy of the system is minimized in equilibrium. We consider a pair of flat bilayers that by an external force/ (per unit area) is kept a distance h apart. This force may also be identified as the disjoining pressure 17[h) . A schematic drawing of this configuration is given in fig. 4.39. We can add the external force to the Gibbs energy and write dG = ^
juidNl + 2ydAs - JAsdh
[4.8.7]
i
With this extra contribution, we now have the modified Gibbs-Duhem (Gibbs law) equation A d
s r =- X N M - / A s d h
I4-8-8'
i
Next we must consider the number of degrees of freedom the bilayers have. Indeed, an unconstrained bilayer might choose to change its area upon interaction with other bilayers. In this case, the bilayer remains without tension, i.e. ( -,c \ Nf—
=2y = 0
[4.8.9]
represents the lowest Gibbs energy for fixed h . Upon integration of [4.8.7] for / = 0 we realize that h is an intensive variable and hence G =
X'"iJVi
[4.8.10]
i
Figure 4.39. Schematic picture of a pair of membranes of which the left one is fixed in space and the right one is kept at a distance h from the first one by an external force / . The area of each membrane is As and the system is kept at a fixed temperature and pressure.
ASSOCIATION COLLOIDS
4.105
Differentiating [4.8.10] and comparison to [4.8.7] leads to a Gibbs-Duhem equation wherein the surface extension work does not occur. The interaction energy for tension free bilayers is found upon integration G(h)-GM = Y JV i (^ i (h)- J u i H) = -fh
/A s dh'
[4.8.11]
i
We will usually be interested in the interaction Gibbs energy at a separation h per unit area, which is present at separation h , and therefore we write
1
As(h)
As(h)
As(h)
At sufficiently low surfactant concentrations the chemical potential of the solvent is essentially constant, and therefore we only have to monitor the chemical potential of the surfactant that is in equilibrium with tensionless membranes separated at a distance h , i.e. // s u r f (h), in order to know whether bilayers attract or repel each other. For example, if the chemical potential increases upon decreasing h this means that repulsion prevails.
Figure 4.40. Gibbs energy of interaction between two tensionless surfactant bilayers composed of CjgEx surfactants separated by a distance h : a) for various values of x as indicated, b) for c 18 E 20 a t various values of the Xo\N a s indicated. The lower the temperature, the more negative this value. In fig. 4.40 we show some typical interaction curves for non-ionic surfactants of the type C lg E x . In fig 4.40a we give results for a series of surfactants for fixed tail length and for a set of head group lengths. The goal of this plot is to show that the range of interactions increases with increasing head group length, whereas at strong compression the interaction energy becomes independent of the head size. This means that when the head group is short, the repulsion is stronger. In other words, bilayers composed of long-headed surfactants are slightly softer. Above we discussed that with increasing temperature the thermodynamic stability of the non-ionic surfactant bilayers can be lost. It was argued that when this is the case
4.106
ASSOCIATION COLLOIDS
the lamellar phase with attractive bilayers should prevail. In fig. 4.40b we give an example of interacting tensionless bilayers of the C lg E 2 o type for a series of xow values. In line with expectations it is found that at high temperatures, i.e. at %ow > ~0-5 , there appears a minimum in the interaction curve. When this minimum is very deep, this means that sufficiently large bilayers will stick to each other. In other words, they are bound; a lamellar phase is formed with relatively little water between the bilayers. However, when the minimum is shallow, we expect that there can be coexistence of bound bilayers with free-floating vesicles. The intrinsic interactions between charged bilayers also contain the important repulsive contribution from the overlap of the electrostatic double layers. It is customary to evaluate the electrostatic part of the Gibbs energy of interaction by considering the diffuse double layer interaction at a fixed (i.e. independent of separation) diffuse layer potential or charge. In sec. IV.3.5 these oversimplified models were replaced by a more general regulation theory between Gouy-Stern layers. This model appears most appropriate for the interaction between charged lipid layers, but it has so far not been applied. However, the present SCF model is even more to the point because it explicitly accounts for the head group (charge) distributions. Changes in head group conformations may also contribute to reduce the repulsion between the bilayers and finally the head group area is a degree of freedom to take into account. All these degrees of freedom are routinely accounted for in the SCF analysis. CTAB bilayers are stable at relatively high ionic strength. For this reason, we choose this system to illustrate typical interaction curves for charged bilayers. In fig. 4.41a we present the change of the chemical potential ^2surf = jUSUTf (h) - Msurf (°°) a s a measure of the interaction. Recall that the bilayers remain tensionless during the interaction. They can do this only by adjusting the area per molecule upon bilayer interaction. The corresponding relative change of the area per molecule is presented in the accompanying graph 4.41b. For weakly interacting bilayers a purely electrostatic interaction is observed. The repulsion
Figure 4.41. a) The change of the chemical potential of the surfactant as a function of the bilayer separation h of tensionless bilayers composed of CTAB with added volume fraction of 1:1 electrolyte as indicated, b) The relative change of the area per surfactant molecule SaT = [a{h] — a(-x^)]/a(^j) as a function of the distance between the corresponding tensionless bilayers.
ASSOCIATION COLLOIDS
4.107
is simply exponential and the Debye length is the characteristic length scale. The changes in the area per molecule are exponentially small and effectively these can be treated as being constant. As expected, upon stronger compression h < 30 the repulsion deviates from the exponential dependence. The surface charge regulates and the mechanism for that is a relatively large increase in the molecular area.
4.8c Liquid crystalline
phases
Above we have shown that with increasing concentration the spherical micelles gradually loose translational entropy and that this effect is sufficient to explain the transition from spherical to cylindrical micelles. Both phases are isotropic. However, when inter-aggregate interactions begin to contribute to the free energy of the system, i.e. at sufficiently high surfactant concentration, a first-order transition to macroscopically aligned objects, e.g. cylindrical aggregates (hexagonal phase) may occur. The aligned solution of the cylindrical aggregates allows for an increase of the volume fraction of packing, or equivalently, it decreases the interaction (Gibbs) energy between the aggregates. This Gibbs energy gain is sufficient to maintain a local cylindrical micelle shape even though the spherical shape may intrinsically be more favourable. For long it was believed that in the hexagonal phase the cylindrical micelles were essentially infinitely long. This seems not necessarily to be the case. The size distribution of cylindrical micelles of finite size (with spherical endcaps) will depend on the intrinsic (in) stability of the cylindrical micelles and on the average volume fraction of packing. There is little theory for the formation of the hexagonal phase. When the packing volume fraction is further increased, we expect a transition from the hexagonal to the lamellar phase. Lamellae may be packed essentially up to a volume faction of unity. We thus witness the gradual loss of curvature of the micellar objects. In between the lamellar phase and the cylindrical one, there may be a cubic phase. The route to explain these geometrically complex phases is to analyze the mechanical parameters of the bilayers as a function of the spacing between the bilayers. Apparently pushing intrinsically repulsive bilayers to close proximity might increase the Gauss bending modulus of these layers such that the negative value gives way to positive ones. Again, there is no work available in the literature to confirm such a hypothesis. In conclusion, we may say that the self-assembly of surfactant in micellar aggregates is a subject that is maturing over the years. The coupling between the self-assembly and the long-range order between the micelles (as in the liquid crystalline phases) remains one of the most important open problems in surfactant science. 4.9 Applications of the modelling We have shown above that much of the surfactant phase diagram may be understood starting from the physical properties of the surfactants, i.e. the molecular architecture
4.108
ASSOCIATION COLLOIDS
and the interactions. Obviously the same also applies to more complicated surfactant systems. The treatment in this chapter emphasized the applicability of SCF theories to a number of relatively "simple" systems. Byway of application, two elaborations aiming at practical applications will now be presented. 4.9a Binary non-ionic ionic surfactant systems Surfactant formulations used in commercial applications are generally blends of surfactants, apart from the fact that most commercial surfactants already are mixtures, even if only caused by polydispersity. Here we consider mixtures of surfactants that differ in the chemistry or chain architecture (say, linear versus branched). A typical example is a mixture of non-ionic and ionic surfactants. The success of the blends of distinct surfactants rests in the synergistic (co-operative) or antagonistic effects or simply when one surfactant can for some reason not do its job, it is taken over by another one. In many cases the advantage of working with mixtures has a dynamic origin, for instance, when one of the constituents acts faster and the other stronger. The prediction of the properties of strongly interacting mixed systems, which is a most important application, is not easy. A widely used model is the regular solution theory (RST) introduced in sec. 1.2.18c. In RST the entropy of mixing is (as in the Flory-Huggins theory) treated as ideal, but the (Gibbs) energy due to the interactions is taken to depend on the composition. The Gibbs energy density of mixing two surfactants A and B is A
9mix = XA l n X A + XB l n XB + PXP,XB
[A.9.1]
where x A is the mole fraction of surfactant A. The RST theory is often invoked as a first step in dealing with non-ideal mixtures. For surfactants this is a particularly complex issue because it is unlikely that the multitude of interactions can be simply merged into just one non-ideality parameter, /?. In a molecularly realistic statistical thermodynamical analysis one must go beyond RST theory. For example, it is possible to account rather accurately for the size differences and implement the origin for the physical interactions between those governing the non-ideality, e.g. when the surfactants are of opposite charge. As a result, one can gain molecular level information on the reasons for deviations from RST. We will now apply the SCF theory to illustrate this by discussing various properties of mixed micelles, such as the c.m.c.s. One may expect that the micellar properties will depend nonlinearly on the composition, that the effect of salt depends on how many ionic surfactants are present, etc. We note that there are detailed quasi-macroscopic approaches to quantify the deviations for the ideal mixing, i.e. to analyze the synergistic effects observed for the c.m.c. of binary surfactant mixtures11. It may be worthwhile to confront the SCF predictions for the mixed system of an ionic and non-ionic surfactant
11
J.C. Eriksson, M. Bergstrom. and M. Persson Russian J. Phys. Chem. 77 (2003) S87.
ASSOCIATION COLLOIDS
4.109
with such quasi-macroscopic approaches, but space does not allow us to go into these details. (i) Mixed spherical micelles Consider the binary surfactant system CTAB and C 12 E 5 in a 1:1 electrolyte solution. As compared with the modelling of the individual systems, there are very few additional FH interaction parameters, viz. the interactions between the O of the nonionic and the charged components in the CTAB salt solution. For simplicity we choose all the new parameters to have athermal interactions, i.e. XQY = 0 for Y = X, Na, Br (where X is the charged group in the CTA surfactant). The number of CTAB molecules per unit volume is given by n CTAB = >CTAB I ^CTAB • where N is the number of segments of the surfactants (i.e. proportional to the molar volume). We define the molar fraction of CTAB as * C T A B = n CTAB /(nCTAB + nc The molar fraction of freely dispersed surfactants is indicated by *CTAB •
E ). tnat in
micelles by X(5pAB • Very close to the c.m.c. of this mixture, the micelles will be spherical. Consistent with this, we first concentrate on the mixed micelles that have a standard state grand potential of £m - 20 kT. This corresponds to a very small volume fraction of micelles
is the natural control parameter in these close to the
c.m.c. systems. Some typical results are collected for this system in fig. 4.42. In fig. 4.42a the c.m.c. (in volume fraction units), which is found as the sum of the two volume fractions of the freely dispersed surfactants, is presented as a function of X
CTAB f° r f ° u r
vames
°f the ionic strength as indicated. For a relatively high volume
fraction of salt cps > 0.004 , the c.m.c. is a decreasing function of the fraction of ionic surfactant. For relatively low ionic strengths, the c.m.c. shows a minimum as a function of this fraction, where the minimum is at relatively high values of -"^TAB • T n e appearance of the minimum stems from the mixing entropy. Relatively small additions of CTAB to the C 12 E 5 surfactant solution give a steep drop of the c.m.c. The area per charge at the micelle (core-water) surface is strongly affected by the composition of the micelle. This results in a large variation in the electrostatic part of the Gibbs energy of the mixed micelle upon composition changes. This electrostatic contribution is the major factor causing the non-ideality in this system. The overall molecular composition in the micelle differs significantly from that of freely dispersed surfactants (and, hence, from the overall composition). Figure 4.42b shows that for all ionic strengths small amounts of CTAB are preferentially taken up by the micelles. The first few charged surfactants can enter the micelle and allow the EO tails to be separated from each other somewhat without the appearance of a strong electrostatic repulsion. The higher the ionic strength the more pronounced the preferential uptake is. This indicates that even at low loading some electrostatic repulsion is already important. Dissolving small amounts of non-ionic surfactant into the ionic
4.110
ASSOCIATION COLLOIDS
Figure 4.42. Micellization data for the C12E5-CTAB surfactant mixture in a 1:1 electrolyte solution. It is assumed that the micelles are spherical and have a grand potential of e m = 20 fcT. a) The overall c.m.c. as a function of the molar fraction of CTAB freely dispersed in the bulk, b) The fraction of CTAB surfactants in the micelles, *CTAB • a s a f unc tion of *CTAB • c ' The overall aggregation number g as a function of *CTAB •T n e °P e n arrow points to the value in the absence of CTAB. d) The c.m.c. versus the aggregation number. The volume fraction of added salt
ASSOCIATION COLLOIDS
4.111
(ii) Stability of linear micelles Let us next consider surfactant concentrations significantly above the c.m.c. At sufficiently high surfactant concentration the overall composition will resemble the composition in the micelles rather than that of the freely dispersed ones. Nevertheless, it remains relevant to know the difference in composition between surfactants in aggregates and those freely dispersed in solution. At these elevated concentrations cylindrical micelles become more prominent. We showed that for the cylindrical C 12 E 5 micelles the endcap energy is approximately 10 kT. Addition of the ionic CTAB surfactant to these micelles will, depending on the ionic strength, modify the wormlike micelles. At very low ionic strength the CTAB prefers spherical micelles. At higher ionic strength, the ionic surfactant is also inclined to form cylindrical micelles. In fig. 4.43 we present a selection of the data of cylindrical micelles composed of a mixture of CTAB surfactants and C 12 E 5 for three values of the volume fraction of added 1:1 electrolyte, (ps =0.006, 0.003 , and 0.001 . The natural control parameter
Figure 4.43. a) The estimate of the endcap energy of cylindrical micelles composed of the binary surfactant mixture CTAB and C 12 E 5 as a function of the fraction of CTAB freely dispersed in solution, b) The aggregation number as a function of *CTAB ' s o u d h'nes f° r the spherical micelle gs (left ordinate); dotted line for the unit length of a cylindrical micelle gc (right ordinate). c) The total volume fraction of surfactants freely dispersed in solution as a function of XQTAB . d) The mole fraction of CTAB in the spherical micelle as a function of X CTAB ' le ft ordinate), and the difference of the mole fraction of CTAB in the spherical micelle and cylindrical micelle A x g p ^ = *CTAB ~ XCTAB a s a f unct ion of *CTAB ( r i S n t ordinate). The ionic strength of the 1:1 electrolyte solution is
4.112
ASSOCIATION COLLOIDS
is the fraction of CTAB freely dispersed in solution *c TAB • In the discussion of these graphs, we pay attention to what happens for very small values of *c TAB anc^ t o t n e trend at larger values of xQTAB . Estimation of the endcap energy e"m as a function of the fraction of CTAB surfactants freely dispersed in solution is presented in fig. 4.43a. The open arrow points to the value found for *CTAB = *-* • ' e - ^or * ne cylindrical micelle of C 12 E 5 only. For all three ionic strengths, we see that small additions of CTAB give an increase of the endcap energy and that this increase is largest for the highest ionic strength followed by a reduction of e*m at larger *CTAB • A high endcap energy is consistent with long micelles and a relatively low surfactant concentration for the sphere-to-cylinder transition. The consequence of a low endcap energy is the suppression of long micelles and a relatively high surfactant concentration before cylindrical micelles appear. With increasing fractions of CTAB freely dispersed in solution, we thus predict an increase in the length of the micelles for very small additions of CTAB followed by the breakdown of cylindrical micelles into spherical ones (major effect). As can be seen from fig. 4.43b, where the aggregation number of the surfactants in the spherical micelle and that of the cylindrical body (per unit length) are presented, it is clear that the aggregation number strongly depends on *Q T A B as well. Again, the open arrows point to the values at vanishing *CTAB • Consistent with the results for the spherical micelles (near their c.m.c.) discussed above, the first incorporated CTAB surfactants will not displace the non-ionic ones. As a result, the aggregation number grows with small additions of CTAB. The trends found for the changes in the aggregation number with increasing values of added CTAB are similar for the spherical and cylindrical micelles. For sufficiently low ionic strength a decrease in aggregation number is predicted, whereas for higher ionic strength the aggregation number goes through a weak minimum before it grows with *CTAB • The maximum in the endcap energy thus correlates with a maximum in the aggregation number. The bulk concentration of surfactants that exists with tensionless cylindrical micelles is also a function of the fraction of CTAB freely dispersed in solution (fig. 4.43c). As expected, with increasing amount of CTAB a decrease of the volume fraction of surfactants in the bulk is observed. As a consequence, the spherical micelles that coexist with the tensionless cylinders must also exist at these lower chemical potentials of the surfactants. Micelles at a low chemical potential tend to have high grand potentials. Here, part of the explanation for the growth of the endcap energy may be found. In fig. 4.43d we finally present the fraction of CTAB in the spherical micelle as a function of the fraction of CTAB freely dispersed in solution and note that the corresponding graph for the fraction of CTAB in the cylindrical body is indistinguishable in this graph from the presented one. To prove this we also present the difference between the composition of CTAB in the spherical micelle and that in the cylindrical body AxJSj.^ = *CTAB ~ -"-CTAB (r'En^ ordinate; dotted lines). Let us first comment on the *CTAB dependence on * Q T A B . In line with expectations, we see that the ionic surfactant is strongly preferentially taken up in the micelle. The mole fraction
ASSOCIATION COLLOIDS
4.113
of the ionic surfactant In the bulk stays behind that in the micelle. This trend is in line with the results discussed above for the micelle near the c.m.c. The preferential uptake of CTAB is made more difficult with decreasing ionic strength. The uptake of charged surfactants in the micelle is only a weak function of the micelle geometry. In line with expectations, the uptake is larger for the spherical symmetry than for the cylindrical one, but the differences are second order. The difference in partitioning of the ionic surfactants between the cylindrical body and spherical ends is therefore not expected to explain the increase in the endcap energy at low fractions of *£ TAB •
4.9b Solubilization qfapolar
compounds
Solubilization of compounds that are virtually insoluble in water is one of the major applications of surfactant micelles. The analysis of solubilization may seem straightforward, but appears fascinatingly rich and complex. The fundamental reason for this is that for these systems two types of phase behaviour merge. The surfactant selfassembly is only first-order-like due to finite size effects, but in addition, the limited solubility of the additive molecules gives rise to a classical first-order demixing. At small amounts of added material, i.e. far from the bulk binodal of the additive, one typically speaks about swollen micelles. At larger added amounts, i.e. nearer to the bulk bimodal, we may enter the microemulsion domain to be discussed in chapter 5. There are no clear demarcation lines between swollen micelles and microemulsions. We will illustrate and outline what is expected for a series of linear alkanes as the additives and limit ourselves to sub-saturation conditions. A quantity of prime interest is the partition coefficient of such compounds, which are defined as the ratio of the concentration of the compound in the micelle and that in the aqueous phase. A unique definition of the partition coefficient is not straightforward because it involves a choice of how to estimate the size of the (core of the) micelle. From the many applications, we chose to present the influence of additives on the micellar geometry and discuss the mechanical properties of swollen bilayers. Our main focus will be on the case when the micelles remain sub-saturated. In the following, we will take C 12 E 5 as the surfactant and limit ourselves to linear alkanes, CN = CH3-(CH2)JV_2-CH3 as the additives. (I) Partition coefficients To define a partition coefficient requires us to specify the size of the core. As the packing volume fraction of the tails in the core of the micelles is not far from unity, we can use the total number of aggregated surfactant molecules in order to find a measure of the core size. Using this, we may compute the partition function by K=
^
/{
^
l(p\
14.9.2]
ASSOCIATION COLLOIDS
4.114
where 0° is the excess amount of a component in the micelle (6° = naN ). Recall that 9 = nc
E
is tne
aggregation number of the surfactant. Here, the volume of the core is
thus estimated by V-_r- ~ (K + 6Z
v
/ 2 where the subindex A refers to the additive.
We can use this Ansatz not only for the partition coefficient in a flat and a cylindrical geometry, but also for the spherical one. One can evaluate the volume of the core more accurately by analyzing the density profiles. We will not do this here. In fig. 4.44a we present the partition coefficient in the limit of low loading of alkanes in micellar aggregates of C 1 2 E 5 surfactants as a function of the length N of the (linear) alkane molecule. We will first focus on a dilute solution (volume fraction of micelles q>m = exp(-15) = 10~7 ) of spherical micelles. For this system the oligomer alkane range is JV = 1 (),•••, 15 and a range of significantly longer alkanes N = 35,•••,40 are selected. To first order the partition coefficient is an exponential function of the length of the alkane, as expected. There are several intricacies worth discussing. In each alkane there are two CH3 end-groups that differ with respect to the interactions with the micelle core from the middle CH2 units. Even upon extreme dilution of additive in the micelle (possible in the cylindrical and lamellar aggregates) the average distribution of the additive inside the micelle depends on JV . The CH3 groups are therefore not always in exactly the same location in the micelle and, hence, there is a deviation from the exponential law (enthalpic effect). Secondly, the aggregate shape has an entropic effect on the partition coefficient K , i.e. with increasing chain length the micellar core imposes some confinement on the coils. Only in the micellar system does
Figure 4.44. a) The partition coefficient K of alkanes in aggregates of C 12 E 5 surfactants as a function of the length N of the alkane CN (JV = 10....,16 and JV = 35,...,40 ); for spherical micelles with grand potential of £ m = 15 kT (open circles with dot) - one alkane molecule per micelle), for tensionless cylindrical micelles (closed symbols) and for tensionless bilayers (open symbols), b) The normalized c.m.c. and normalized aggregation number of spherical micelles of C^Eg surfactants in which a single alkane molecule CN is solubilized as a function of the length of the alkane N. The grand potential of the micelle is e m = 15 kT , the unloaded c.m.c. *= 0.00083601 for which the aggregation number g* = 95.417.
ASSOCIATION COLLOIDS
4.115
the dimension of the core depend on the additive as we fix n^ = 1 in the micelle. In the other two geometries, the amount of additive was chosen sufficiently small so that the micelle was not perturbed. In the oligomeric range, the partition coefficient is found highest for the lamellae and lowest for the spherical micelles. For the polymeric additives, the partition coefficient is highest for the cylindrical micelle and lowest for the lamellae. We thus see that gradually the spherical micelle becomes the better solubilizing geometry, possibly because the confinement is less severe in this case; the radius of the core of the spherical micelle is larger than that of the cylindrical one, which in turn is larger than (half) the bilayer thickness. In fig. 4.44a we show that additives have a measurable effect on the c.m.c. and on the aggregation number, even when the additive molecule is so small that only a little material is transferred into the micelle. The drop in c.m.c. is on the order of 1 % and has a non-monotonic dependence on JV. The aggregation responds stronger to the incorporation of one foreign molecule. With increasing chain length of the additive, the aggregation number increases to approximately 10 % for N - 40 . These numbers will change, of course, when more than one additive molecule per micelle is allowed. (ii) Stability of swollen linear micelles In the previous section, we have considered the ideal partition coefficient defined in the limit of low loading. In practice, however, one cannot easily measure this; typically both the additive/surfactant ratio must be sufficiently large and the surfactant concentration should be significantly above the c.m.c. to measure partition coefficients accurately. However, under these conditions the spherical micelle shape may be lost. For this reason, it is of interest to investigate how alkane additives influence the sphere-to-rod transition in micellar solutions. In fig. 4.45a we present the grand potential of the spherical micelles that coexist with tensionless cylindrical assemblies (micelles) into which gradually more and more alkanes were dissolved. The grand potential is an accurate measure of the endcap energy of cylindrical micelles and it is also correlated to the second c.m.c. (see sec. 4.6.5). Data are given for C l o up to C16 . It is shown that the endcap energy grows with the increasing number of dissolved additives in the micelle. For low loading the increase is linear and almost independent of the length of the additive molecule. The increase of the endcap energy means that the second c.m.c. decreases with increasing loading and that in the phase diagram spherical micelles are less prominently present. A phenomenological explanation for the relative decrease of stability of spherical micelles and the relative stability of cylindrical ones may be found considering the surfactant packing parameter P . The alkanes apparently increase the volume occupied by the tails, whereas the radius of the micelle is still linked to the length of the surfactant molecule. The area per molecule may be only a weak function of the loading
4.116
ASSOCIATION COLLOIDS
Figure 4.45. a) The estimate of the endcap energy (grand potential of the spherical micelle that coexists with the tensionless infinitely long cylindrical micelle) as a function of the number of additive molecules in the spherical micelle n^ . b) The number of incorporated alkanes in spherical micelles as a function of the bulk concentration of alkanes. The bulk saturation value (bimodal) of each alkane is indicated by the vertical lines. Four alkanes CN were used with JV = 10. 16 as indicated. The C ] o graph has closed labels.
and therefore the value of P tends to increase with increasing loading. As a result the cylindrical micelles become more important. The uptake isotherm of additive in the spherical micelles that coexists with long cylindrical micelles is given in fig. 4.45b in double logarithmic co-ordinates. At low absorbed amounts, there is a straight line indicating that the partition coefficient is independent of the loading. At a higher loading, approaching the bulk binodal of the additive (indicated by the vertical lines in fig 4.45b) the absorption isotherm swings upwards. This means that the absorption is co-operative and that the partition coefficient increases with loading. In passing, we note that for very large alkanes anticooperative behaviour is found, which implies that the partition coefficient decreases upon loading (not shown). The uptake isotherms terminate at some level of uptake and this limit increases with the increasing length of the alkane. The isotherms do not continue because the swollen spherical micelle becomes structurally unstable. This type of instability does not occur for sufficiently long alkane additives (e.g. for C16 and up). Below we will evaluate what happens with the uptake isotherm near the saturation value for the alkane in water in more detail. It is illustrative to show the radial density profiles of swollen micelles. We choose for this the Cj 6 swollen micelles because these molecules swell the C 12 E 5 surfactant micelles without destroying them. In fig. 4.46 a selection of these profiles for n£ = 10,100,1000 and 10000 are presented. The alkane additive obviously is situated all over the core of the micelle. The surfactant has a dip in density at the position where the additive is situated leading to a more or less homogeneous density of apolar molecules in the core, very similar to the value found in the absence of additives. It is also interesting to consider the corresponding grand potential density profiles for these systems. We mention once again that the grand potential density is not uniquely defined. However, when the apolar centre is made up of a homogeneous oil phase, one
ASSOCIATION COLLOIDS
4.117
Figure 4.46. a) Collection of radial density profiles of C 16 -swollen spherical micelles composed of C^Eg surfactants for four values of the number of absorbed additive molecules as indicated. The overall surfactant volume fraction profiles are solid lines; those of the additives are dotted. Lines with corresponding labels correspond to each other. The spherical micelles have a grand potential of 15 kT . b) The corresponding grand potential density profiles in units kT/l3 . For three lines, an upward shift is implemented for better visibility: of 0.01, 0.02 and 0.03 for the crosses, the top-down triangles and the open circles respectively.
must observe a well-defined osmotic pressure inside the micelle (emulsion droplet). Inspection of fig. 4.46b shows that for n^ = 10 , 100 in the core of the micelle, there are still gradients in grand potential density, and for larger values there is a tendency that the grand potential density levels off to some (well-defined) homogeneous value corresponding to a homogeneous osmotic pressure, which is found when the interfacial surfactants no longer reach the centre of the micelle. As the grand potential of the spherical micelles is fixed to 15 kT , it is clear that the osmotic pressure in the swollen micelle is a decreasing function of the radius. (til) Bending moduli of loaded bllayers At some higher surfactant concentration, the C 12 E 5 surfactant forms a lamellar phase. It is of interest to investigate how the mechanical parameters of the surfactant layers change when they accommodate apolar alkanes in their hydrophobic domain. The procedure to compute these quantities has been explained several times before and it suffices here to show the result of C 12 E 5 bilayers that are exposed to gradually higher amounts of C10 . The fraction of the additive in the non-ionic bilayer is computed as / „ = 6Z R9Z + 0Z _ ) where 9° = naxN is the total amount of a c
10
c
10
u
10
C
12t5
species per unit area. In fig. 4.47 it can be seen that a small loading with decane increases the mean bending modulus, but does not affect the saddle splay modulus significantly, whereas at higher loadings Jc > 0.2 the mean bending modulus starts to decrease and the Gauss bending modulus increases sharply. If / c could have been increased up to 0.35 we would have made the Gauss bending modulus positive. However, freely dispersed C 12 E 5 bilayers did not allow more decane to be dissolved per unit area. We
4.118
ASSOCIATION COLLOIDS
Figure 4.47. The total curvature energy of spherical vesicles e^ (left ordinate) and the mean and Gauss bending modulus as a function of the weight fraction of Cj 0 in the C^Eg lamellar aggregates.
expect that when interacting bilayers are forced to take up more decane that the Gauss bending modulus will become positive. That is why the decane-C12E5 system must have a pronounced L3 (or sponge) phase. The fact that apolar compounds tend to increase the Gauss bending modulus (making it less negative) of lipid bilayers was also discussed by Claessens11. It is known in the literature that apolar compounds promote the fusion of vesicles, and less negative values of the saddle splay modulus thus allow vesicles to fuse more easily. In biological systems the physical characteristics of bilayers are also controlled by incorporation of additives such as cholesterol. One reason for a biological system to play with the amount of additives is to maintain a marginal topological stability of the bilayer. A marginal stability of the bilayer may allow large perturbation of the local topology without too much energy input. 4.10 Kinetic aspects of surfactant solutions near the c.m.c. The main message of this chapter was the application of a versatile model that helps to understand the various structures that association colloids can assume under equilibrium conditions. This theory has not yet been applied to account for the rates of the transitions between such states, although experimental evidence is available for some cases. By way of illustration we now briefly describe the kinetics of (de-) micellization. For micelle formation the main kinetic questions are: (i) in which way a micelle is generated upon a change of conditions in favour of their formation, and (ii) how they decay, say, upon dilution. These kinetics can be experimentally investigated21 and such studies are informative about equilibrium micelles as well. The central message is that 11
M.M.A.E. Claessens, loc.cit. E.A.G. Aniansson, S.N. Wall, M. Almgren, H. Hoffmann, I. Kielmann, W. Ulbricht, R. Zana. J. Lang, and C. Tondre, J. Phys. Chem. 80 (1976) 905. 21
ASSOCIATION COLLOIDS
4.119
there are two main relaxation processes, characterized by the relaxation times rl and r2 . They differ by as much as two to three orders of magnitude ( Tj « r 2 )• For c = 0.01 M SDS Tj =15 //s and r 2 = 1.8 ms. The fast process has been assigned to the association-dissociation equilibrium of isolated surfactants to/from micelles. Typically it is found that the association reaction constant k+ is nearly diffusion controlled, whereas the rate of dissociation constant k~ (removal of one surfactant from a micelle) depends strongly on the hydrophobicity of the surfactant. This is similar to the rates of adsorption in, and desorption from, an adsorbed monolayer (sec. III.4.5). The c.m.c. is related to the ratio k~//c + = c.m.c. For SDS the c.m.c. = 0.0083 mol, and k~ = l x l O 7 s" 1 , and k+ = 1.2 xlO 9 m o l ^ s " 1 . The ability of the surfactant solution to respond to changes in conditions is intimately linked to the equilibrium fluctuations in the size (aggregation number). Detailed analysis shows that for the fast process Tx=—^ + k cri (c.m.c. x g)
[4.10.1]
where g is the most probable aggregation number and a is the standard deviation of the micelle size distribution function. The slower process establishes the new size distribution, i.e. the time r 2 is needed to equilibrate the number of micelles and the distribution of micelle sizes. It is given by
^^Jj^rfEzEEfilV1 c.m.c. R^y
y
[4.10.2]
c.m.c. g)
As the slow relaxation time ( r 2 ) is attributed to the micellization-dissolution equilibrium (micellar lifetime), it is related to the breakdown of entire micelles. The kinetic properties of micelles with the least probable aggregation number determine the resistance Rj. Interestingly, through the concentration dependence of r 2
o n e ma
Y
obtain the relative fluctuations in the micelle aggregation number a^ I g . Finding two such relaxation times fits very well with the probability distribution of micelle size (fig. 4.5a) demonstrating a deep minimum at intermediate sizes. The implication is that formation and degradation via such intermediates has a high activation of Gibbs energy. To find estimates for this, i.e. find reasonable values for Rj, one needs to have an idea how the critical nucleus of the formation of the micelle looks like. This may depend strongly on physicochemical conditions such as ionic strength and temperature. Basically the modeling in this chapter can help to get insight into this state. There are many other relevant kinetic aspects in surfactant solutions, too many to be covered in any detail in this chapter. A few examples will suffice: the time needed for a surfactant molecule to change from one to the other monolayer in surfactant bilayers (flip-flop), the shape fluctuations of bilayers, the kinetics of the phase change from the lamellar Lff to the L 3 (sponge) phase, the kinetics of the formation of linear micelles
4.120
ASSOCIATION COLLOIDS
from concentrating the spherical micelles, the closing of disks to form (spherical) vesicles, the uptake (release) of additives in micelles, etc. Similarly, as in the wellknown micellization/demicellization example, for all these cases the analysis requires detailed insight into the equilibrium characteristics. We showed above that for many topics this type of information seems to be available. We therefore expect that the kinetic phenomena in surfactant solutions will remain a fruitful field of research in the near future. 4.11 Outlook Surfactant self-assembly is a beautifully rich and, therefore, complex topic. Many aspects are now at least in first order understood, partly because the molecular modelling has advanced to a rather sophisticated level. The most detailed statistical technique is molecular dynamics. For example, all that one wants to know about the lipid bilayer membrane on the nm length scale and the ns timescale is in principle available from modern MD simulations. All-atom versions of this method are, however, still too time-consuming to elaborate sufficiently large systems and obtain information on structural, mechanical, thermodynamic and eventually dynamic information on experimentally relevant phenomenon (in terms of length scales and time scales) in many self-assembling systems. Coarse-grained elaborations, which are now actively developed, will become important in the near future. In such procedures, one has to drop uninteresting degrees of freedom and maintain the interesting ones. With these less detailed approaches there is good hope for obtaining information on the formation and fusion of vesicles, nucleation of pores in bilayers, quantification of shape fluctuations in lamellae, etc. So, many challenges remain. The self-consistent field theory may be used to obtain systematic information on structural, mechanical and thermodynamic data on many self-assembly issues. However, one should keep in mind that this is not a rigorous approach because a mean field approximation is implemented. The success of the SCF theory for densely packed surfactant layers may be understood from the fact that each surfactant molecule continuously interacts with many neighbours. As a result, the actual surroundings, as accounted for in an exact theory, do not differ very much from the averaged surrounding. As a result, we find only minor differences between simulations and SCF calculations. Analytical quasi-macroscopic approaches are useful mainly to identify the physics of various aspects of self-assembly. In combination with computer-guided tools, these methods will help to upgrade surfactant science from a merely experimental to a theory-guided discipline. There are many specific molecular details that influence the self-assembly characteristics, but fortunately there are also many generic aspects. The practical success of the packing parameter P is a proof. Whether these generic features can also help us to
ASSOCIATION COLLOIDS
4.121
understand micelles of different internal structure, such as complex coacervates, or find regularities in surfactant self-assembly in supercritical solutions, still remains to be seen. Surfactants are of low molecular weight as compared with polymer analogues. In polymer self-assembly, there are hopes that scaling laws may be found that help to find regularities in self-assembly as well. In any case, molecular realistic SCF modelling can be used in most of these systems, which will help to make progress in these and related systems. A suitable description of systems that combine self-assembly with long-range order (as occurring in the lamellar or hexagonal phases) is one of the major open problems in the field. Perhaps even more challenging are many open issues related to dynamics. The relevance for biological problems in combination with the promise
for
nanotechnology will keep self-assembly in the focus of modern science for a long time to come. 4.12 General references
4.12a General Physics of Amphiphiles: Micelles, Vesicles and Microemulsions, V. Degiorio and M. Corti, Eds., North Holland (1985). (A book with many interesting topics that still is worth inspection.) D.F. Evans, H. Wennerstrbm, The Colloidal Domain where Physics,
Chemistry,
Biology, and Technology Meet, VCH Publishers (1994). (This introductory text on colloid science contains much information on surfactant self-assembly.) Micelles, Membranes, Microemulsions and Monolayers, W.M. Gelbart, A. BenShaul, and D. Roux. Eds., Springer-Verlag (1994). Interesting collection of 12 state-ofthe art chapters written by experts on various aspects of association colloids. Topics range from statistical thermodynamics of amphiphile self-assembly, micellar growth (worm-like micelles), micellar liquid crystals, geometric foundation of mesomorphic polymorphism, effects of fluctuations in lamellar phases (both theory and experiments), the structure of microemulsions (theory and experiments), interfacial tension (theory and experiments), critical behaviour of surfactant solutions and structure and phase transitions in Langmuir monolayers.) R.G. Laughlin, The Aqueous Phase Behavior of Surfactants,
Academic Press
(1994). (Large collection of phase diagrams of surfactant in aqueous systems. This book gives, besides a huge collections of experimental data, a good introduction into the phase rule and the role of phase science in physical chemistry.)
4.122
ASSOCIATION COLLOIDS
J.R. Lu, R.K. Thomas and J. Penfold, Surfactant Layers at the Air/Water Interface: Structure and Composition, Adv. Colloid Interface Set 84 (2000) 143-304. (Review, 316 refs. demonstrating the potentialities of neutron scattering, and critical comparison of the results with those from other techniques.) P. Mukerjee, The Nature of the Association Equilibria and Hydrophobic Bonding in Aqueous Solutions of Association Colloids, Adv. Colloid Interface Sci. (1967) 24175. (Review with 130 refs.; historically interesting.) A.I. Rusanov, The Mass Action Law Theory of Micellar Solutions, Adv. Colloid Interface Sci. 45 (1993) 1-78. (Review, 69 refs., mostly theoretical; thermodynamics, counterion binding.) S. Safran, Statistical Thermodynamics of Surfaces, Interfaces and Membranes. Frontiers in Physics, D. Pines, Ed. Addison-Wesley Publishing Comp. (1994). (Interesting monograph on various surfactant systems from a physicists point of view, e.g., surfactant layers are seen as elastic sheets for which the bending moduli are the phenomenological coefficients.) C. Tanford, The Hydrophobic Effect. Formation of Micelles and Biological Membranes, 2nd ed. New York: Wiley and Sons Inc. (1980). (This book is 25 year after its appearance still full of interesting data on relevant topics such as the solubility of hydrocarbons in water, the anomalous entropy and heat capacity when apolar compounds are dissolved in water ("the hydrophobic effect"), the thermodynamics of micelle formation, etc. It also gives an introduction to what was known about the biological membrane at that time.) C. Tondre, C. Caillet, Properties of the Amphiphilic Films in Mixed Cationic/ Anionic Vesicles: a Comprehensive view from a Literature Analysis, Adv. Colloid Interface Sci. 93 (2001) 115-34. (Review, 78 refs., much experimental information on mixed vesicles.) I.F. Uchegbu, A.T. Florence, Non-ionic Surfactant Vesicles (niosomes): Physical & Pharmaceutical Chemistry, Adv. Colloid Interface Sci. 58 (1995) 1-55. (Review, 100 refs., emphasis on factural information.) R. Zana, Aqueous Surfactant-Alcohol Systems: a Review. Adv. Colloid Interface Sci. 57 (1995) 1-64. (Review, 183 refs., Influence of various alcohols on cm.a, micellar size, shape, dynamics and interactions; phase diagrams, microemulsions.)
ASSOCIATION COLLOIDS
4.123
R. Zana, Dimeric and Oligomeric Surfactants. Behavior at Interfaces and in Aqueous Solution: a Review. Adv. Colloid Interface Sci. 97 (2002) 205-253. (Review, 206 refs. geminis and oligomers, below and above the c.m.c., structure, dynamics, rheology, interactions.) A. Zapf, R. Beck, G. Platz, and H. Hofffmann, Calcium Surfactants: a Review, Adv. Colloid Interface Sci. 100-102 (03)349-80. (Review, 103 refs. of the phase behaviour, Krafft point properties, swelling and suppression of double layer interaction phenomena.) 4.12b Relevant series Surfactant Science Series: A series of over a hundred books, each with a guest editor. Senior editors, M.J. Schick and A.T. Hubbard, Marcel Dekker. The series reflects the progress and richness of the field. Some of the more recent issues do not deal with surfactants and those dealing with surfactants mostly address surfactants at interfaces, their synthesis, analysis and applications. Volumes in these series, dealing with surfactants in solution include: 1, 23 Non-ionic Surfactants, M.J. Schick, Ed.; 2 Solvent Properties of Surfactant Solutions, K. Shinoda, Ed.; 4 Cationic Surfactants, E. Jungermann, Ed.; 7 Anionic Surfactants, W.M. Linfield, Ed.; 11 Anionic Surfactants; Physical Chemistry of Surfactant Action, E.H. Lucassen-Reynders, Ed.; 22 Surfactant Solutions: New Methods of Investigation, R. Zana, Ed.; 37 Cationic Surfactants: Physical Chemistry, D.N. Rubingh and P.M. Holland, Eds.; 46 Mixed Surfactant Systems, K. Ogino and M. Abe, Eds.; 55 Solubilization in Surfactant Aggregates, S.D. Christian and J.F. Scamehorn, Eds., 62 Vesicles, M. Rosoff, Ed.; 64 Surfactants in Solution, A.K. Chattopadhyag, and K.L. Mittal, Eds.; 74 Novel Surfactants, K. Holmberg, Ed.; 109 Adsorption and Aggregation of Surfactants in Solution, K.L. Mittal and D.O. Shah, Eds. Surfactants in Solution A series of conference proceedings edited by K.L. Mittal. Useful because of its large collection of experimental data but contributions may only marginally be reviewed. 4.12c Tabulation P. Mukerjee, K.J. Mysels, Critical Micelle Concentrations of Aqueous Surfactant Systems. Natl Bureau of Standards. NSRDS-NBS-36 (1971). (Critically evaluated c.m.c. data, referring to the older literature; data on well-defined non-ionics are underexposed.)
This Page is Intentionally Left Blank
5
MlCROEMULSIONS
Thomas Sottmann and Reinhard Strey 5.1
Introduction and definitions
5.1
5.2
Phase behaviour
5.4
5.2a
Binary systems
5.4
5.2b
Ternary systems
5.5
5.2c
Phase inversion
5.6
5.2d
The optimum state
5.2e
Efficiency
5.13
5.2f
Monomeric solubility
5.15
5. 2g
Traj ectory of the middle phase
5.17
5.2h
Scaling
5.18
5.2i
Straight-line binodal curves
5.3
5.4
5.5 5.6
5.7
5.9
5.20
Microstructure
5.24
5.3a
Methods
5.26
5.3b
Transmission electron microscopy
5.27
5.3c
Scattering techniques
5.32
5.3d
SANS
5.33
5.3e
NMR diffusometry
5.40
5.3f
Electric conductivity
5.43
5.3g
Overview of microstructure
5.45
5.3h
Length scales
5.46
Ultra-low interfacial tensions
5.55
5.4a
Ultra-low interfacial tensions due to amphiphile adsorption
5.56
5.4b
Ultra-low interfacial tensions versus phase behaviour
5.58
5.4c
Wetting behaviour
5.60
5.4d
Interfacial tension curves
5.61
5.4e
Scaling of interfacial tensions
5.63
Theoretical description
5.65
5.5a
5.67
Membrane theories
Applications
5.76
5.6a
Technical-grade mixtures of non-ionic surfactants
5.77
5.6b
Alkylpolyglucoside microemulsions
5.79
5.6c
Ionic microemulsions
5.83
5.6d
Microemulsions with non-ionic and ionic surfactants
5.88
5.6e
Amphiphilic block-copolymers as efficiency booster
5.90
General references
5.95
This Page is Intentionally Left Blank
5 MICROEMULSIONS THOMAS SOTTMANN AND REINHARD STREY
S. 1 Introduction and definitions Microemulsions are macroscopically isotropic mixtures of at least three components: water, oil and surfactant. They are single thermodynamically stable phases, different from ordinary emulsions (chapter 8). Microscopically, the surfactant molecules form an extended interfacial monolayer separating the water from the oil molecules. The preferential adsorption of the surfactant reduces the interfacial tension between the polar and non-polar solvent effectively to zero, which, in turn, permits thermal energy to disperse the two incompatible solvents into each other. The general features of the phase behaviour of microemulsions are best introduced by considering the following simple experiment. A simple experiment: We take a test tube with equal amounts of water and oil. As water and oil do not mix, we see two phases, water (A) forming the bottom phase, oil (B) forming the phase on top. This situation is shown by the test tube furthest on the left in fig. 5.1. When we add a surfactant, it has, in principle, three options. It can dissolve in the water phase (fig. 5.1, tube I), it can dissolve in the oil phase (fig. 5.1, tube II) or it can make up its own phase (fig. 5.1, tube III). These situations are frequently observed and are denoted by the three Winsor states (I, II, III), after Winsor, who was the first to study this behaviour systematically11. The surfactant rich phase is called the microemulsion. We will explain further how to select the components to achieve a desired microemulsion type, which structures and properties to expect and provide hints for applications of microemulsions. As illustrated in fig. 5.1, the first observation dealing with mixtures of water, oil and surfactant is the spontaneous appearance of different phases. Therefore studying the phase behaviour and the construction of phase diagrams is the first step. A phase diagram may be viewed as the road map for the researcher and helps him to reach his destination or goal. For example, the Gibbs triangle on the left-hand side of fig. 5.1 indicates that at intermediate temperatures a hydrophilic surfactant system is over wide composition regions in the Winsor I (2)-state. The Gibbs triangle on the right" P.A. Winsor. Solvent Properties of Amphiphilic Compounds. Butherworth & Co. (1954). Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
5.2
MICROEMULSIONS
Figure 5.1. Approaching microemulsions at room temperature ( T = 25°C). Top: test tubes of, left, water and n-octane, a = 0.413 , wc = 0 ; tube I: H2O -n-octane- C 10 E 5 , a = 0.413 , u>c = 0.05, Winsor I (2); tube II, H2O -n-octane- C 10 E 3 , a = 0.413 , wc = 0.05 , Winsor II ( 2 ); tube III, H 2 O-n-octane-Cj 0 E 4 , or = 0.413, wc = 0.05 , Winsor III (3). Bottom: isothermal Gibbs triangles of the three systems at T= 25°C. a and WQ are mixing ratios, see [5.2.1 and 2].
hand side of fig. 5.1 shows a hydrophobic surfactant system displaying a Winsor II ( 2 )behaviour. The lowest triangle in fig. 5.1 shows the extended three-phase triangle typical for the Winsor III (3) state of a balanced surfactant system, which corresponds to the optimal state of each mlcroemulsion system, as we will explain in the following. The selection of the components and their concentrations, but also of the state variables temperature T and pressure p, is crucial for the type of microemulsion that is obtained and its properties upon dilution. In general, microemulsions can contain any number of polar, non-polar and amphiphilic components. Thus, the choice of components is abundant. However, they may be
MICROEMULSIONS
5.3
grouped into water-soluble, oil-soluble and components that adsorb at the interface and determine the properties of the amphiphilic film. Winsor11 studied microemulsions of five-component systems, including ionic surfactants. He induced a phase progresssion like those seen in fig. 5.1, e.g. by varying the salinity. The key observation was that of phase inversion, i.e. the propensity of progressing gradually from the situation depicted in tube I to that in tube II. The effect of any other component mentioned above may be judged by studying the phase behaviour quantitatively. However, Kahlweit et al. showed that reduction to the three components water (A), oil (B) and a single surfactant (C) suffices as the road to success. That microemulsions can be achieved in ternary systems and that in these systems the phase inversion can be induced by tuning the temperature was first demonstrated by Shinoda3'. He also showed that the Winsor III situation of tube III, occurs between I and II. In systematic studies, Kahlweit et al. demonstrated that this phase behaviour is generic, i.e. it is similar for a large variety of different systems. Microemulsions may contain water and oil in any ratio. They are isotropic liquids of low viscosity. The interfacial tension between a microemulsion and a water or oil-rich excess phase is, in general, very low. Shinoda and Friberg4' showed that the interfacial tensions become ultra low within the Winsor III region. Schechter et al.5' found an interrelation between the efficiency of the surfactant in solubilizing water and oil and the value of the water/oil interfacial tension. Lindman et al.6) emphasized the bicontinuity of the most efficient microemulsions observed as the phase inversion is traversed. Strey7' used small-angle neutron scattering (SANS) to quantitatively determine the gradual change in type and size of the microstructure and emphasized that they are continuously connected in phase space. He showed that the minimum of the interfacial tension correlates with the bicontinuity and the largest length scale of the microstructure. In the next section, we will show by example of a simple water-oil-short chain surfactant system how the phase progression works in principle. From this progression it will become clear that the observations in fig. 5.1 are isolated parts of a general picture. Then the transition to medium chain and long chain surfactants is made giving rise to enormous increases of efficiency of the surfactant. In the end, we shall show that 11
P. A. Winsor, loc. cit. M. Kahlweit and R. Strey, Angew. Chem. Int. 24 (1985) 654; M. Kahlweit, R. Strey and G. Busse, J. Phys. Chem. 94 (1990) 3881; M. Kahlweit, R. Strey and G. Bussc, Phys. Rev. E47 (1993) 4197. K. Shinoda. In Solvent Properties of Surfactant Solutions; K. Shinoda, Ed.; Marcel Dekker, 1967; p27. 41 K. Shinoda and S. Friberg, Adv. Colloid Interface Set 4 (1975) 281. 51 R.S. Schechter, W.H. Wade, U. Wcerasooriya, V. Weerasooriya and S. Yiv, J. Dispersion Sci. Techn. 6(1985) 223. B. Lindman, K. Shinoda, U. Olsson, D. Anderson, G. Karlstrom and H. Wennerstrom, Colloids Surfaces 38 (1989) 205. 71 R. Strey, Colloid Polym. Sci. Ill (1994) 1005. 21
5.4
MICROEMULSIONS
the phase boundaries themselves already contain much important information, which is easily gathered. We just have to learn to see it. 5.2 Phase behaviour Meaningful experiments on microemulsions can only be conducted if the phase behaviour is well known. Often it turns out that after the study of the phase behaviour the examination of microstructure, interfacial tension and solubilization capacity can be kept at a minimum because many properties are similar in corresponding parts of the phase diagram. Thus, we start by considering the general patterns of the phase behaviour of microemulsions. As model systems, ternary mixtures of the water (A)-oil (B)-non-ionic surfactant (C) type are chosen. Furthermore, we confine ourselves to the effect of temperature on the phase behaviour since the effect of pressure is rather weak11. One successful approach to understand the already complex behaviour of this type of microemulsions is to consider first the phase diagrams of the three binary systems, water (A) - oil (B), oil (B) - surfactant (C) and water (A) - surfactant (C). 5.2a Binary systems Figure 5.2 schematically shows the phase diagrams of the three binary mixtures. The simplest phase diagram of the three is that of the water (A) - oil (B) system (fig. 5.2, left). As water and oil are immiscible the upper critical point of that miscibility gap lies far above the boiling point of the mixture. The phase diagram of the binary oil (B) non-ionic surfactant (C) system is just as simple. It shows a lower miscibility gap with an upper critical point cp a (see fig. 5.2, right) often near ambient temperature. The exact location of cp a depends on the nature of both oil and surfactant. The lower it is, the more hydrophilic the oil and the more hydrophobic the surfactant is. Sometimes it lies below the melting point of the mixture (for details see e.g. fig. 6 in2)). The phase diagram of the binary water (A) - non-ionic surfactant (C) system is the most complex of the three. Far below the melting point of the mixture one finds a lower miscibility gap, which plays no role in further considerations. At ambient temperatures and above the critical micelle concentration (cm.a), the surfactant molecules form association colloids. Additionally, concentrated and dilute liquid crystalline phases can be found31. At higher temperatures most of the systems show an additional upper (closed) miscibility gap with a lower critical point cp». The importance of this miscibility gap for the formation of microemulsions is pointed out in fig. 5.2, centre, from which both the liquid crystalline phases and the lower miscibility gap are omitted.
11
C.L. Sasscn, A.G. Cassicllcs, T.W. dc Loos, and J. dc Swaan-Arons, Fluid Phase Equilibria 72 (1992) 173: G.M. Schneider, Pure Appl. Chem. 55 (1983) 479. 21 M. Kahlweit R. Strcy, loc. cit. 31 R.G. Laughlin, Aqueous Phase Behavior of Surfactants, Academic Press (1994): R. Strcy, Ber. Bunsenges. Phys. Chem. 100 (1996) 182.
5.5
MICROEMULSIONS
Figure 5.2. A schematic view of the three binary phase diagrams, water (A) - oil (B), oil (B) non-ionic surfactant (C), water (A) - non-ionic surfactant (C). The most important features are the upper critical point cpa of the B-C phase diagram and the lower critical point cpp of the binary A-C diagram. Thus, at low temperatures water is a good solvent for the surfactant, whereas at high temperature the surfactant becomes increasingly soluble in the oil. The thicker lines denote the phase boundaries; the thinner ones are the tie-lines. From fig. 5.2 centre and right, it can be anticipated that the phase behaviour of the ternary system with temperature is a result of the interplay between the lower miscibility gap of the B-C mixture and the upper miscibility gap of the A-C mixture. At low temperatures, the non-ionic surfactant is preferentially soluble in water, at high temperatures in oil. Thus, an increase in temperature turns a non-ionic surfactant from hydrophilic into hydrophobic. Keeping this in mind, we will now discuss the phase behaviour of the ternary system in detail. 5.2b Ternary systems A ternary system is described by four thermodynamic variables, namely p, T and two composition variables. Because in these mixtures the effect of pressure is weak compared with that of temperature, one may dispense with it. In fact, p is usually kept
Figure 5.3. Phase prism in the temperaturecomposition space. The two grey highlighted sections represent common methods for representing the phase behaviour. The T{wc) sec. at constant a typifies the so-called isoplethal method for conveniently studying the occurring phase. From such sections, the Gibbs triangles (exact isothermal representations) can be constructed. We emphasize and demonstrate below the utility of the isoplethal method for multi-component systems.
5.6
MICROEMULSIONS
constant. Then the phase behaviour can be exactly represented in an upright phase prism with the Gibbs triangle A-B-C as the base and T as the ordinate with the three binary systems of fig. 5.2 forming the sides of the prism. A schematic prism is shown in fig. 5.3. Each point in the phase prism is unambiguously defined by T and two composition variables. It has been proven useful11 to choose the mass fraction of the oil in the mixture of water and oil a=
mR mA+mB
[5.2.1]
and that of the surfactant in the mixture of all three components wr =
m^ ^ mA+mB+mc
[5.2.2]
as the variables. Knowing the densities of the components for calculating the volumes, the volume fractions
5.2c Phase inversion The procedure to determine the phase inversion is illustrated in fig. 5.4. It shows a T[wc) section of the system H2O-n- dodecane-n-butyl monoglycol ether (C 4 Ej) at a constant oil-to-water-plus-oil mass fraction a = 0.3 2 ) . Depending on the problem addressed, any other a (i.e. 0 . 1 < a < 0 . 9 ) can be selected. In order to obtain rapid phase separation, we choose here the rather inefficient short chained C 4 Ej. In general, C n E x surfactants are favourable model surfactants since the hydrophilicity, (or hydrophobicity), as well as the efficiency of the system can be gradually tuned by varying n and x . As can be seen from fig. 5.4, in such a T(ix>c) section, the phase boundaries 11 2)
M. Kahlweit and R. Strcy, loc. cit. S. Buraucr, T. Sachert, T. Sottmann and R. Strcy, Phys. Chem. Chem. Phys. 1 (1999) 4299.
MICROEMULSIONS
5.7
Figure 5.4. T(U>Q) section of the system H2O-/1- dodecane- C4Ej at a constant oil-to-waterplus-oil mass fraction a = 0.3 . At low temperatures a surfactant-rich water phase (a) coexists with an excess oil phase (b) (denoted as 2). At high temperatures the situation is inverted (denoted as 2 ); a surfactant-rich oil phase (b) coexists with an excess water phase (a). In between, i.e. in the phase inversion regime, three phases (3) can be observed. The point X defines the minimum amount of surfactant U>Q needed at the temperature T to solubilize both water-oil in a single phase (1). Therefore, wc is a measure of the efficiency of the surfactant. resemble the shape of a fish. Starting with the binary H 2 O -n-dodecane system two phases, a pure water and a pure n-dodecane phase, coexist over the entire experimentally accessible temperature region. Small amounts of added C 4 Ej molecules dissolve monomerically in the two phases. Being amphiphilic, the C n E x molecules preferentially adsorb at the macroscopic interface. At a mass fraction wc 0 both excess phases and the macroscopic interface are saturated with C^Ej so that the C^Ej molecules cannot help but form aggregates. As a function of temperature different aggregates can be found. Since at low temperatures water is a good solvent for the surfactant (see fig. 5.1, test tube I), oil-swollen micelles in a continuous water phase (a) coexist with an oil-excess phase (b). This situation is denoted as 2 or Winsor I. At high temperatures the inverted situation is found. Here, oil is the better solvent for the surfactant than water (see fig. 5.1, test tube II). Thus, water-swollen micelles in a continuous oil phase (b) coexist with a water excess phase (a) (2 or Winsor II). At intermediate temperatures the surfactant is almost equally soluble in both solvents. Here, three phases (3 or Winsor III), i.e. a surfactant-rich bicontinuously structured (for details see sec. 5.3) phase (c), an excess oil and water phase coexist. Increasing wc further, the volume of the third phase increases (see test tubes in fig.
5.8
MICROEMULSIONS
5.4) until the three-phase body meets the one-phase region at point X. Thus, the X point defines both the minimum amount of surfactant wc needed to solubilize water and oil, i.e. the efficiency of the surfactant, as well as the corresponding temperature T, which is a measure of the phase inversion temperature (p.i.t.). At even higher surfactant mass fractions, the one-phase region widens. Experimentally the X point can be determined by extrapolations of the phase boundaries from 2 (turbid) to 1 (clear). However, the exact determination of the entire T(wc) section can be time-consuming, since two- and three-phase states are both turbid and therefore difficult to distinguish without awaiting the phase separation. The body of heterogeneous phases, i.e. the phase behaviour inside the entire phase prism, can be obtained by stacking isothermal Gibbs triangles on top of each other. The Gibbs triangles themselves can be constructed for each temperature from experi-
Figure 5.5. Gibbs triangles of the system H2O-n- dodecane- C4Ej at nine different temperatures. The full lines denote the phase boundaries, the dashed lines schematically represent the tie-lines and the filled squares mark the critical points. Note that the tip of the three-phase triangle corresponds to the X point. At T = 52°C the surfactant C4Ej is at its optimal state, as reflected by the almost symmetrical three-phase triangle. (Redrawn from S. Burauer et al., loc. cit.)
MICROEMULSIONS
5.9
mental T(wc) sections by reading off the composition of the phase transitions at the respective temperature. In fig. 5.5 the Gibbs triangles of the system H2O -n-dodecaneC4Ej are shown at nine different temperatures. They are constructed from ten T(wc) sections as well as data for the binary H2O C 4 Ej system11. From fig. 5.5 one can see how the phase behaviour evolves with increasing temperature. At low temperatures (T = 22°C) an extended central miscibility gap exists, which originates from the binary water-oil system. The inclination of the tielines results from the preferred solubility of C4Ej in water and indicates that a surfactant-rich water phase (a) coexists with an oil-excess phase (b). At Tj = 31.23°C the three-phase triangle evolves from the lower critical tie-line, which connects a critical phase (a) at the critical endpoint cepn (shown as a filled square) with phase (b). With increasing temperature the critical phase separates into two phases, one increasingly water-rich (a) and the other increasingly surfactant-rich (c). As a result, the three-phase triangle opens with a small 2 phase region at the water-rich side and a large 2 phase region at the oil-rich side (shown at T = 37°C). Increasing the temperature further, the 2 phase region shrinks, while the 2 phase region grows and touches the binary waterC 4 E, side of the Gibbs triangle at 49.2°CU. Near T = 52°C the extension of the threephase triangle reaches a maximum. Here the triangle is almost symmetric which, in turn, implies the solubilization of equal volumes of water and oil, implying that the surfactant is at its optimum state. Looking at the tip of the three-phase triangle (X point), I.e. the composition of the surfactant-rich middle phase (c), it shifts with increasing temperature from the water-rich to the oil-rich side until it meets the upper critical endpoint cep a at Tu = 72.89°C. Here, the upper critical tie-line connects a critical surfactant-rich oil phase (b) (shown as a filled square) with an excess water phase (a). Above Tu at T = 82°C, just the extended central miscibility gap exists. Here the positive inclination of the tie-lines indicates that a surfactant-rich oil phase (b) coexists with a water-excess phase (a). From the phase progression in fig. 5.5, it is clear that the phase inversion occurs over an extended temperature range, which makes the exact location of "the" phase inversion temperature difficult. Identifying the optimal state, on the other hand, is unambiguous. 5.2d The optimum state As already discussed in connection with fig. 5.5, the optimum state of a surfactant for a given microemulsion manifests itself in a symmetric three-phase triangle. Thereby, the tip of this symmetric triangle sets the lowest possible amount of surfactant to solubilize equal volumes of water and oil. Thus, among the infinite number of X-points, which will be discussed below in connection with the trajectory of the phase (c), this point (the Xm-point) is a special one. Henceforth, we will drop the
11
A.G. Aizpiri, F. Monroy, C. Delcampo, R.G. Rubio, and M. D. Pena, Chem. Phys. 165 (1992) 31.
5.10
MICROEMULSIONS
subscript m and refer to the X-point, meaning its value at symmetry unless otherwise specified. The X-point is the pivot point for microemulsion phase behaviour and our understanding of microemulsions in general. In particular fixing the temperature such that the X-point is part of the Gibbs triangle permits comparing systems with different surfactants and oils. This situation (the X-point), however, is obtained experimentally both from an isothermal cross section through the phase prism at the mean temperature of the three-phase body Tm = (Tu + Tj) / 2 or from an isoplethal cross section at oil-to-water plus oil volume fraction of > = 0.5 . One of the central questions of microemulsion formulation has been, and still is, the quest for high efficiency, i.e. finding the minimum amount of surfactant for solubilizing oil in water or vice versa. In fig. 5.6 we demonstrate how the X-point, and with it the whole three-phase triangle, are affected by increasing the chain length of the surfactant. Starting with the H2O-n- octane-C8E5 system (fig. 5.6, top) and comparing it with the triangle (at T = 52°C) of the C4E[ -system in fig. 5.5, one can see that the surfactant fraction wCm at the X-point decreases from wCm ~ 0.63 to wCm ~ 0.28. The reason for this enormous reduction of wc m is the increasing amphiphilicity of the surfactant, which is a result of more than doubling the sizes of both the hydrophobic and hydrophilic parts. C4Ej molecules adsorb in a rather disordered fashion at the microscopic interface between H2O and n-dodecane. Thus, a solution of H 2 O, ndodecane and C4Ej is only weakly structured and can rather be described as a near tri-critical mixture of three components. The features of such mixtures can even be mimicked as ternary regular mixtures11. Increasing the amphiphilicity, the surfactant molecules are forced into the microscopic water/oil interface, which leads to a better defined interfacial film and, hence, to a strongly structured mixture (for details see sec. 5.3), i.e. a microemulsion. Increasing the hydrophobic chain length of the surfactant from C8E5 by two to Cj 0 E 5 (see fig. 5.6, centre), the mean temperature Tm is lowered from Tm = 61.5°C to Tm = 44.6°C. Comparing both Gibbs triangles with increasing amphiphilicity, a further decrease of the X-point to wCm = 0.14 and the striking appearance of the lamellar mesophase (L a ) is observed. The latter extends deep into the water and oil corners of the Gibbs triangle. Thereby, the La phase intrudes into the two-phase regions, the 2 on the water-rich side and 2 on the oil-rich side, respectively, and additional three-phase triangles result. It appears as if certain ratios of surfactant and oil, or surfactant and water, particularly favour the existence of the La phase21. However, in many applications an attempt is made to avoid formation of mesophases as these are often highly viscous and tend to complicate the handling of water-oilsurfactant systems.
11
M. Kahlweit, R. Strcy, P. Firman, D. Haasc. J. Jen, and R. Schomacker, Langmulr 4 (1988) 499. 21 U. Olsson, U. Wurz, and R. Strey, J. Phys. Chem. 97 (1993) 4535.
MICROEMULSIONS
5.11
Figure 5.6. Isothermal Gibbs triangles of the systems H 2 O-n-octane-CgEg , CjgEg , and C ) 2 E 5 determined in each case at the mean temperature Tm = (Tu + Tj) / 2 . All three systems show an almost symmetrical three-phase triangle. With increasing hydrophobic chain length of the surfactants, the height of the triangle decreases, i.e. the X-point is shifted to lower surfactant mass fractions. Note that the lamellar mesophase (L a ), which is absent in the C 8 E 5 system, extends over wide regions in the CJQEQ and especially in the Cj2Eg system.
Passing on to the efficient H2O-n- octane-C12E5 -system (fig. 5.6, bottom), the trends observed going from the C8E5 to the C10E5 -system continue, although an even more striking extension of the lamellar phase can be observed. The mean temperature is lowered to Tm = 32.85°C and the X-point is shifted to wCm = 0.05 through which the three-phase triangle shrinks further towards the water-oil binary axis. Towards larger surfactant mass fractions, the existence of the one-phase bicontinuous mieroemulsions is limited to a very small region by the lamellar phase. Comparing the three Gibbs triangles one can easily see that with increasing chain length of the surfactant, the extension of the lamellar phase grows faster than the X-point is shifted to smaller
5.12
MICROEMULSIONS
values of wc . Therefore, the regime of the one-phase mieroemulsion is increasingly reduced. For surfactants with even longer alkyl chains, as for example C 14 E 5 , the lamellar phase touches the X-point and the one-phase mieroemulsion disappears. Thus, this road to the formulation of highly efficient microemulsions seems to be a dead end. However, we found a new and successful path using amphiphilic block copolymers as efficiency boosters of microemulsions11 (for details see sec. 5.6e). The increasing complexity of the Gibbs triangles of the C10E5- and C12E5-systems make it clear that any detailed study of the optimal point via isothermal Gibbs triangles is in fact time-consuming. The study of each of the systems required several months. As mentioned above, an alternative and more rapid method is recording isoplethal T{wc) sections at an oil-to-water plus oil volume fraction of 0= 0.5 , which involves the reasonable assumption that X{<j>= 0.5) = X m . In this fashion, the optimal point (X-point) can be determined by extrapolating the phase boundaries from 2 to 1 and from 1 to 2, which makes the exact determination of the three-phase region dispensable. Figure 5.7 show such sections for the H2O-n-octane-C6E2-, C g E 3 -, C 10 E 4 -, and C 12 E 5 -systems. In principle, the same trends as observed in the isothermal Gibbs triangles are visible in the T(wc) sections, although in the latter both the hydrophobic chain length and the size of the hydrophilic head group of the surfactant are increased. With increasing amphiphilicity, the X-point shifts from wCm = 0.33 for the C6E2 (fig. 5.7, top) to wCm = 0.05 for the C12E5 -system (fig. 5.7, bottom). Simultaneously, the lamellar mesophase (surrounded by a two-phase coexistence region (not shown)), appears, which is not present in the C6E2 -system, but which does occur in the C8E3 system where it is embedded in the homogeneous mieroemulsion phase. With further increasing amphiphilicity of the surfactant, it extends almost across the entire onephase region. While the Gibbs triangles contain additional information about the phase behaviour beyond the <j> = 0.50 ratio, the T{wc) section contains the variation of the phase behaviour with temperature. Thus, one can see that the phase sequence follows the same principles for short chain and long chain surfactant systems. However, the latter group of systems show, besides an extended lamellar phase that is found to be more stable at low than at high temperatures, other ordered mesophases, including the cubic (V,) phase at low temperatures in the C12E5-system. Looking at the phase transitions as a function of temperature, one realizes that the ability of the surfactant to solubilize water and oil runs through a maximum at the mean temperature Tm and decreases rapidly below and above this temperature. This suggests the possibility of defining the efficiency of a surfactant by the lowest amount of surfactant t« C m required to solubilize equal volumes of water and oil at the temperature T .
11 B. Jakobs, T. Sottmann, R. Strey, J. Allgaier, L. Willner, and D. Richter, Langmuir 15 (1999) 6707.
MICROEMULSIONS
5.13
Figure 5.7. Isoplethal T(wc) sections through the phase prism of the systems H2O-n- octaneCgE2 , CgE3 , CJQE 4 , Cj2E5 at an oil-to-water plus oil volume fraction tf>= 0.5. In each case, only the phase boundaries for wc > wc m are measured sufficient to determine the respective X - point. Increasing both the hydrophobic chain length and the size of the hydrophilic head group of the surfactants, the X-point shifts to lower values of WQm , i.e. the efficiency increases. Thereby, the stability range of the bicontinuous one-phase microemulsion shrinks dramatically due to the increased extension of the lamellar mesophase (L a ). Vj refers to the low temperature cubic phase.
5.2e Efficiency In technical applications a formulation should be as efficient as possible, I.e. the amount of surfactant needed should be a minimum. For this, the X-point provides an excellent criterion. To demonstrate this, we show in fig. 5.81! a synopsis at the X-points of fourteen different H2O-ri-octane-CnEx-systems in a Tm{wCm) 11
plot. The hydro-
S. Burauer, T. Sachert, T. Sottmann and R. Strey, Phys. Chem. Chem. Phys. 1 (1999] 4299.
5.14
MICROEMULSIONS
Figure 5.8. X-points of the systems H 2 O-n-octane-C n E x at an oil-to-water phase, oil volume fraction 0 — 0.5 (redrawn from11). The individual systems are characterized by the (n,x) pairs. While an increase of n leads mainly to a decrease of WQm , an increase of x mainly increases T m , which we suggest to take as a more exact measure of the p.i.t.
phobic chain length n is varied between 6 and 12, the number of oxyethylene groups x between 2 and 7. As mentioned, for n> 14 the phase behaviour changes qualitatively and no X-point can be found anymore due to the vast extension of the lamellar phase. Generally, the following systematic trends can be observed: - With increasing n, wc m decreases strongly, i.e. the surfactant becomes more efficient. Since the surfactants coneomitantly become more hydrophobic, the X-points shift to lower temperatures. - With increasing x, the points shift to higher temperatures due to an increasing hydrophilicity of the surfactant and wc m increases slightly. The whole grid of the X-points (shown in fig. 5.8 for n-octane) varies systematically with the chain length k of the n-alkane (C k H 2k+ [ not shown in fig. 5.8 for clarity). From a compilation by Kahlweit et al. (see fig. 12 in21) we can see that; - with increasing k, the X-points shift to higher temperatures and wc m increases, i.e. the surfactant becomes less efficient. This trend can be attributed to the phase behaviour of the binary n-alkane-CnEx-system (see fig. 5.2, right). With increasing oil chain length, the critical temperature of the lower miscibility gap Ta shifts to higher temperatures. More recently, the corresponding trends of the X-points with k were observed using other classes of oils, namely polymerizable n-alkyl methacrylates , n-alkyl ether and esters as well as triglycerides.
11
S. Burauer, et al., loc. clt. M. Kahlweit, R. Strey, P. Firman, D. Haase, J. Jen, and R. Schomacker, Langmuir 4 (1988) 499. 31 O. Lade, K. Beizai, T. Sottmann, and R. Strey, Langmuir 16 (2000) 4122. 21
MICROEMULSIONS
5.15
5.2/ Monomeric solubility In general, one can distinguish between surfactant molecules, which reside at the microscopic water/oil interface and molecules, which dissolve monomerically in either oil or water. If three phases are present, surfactant molecules dissolve in the excess phases as well as in the microemulsion sub-phases, i.e. in the oil- and water-rich domains of swollen micelles or bicontinuous microemulsions. The significance of this fact is not only that part of the surfactant is lost in this way but also that the associated osmotic pressure seems to be an important variable influencing the driving force for microemulsion formation. Experimentally, wc
m o na
the
monomeric
solubility
of
the
surfactant
in
the
water
can easily be determined from surface tension measurements11. On the other
hand, the direct determination of the monomeric solubility of the surfactant in the oil LD Cmonb is more difficult and time consuming. An interesting method to obtain ix>cm o nb is provided by the macroscopic phase behaviour through the determination of the mass fraction of surfactant wco
(see fig. 5.4), i.e. the monomerically dissolved
surfactant in both excess phases. Therefore, the volume fraction of the middle phase V wr-wrri —C- = 0c=—± ±*L Vtot wc - wc0
[5.2.3]
has to be measured as a function of the mass fraction of surfactant wc at a constant a and the corresponding temperature T 2 ) . Here Vc is the volume of the surfactantrich middle phase and Vtot the total volume of the mixture in the test tube. By plotting @c versus wc,
one obtains wc
0
at 0C = 0 and wc at @c = 1 . The monomeric
solubility in the oil is then calculated from
= C m
' °n'b
^C.O + " c . m o n > a - " W - l ]
[52A]
"c.O+«
Figure 5.9 shows the monomeric solubility i^cmonb
m
"-octane at the mean
temperature Tm calculated according to [5.2.4]. Thereby, the monomeric solubility ^Cmona
i n w a t e r
w a s s e t e
qual
t o 0 0 3
. 0.02, 0.01, 0.006 and 0.002 for C 6 E 2 ,
CgE 3 , C 6 E 4 , C 7 E 3 and C g E x , respectively. For longer chain surfactants,
i»Cmona<
0.001 was neglected 31 . Furthermore, a grid of lines was drawn through the data point at constant n and x to guide the eye in an attempt to even out the unavoidable experimental error.
11
M. Kahlweit, R. Strey, and G. Busse, J. Phys. Chem. 94 (1990) 3881. S. Burauer, T. Sachert, T. Sottmann, and R. Strey, Phys. Chem. Chem. Phys. 1 (1999) 4299; S. Burauer, T. Sottmann, and R. Strey, Tenside Surf. Det. 37 (2000) 8; H. Kunieda, K. Shinoda, J. Colloid Interface Set 107 (1985) 107. 31 M. Kahlweit, R. Strey, and P. Firman, J. Phys. Chem. 90 (1986) 671; M. Kahlweit, R. Strey, and G. Busse, Phys. Rev. E47 (1993) 4197. 21
5.16
MICROEMULSIONS
Figure 5.9. Monomeric solubility U>Q m o n k of fourteen different surfactants in n-octane at the mean temperature Tm (redrawn from obtained from the determination of wco (see fig. 5.4). Note how parallel the T m(">C.mon.b) S r i d follows the pattern of the Tm(u>Cm) grid.
As can be seen, the Tm{wCmonb) grid shows the same pattern as the Tm{wCm) grid, i.e. the monomeric solubility ii> Cmonb in n-octane decreases with n and increases slightly with x. Although the differences in monomeric solubility were used before to distinguish between weak and strong amphiphiles21, these findings prove that both monomeric solubilities are indeed correlated with the efficiency of the surfactant to solubilize water and oil. Furthermore, the suggestion that the monomeric solubility, in particular the water-oil partition coefficient of the surfactant Kwo = wc m o n a / u ) C m o n b , plays an important role for the description of microemulsion as a whole31, which seems also to be confirmed. While for technical applications the efficiency wc m of the surfactant is the important parameter, in theoretical descriptions of microemulsions, the properties of the amphiphilic film forming the internal interface also play an important role . One important property is that the surfactant molecules, which reside at the microscopic water/oil interface, Ujg = wc -
WAWc mon a 1
- _ ^B^C.mon.b ~ u;C.mon,a * " mC,mon.b
[5 2 5 ]
sets the size of the specific area of the interface AIV ~ u)g, the characteristic length
11
S. Burauer, et al., loc. cit. M. Kahlweit, R. Strey, and P. Firman, J. Phys. Chem. 90 (1986) 671; M. Kahlweit, R. Strey, and G. Busse, Phys. Reu. E47 (1993) 4197. 31 M. Kahlweit, R. Strey, and G. Busse, J. Phys. Chem. 94 (1990) 3881; H. Schott, J. Pharm. Set 84 (1995) 1215. 41 S.A. Safran, D. Roux, M.E. Cates, and D. Andelman, Phys. Reu. Lett. 57 (1986) 491; S.A. Safran in Structure and Dynamics of Strongly Interacting Colloids and Supramolecular Aggregates in Solutions; S.-H. Chen, J.S. Huang, and P. Tartaglia, Eds., Kluwer Academic Publishers (1992) p 237. 21
MICROEMULSIONS
5.17
£, — (uig)"1 " of the structures and the interfacial tension y ab = yw° ~ (Lug)2 between water- and oil-rich phases 2 (for details see sees. 5.3 and 5.4). Here wA and wB are the weight fractions of water and oil, respectively. Alternatively, at the X -points the surfactant mass fraction of molecules residing in the microscopic water/oil interface ffig can be calculated31 from w°c=wc-wco—-f-
[5.2.6]
The facts presented so far show that the microemulsion phase behaviour, i.e. the efficiency wcm , the mean temperature Tm and the monomeric solubilities depend sensitively and systematically on the chemical nature of the components. However, general patterns may be recognized that many microemulsion systems share, independent of whether they are made of specially purified substances (figs. 5.6 and 7) or technical grade compounds (see sec. 5.6a). These striking similarities indicate that, as in the corresponding state description of real gases, suitable general parameters exist, which scale the phase behaviour of all microemulsion systems. In order to find the relevant parameters, the central properties of the phase behaviour, that is the location and extension of the three-phase body, have to be considered. The latter is mainly determined by the trajectory of the X-point composition, that is the composition of the middle phase microemulsion as it traverses through the three-dimensional phase space as a function of the volume fraction of the oil in the mixture of water and oil
5.18
M1CROEMULSIONS
Figure 5.10. Trajectories of the surfactant-rich middle phase, i.e. the X -point for the systems H 2 O-n-octane-C 6 E 2 , CgE 3 , C 10 E 4 , C 12 E 5 . (a) Projection onto the QQ-
the volume fraction of surfactant 0g, which resides at the internal interface of the middle phase, is plotted as a function of the oil-to-water-plus-oil volume fraction <j> in fig. 5.10, left. As can be seen, the trajectory extends from the lower critical endpoint (cepo , filled symbol) at Tj on the water-rich side to the upper critical endpoint (cep a , filled symbol) at Tu on the oil-rich side. Thereby, each individual trajectory has a nearly parabolic shape with a maximum volume fraction of surfactant 0g at almost 0 = 0.5. Increasing both the hydrophobic chain length n by 2 and the size of the hydrophilic head group of the surfactant x by 1 shifts the
T. Sottmann and R. Strcy, loc.cit..
MICROEMULSIONS
5.19
Figure 5.11. Representation of the reduced projected trajectories of the surfactant-rich middle phase for the systems H2O-n-octane-CgE2 , C g E 3 , C 1Q E 4 , C 12 E 5 (redrawn from ). (a) Scaling of the (*Q(0) trajectories by the maximum volume fraction of surfactant 0g of each trajectory, (b) Scaling of the T(0) trajectory. Subtracting the mean temperature Tm and normalizing the temperature axis by AT/2 lets all trajectories collapse into a single curve. Symbols used are the same as in fig. 5.10. projected onto the base of the phase prism. Normalizing each 0g(0) projection by its maximum ) suggests that there must be a corresponding state of different microemulsion systems. Only three empirical parameters are needed to scale the trajectory of the middle phase X (
T. Sottmann and R. Strey, J. Phys. Condens. Matter 8 (1996) A39.
5.20
MICROEMULSIONS
phase behaviour of non-ionic microemulsions and, in particular, with the salient features of the three-phase body (i.e. the Winsor III regime). For theoretical considerations, as well as technical applications, we want to focus now on the phase behavior far on the water-rich and oil-rich side of the phase prism (the Winsor I and II regimes). 5.21 Straight-line binodal curves The phase behaviour of water- and oil-rich microemulsions can be studied most conveniently by considering vertical sections through the phase prism at a constant surfactant-to-water plus surfactant mass fraction wCa=
^ m
A +
m
[5.2.7] C
at a constant surfactant-to-oil plus surfactant mass fraction wch=
m
^ B+mc
[5.2.8]
Starting from the binary systems A-C or B-C, the temperature extension of the onephase regions is measured as a function of the mass fraction of oil wB or water wA , respectively. Figure 5.12 (left) shows for the system H2O-n-octane-C10E5 three such T-wB sections at wCA= 0.05, 0.10 and 0.15 on the water-rich side of the phase prism. The corresponding sections T-wA on the oil-rich side are shown in fig. 5.11 (right).
Figure 5.12. Vertical T — IX>B(UJA) sections through the phase prism of the system H2O-noctanc-C 10 E 5 starting with wCa(wcb)= 0.05, 0.10, 0.15. Water-rich side (left): Starting from the binary system with increasing mass fraction of oil wB the oil cmulsification boundary (2—>1) ascends while the near-critical phase boundary U^>2) descends. The inverse temperature behaviour is found on the oil-rich side (right). With an increasing fraction of water UJA the water emulsification boundary (1—>2) descends, whereas the near-critical phase boundary (2-»l) ascends. Note that both on the water- and oil-rich side the critical phase boundaries pass through an extremum.
MICROEMULSIONS
5.21
Let us first consider the phase boundaries on the water-rich side. The first point at high temperatures is a point on the cloud point curve of the binary water-C1QE5 system (see fig. 5.2). The 1—>2 phase boundary is therefore a near-critical boundary, which descends steeply upon the addition of n-octane. As can be seen in fig. 5.12 (left) this boundary runs through a minimum as the weight fraction of oil wB is increased further. Simultaneously, the 2—>1 phase boundary ascends monotonically with increasing wB . It marks, for a given temperature, the maximum amount of oil solubilizable in a one-phase, oil-in-water (o/w) microemulsion and is therefore called the emulsification failure boundary (efb). With increasing temperature, the capability of the surfactant to solubilize oil is strongly increased. Near the lower critical endpoint temperature Tx, the one-phase region closes like a funnel. It terminates at the intersection of the lower oil emulsification failure and the upper nearcritical phase boundary. Comparing the three T - wB sections measured at different wc a , it can be seen that the shape of the phase boundaries remains almost unchanged, whereas the efb and, with it, the point of intersection of the 1—»2 and 2—>1 phase boundaries varies systematically with wc a . Thus, as expected, an increasing amount of surfactant, i.e. an increasing wCa , enables the solubilization of more oil. At the oil-rich side, in T-wh sections starting at different ii>cb (see fig. 5.12, right), inverse but similar phase behaviour can be observed with respect to the temperature. Thus, the near-critical phase boundary 2—>1 starts at low temperatures from the lower n-octane-C10E5 miscibility gap (below < 0°C) and ascends steeply upon the addition of water. Increasing UJA , this boundary runs through a maximum and then decreases down to the upper critical endpoint temperature T u . The emulsification failure boundary 1—>2 starts at high temperatures and low values of wA , which means that only small amounts of water can be solubilized in a water-in-oil (w/o) microemulsion at temperatures far above the phase inversion. Increasing amounts of water can be solubilized by decreasing the temperature, i.e. by approaching the phase inversion. At Tu the efb intersects the near-critical phase boundary. Here, the funnel-shaped, one-phase region closes. Comparing the three T - wA sections, it can be seen that with increasing amount of surfactant, i.e. increasing wc b , more water can be solubilized. However, contrary to the sections on the water-rich side, the efb, and with it the point of intersection of the 1—>2 and 2—>1 phase boundaries, does not exactly scale with wc b . The reason is the relatively large monomeric solubility u ; C m o n b of the C 10 E 5 in n-octane (see fig. 5.9). From the above, it can be concluded that such T -wB(wA) sections provide an easy method to determine two of the relevant scaling parameters (Tj and Tu ) of the threephase body (see fig. 5.1 lb), as well as the emulsification failure boundaries, which are of particular interest for technical applications. Furthermore, weak and strong surfactants systems can be distinguished considering the shapes of the near-critical phase
5.22
MICROEMULSIONS
Figure 5.13. Water-rich and oil-rich corners of the Gibbs triangle for the system H2O-n octane-C 10 E 5 . The diagram on the left is determined at T = 38°C < Tj , while the one on the right is measured at T = 49°C, i.e. slightly above T u . The full lines denote the phase boundaries, the dashed lines represent the tie-lines schematically. Note that both triangles show two two-phase regions, the central water-oil miscibility gap and the loops, which are separated by a one-phase region. boundaries 11 . While for weak surfactant systems the boundary decreases monotonically to Tj (water-rich side), it increases respectively up to Tu (oil-rich side) in strong surfactant systems, the near-critical curve passing through a minimum (water-rich side) or maximum (oil-rich side). As will be shown in fig. 5.13, the extrema are consequences of additional two-phase regions In the form of closed lobes appearing at temperatures below Tx and above Tu in the Gibbs triangle21. In recent theoretical considerations, these lobes play an important role in explaining the origin of the threephase body (see sec. 5.3) 31 . As has already been shown in connection with fig. 5.5, Gibbs triangles can be constructed from vertical sections through the phase prism by reading off the composition of the phase transitions at the respective temperature. In fig. 5.13 the water-rich and oil-rich corners of the Gibbs triangle of the system H 2 O-n -octaneC 1 0 E 5 are shown at T = 38°C < Tj (left) and T = 49°C > Tu (right). They are constructed from six T-wB and five T - wA sections, respectively. Contrary to the phase diagrams of the system H 2 O-n-dodecane-C 4 E 1 (see fig. 5.5, T = 22°C and T = 82°C), an additional two-phase region (which has the shape of a
11
M. Kahlweit, R. Strey, and G. Busse, Phys. Rev. E47 (1993) 4197. M. Kahlweit, R. Strey, and G. Busse, J. Phys. Chem. 94 (1990) 3881; M. Kahlweit, R. Strey, and G. Busse, Phys. Rev. E47 (1993) 4197; P.K. Kilpatrick, C. A. Gorman, H.T. Davis, L.E. Scriven, and W. G. Miller, J. Phys. Chem. 90 (1986) 5292; R. Strey, Ber. Bunsenges. Phys. Chem. 100 (1996) 182. 31 T. Tlusty, S.A. Safran, R. Menes, and R. Strey, Phys. Rev. Lett. 78 (1997) 2616; T. Tlusty, S. A. Safran, and R. Strey, Phys. Rev. Lett. 84 (2000) 1244; A.G. Zilman, S.A. Safran, Phys. Rev. E66 (2002) 051107. 21
MICROEMULSIONS
5.23
lobe and two critical points) can be observed at temperatures just below Tj and above Tu . The importance of these lobes cannot be overemphasized because they are the crown witnesses of the origin of phase inversion. For example, consider on the lefthand side of fig. 5.13, a sample with wc a = 0.10 to which rt-octane is added at a constant temperature T of 38°C < Tj. This path would produce the phase sequence 1—>2—>1—>2. Thereby, the tie-lines (shown as dashed lines) within the lobe run essentially parallel to the H 2 O-C 10 E 5 side and hence indicate that a C 10 E 5 -rich upper phase coexists with a water phase containing much less surfactant (i.e. 2 ). The 2-state is limited by a binodal, which is in this part of the Gibbs triangle a straight line. Extrapolating this almost straight line to the binary H 2 O-C 10 E 5 side yields approximately the monomeric solubility of the surfactant in water wc m o n a . The slope of this line is a measure of the oil-to-surfactant ratio and corresponds to the size of the R-octane-swollen C 10 E 5 micelles (see sec. 5.3, [5.3.24]). Now consider a sample on the oil-rich side with wcb = 0.10 to which water is added at 49°C > Tu . The phase sequence 1—>2—>1—>2 can be observed in fig. 5.13, right. Within the lobe, a C 10 E 5 -rich lower phase coexists with an oil phase with a low surfactant content (2). The monomeric solubility of the surfactant in oil wc m o n b can be obtained from extrapolating the straight-line binodal to the binary n-octane-C10E5 side. The negative slope of the binodal indicates that water-swollen C 10 E 5 micelles are formed. Their sizes are given by the absolute values of the slopes. Thus, importantly, isothermal phase diagrams of the water (below T^) and oil-rich corner (above Tu ) of the Gibbs triangle provide quantitative values for the monomeric solubility of the surfactant in water and oil on the one hand and the size of the oil- or water-swollen micelles, on the other hand. We are now in a position to see that the origin of the three-phase body in strong surfactant systems (compare the phase behaviour of the weak surfactant system shown in fig. 5.5) is contained in fig. 5.13. As one raises the temperature on the water-rich side, the critical point on the oil-rich side of the lobe approaches the central miscibility gap to touch it at Tj. Thus, three phases are formed, whereby the two phases a and c are near critical and initially of equal composition. As the tie-lines of the lobe are essentially perpendicular to those of the central miscibility gap, the composition of the two phases a and c diverges rapidly upon further increasing temperature. The inverse situation holds for the oil-rich side. As one decreases the temperature, the critical point on the water-rich side of the lobe approaches the central gap to touch it at Tu , which also leads to the formation of three phases, two phases b and c of identical composition and a water excess phase a. The origin of the lobes, i.e. the separation of a micellar phase into two phases, which become homogeneous again upon swelling with a solute, was not understood for a long time. Safran et al. attributed the origin of the lobes to the demixing of a connected network of swollen cylindrical micelles into a dense connected network in
5.24
MICROEMULSIONS
equilibrium with a dilute phase11 (see sec. 5.3). This description furthermore explains why the lobes appear only in strongly structured and not in weakly structured microemulsion systems. 5.3 Microstructure In sec. 5.2 we have pointed out that although the overall phase behaviour of microemulsions is generic, the origin of the three-phase region depends on the chain length of the surfactant used. While in mixtures containing only short-chain surfactants the three-phase region can be mimicked by the regular solution theory21 (see sec. I, 2.18c), the three-phase region in mixtures of water, oil and surfactants with long alkyl chains is a consequence of the microstructure formed on a microscopic level. Because of their strong amphiphilicity, long-chained surfactant molecules are forced into the microscopic water/oil interface leading to topologically ordered interfacial films in solutions, i.e. the "real" microemulsions31. The nature and properties of these microscopic interfacial films are essential for microemulsions as a whole and, in particular, for the most interesting feature of microemulsions, viz. the variety of microstructures. One finds droplet and wormlike microemulsions, sample-spanning networks and bicontinuous structures. Furthermore, liquid crystalline mesophases, such as the cubic (V), hexagonal (H) and lamellar phases (L a ), exist and compete with these complex fluids. It has been realized that the main parameter determining the microstructure is the local curvature of the amphiphilic interfacial film41. Thus, controlling the curvature is the ultimate goal permitting to choose any desired structure. Mathematically, the curvature at every point on a surface is given by the two principal radii of curvature R^ and R2 defining the principal curvatures C l =-iK
and
c2=-±K
l
[5.3.1]
2
Then, the first or mean, and second or Gauss curvature, of the amphiphilic film are defined by J = (cl+c2)
and
K = CjC2
[5.3.2] 5)
respectively. We count c ; as positive if the amphiphilic film tends to enclose the oil and 11 T. Tlusty, S.A. Safran, R. Menes, and R. Strey, Phys. Rev. Lett. 78 (1997) 2616; T. Tlusty, S.A. Safran, and R. Strey, Phys. Rev. Lett. 84 (2000) 1244; A.G. Zilman, S.A. Safran, Phys. Rev. E66 (2002) 051107. 21 M. Kahlweit, R. Strey, P. Firman, D. Haase, J. Jen, and R. Schomacker, Langmuir 4 (1988) 499. 31 S.A. Safran, in Structure and Dynamics of Strongly Interacting Colloids and Supramolecular Aggregates in Solutions, S.-H. Chen, J.S. Huang, and P. Tartaglia, Eds., Academic Publ. (1992) p. 237. 41 R. Strey, Colloid Polym. Set 272 (1994) 1005. 51 Note that sometimes the quantity H = i (Cj + c2) is used; H = J12 .
MICROEMULSIONS
5.25
negative if it tends to enclose the water (see sec. I, 2.23). The actual curvature J is closely related to the most important parameter J o in Helfrich's membrane bending Helmholtz energy11 (see [III. 1.15.1 and 2] and sec. 5.5). J o is that curvature the interfacial film will adopt if no external forces, thermal fluctuations or conservation constraints exist. J o is called the spontaneous (preferred) mean curvature. The variation of the spontaneous curvature J o of a non-ionic surfactant film can be deduced to a good approximation from the variation of the phase behaviour. Macroscopically, we can observe that in ternary mixtures of water (A) - oil (B) - non-ionic surfactant (C), the surfactant dissolves mainly in water at low temperatures, whereas at high temperatures it prefers the oil. Looking at these mixtures microscopically, we find at low temperatures an amphiphilic film that forms a sphere where the hydrophobic side points to the inside of the sphere (oil-swollen micelles), i.e.JQ > 0. On the other hand, at high temperatures the film forms spheres having their hydrophilic side on the inside of the sphere (inverse or water-swollen micelles), i.e. Jo < 0. Locally, planar amphiphilic films, I.e. Jo = 0, which are the optimum state of the microemulsion, can be found in between at the mean temperature Tm . Figure 5.14 illustrates how for a non-ionic surfactant film these three situations, forming the microscopic water/oil interface, are related to the molecular structure of the surfactant. At low temperatures the size of the surfactant head group is larger than that of the hydrophobic chain, leading to an amphiphilic film curved around the oil (fig. 5.14, bottom). Increasing the temperature, the size of the surfactant head group shrinks in line with its decreasing solubility, whereas the size of the hydrophobic chain increases due to the increasing number of chain conformations and the increasing penetration of oil molecules. These trends lead to a gradual change of the curvature of
Figure 5.14. Spontaneous curvature of a nonionic surfactant film forming the microscopic water/oil interface as a function of temperature. In order to illustrate this behaviour, a wedge shape has been chosen for the molecules.
11
W. Helfrich, Z. Naturforsch. 28c (1979) 693.
5.26
MICROEMULSIONS
the amphiphilic film from Jo > 0 to J o < 0 (fig. 5.14, top), i.e. from oil in water (o/w) to water in oil (w/o) via the locally planar amphiphilic film, that is J o = 0 (fig. 5.14, centre). It should be noted that the temperature is not the sole parameter on which the spontaneous curvature depends, but due to the reversibility of the process, temperature is an extremely useful and, of course, thermodynamically relevant tuning parameter. Other parameters are pressure, salt and co-surfactant concentration. However, in order to determine the variety and length scale of the microstructure and, with it the underlying curvature of the amphiphilic film quantitatively, several different experimental methods have to be employed. 5.3a Methods In general, microstructure is suitably examined by microscopy. Because of the small size of the microemulsion nanostructures, only transmission electron microscopy is suitable to visualize these structures. The three different methods, which can be employed to dissect the liquid mixtures, are the well-established Freeze FractureElectron-Microscopy (FFEM)1'2'31 and Cryo Direct Imaging (Cryo-DI)4'51 as well as the more recently developed Freeze Fracture Direct Imaging (FFDI)61 technique. A combination of these three methods provides detailed, local information about the occurring types of the structure. Statistical information about frequently occurring distances in microscopic heterogeneous systems can be obtained from scattering techniques (see sec. 1.7). For example, microemulsions scatter light, which renders them opalescent. The stronger the intensity of the scattered light is, the larger the length scale of the structure. However, as the wavelength of the light is relatively large compared with the typical length scales in microemulsions, static light scattering (SLS)71 provides in general only unspecific information. Somewhat more useful is dynamic light scattering (DLS)81, which yields the diffusion coefficient of the structural domains. Considerably more detailed information can be gained from small angle X-ray (SAXS)9) and small angle neutron scattering (SANS)101 technique with wavelengths, which fall inside the range of the microstructure and thus permit the determination of the (average) microstructure
11
W. Jahn, R. Strey, in Physics of Amphiphilic Layers, J. Meunier, D. Langevin, and N. Boccara, Eds., Springer (1987) p. 359. 21 W. Jahn, R. Strey, J. Phys. Chem. 92 (1988) 2294. 31 J.F. Bodet, J.R. Bellare, H.T. Davis, L.E. Scriven, and W.G. Miller, J. Phys. Chem. 92 (1988) 1898. 41 M. Almgren, K. Edwards, and J. Gustafsson, Curr. Opin. Coll. Interface Set 1 (1996) 270. 51 Y. Talmon, in Modern Characterization Methods of Surfactant Systems, B.P. Binks, Ed., Marcel Dekker (1999) p. 147. 61 L. Belkoura, C. Stubenrauch, and R. Strey, Langmuir 20 (2004) 4391. 71 B. Chu, Laser Light Scattering, Academic Press (1974). 81 B. J. Berne, R. Pecora, Dynamic Light Scattering, Wiley (1976). 91 F. Lichterfeld, T. Schmeling, and R. Strey, J. Phys. Chem. 90 (1986) 5762. 101 S.H. Chen, Ann. Rev. Phys. Chem. 37(1986) 351.
MICROEMULSIONS
5.27
length scales quantitatively. However, since scattering techniques always provide only the Fourier transform of the spatial autocorrelation function, a quantitative interpretation of the scattering data requires knowledge of the shapes of the microstructure. Thus, scattering techniques and microscopy complement one another. Furthermore, indirect methods like NMR-self-diffusion1 and electric conductivity3 measurements provide valuable information on the connectivity of the microstructure and the transition from one type of structure to another. Each of the techniques to be described below provides a piece in the puzzle of the structure of microemulsions, which will be developed in this section choosing the ternary non-ionic microemulsion system H2O-n-octane-C12E5 . 5.3b Transmission electron microscopy The direct imaging of the microemulsion structure via electron microscopy provides detailed pictures of the zoo of differently shaped amphiphilic films. However, one has to solve the problem of the fixation of the liquid mixtures. Two well-established techniques are Freeze Fracture Electron Microscopy (FFEM) and Cryo-Direct Imaging (Cryo-DI). In FFEM the samples are rapidly frozen, fractured, shadowed with metal, and replicated with a thin carbon film. The replica of the fractured surface, the morphology of which Is controlled by the sample's microstructure, is then studied by a transmission electron microscope (TEM). In contrast to FFEM, in conventional CryoDI, thin films of the sample are rapidly frozen but immediately, without replication, transferred to a low-temperature stage within the microscope and imaged directly. In this case, additional problems are encountered in the preparation of the films because they need to be thin enough to allow for the electrons to transverse the sample. The widely used method - the blotting of the sample prior to vitrification - does not only lead to unavoidable concentration changes and size segregation, but also to shearing of internal liquid structures. To overcome these problems of the direct imaging technique, the Freeze Fracture Direct Imaging (FFDI) method was more recently developed. FFDI is a hybrid of FFEM and Cryo-DI. The key advantage is that instead of replicating the fracture, it is the fractured frozen sample itself, which is looked at with the microscope (true direct imaging). Thus, all three methods use cryofixation to solidify the microemulsion structure. Having in mind that the phase behaviour, as well as the curvature of the amphiphilic film (see fig. 5.14) and with it the microstructure of most microemulsions, is strongly temperature-dependent, one has to ensure that the cooling should be as fast as possible (>10 4 K/s) compared with the reorganization kinetics of the microstructure. Despite all the sources of error, which could be encountered during the preparation of 11
B. Lindman, K. Shinoda, U. Olsson, D. Anderson, G. Karlstrom, and H. Wennerstrom, Colloids Surfaces 38 (1989) 205. 21 B. Lindman, U. Olsson, Ber. Bunsenges. Phys. Chem. 100 (1996) 344. 31 H.F. Eicke, J.C.W. Shepherd, and A. Steinmann, J. Colloid Interface Set. 56 (1976) 168.
5.28
MICROEMULSIONS
Figure 5.15. Micrographs of microemulsion droplets of the o/w type in the system H2O -noctane-C 12 E 5 prepared near the emulsification failure boundary at wc a = 0.022, wB = 0.040 and T = 26.1°C (redrawn from11). Left: freeze fracture direct imaging (FFDI) picture showing dark spherical oil droplets of a mean diameter (d) = 2 4 + 9 nm in front of a grey aqueous background. Note that each oil droplet contains a bright domain of elliptic shape, which is interpreted as voids. Right: the freeze fracture electron microscopy (FFEM) picture supports the FFDI result. Each fracture across droplets, which contain bubbles, shows a rough, fractured surface.
TEM-samples, reliable pictures of microemulsion microstructures can be obtained. Let us first consider micrographs of microemulsions prepared at low temperatures (T
S. Buraucr. Inaug. Diss., Elektronenmikroskopie Komplexer Fluide, TENEA, Berlin (2002).
MICROEMULSIONS
5.29
tically shaped bright domains, which can be seen in each n-octane droplet in the FFDI picture. Since the number density and size of the voids are adjustable over broad ranges by changing the initial formulation conditions of the microemulsions, this mechanism for creating bubbles seems to be promising for various technical applications. As can be seen in fig. 5.15, right, the freeze fracture electron microscopy (FFEM) picture of the same sample supports the FFDI result. Oil-in-water droplets of comparable size are visible. In fractures across the droplets, one can see that the bubbles in part exhibit planar fractures, but in other places show a rough surface with overhangs. The inverted structures, that is to say, amphiphilic films curved around the water, are expected at high temperatures. In fig. 5.16 micrographs of a sample are shown, prepared within the one-phase region near the emulsification failure boundary11 on the oil-rich side of the ternary mixture H20/NaCl-n-octane-C12E5 (wcb= 0.050, wA = 0.100, e= 0.006 and T = 36.3°C). The mass fraction of NaCl (E) in the mixture of H2O and NaCl is denoted by e=
[5.3.3] mA + m E
Salinity, measured in terms of e, is a useful parameter for determining the transition occurring from discrete aggregates to a continuous network via electric conductivity measurements with decreasing temperature (see sec. 5.3f). Both pictures, that is, the FFDI on the left and the FFEM on the right-hand side of fig. 5.16, support each other in
Figure 5.16. Micrographs of microcmulsion droplets of the w/o type in the system H2O/NaCln-octane-Cj2Eg prepared near the emulsification failure boundary at m^^ = 0.050, wA = 0.100, £= 0.006 and T = 36.3°C (redrawn from11). Left: freeze fracture direct imaging (FFDI) picture showing bright water droplets of a mean diameter (d) = 44 ± 13 nm against a dark oily background. Right: the freeze fracture electron microscopy (FFEM) picture supports the FFDI result. The mean diameter of the water droplets is (d) = 47 ± 8 nm. 11
L. Belkoura, C. Stubenrauch, and R. Strey, Langmuir 20 (2004) 4391.
5.30
MICROEMULSIONS
proving the existence of the inverted structures. The image obtained with the FFDI technique shows bright water droplets in the dark n-octane matrix, which itself is textured. For the size of the water droplets a mean diameter of (d) = 44 ± 13 nm is found. Looking at the FFEM picture, one clearly sees not only qualitatively but also quantitatively the same overall structure. Here, the mean diameter of the water droplets amounts to (d) = 47 ± 8 nm, which is in nearly perfect agreement with the FFDI picture. Furthermore, one should notice that the texture of the oil domains stems from the shadowing of the fractured surfaces with tantalum (Ta)-tungsten (W)li2). It is caused by the differing nueleation probabilities and surface mobilities on the various substrates and should not be mistaken as real mierostructure. In between these two borderline structures of discrete aggregates of oil In water and vice versa, the mierostructure has to invert in tandem with the phase behaviour. This happens at and around the phase inversion temperature (p.i.t). In sec. 5.2 we have learned that the phase inversion manifests Itself as a trajectory of the middle phase (X(_(jj)), which moves on an ascending curve from the water-rich to the oil-rich side of the phase prism. Therefore, it is interesting to examine which mierostructure accompanies the inversion process along this path. Figure 5.17 shows freeze fracture electron microscopy (FFEM) pictures documenting the change of the mierostructure along the trajectory of the middle phase of the system H2O-n-octane-C12E5 at (* = 0.1 to 0.9 in steps of 0.1. The sample's composition was located as close as possible to the Xpoint, i.e. at wc<wc <wQ+ 0.015, to have a finite temperature interval for preparation. The following general trends can be observed. The area fraction of oil-rich domains, distinguishable from the water-rich domains by their texture, increases as expected from if> - 0.1 to
11 21
W. Jahn, R. Strey, J. Phys. Chem. 92 (1988) 2294. S. Burauer, L. Bclkoura, C. Stubcnrauch, and R. Strey, Colloids Surfaces A228 (2003) 159.
MICROEMULSIONS
5.31
Figure 5.17. FFEM images of freeze-fractured microemulsions of the system H2O -n-octaneC 12 E 5 prepared along the trajectory of the middle phase (see fig. 5.10), i.e. for 0= 0.1-0.9 in steps of 0.1. Starting at 0 = 0.5, a saddle-shaped bicontinuous microstructure can be found. Turning to the water-rich side (tj> < 0.5 ), the structure of the oil domains changes to more and more branched tubes, which means that the amphiphilic film is curved towards the oil ( J > 0 ). Going to (j>> 0.5 , i.e. the oil-rich side the structure is reversed ( J < 0 ) with the tubes consisting of water instead of oil. (Redrawn from Burauer et al. , with permission.)
Turning to the water-rich side (0
11
S. Burauer et al., loc. cit.
5.32
MICROEMULSIONS
the water-rich side cepo , illustrate the transition between connected and discrete structures. Upon following the trajectory of the middle phase towards the oil-rich side (>> 0.5 ), upon which T changes to higher values, one finds inverse structures consisting of water tubes in oil. The water tubes form networks in a continuous oil phase, indicating a negative mean curvature J < 0 . Similarly, as on the water-rich side, at low fractions of the dispersed components
[5.3.4]
2
where 0 is the scattering angle, X the wavelength of the incident radiation and n the index of refraction of the medium. The most commonly used forms of radiation to study microemulsions are light, X-rays and neutrons. For light, n = 1.33 in water, but for X-rays and neutrons n is very close to unity. The scattering vector q may be conceived as a measuring stick for the reciprocal length scales.
MICROEMULSIONS
5.33
To determine the length scale E, of a microstructure with scattering experiments, a q-range is needed, which extends an order of magnitude on each side of q = it I£. This means that in a typical small angle neutron (SANS) and small angle X-ray scattering (SAXS) experiment where q < 6 nm" 1 , the maximum spatial resolution achievable is no more than 0.5 nm, whereas in a typical light scattering (SLS) experiment q < 0.03 nm" 1 , the resolution is no better than 100 nm. Knowing that the characteristic length scales of microemulsions vary between 1 and 100 nm, only SANS and SAXS are appropriate techniques to study the microemulsion microstructure. A further, but rather indirect, method for determining the size of a colloidal particle is dynamic light scattering (DLS) where hydrodynamic properties of the samples are studied. In the dilute limit this technique yields the self-diffusion coefficient D of the particle. The Stokes-Einstein relation D=
[5.3.5] 6itT]ah
can be used to relate D to the hydrodynamic radius a h , provided the solvent viscosity 77 is known. Apart from the spatial resolution the nature of the interaction between the radiation used and matter is important, that is the contrast. In the case of light and X-rays, the interaction is between the electric field of the radiation and the electrons. As neutrons do not carry an electrical charge, in almost all situations they exclusively interact with the nuclei. This means that for X-rays the scattering intensity of hydrocarbons and water are not significantly different, but for neutrons the scattering intensity can be made to vary greatly by substitution of the protons by deuterons (contrast variation, see below). In particular, the latter feature has led to an intensive use of SANS for studying the microstructure of microemulsions. SAXS, on the other hand, has been of comparatively little use because of its limited contrast. We recall that in chapter IV.5 many illustrations are given for concentrated colloids. 5.3d SANS In a SANS experiment, a parallel beam of neutrons ( X = 0.1-2 nm), of intensity IQ incident on a flat sample cell containing the microemulsion, is scattered in a small cone around the forward direction. The measurement of the scattered neutron intensity Js at an angle 6 is carried out with a detector at a distance I away from the sample. Afterwards, the raw data are reduced and made absolute by measuring the scattering of a standard. This quantity is referred to as the scattering cross section per unit volume d£/d/2(q) or, alternatively, as the absolute intensity I{q). While the realization of a SANS experiment is rather straightforward, the quantitative analysis and interpretation of the scattering data are highly complicated. In principle, the data analysis can be performed in two ways. The first way is to compare the scattering data with calculated scattering curves of model structures. In the
5.34
MICROEMULSIONS
second procedure, the scattering data are smoothed, corrected for instrumental broadening from geometry and wavelength effects, and Fourier transformed 12 ' 3141 . The Fourier transformation yields the radial distribution function, or pair correlation function g[r), and the scattering length density profile p(r), which allows the determination of the basic geometry even for inhomogeneous particles. Generally, the latter procedure is preferred if it is unknown what structure is expected. However, having visualized the microstructure via transmission electron microscopy, we are now able to analyze the data by comparing them with the calculated scattering curve of the corresponding structure. Let us first consider the scattering curves of oil-in-water (o/w) and water-in-oil (w/o) droplet microemulsions of the ternary mixture D 2 O -n-CgD 1 8 -C 1 2 E 5 . Thereby, the scattering of the amphiphilic film was obtained by using deuterated solvents having almost equal scattering length densities, resulting in the so-called film contrast. This contrast allows the precise determination of the radius of the swollen micelles as well as their dispersity. Figure 5.18 shows the scattering curve of oil-swollen micelles in water prepared within the one-phase region near the emulsification failure boundary on the water-rich side (wCa= 0.0707, wB = 0.0561, and T = 18.0°C). The corresponding scattering curve of water-swollen micelles in oil prepared near the emulsification failure boundary on the oil-rich side ( t u C b = 0.1260, wA = 0.0959 and T = 46.0°C) is given in fig. 5.18(b). It should be noted that the temperatures of both samples have almost the same distance (±14 K) from the phase inversion temperature T m » 32°C. Looking for a moment only at the experimental data, it can be seen that both curves have a shape, which is typical for polydisperse spherical objects in film contrast. Starting at low values of q, the absolute intensity l(q) is high and almost constant. With increasing q the intensity decreases, runs through a minimum and a maximum until it decreases as q" 4 e~? £ 5) , and reaches the incoherent background intensity Ancoh a * l a r g e values of q. While the scattering curve of the o/w-microemulsion shows a distinct minimum and maximum, the curve of the w/o-microemulsion displays only a plateau, which points to a larger polydispersity of the micelles. Knowing from transmission electron microscopy that the microstructure of these microemulsions consists of almost spherical droplets, one can try to describe the experimental data by the scattering intensity of discrete aggregates 7(q) = JVP(q)S(q) + / i n c o h
11
[5.3.6]
O. Glatter, J. Appl. Crystallogr. 10 (1977) 415. O. Glatter, in Small Angle X-ray Scattering, O. Glatter, O. Kratky, Eds., Academic Press (1982) 119. 31 J. Brunner-Popela, and O. Glatter, J. Appl. Crystallogr. 30 (1997) 431. 41 R. Strey, O. Glatter, K.V. Schubert, and E.W. Kaler, J. Phys. Chem. 105 (1996) 1175. 51 M. Gradzielski, D. Langevin, L. Magid and R. Strey, J. Phys. Chem. 99 (1995) 13232. 21
MICROEMULSIONS
5.35
Figure 5.18. Small angle neutron scattering curves for droplet microemulsions of the ternary mixture D 2 O-n-CgD]g-Cj 2 Eg prepared under film contrast conditions at the corresponding emulsification failure boundary on a double-logarithmic plot. Panel a: o/w microemulsion (a = 0.0601, a>c = 0.0667 and T= 18.0°C). Panel b: w/o microemulsion {a = 0.8918, wc = 0.1139 and T = 46.0°C). The parameters r , a and t will be explained in the text below. Both scattering curves have qualitatively the same shape. However, the distinct minimum and maximum, which can be found in the o/w curve, converge at a plateau of the w/o curve. Both curves can be described almost quantitatively using the particle form factor of polydisperse shells 1 ' 2 and the Percus-Yevick structure factor3 . From these results and the TEM-pictures, it can be concluded that nearly spherical swollen micelles exist at the respective emulsification failure boundary. Here, JV is the number density of aggregates, P(q) the particle form factor, S[q) the interparticle structure factor (see sec. I.7f) and i i n c o h the incoherent background intensity. Using the particle form factor of polydisperse shells", one obtains 1{q]
= , + t ? C / 2 U c ( y ) 2 e x P ' - q 2 t 2 ) ( / 1 ( q ) + / 2 (q) + / 3 (q) + /4(q))S(q) + I incoh i + o~ i r0 a c r 0 q
[5.3.7]
where /l(q) = ^ q 2 t 4 (l + cos2qr0 exp(-2cr2q2))
[5.3.7a]
/ 2 (q) = qt2 (r0 sin2qr0 + 2qcr2 cos2qr0)exp(-2cr2q2)
[5.3.7b]
/ 3 (q) = | r o ( l - c o s 2 q r 0 e x p ( - 2 a V ) )
[5.3.7c]
2
/ 4 (q) = — (l + 4qr 0 sin2qr 0 exp(-2(T 2 q 2 )+cos2qr 0 (4cr 2 q 2 -l)exp(-2cr 2 q 2 )) [5.3.7d]
11
M. Gradzielski, et al. loc. cit. R. Strey, Colloid Polymer Sci. 272 (1994) 1005. 31 J.K. Percus, G.J. Yevick, Phys. Rev. 110 (1958) 1. 41 N.W. Ashcroft, J. Lekner, Phys. Rev. 145 (1966) 83. 21
5.36
MICROEMULSIONS
Equation [5.3.7] is deduced by assuming the amphiphilic film to be penetrated by the solvent molecules giving rise to a Gauss scattering length density profile characterized by a thickness parameter t. Furthermore, a droplet polydispersity is included as a Gauss distribution with mean radius r0 and standard deviation a (see sec. 1.3.7a). In [5.3.7] the contrast Ap is the difference in scattering length density between the solvents and the amphiphilic film, the volume vc and the area per surfactant molecule ac , and its volume fraction in the internal interface being 0g . Even if for these moderately concentrated samples the absolute scattering intensity l(q) mainly reflects the particle form factor, we used the interparticle structure factor of monodisperse hard spheres according to Percus-Yevick11. As a function of the dispersed volume fraction of swollen micelles 0di d ~ 2r0 + vc/ac,
and the hard sphere diameter
one obtains21
S(q) = ^
[5.3.8] l-JVc(q)
where
\( 2 A (sin qd -qd cos qd) + B NC{Q) =
1 < q d 3 t>dispA\ 24 o 3^3" 3
2
L d
+ 4
(
1
6 "I . ( 2^2" S l n q d ~ 1
V 9
d
21 '
\
—=—^--1 qdcosqd + 2sinqd \\q2d2 ) qd\
)
12 24 ^ 1 2^2"+ ^ ^ qdcosqd
V 9
d
1
d
)
J [5.3.8a]
with
B = 36^f^4-
[5 3 8bl
--
(l-^disp)
Figure 5.18 shows the fit I(q) of the scattering curves according to [5.3.6] (full curves) together with the used form factors for the polydisperse shell P(q) and the PercusYevick structure factor S(q) (dashed curves). An almost quantitative agreement between the experimental data and the model is observed. Only small but systematic deviations are found in the q-range of the shoulder, presumably caused by using the interparticle form factor S(q) of monodisperse, instead of polydisperse, hard spheres. The individual fit parameters have a different sensitivity in different regions of the scattering curve. The radius r0 is obtained from the position of the minimum, a from the smearing-out of the minimum/maximum region and t from the decay at large values of q. As the temperatures of both samples have almost the same distance from the phase inversion temperature, the values of the parameter can be compared quantitatively. One finds that the radius rQ of the oil-swollen micelles is larger than that of the water-swollen micelles. However, p = o7r Q , which is a measure of the polydispersity, is considerably larger for the latter micelles. Similar differences in polydispersity 11 21
J.K. Percus, G.J. Yevick, Phys. Rev. 110 (1958) 1. Also see chapter IV.5. N.W. Ashcroft, J. Lekner, Phys. Rev. 145 (1966) 83.
MICROEMULSIONS
5.37
between oil- and water-swollen micelles have been found for other microemulsion systems1'2'. While the length scales of the microstructure will be described below In connection with the respective tuning parameter, as for the temperature, we confine ourselves here to the discussion of the shapes of the microstructures. Having studied the microstructure of o/w- and w/o-microemulsions by TEM, as well as SANS, we can conclude that at temperatures far below the p.i.t. nearly spherical oil-swollen micelles are observed, whereas at temperatures far above the p.i.t. the inverse, spherical waterswollen micelles can be found. In order to examine how the microstructure varies along the phase inversion process, SANS experiments are performed along the trajectory of the middle phase (X(
11
P. Sicoli, D. Langevin, and L.T. Lee, J. Chem. Phys. 99 (1993) 4759. T. Hellweg, A. Brulet, and T. Sottmann, Phys. Chem. Chem. Phys. 2 (2000) 5168. 31 R. Strey, J. Winkler, and L. Magld, loc. cit. 41 M. Teubner, R. Strey, J. Chem. Phys. 87 (1987) 3195. 21
5.38
MICROEMULSIONS
As can be seen, the scattering curve of the symmetrical microemulsion at 0 =0.5 shows a characteristic structure peak. At large values of the scattering vector q, the scattering intensity I(q) decreases as q~4e~q ' until the incoherent background linco^ is reached1'21. Both with decreasing and increasing values of 0, one finds a weakening of this peak, while at the same time the scattered intensity towards small q increases instead of becoming constant. It is characteristic that the position of the peak remains almost constant for 0 between 0.2 and 0.7, indicating that the periodicity of the structure Is nearly invariant. Upon further decreasing or increasing 0, that is, moving towards the critical endpoints the peak disappears, but a shoulder remains in the scattering curve as an indication of persisting structure. Furthermore, critical scattering develops near q = 0, that is critical scattering and scattering from the microstructure are superimposed. Thus, the existence of the shoulder in the scattering curve recorded near the critical endpoint helps to discriminate between weakly and strongly structured microemulsions as the lobes aQ in the phase behaviour (see fig. 5.13). While for weakly structured microemulsions the scattering peak moves to zero as the wetting transition is approached3'4'51, the scattering spectra of strongly structured microemulsions show the persistence of local structure through the wetting transition . The SANS results are in line with the FFEM pictures, which have shown the occurrence of a dispersion of bicontinuous domains as the critical endpoints were approached. Knowing from transmission electron microscopy that the microstructure of, in particular, the microemulsion at 0.2< 0 <0.7, is bicontinuous, the scattering curves can be described by the Fourier transformation of a correlation function, which combines the alternation of the water-containing and oil-containing domains and the absence of long-range order. Accordingly, the absolute intensity I(q) is given by the Teubner-Strey formula81 8?rc24>a
~
a2+Clq2+C2q4
+
^
^
^
where Ap is the scattering length density difference between the two sub-phases a and b, <pa and
MICROEMULSIONS
5.39
density. The correlation length £TS and the periodicity d TS of the bicontinuous structure are then defined by
and
*-(&r-feir respectively. In order to quantify the amphiphilic strength, it has been proven useful to formulate an amphiphilicity factor / a = r^— V 4a 2 c 2
[5.3.12]
which is typically close to -1 for strongly structured microemulsions. More details are discussed in refs. L2) and 3). The full curve in fig. 5.19 emphasizes that, in particular, the peak of the scattering curve at
[5.3.13]
which sets the overall length scale £ in microemulsions, that is £~ [A/V)~l. Here, ac and vc denote the area and volume per surfactant molecule, respectively, and 0° the volume fraction of surfactant at the interface. Taking into account the diffuseness of the locally flat interfacial film, Strey et al.5) showed that the convolution of a Gauss scattering density profile of standard deviation t describes the scattering intensity at large q by
11
K.V. Schubert, R. Strey, S. Kline, and E.W. Kaler, J. Chem. Phys. 101 (1994) 5343 G. Gompper, M. Schick, Self-assembling Amphiphilic Systems, Academic Press (1994). 31 H. Lcitao, M.M. Telo da Gama, and R. Strey, J. Chem. Phys. 108 (1998) 4189. G. Porod, in Small-angle X-ray Scattering, O. Glatter, O. Kratky, Eds., Academic Press (1982) pp. 18. 51 R. Strey, J. Winkler, and L. Magid, J. Phys. Chem. 95 (1991) 7502).
5.40
MICROEMULSIONS
lim[/(q)] = 2^(Ap)2 ~ q " 4 exp(-q 2 t 2 ) + I incoh q^°= V
[5.3.141
In fig. 5.19 the description of the data at large q by [5.3.14] is shown as a full curve for the scattering curve at
= 2 / r 2 ^ (Apf
[5.3.151
0
is used in [5.3.14]. The specific interfacial area A/V is then given directly by the scattering intensity at large q according to A -# a 0hq 4 exp(q 2 t 2 )f ^ ^= |llm[J(q)]-J tacoh J Q ^
[5.3.16]
Thereby, Q can be determined either numerically from the given spectra by integration or analytically, provided the scattering length density difference Ap is known and the spectrometer is well calibrated. In conclusion, the SANS experiments confirm the pattern of structure changes in non-ionic microemulsions found by TEM. With increasing temperature, the structure changes gradually from discrete o/w-micelles via a bicontinuous network to discrete w/o-micelles. Additionally, SANS provides the length scale of the microstructure by analyzing the scattering curves in the appropriate way. However, before discussing the variation of the length scale in detail, let us consider briefly two more indirect, albeit useful, methods for studying the nature of the microstructure. 5.3e NMR diffusometry NMR diffusometry is an elegant method for obtaining insight into the structure of microemulsions by measuring the self-diffusion coefficients of the various components of the system. Among the different NMR techniques, the Fourier transform pulsed field gradient spin-echo NMR (FT PGSE NMR) approach has been proven to be superior11. In this technique, the displacements of nuclear spins in a controlled magnetic field gradient are monitored and the contributions of different components are resolved by Fourier transformation of the NMR signal21. Thus, this technique easily discriminates between droplet and bicontinuous microemulsions31 by simultaneously measuring the transport properties on the two sides of the surfactant monolayer, as well as in the amphiphilic film itself. In contrast, the closely related technique of measuring electric
11 21 31
B. Lindman, U. Olsson, Ber. Bunsenges. Phys. Chem. 100 (1996) 344. O. Soderman, P. Stilbs, Prog. Nucl. Mag. Reson. Spectrosc. 26 (1994) 445. U. Olsson, K. Shinoda, and B. Lindman, J. Phys. Chem. 90 (1986) 4083.
MICROEMULSIONS
5.41
conductivity (see below) can only discriminate between water-continuous and discontinuous structures. In the simplest version, the PGSE experiment consists of two equal and rectangular gradient pulses of magnitude g and length 5, sandwiched on either side of the 180° pulse, in a simple Hahn echo experiment. For molecules undergoing free diffusion characterized by a diffusion coefficient of magnitude D, the echo attenuation due to diffusion is given by I = I0exp[-(ymgS)2(A-^)D^
[5.3.17]
where A presents the distance between the leading edges of two gradient pulses, ym is the gyromagnetic ratio of the monitored spin, and IQ denotes the echo intensity in the absence of any field gradient. By varying g, S or A , the diffusion coefficient D is obtained by fitting [5.3.17) to the observed intensities. Having determined the self-diffusion coefficients of water (DA) and oil (DB) in a microemulsion, the connectivity of the solvent domains and, therefore, the shape of the microstructure can be easily obtained from the relative values of the self-diffusion coefficients. For the three limiting cases, o/w, w/o-droplets and bicontinuous structures, we expect the following: o/w: DB « DA w/o: DA « DB bicontinuous: DA = DB Furthermore, for discrete o/w- and w/o-droplet structures, the hydrodynamic radius of the droplet can be estimated from the diffusion coefficient of the dispersed phase using the Stokes-Einstein relation (see [5.3.5]). This holds because the droplets have smaller extensions than the distance monitored in a PGSE experiment. Therefore, such an experiment is not sensitive to the molecular displacement within the droplets, but only to the translation of the entire droplet. Additional information about the shape of discrete aggregates can be obtained by considering the obstruction effects of the droplets with respect to the diffusion of molecules of the continuous medium. Then the diffusion coefficient simply obeys O = D0(l-MdisP+-)
I5-3-18!
where ks is a dimensionless constant, which depends on the aggregate geometry and the interactions between them . Furthermore, one has to consider that if some of the continuous medium molecules diffuse with the aggregate, this will slow down the B. Jonsson, H. Wennerstrom, P.G. Nilsson, and P. Linse, Colloid and Polymer Sci. 264 (1986) 77.
5.42
MICROEMULSIONS
Figure 5.20. Variation of the HDO (shown as open symbols) and n-octane (shown as filled symbols) self-diffusion coefficients DA and Dg in the system HDO-n-octane-C^E^ as a function of 0, i.e. of temperature, at a constant surfactant mass fraction of WQ = 0.07 (redrawn from11). The curves for DA and D B show a sigmoidal shape, DA decreases by one order of magnitude with increasing 0 , whereas D B increases, indicating a continuous change of the free diffusion path close to the inflection points of the curves.
diffusion according to21 D
= PfreeDfree+PagDag
I 5 - 3 - 19 !
Herein, p free and p are the fractions of the molecules, which can diffuse freely and with the aggregates, respectively. Dfree and D are the corresponding diffusion coefficients. Keeping this in mind, NMR diffusometry provides another piece of evidence for solving the puzzle of the structure of microemulsions. Figure 5.20 shows the variation of the self-diffusion coefficients of HDO (DA open symbols) and n-octane (D B filled symbols) as a function of the oil-to-water plus oil volume fraction
M. Kahlweit, R. Strey, D. Haase, H. Kuneida, T. Schmeling, B. Faulhaber, M. Borkovec, H.F. Eicke, G. Busse, F. Eggers, T. Funck, H. Richmann, L. Magid, O. Soderman, P. Stilbs, J. Winkler, A. Dittrich, and W. Jahn, J. Colloid Interface Sci. 118 (1987) 436. B. Jonsson, B. Lindman, K. Holmberg, and B. Kronberg, Surfactants and Polymers in Aqueous Solution, John Wiley (1998]. 31 K. Shinoda. H. Saito, J. Colloid Interface Sci. 26 (1968) 70.
MICROEMULSIONS
5.43
side, has been proven useful to characterize the variation of the microstructure 12 ' 341 . Looking at fig. 5.20, it can be seen that at low values of 0 the water (HDO) diffusion is rapid while the n-octane diffusion is slow. The opposite situation holds for high values of 0, while at intermediate values of <j> the diffusion of both components is relatively rapid and the diffusion coefficients are equal in magnitude. Thus, the data again imply a change in the structure from oil-in-water (o/w) droplets via a bicontinuous structure to water-in-oll (w/o) droplets with increasing 0, i.e. increasing temperature T. Remarkably, a large range of bicontinuity is observable around the phase inversion. The crossover point (where the diffusion coefficients of water and oil are equal) is found at 0 ~ 0.50 and here the microemulsion is truly bicontinuous (compare the FFEM picture at 0 = 0.5 in fig. 5.17). We note that the dividing film between the water and the oil component is itself sample spanning. So, actually the structure is even "tricontinuous". Thus, the results obtained by NMR-diffusometry confirm and deepen our knowledge of the microstructure of microemulsions. 5.3f Electric conductivity Another method for studying the continuity of microstructures is measuring the electric conductivity of the respective microemulsion samples. Despite the rather simple experimental set-up, measurements of the electric conductivity enable the discrimination between water-continuous and discontinuous structures. Applied to microemulsions, this means that if such a thermodynamic mixture is driven through the phase inversion, i.e. the structure changes from water to oil-continuous, the electric conductivity has to show a characteristic jump. This behaviour makes the technique advantageous for the technical formulation of emulsions51 with unknown or difficult to control ingredients, e.g., as in reaction vessels. Scientifically this technique was mainly used to study the transition from discrete to continuous structures, that is percolation In oil-rich, ionic microemulsions6'71. In order to complete our understanding of the microstructure patterns of microemulsions, let us consider the variation of the electric conductivity K along the macroscopically isotropic "one-phase channel" of the system H2O/NaCl-n-octaneC 12 E 5 at a constant wc = 0.07 and £=5.8xl0~ 5 , where s is given by [5.3.3]. An electrolyte like NaCl has to be added to obtain a well-defined electric conductivity different from zero. Such a relatively small amount has only negligible influence on the
11
M. Kahlweit et al., loc. cit. F. Lichterfeld, T. Schmeling, and R. Strey, J. Phys. Chem. 90 (1986) 5762. 31 U. Olsson, K. Shinoda, and B. Lindman, J. Phys. Chem. 90 (1986) 4083. 41 W. Jahn, R. Strey, J. Phys. Chem. 92 (1988) 2294. 51 J.L. Salager, M. Minana-Perez, M. Perez-Sanchez, M. Ramirez-Gouveia, and C.I. Rojas, J. Dispersion Set Techn. 4 (1983) 313. 61 H.F. Eicke, J.C.W. Shepherd, and A. Steinmann, J. Colloid Interface Set 56 (1976) 168. 71 H.F. Eicke, R. Hilfiker, and M. Holz, Helv. Chim. Acta 67 (1984) 361. 21
5.44
MICROEMULSIONS
phase behaviour of non-ionic microemulsions11. The experiments are performed temperature-wise from the lower to the upper boundary of the homogeneous channel for various
Figure S.21. Electric conductivity K as a function of temperature in the system PLjO/NaCl-noctane- Cj 2E5 for various oil-to-water plus oil volume fractions 0. The experiments were performed within the one-phase channel at a constant surfactant mass fraction of wc = 0.07 and £ = 5.8 x 10~^ . The electric conductivity decreases by many orders of magnitude from the water-rich to the oil-rich side. As the temperature is increased from the lower to the upper boundary of the homogeneous channel at a constant
M. Kahlweit, R. Strey, P. Finman, D. Haase, J. Jen, and R. Schomacker, Langmuir 4 (1988) 499.
MICROEMULSIONS
5.45
it can be seen that from the water-rich side up to Q~ 0.768, the structure of the microemulsion is water-continuous over the entire one-phase temperature interval. A further increase in tp leads to a strong decrease of the electric conductivity and thus to more and more water-discontinuous, i.e. isolated droplet structures, in particular at the high temperature end of the one-phase temperature interval. Comparing the
Figure 5.22. Overview of the microstructure of non-ionic microemulsions as deduced from TEM, SANS, NMR-diffusometry and electric conductivity (schematic) together with the underlying phase behaviour. The shaded regions represent the oil and the white regions represent water, (a) T[WQ) section at a constant 0= 0.5. The variation of the mean curvature of the amphiphilic film with temperature becomes apparent by the change of microstructure from o/w to w/o-droplet structures. In between, and around the p.i.t. bicontinuous structures can be found at low wc , whereas the lamellar phase La exists at higher wc (redrawn from ref. 1!). (b) T(0) section through the phase prism at a constant wc > WQ ("Shinoda cut"). Within the homogeneous channel, the microstructure changes from discrete w/o-structures on the waterrich side and low temperatures to discrete o/w-structures on the oil-rich side and high 91
temperatures (redrawn from ref. 11
).
R. Strey, Colloid and Polymer Sci. 272 (1994) 1005. M. Kahlweit, R. Strey, and R. Schomacker, in Reactions in Compartmentalized Liquids, W. Knoche, R. Schomacker, Eds. Springer (1989) 1.
5.46
MICROEMULSIONS
figs. 5.4 and 5.7) into a schematic representation drawing the possible microstructures into the extended one-phase region behind the X-point (redrawn from ref."). Starting at the X-point, i.e. near the three-phase region, the structure of the microemulsion is bicontinuous with a zero mean curvature of the amphiphilic film ( J = 0), but a negative Gauss curvature (K < 0). As the surfactant mass fraction wc increases, the length scale of the structure becomes smaller because the total area of the internal interface increases. At high surfactant concentrations, the lamellar phase is observed with a zero curvature structure, i.e. J = 0 and K = 0. Moving both wc and temperature-wise away from the X-point a transition to oil-in-water and water-in-oil droplets is found at low and high temperatures, respectively, with the droplet size decreasing as one moves further away from the X -point. The change of structure as a function a of the oil-to-water plus oil volume fraction
K. Shinoda, H. Saito, J. Colloid Interface Set 26 (1968) 70. T. Sottmann, R. Strey, and S.H. Chen, J. Chem. Phys. 106 (1997) 6483.
MICROEMULSIONS
5.47
or oil domain and identify it by half the periodicity £ = ^E§-
[5.3.20]
we obtain a useful and characteristic measure of the length scale E, of bicontinuous microemulsions. For thermodynamic reasons and for the interpretation of interfacial tensions, the maximum length scale £ , i.e. the length scale of the microstructure exactly at the X-point, is of interest. Considering that the characteristic length scale is proportional to the inverse fraction of surfactant molecules in the internal interface (#g , £ can be obtained by extrapolation ( £ = £0g /0g ). In fig. 5.23a, the maximum length scale £ is plotted versus the interfacial volume fraction of surfactant >£ at the X-point of the 18 systems mentioned above on a double-logarithmic scale. As can be seen, £ varies over more than one order of magnitude from 5 to 60 nm. The negative slope of -1.12 confirms that the length scale is indeed inversely proportional to the internal volume fraction of surfactant, i.e. £ is the larger, the smaller
[5.3.21]
Figure 5.23. Maximum length scale cf and associated specific interfacial area A/V obtained by analyzing the SANS-curves of 18 different bicontinuous D 2 O-n-alkane-C n E x microemulsions prepared close to the X -point at >= 0.50. (a) Length scale g obtained by analyzing the scattering peaks as a function of the volume fraction of surfactant ~0Q forming the internal interface, (b) £ plotted versus the right-hand side of [5.3.21] to determine the prefactor a, using the specific area of the internal interface A/V obtained from the large-q part of the scattering curves (see fig. 5.19).
5.48
M1CROEMULSIONS
where the absolute value of the prefactor a is different between the various models. The model of Debye et al.11 predicts a = 4 , the Voronoi tessellation of Talmon and Prager21 a = 5.84 and the model of cubes by De Gennes and Taupin31 yields a = 6 . As the specific area of the internal interface AIV can be determined independently by analyzing the large-q part of the scattering curve (see fig. 5.19 and [5.3.16]), the "exact" value of a can be found experimentally by plotting t, versus the right-hand side of [5.3.21] (with 0=0.5), i.e. (4A/V)' 1 using the extrapolated specific area of the internal interface A/Vat the X-point. The result is shown in fig. 5.23b differentiating between measurements performed in 1.0 mm (open symbols) and 0.2 mm (filled symbols) sample cells. Such thin cells are used in order to minimize multiple scattering, which could lead to deviations in the large-q part of the scattering curves and hence to significantly larger values of A/V. As can be seen, both data sets fall onto a straight line passing through the origin. The marginally different slopes are a consequence of the multiple scattering, which overestimates A/V . The measurements performed in the 1.0 mm cell result in a prefactor of a{ 0 = 7.5, while the prefactor determined for the 0.2 mm cells amounts to a 0 2 = 7.0. Both experimental values compare very favourably with any of the models. (ii) Area per surfactant molecule Figure 5.23b has been constructed using the experimentally determined values for A/V
subject to the relatively large errors in the large-q part of the individual
scattering curves caused by the low values of the absolute scattering intensities. However, the high intensity at the position of the peak and from it the maximum length scale £, is more precisely known. The mutual support of the data points, in conjunction with [5.3.21], suggests turning the argument around to obtain more precise values for the area per surfactant molecule a c from
a
0(1 - 0) IV. -aZX^ZlJL
[5.3.22]
where all quantities on the right-hand side are known. Using the presumably most reliable prefactor aQ 2 = 7.0 found for the 0.2 mm cells, a c is plotted versus the head group size x of the C n E x surfactants in fig. 5.24. This is possible because the area per surfactant molecule is neither dependent on the alkyl chain of the surfactant n nor the oil chain length k. As can be seen, all individual points fall nearly on a straight line. Accordingly, the area per surfactant molecule of the C n E x type may rather precisely be calculated from the regression
11
P. Dcbyc, H.R. Anderson, and H. Brumberger, J. Appl. Phys. 28 (1957) 679. Y. Talmon, S. Prager, J. Chem. Phys. 69 (1978) 2984. 31 P.G. Dc Gennes, C. Taupin, J. Phys. Chem. 86 (1982) 2294. 21
MICROEMULSIONS
5.49
Figure 5.24. Variation of the area per surfactant molecule OQ as a function of the surfactant headgroup length x for bicontinuous D2O -nalkane-C n E x microemulsions. Note that OQ increases almost linearly with x, while it is independent of the length of the alkyl chain n of the surfactant as well as of the oil chain length k.
a c = 0.284 + 0.0609 x
(nm2)
[5.3.23]
As a result of the consideration of the multiple scattering, the areas per surfactant molecule are slightly (less than 5%) smaller than those presented in ref.11. In this connection, it is interesting to note that the areas per surfactant molecule at the air-water interface determined by surface tension21 and neutron reflectivity measurements31 are still somewhat (but systematically) lower41. There may be various reasons for this difference to be real. One option is that the liquid-air surface tension refers to a flat, that is, projected area, while the Porod analysis traces out the actual, possibly quite undulating, surfactant interface. Also, at the air-water interface, the interfacial tension is much higher (of the order of 30 mN/m) because the contribution to the film pressure by the penetrating n-alkane molecules is absent. (Iii) Variation of the length scales Having determined the maximum length scale of the microstructure £, at the optimum point of microemulsions, let us now concentrate on the variation of the characteristic length scale £ with the respective tuning parameter. To be consistent, we once again choose the temperature T (as the tuning parameter). The variation of the length scale
11
T. Sottmann, R. Strey, and S.H. Chen, J. Chem. Phys. 106 (1997) 6483. M.J. Rosen, A.W. Cohen, M. Dahanayake, and X.Y. Hua, J. Phys. Chem. 86 (1982) 541. 31 J.R. Lu, Z.X. Li, T.J. Su, R.K. Thomas, and J. Penfold, Langmuir 9 (1993) 2408. These results may also be compared with those of fig. III.4.31, where n is varied at fixed x, resulting in an odd-even disparity. 21
5.50
MICROEMULSIONS
Figure 5.25. Temperature dependence of the characteristic length scale £ for the three systems, H 2 O-n-octane-C 8 E 3 , C 10 E 4 and C 12 E 5 . All three £(T)- curves are inverted wedgeshaped and run through a maximum at their respective phase inversion temperatures. The length scales determined by SANS are shown as filled symbols, the length scales calculated from composition, i.e. [5.3.24] as open symbols. The drawn curves are calculated from [5.3.28]. VC < * i + ^ "
r0 = 3 - £ 0 ac
l
+ p
2
_2__LLP^
[5.3.24]
under the assumption that the droplets are spherical and the area per surfactant molecule a c is known (see [5.3.23]). Here, 0i denotes the volume fraction of the solubilized component and p = a I r0 is the polydispersity of an assumed Gauss distribution of radii. If the interfacial tension yab between the microemulsion and excess phase (see sec. 5.4), i.e. water and oil, is known, p can be computed from11
11
M. Gradzielski, D. Langevin, and B. Farago, Phys. Rev. E53 (1996) 3900.
MICROEMULSIONS
5.51
In fig. 5.25 the determined length scales E, are plotted logarithmically versus the temperature for the three systems, H2O-n-octane-C8E3 (top), C1QE4 (centre) and C 12 E 5 (bottom). As one can see,
11
R. Strey, Colloid Polym. Set 272 (1994) 1005.
5.52
J =c
MICROEMULSIONS
2(T -T) ^ '-
[5.3.26]
"c Here, the prefactor c is the temperature coefficient of the mean curvature J . Equation [5.3.26] accurately characterizes the borderline cases, at T~Tm and T«Tm, T » Tm . While for temperatures around the mean temperature Tm the second term in the denominator is negligible and, hence, J = 2c(Tm - T), for temperatures far away from T m , J amounts to 2ac/vc. Thus, using [5.3.26] the experimental data of the system H2O -n-octane-C 12 E 5 are quantitatively described over the whole temperature regime.
Figure 5.26. Temperature dependence of the mean curvature J for the three systems, H2O n-octane-C 8 E 3 , C 10 E 4 and C 12 E 5 . The data, which originate from SANS, are shown as filled symbols, the data calculated from the composition as open symbols. The solid line is an empirical description of the variation of the mean curvature J according to [5.3.26]. The dashed lines are the calculated principal curvatures ct and c 2 (see [5.3.27]). Note that while the J[T) curve is almost linear for the C 12 E 5 -system, it becomes increasingly sigmoidal for the C 10 E 4 and C 8 E 3 systems. (Redrawn from R. Strey, loc. clt.)
MICROEMULSIONS
5.53
Looking at the J{T) curve of the C 1 0 E 4 and C 8 E 3 systems, one can see that with decreasing surfactant chain length the deviations from the linear trend become increasingly pronounced. However, [5.3.26] still approximates the experimental data of these systems almost quantitatively by an increasingly sigmoidal curve. Concomitantly, the temperature coefficient c increases with decreasing surfactant chain length, having the numerical values 0.0165, 0.022 and 0.035 K"1 nm" 1 , for C 1 2 E 5 , C 1 0 E 4 and C 8 E 3 , respectively. Knowing the mean curvature J as well as the shape of the microstructure, the temperature variation of the two principal curvatures cl and c 2 , which compose J (see [5.3.2]), can be generated. Both principal curvatures remain finite for bicontinuous microemulsions. At the mean temperature of the three-phase body T m , J becomes zero and Cj = - c 2 . Furthermore, TEM, NMR diffusometry and electric conductivity provide evidence that near Tj locally cylindrical oil domains exist, while near T u cylindrical water domains can be found11. Therefore, one has to postulate that Cj = 0 near Tu and c 2 = 0 near Tj . Additionally, the ci s have nearly equal positive values for temperatures far below T m (o/w droplet microemulsions) and equal negative values at temperatures far above T m (w/o droplet microemulsions). Combining these observations, one might expect the principal curvatures to be given by Cj=c
T - T
and
c2 = c
l + c-^-IT -T) a C
l
T-T
-
[5.3.27]
l + c-S-lT,-T) a C
In fig. 5.26, 2Cj and 2c 2 are shown as dashed lines. As postulated, the CjS are equal and of opposite sign at T m , pass through zero at Tj and Tu , respectively, and become almost equal at temperatures far away from Tm . The fact that at T m the principle curvatures c ( remain non-zero, while the mean curvature J passes through zero, leads to a distinct maximum of the length scale £ . Thus, instead of simply assuming £= ( J / 2 ) " 1 , which would let £, diverge, a characteristic length scale t;, defined empirically as £=
.
3
,
[5.3.28]
~ K + c 2 | + ^ ( c ? + c | ) + + 4^( c 4 +c 4) leads to a convenient description of the temperature variation of £. As already seen in fig. 5.25, [5.3.28] fits equally well the observed length scales both in the droplet and in the bicontinuous region.
11 M. Kahlweit, R. Strey, D. Haase, H. Kunieda, T. Schmeling, B. Faulhaber, M. Borkovec, H.F. Eicke, G. Busse, F. Eggers, T. Funck, H. Richmann, L. Magid, O. Soderman, P. Stilbs, J. Winkler, A. Dittrich, and W. Jahn, J. Colloid Interface Sci. 118 (1987) 436.
5.54
MICROEMULSIONS
(v) Scaling of length scales and curvatures Since the variation of the characteristic length scale £ with the temperature has the same inverted wedge shape for the three non-ionic microemulsion systems studied (see fig. 5.25), it suggests itself to scale the length scales similar to the trajectories of the middle phase (see fig. 5.11) by appropriate parameters onto each other. Looking at the ^(T)-curves, in particular, in the region of the maximum, it becomes obvious that the steepness of the £(T) -curves correlates directly with the height of the three-phase body AT = (Tu - Tj). Following the scaling description of the trajectory of the middle phase, the
Figure 5.27. Scaling of the temperature variation of the characteristic length scale £ (a) and the mean curvature J (b) for the systems H2O-n-octane-CgE3 (triangles), CJQE 4 (squares) and C[2Eg (spheres). Following the scaling description of the three-phase body (see sec. 5.2h, fig. 5.11), the temperature axes are reduced by subtracting the mean temperature of the threephase body Tm and normalizing by AT/2 . Multiplying £ by the volume fraction of surfactant 0(5 residing at the microscopic water/oil interface of the optimal microemulsion makes all £,(T) curves collapse into a single curve. An almost quantitative scaling of the J{T) -curves is obtained by dividing the mean curvature J by 0g .
MICROEMULSIONS
5.55
Together with the scaling behaviour of the three-phase body, the scaling of the characteristic length scales £ and the mean curvatures J suggests the existence of a corresponding state of microemulsion systems. In that case, the scaling of both the phase behaviour and the microstructure by the same three parameters,
P.A. Winsor, Solvent Properties of Amphiphilic Compounds, Butterworth & Co.(1954). L.M. Prince, Microemulsions: Theory and Practice, Academic Press (1977). 31 P.G. De Gennes, C. Taupin, J. Phys. Chem. 86 (1982) 2294. 41 M.-J. Schwuger, K. Stickdorn, and R. Schomacker, Chem. Rev. 95 (1995) 849. 51 M. Kahlweit, R. Strey, D. Haase, and P. Firman, Langmuir 4 (1988) 785. 61 D.S. Ambwani. T. Fort, in Surface and Colloid Set Vol. II, R. J. Good and R. R. Stromberg, Eds.. Plenum (1979). D. Langevin, J. Meunier, in Photon Correlation Spectroscopy and Velocimetry, H.Z. Cummins. P. Pike. Eds.. Plenum (1977).
5.56
MICROEMULSIONS
case of a long cylindrical droplet, the interfacial tension yah depends only on the angular velocity, the density difference between the two phases and the radius of the droplet11. Although comparatively simple to use, some experimental details have to be considered to yield accurate results. A more detailed description of the measurement of low and ultra-low water/ oil interfacial tensions with the spinning drop technique are discussed elsewhere21. 5.4a Ultra-low interfacial tensions due to amphiphile adsorption The most evident way to introduce the general features of interfacial tensions in fluid mixtures containing a mieroemulsion phase is to consider a simple experiment. We take equal volumes of water (A) and oil (B), i.e. <j> = 0.5, and measure the interfacial tension yah between these two phases upon the addition of surfactant. To be consistent with sees. 5.2 and 5.3, a non-ionic was selected as the surfactant. The result is shown schematically in fig. 5.28 (solid line). The starting point is the pure water-oil system in the absence of surfactant (see fig. 5.28, left test tube). Here, the water/oil interfacial tension is on the order of 50 mNm"1. Adding small amounts of surfactant,
Figure 5.28. Schematic representation of the water/oil interfacial tension yab (full line) as function of the non-ionic surfactant mass fraction WQ at the mean temperature of the threephase body T m . Starting from equal volumes of pure water (A) and oil (B), I.e. 0 = 0.50 the interfacial tension yab decreases from 50 mNm" 1 to values of yab as low as 10~ 4 mNm~ 1 . Crossing the monomeric solubility of the surfactant in the water-rich and oil-rich phase, i.e. wc 0 , yab remains constant. The test tubes illustrate the situation without surfactant (left), with only partially screened water/oil contact (centre) and at WQ > WQ Q , where the mieroemulsion phase (c) is formed (right). 11 21
B. Vonnegut, Rev. Set Instr. 13 (1942) 6. T. Sottmann, R. Strey, J. Chem. Phys. 106 (1997) 8606.
MICROEMULSIONS
5.57
the molecules will either adsorb at the water/oil interface or be monomerically distributed between the water-rich (a) and oil-rich (b) phase. This situation is shown schematically in the middle test tube of fig. 5.28. For non-ionic surfactants, the monomer distribution is in favour of the oil-rich phase, i.e. wr _,„_ , « wr „„_ h (see sec. 5.2f). A further increase of wc leads to a complete saturation of the water-rich and oil-rich phase, as well as the water/oil interface with surfactant molecules at a mass fraction denoted as wco (see sec. 5.2c). Thus, the unfavourable water/oil contact is nearly perfectly screened, which lets the interfacial tension drop to values from a few mNm"1 to ultra-low values of 10~3 -10" 4 mNm" 1 . The latter are found at the mean temperature of the three-phase body Tm of efficient microemulsion systems". Above wc 0 , the excess surfactant molecules form aggregates in either the water, oil or a third phase (between the temperatures T{ and Tu ) as shown schematically in the right test tube of fig. 5.28. Thus, the monomeric concentration of the surfactant in the water-rich and oil-rich phases as well as the water/oil interfacial tension, stays practically constant. As in aqueous solutions of surfactants (see sec. III.4.6b), the distinct discontinuity of the slope of the yah{ log wc) -curve at it>co is indicative of the onset of aggregation. Below wc 0 , the interfacial concentration Fc of the surfactant is according to Gibbs21 given by
c
1 dyah ~~2.303RT 31ogu;c
T
For most surfactants the slope of the curve becomes practically constant already at concentrations well below wco (see dashed line in fig. 5.28), concomitantly yab continues to decrease rather steeply. In other words, the strong adsorption of the surfactants makes the water/oil interface become practically saturated well below LO C Q 3) . Having the interfacial concentration Fc in a saturated water/oil monolayer at hand, the area per molecule
can be determined; see sec. III.4.6c and refs.4 . As mentioned above, reliable values of a c in the water/oil interface could otherwise only be obtained experimentally by analyzing the more demanding SANS measurements61 (see sec. 5.3h). Comparing the a c values obtained by the two methods, good agreement is found. "w.H. Wade, J.C. Morgan, R.S. Schechter, J.K. Jacobson, and J.L. Salager, Soc. Petroleum Eng. 18(1978)242. 2) J.W. Gibbs, The Collected Works oJJ.W. Gibbs., Longmans Green, London (1928). 31 M. Kahlweit, G. Busse, and J. Jen, J. Phys. Chem. 95 (1991) 5580. 41 R. Aveyard, B.P. Binks, and P.D.I. Fletcher, in The Structure, Dynamics and Equilibrium Properties of Colloidal Systems, DM. Bloor, E. Wyn-Jones, Eds., Kluwer Academic Publishers (1990), pp. 557. 51 M.J. Rosen, D.S. Murphy, Langmuir 7 (1991) 2630. 61 T. Sottmann, R. Strey, and S.H. Chen, J. Chem. Phys. 106 (1997) 6483.
5.58
MICROEMULSIONS
5.4b Ultra-low interfacial tensions versus phase behaviour As we have seen before, a prerequisite of low water/oil interfacial tensions is the complete saturation of the water-rich and oil-rich phase, as well as the water/oil interface by surfactant molecules. Of course, this prerequisite is fulfilled if one of the phases considered is a microemulsion. Ultra-low interfacial tensions, however, may be obtained if a fluid mixture containing a microemulsion phase is driven through phase inversion11. In order to understand this behaviour, let us consider the temperature dependence of interfacial tensions in the system water-oil-non-ionic surfactant. As it turns out, the general form of the interfacial tension curves can be understood from the general principles of the phase behaviour. Figure 5.29 shows the phase prism (centre), which is obtained by stacking isothermal Gibbs triangles on top of each other (see sec. 5.2c), and its link to the low interfacial tensions (right), schematically. As discussed in sec. 5.2c, non-ionic surfactants mainly dissolve in the aqueous phase at low temperatures and form an oil-inwater (o/w) microemulsion (lower test tube). Increasing the temperature, one observes
Figure 5.29. Schematic representation of the phase behaviour (centre) and interfacial tensions (right) as the function of the temperature for the system water-oil-non-ionic amphiphile. The minimum of the water/oil interfacial tension fab at Tm is a consequence of the phase behaviour. Increasing the temperature, the aqueous phases separates into the phases (a) and (c) at the critical endpoints cepg, whereas the phases (b) and (c) merge into a single, oil-rich phase at cep a . Thus, the interfacial tensions yac and ybc show an opposite temperature dependence, becoming zero at 7} and Tu , respectively. Note that the interfacial tension is plotted on a log scale. The three test tubes on the left-hand side show the variation of the phase volumes.
11
J.C. Lang, B. Widom, Physica A81 (1975) 190.
MICROEMULSIONS
5.59
that this o/w microemulsion splits into two phases (a) and (c) at the temperature TJ of the lower critical endpoint cep^ . Subsequently, the composition of the lower water-rich phase (a) moves toward the water corner, while the composition of the surfactant-rich middle phase (c) moves toward the oil corner of the phase prism. At the temperature Tu of the upper critical endpoint cep a , a water-in-oil (w/o) microemulsion (upper test tube) is formed by merging the two phases (c) and (b). From the above, it is clear that three different interfacial tensions have to be considered. At temperatures below 7} and above Tu , only the water/oil interfacial tension yab occurs, which refers to the interface between an o/w-microemulsion and an oil-rich excess phase and that between a w/o-microemulsion and a water-rich excess phase, respectively. Between T; and Tu , i.e. within the three-phase body, the interfacial tensions between the water-rich and surfactant-rich middle phase yac and between the oil and surfactant-rich phase ybc exists, in addition. The water/oil interfacial tension y ab , however, can only be measured, if most of the surfactant-rich middle phase (c) is removed, with the rest floating as a lens at the water/oil interface (see sec. 5.4c). Although all three interfacial tension curves have been determined in literature1 2 3 , the qualitative shape of the curves can already be deduced from the phase behaviour. As can be seen in fig. 5.29, right, the interfacial tension yac starts at 7] from zero (the two phases (a) and (c) are identical) and increases monotonically with increasing temperature. On the other hand, the interfacial tension ybc decreases (monotonically) with increasing temperature and vanishes at Tu because the two phases (c) and (b) become identical there. This opposite temperature dependence of yac and ybc results in a minimum in the sum of the two, yac + yhc . In order to assure the stability of the water/oil interface, j/ab < yac + ybc
[5.4.3]
must hold41. Otherwise, a thin layer of the middle phase would penetrate between the water-rich and oil-rich excess phase, which is the case for well-structured microemulsions only near the critical endpoints 561 (see sec 5.4c). Consequently, yab also has to pass through a minimum. This means that if yab, 7} and Tu are known, the relative location of the individual yac and ybc curves can be established. Near the critical endpoint temperatures 7} and Tu , even a quantitative description of the interfacial tensions yac and ybc can be obtained applying the scaling laws
11
J.C. Lang, B. Widom, loc. cit. A.M. Cazabat, D. Langevin, J. Meunier, and A. Pouchelon, J. Phys. Lett. 43 (1982) L89. 3) K. Bonkhoff, A. Hirtz, and G.H. Findenegg, Physica A172 (1991) 174. 41 B. Widom, Langmulr 3 (1987) 12. 51 M. Kahlweit, R. Strey, and G. Busse, Phys. Rev. E47 (6) (1993) 4197. 61 K.-V. Schubert, R. Strey, S. Kline, and E.W. Kaler, J. Chem. Phys. 101 (1994) 5343. 21
5.60
MICROEMULSIONS
U
yac =
^ac 1 ~
IU
and
ybc = ybc
u~ I
[5 4 4]
where / / = 1.26 1 2 1 and y^c and XQC are the critical amplitudes. 5.4c Wetting behaviour So far we have seen that short- and long-chain non-ionic surfactants have, in principle, a similar behaviour in mixtures of water and oil. There are, however, some differences between the mixtures made of short-chain or long-chain surfactants. For example, the phase behaviour of the latter systems shows an extended lamellar phase (see sec. 5.2d) and leads to the formation of the so-called lobes (see sec. 5.2i). In particular, the difference in scattering behaviour near the critical endpoints (see sec. 5.3d) suggests considering these as weakly and strongly structured mixtures, respectively. Another major difference is the macroscopically observable difference in the wetting behaviour of the surfactant-rich middle phase (c) at the water/oil-interface. Consider a test tube containing a ternary microemulsion at Tm at a surfactant mass fraction of W C m /2 • Then the mixture separates into three liquid phases. If the middle phase (c) is removed until only a small drop is left, this drop either forms a lens between the water/oil-interface [i.e. the drop does not wet the interface, see photographs in fig. 5.30) or spreads (complete wetting). For example, weakly structured mixtures containing the surfactant C4Ej always display wetting middle phases, while rather strongly structured mixtures with C 6 E 3 and surfactants with longer chains have middle phases that do not wet. Transitions in this wetting behaviour have been observed in several systems as a function of the chain length of the oil 341 , the oil-to-water ratio51, the amphiphilicity of the surfactant 671 or temperature . Thermodynamically, the ujetting/non-wetting transition is strongly connected to the interfacial Gibbs energy, which has to be spent to build up the a/b-interface, on the one hand, and the two a/c and b/c-interfaces, on the other hand, see III, sec. 5.2. To make this clear, fig. 5.30 shows the variation of the three interfacial tensions near the lower critical endpoint cep b , schematically. Very close to the critical endpoint, the middle phase still wets the macroscopic a/b interface, so that for TJ < T < Tw , yab > yac + ybc . At the wetting transition temperature Tw , yah = yac + ybc , while for T > T^ ,
11
B. Widom,toe.cit. K. Bonkhoff, A. Hirtz. and G.H. Findenegg, Physica A172 (1991) 174. D. Langevin, in Structure and Dynamics of Strongly Interacting Colloids and Supramolecular aggregates in Solutions, S.H. Chen, J.S. Huang, and P. Tartaglia, Eds., Kluwer Academic (1992) pp. 325. 41 K.-V. Schubert, R. Strey, S. Kline, and E.W. Kaler, J. Chem. Phys. 101 (1994) 5343. 51 M. Gradzielski, D. Langevin, and T. Sottmann, J. Chem. Phys. 104 (1996) 3782. 61 M. Kahlweit, R. Strey, M. Aratono, G. Busse, J. Jen, and K.-V. Schubert. J. Chem. Phys. 95 (1991) 2842. 71 K.-V. Schubert, R. Strey, J. Chem. Phys. 95 (1991) 8532. 81 D.H. Smith, G.L. Covatch, J. Chem. Phys. 93 (1990) 6970. 21
MICROEMULSIONS
5.61
temperature Figure 5.30. Behaviour of the interfacial tensions in the vicinity of the lower critical endpoint temperature 7] . For T| < T < Tw , the middle phase (c) wets the macroscopic a/b interface so that yah > Yac + 7bc • At the wetting temperature T w , yab = yac + ybc, while for T > Tw j/ab < ^ac + ybc p o r j^e i a t t e r temperatures, the phase (c) floats as a lens at the a/b interface, as shown by the photographs. In addition, the appearance of the cylindrical oil-rich drop in the spinning drop tensiometer is shown schematically. yab < yac + ybc
ancj tne
interface between the water-rich and oil-rich excess phase is
not wet by the middle phase, as can be seen in the photograph. Experimentally, the gap between Tj and Tw Is very small, 0.1 K or less, for strong surfactants1 . The analogous sequence is found near the oil-rich critical endpoint at Tu . In fig. 5.30 we included as an experimentally observed manifestation of the wetting behaviour the appearance of the cylindrical drop in the spinning drop tensiometer. The presence of the small, non-wetting blisters of the surfactant-rich middle phase (c) on the cylindrical drop are essential for precise measurements of the water/oil interfacial tension in the three-phase region. They act as a pool of surfactant compensating the lack of molecules due to the adsorption on heterogeneous surfaces. 5.4d Interfacial tension curves Among the three Interfacial tensions yac , yhc , and yab observable In fluid mixtures containing a microemulsion phase, particularly the water/oil interfacial tension, yab 11
M. Gradzielski, et al., loc. cit.; K.-V. Schubert, et al., loc. cit.
5.62
MICROEMULSIONS
plays an important role in technical applications, as for example in cleaning processes and enhanced oil recovery. Thus, much work has been carried out to obtain the variation of yab as a function of the relevant tuning parameter. The parameters used are salinity 121 or alcohol to surfactant ratio31 for ionic surfactants and temperature for non-ionic surfactants 4561 . Because of the fundamental link between interfacial tension and phase behaviour discussed above, irrespective of the parameter used to drive the system through the phase inversion; the shape of the interfacial tension curves is always similar.
Figure 5.31. Water/oil interfacial tension ya]° in the water-n-octane-CjQE^ system as a function of temperature. Data obtained by us ' (circles), Fletcher et al. (triangles) and Kahlweit et al.81 (reversed triangles) with the spinning drop and by Langevin et al. ' with surface light scattering, are shown for comparison. The full line, drawn to guide the eye, is calculated from the interfacial tension model 9101 (see sec. 5.5). The test tubes illustrate to which interface the measurement refers. 11
R. Aveyard, B.P. Binks, S. Clark, and J. Mead, J. Chem. Soc. Faraday Trans. 82 (1986) 125. B.P. Binks, J. Meunier, O. Abillon, and D. Langevin, Langmulr 5 (1989) 415. D. Langevin, in Structure and Dynamics of Strongly Interacting Colloids and Supramolecular aggregates in Solutions, S.H. Chen, J.S. Huang, and P. Tartaglia, Eds., Kluwer Academic (1992) pp. 325. 41 R. Aveyard, T.A. Lawless, J. Chem. Soc. Faraday Trans. 82 (1986) 2951. 51 L.T. Lee, D. Langevin, J. Meunier, K. Wong, and B. Cabane, Progr. Colloid Polym. Sci. 81 (1990) 209. 61 T. Sottmann, R. Strey, J. Chem. Phys. 106 (1997) 8606. 71 P.D.I. Fletcher, D.I. Horsup, J. Chem. Soc. Faraday Trans. 88 (1992) 855. 81 M. Kahlweit, J. Jen, and G. Busse, J. Chem. Phys. 97 (1992) 6917. 91 R. Strey, Colloid Polym. Sci. 272 (1994) 1005. 101 H. Leitao, A.M. Somoza, M.M. Telo da Gama, T. Sottmann, and R. Strey, J. Chem. Phys. 105 (1996) 2875. 21
MICROEMULSIONS
5.63
As an example, the variation of the interfacial tension yah with the temperature is shown in fig. 5.31 for the system H2O-n-octane-C10E4. In order to assess the reliability of the different techniques mentioned above (for details see also chapter III.l), different sets of literature data are compared. Here, a log scale is appropriate for the interfacial tension because of the strong variation over several orders of magnitudes. As one can see, the data obtained by us11 (circles), Fletcher et al.2) (triangles) and Kahlweit et al.3) (Inverted triangles) with the spinning drop and by Langevin et al. with surface light scattering all support each other. Each data set exhibits the above-discussed minimum of the interfacial tension fab in the three-phase region around Tm = (Tu + Tp/2 . Here, the interfacial tension becomes ultra-low. Interestingly, the full curve, calculated from a theoretical description based on the analysis of the interfacial tension measurement in terms of bending energy561 (see sec 5.5, [5.5.11]) describes the data points within experimental error. At this stage, the full curves should be viewed as a guide to the eye, below we will analyze them quantitatively. Let us now consider, as for the phase behaviour (see fig. 5.7), length scales (see fig. 5.25) and curvatures (see fig. 5.26), the variation of the interfacial tension curve with the surfactant chain length. To this end, In fig. 5.32 the yah[T) curves for the four representative systems, H2O -n-octane- C 6 E 2 , C g E 3 , C 10 E 4 and C 12 E 5 , are shown. As can be seen, the interfacial tension curves shift to higher temperatures, as do the three-phase bodies (see sec. 5.2d). Even more striking is the strong decrease of the minimum of the interfacial tensions fab with increasing chain length of the surfactant shifting from system to system by one order of magnitude to lower values. But, although the curves sharpen as the surfactant chain becomes longer, the shape remains similar. 5.4e Scaling of interfacial tensions I The similarity in shape of the yah (T) -curves is more general than found for the four systems shown above. An extended set of 19 systems altogether , suggests the possibility of normalizing the measured quantities similarly as in the case of the trajectories of the middle phase (see fig. 5.11), the length scales £ and curvatures J (see fig. 5.27). In line with the scaling of the £(T) curves, the steepness of the interfacial tension curves around the mean temperature of the three-phase body Tm can be seen to correlate directly with the height of the three-phase body AT = (Tu - T{). Thus, after
11
T. Sottmann, R. Strey, loc. cit. P.D.I. Fletcher et al., loc. cit. 31 M. Kahlweit, et al. loc. cit. 41 D. Langevin, loc. cit. 51 R. Strey, loc.cit. 61 H. Leitao, et al. loc. cit. 71 T. Sottmann, R. Strey, J. Chem. Phys. loc.cit. 21
5.64
MICROEMULSIONS
Figure 5.32. Temperature dependence of the water/oil interfacial tensions y a b for some representative H 2 O-n-octane-C n E x systems, increasing the total surfactant chain length from CgE2 to C^Eg . Note that the minimum of the interfacial tension curves decreases with increasing surfactant chain length. The shift on the temperature scale stems from the shift of the phase behaviour (see sec. 5.2d). centering the yah[T) curves by subtracting T m , the temperature axis can be normalized in the same manner by AT / 2 . For the normalization of the interfacial tension scale, one may follow Volmer 1 ' 2 ' who argued that colloidal dispersion should become thermodynamically stable if the interfacial Gibbs energy times the area of the colloidal object is equal to kT, i.e. rab^»kT
[5.4.5]
Interestingly, many aspects of the stability of microemulsions can be explained by this simple argument 341 . Since, in particular, the maximum characteristic length scale g is inversely proportional to the interfacial volume fraction of surfactant
rah~{0cf 11
M. Volmer, Z. Phys. Chem. 125 (1927) 151. M. Volmer, Z. Phys. Chem. 206 (1957) 181. 31 M. Kahlweit, H. Reiss, Langmuir 7 (1991) 2928. 41 M. Kahlweit, J. Jen, and, G. Busse, J. Chem. Phys. 97 (1992) 6917. 21
[5 4 61
--
MICROEMULSIONS
5.65
Figure 5.33. Scaling of the temperature variation of the water/oil interfacial tension y ab for the systems H2O-n-octane-CgE2 (inverted triangles), CgEg (triangles), CJQE^ (squares) and C| 2 E 5 (spheres). Following the scaling description of the three-phase body, length scales and curvatures, the temperature axis is reduced by subtracting the mean temperature of the threephase body Tm and normalizing by AT/2. Dividing yab by the volume fraction of surfactant (*g, residing at the microscopic water/oil interface of the optimum microemulsion makes all fab(T) -curves collapse onto a single curve.
which was confirmed experimentally1 2). This finding suggests normalizing the interfacial tension by (0Q) 2 . Figure 5.33 shows yah{T) curves of the four systems considered reduced in this way. Indeed, all four interfacial tension curves collapse onto one single curve with only some rather small, but systematic deviations, in particular, in the three-phase-region remaining. As we will see below (see sec. 5.5), these deviations can be eliminated by an even more quantitative scaling obtainable by a more detailed analysis in terms of bending energy . Here, we like to emphasize the modelfree normalizations of various properties of microemulsions sufficient to obtain a corresponding state type representation of the experimental data (see figs. 5.11, 5.27 and 5.33). The identified relevant parameters are ^g , Tm and AT. 5.5 Theoretical description The theoretical description of microemulsions, or more general complex fluids, started 11
M. Kahlweit, R. Strey, D. Haase, and P. Firman, Langmuir 4 (1988) 785. M. Kahlweit, R. Strey, and G. Busse, J. Phys. Chem. 94 (1990) 3881. 31 W. Helfrich, Z. Naturforsch. 28c (1973) 693. 21
5.66
MICROEMULSIONS
in the 1970s. How the individual theories are constructed depends on the length scale on which phenomena are to be explained and predictions to be made11. Effectively, three different theories were developed: microscopic models, Landau theories and membrane models. The natural starting point for a theoretical description of complex fluids is a microscopic model, in which the positions of all the atoms of the different molecules and their microscopic interactions are taken Into account. Unfortunately, such a description is too complex to calculate the macroscopic properties of microemulsions. Thus, the models have to be simplified considerably. One way is with the use of lattice approximations. For instance, Widom and Wheeler231 mimicked water, oil and surfactant molecules by spins, which are arranged on the sites of an Ising lattice. Thereby, a considerable simplification is obtained by summarizing the local interactions of the different species In terms of a Hamiltonian. In the early 90s, lattice models were used by Gompper and Schick to describe some features of the phase behaviour41 and scattering properties51 of microemulsions. Matsen et al. found rather good agreement with experiments on non-ionic surfactant systems61. In chapter 4, a large number of illustrations of this approach for non-ionic micelles and other associates can be found. Ginzburg-Landau models describe a ternary microemulsion by the local concentration of the three species of molecules, water, oil and surfactant, on a somewhat larger length scale than the microscopic one. Here, the interaction between the molecules is described by a Helmholtz energy functional, which depends strongly on the range of the phenomena the model is expected to describe. If a ternary microemulsion is studied for which the oil and water concentrations do not differ much, a model with one single order parameter suffices71. As shown by Teubner and Strey, a one-component, scalar order parameter theory with a negative gradient term is sufficient to quantitatively describe the scattering of bicontinuous microemulsions (see sec. 5.3d, [5.3.9]). However, if the whole range of concentrations between the binary and the ternary system is to be predicted, a model with two or three order parameters is required8 91. On an even larger length scale, the structure of the microscopic interface between water and oil can no longer be resolved. The microscopic interface, respectively amphiphilic film formed by the surfactant molecules, must now be described by twodimensional surfaces (see sec. 5.3, chapter 6 and sec. 1.2.23). Over the past three decades, it turned out that most of the macroscopic and microscopic properties of 11
G. Gompper, M. Schick, SelJ-assembling Amphiphilic Systems, Academic Press (1994). J.C. Wheeler, B. Widom, J. Am. Chem. Soc. 90 (1968) 3064. 31 B. Widom, J. Chem. Phys. 84 (1986) 6943. 41 G. Gompper, M. Schick, Phys. Rev. B41 (1990) 9148. 51 G. Gompper, M. Schick, Phys. Rev. Lett. 62 (1989) 1647. 61 M.W. Matsen, M. Schick, and D.E. Sullivan, J. Chem. Phys. 98 (1993) 2341. 71 G. Gompper, M. Schick, Phys. Rev. Lett. 65 (1990) 1116. 81 G. Gompper, S. Klein, J. Phys. II France 2 (1992) 1725. 91 D. Roux, C. Coulon, and M. E. Cates, J. Phys. Chem. 96 (1992) 4174. 21
MICROEMULSIONS
5.67
microemulsions can be described or even predicted in terms of these so-called membrane theories121. Therefore, in the following the basic features of these theories will be described in more detail. 5.5a Membrane theories Basically, all membrane-bending theories foot on Helfrich's mechanical approach3 , which was originally intended for vesicles. The two-dimensional amphiphilic film is described as an ensemble of fluctuating surfaces whose shapes are determined by a macroscopic bending Helmholtz energy, which per unit area amounts to
(see [III. 1.15.1] and sec. III.4.7). Here, kY is the first, or mean bending modulus, (also called bending rigidity) and k2 is the saddle-splay or Gauss modulus. The quantities J and K have been defined in [5.3.2]. The first term on the r.h.s. of [5.5.1] represents the change in Helmholtz energy if the mean curvature of the membrane deviates from the spontaneous curvature J o , which is an intrinsic property of the given membrane in a particular environment. The second term represents the part of the Helmholtz energy, which depends on the Gauss curvature and, hence, on the topology of the membrane. We repeat the caveat from sec. III. 1.15 that in the literature various alternative symbol conventions can be found, see p. III. 1.79. In some cases, there are also different prefactors; for instance, when [5.5.1] is written in terms of H (= — (Cj + c2)) instead of J, the prefactor must be 2 instead of - . The first approach in describing the fascinating properties of microemulsions in terms of membrane-bending theories was proposed by de Gennes and Taupin41. They argued that a lamellar phase of a water-oil-surfactant mixture should melt into a bicontinuous structure when the spacing d between neighboring monolayers reaches approximately the persistence length of membranes
Here, S can be considered as the effective thickness of the amphiphilic film, which equals vQl aQ. A similar physical picture of bicontinuous microemulsions as random structures with length scales of the order £k was given by Safran et al. 56) , in which the bending energies of such structures were described by using a length scale-dependent renormalized bending energy. This model explained for the first time the existence of 11
S.A. Safran, T. Tlusry, Ber. Bunsenges. Phys. Chem. 100 (1996) 252. D.C. Morse, Curr. Opin. Coll. Interface Sci. 2 (1997) 365. 31 W. Helfrich, Z. Naturforsch. 28c (1973) 693. 41 P.G. de Gennes, C. Taupin, J. Phys. Chem. 86 (1982) 2294. 51 S.A. Safran, D. Roux, M.E. Cates, and D. Andelman, Phys. Rev. Lett. 57 (1986) 491. 61 M.E. Cates, D. Andelman, S.A. Safran, and D. Roux, Langmuir 4 (1988) 802. 21
5.68
MICROEMULSIONS
large regions of coexistence between a bicontinuous and excess water and excess oil phase. However, as the effect of the saddle-splay modules k2 was ignored and an essentially random surface, i.e. bicontinuous structure, was presupposed, discrimination between structures of different topologies was impossible. Subsequent observations and analyses by Porte et al. 1 ' and Anderson et al. 2 suggested a rather different picture of the lamellar to bicontinuous transition. They proposed a topology-driven transition in which the lamellar phase becomes unstable with respect to a bicontinuous structure of nearly minimal surfaces due to the change in the value of fc2 . Accordingly, the bicontinuous structure should be stable only over a narrow range of values of k2 ~ 0 , but for all values of the specific interface AIV , i.e. for any surfactant concentration. Although ignoring any Helmholtz energy contribution arising from the renormalization of k2, the model yields a qualitative description of the phase behaviour of pseudo-binary water-surfactant mixtures. It fails, however, in describing the lamellar to bicontinuous transition upon changing the surfactant concentration. Another promising approach was suggested independently by Golubovic 3 , Morse 4 and later on confirmed by Gompper and Kroll 5 ' via Monte Carlo simulations of dynamically triangulated surfaces of fluctuating topology. While dropping the assumption of random surfaces, they incorporated the effects of thermal fluctuations of the amphiphilic film in Helfrich's bending energy (see [5.5.11) by using instead of the rigidities fcj and k2 , the renormalized, length scale-dependent parameters SA"T k
^
itt) = Ko-^^-§
^kT
^
F
M£) = *2,o+-g7 1 ^
I5-5-3!
with £ being the length scale of the microstructure and Jcj0 and k2 0 , the so-called bare rigidities. Originally derived for amphiphilic films formed by bilayers, it holds apparently also for monolayer structures. Within this description, the swelling of the lamellar phase will generally lead to a fluctuation-induced melting transition into a bicontinuous or a vesicle phase, depending upon the value of the ratio k2/ k^. Thereby, the melting transition occurs at fc2(£) = 0, which defines a topological persistence length
^=* e x p i~~^n
[5541
--
Using the scale-dependent rigidities leads to the prediction of a bicontinuous phase that is only stable over a limited range of surfactant concentrations. While it coexists with the lamellar phase at high concentrations, a coexistence with an excess water 11
G. Porte, J. Appell, P. Bassereau, and J. Marignan, J. Phys. 50 (1989) 1335. D. Anderson, H. Wennerstrom, and U. Olsson, J. Phys. Chem. 93 (1989) 4243. 31 L. Golubovic, Phys. Rev. E50 (1994) 2419. 41 D.C. Morse, Phys. Rev. E50 (1994) 2423. 51 G. Gompper, D.M. Kroll, Phys. Rev. Lett. 81 (1998) 2284. 21
MICROEMULSIONS
5.69
phase (in the bilayer case) or excess water and oil phases (in the monolayer case) is predicted accurately at low concentration. Furthermore, Monte Carlo simulations show that the instability of the bicontinuous phase runs as a parallel line to the instability of the lamellar phase. A series of other membrane approaches were suggested, which will now be mentioned briefly. Wennerstrom and Olsson focused on the effect of anharmonic contributions to the bending energy, i.e. contributions that are higher than second order in the principal curvatures. Comparing the consequences of the two completely different contributions to the bending Helmholtz energy, one would expect that fluctuations dominate in very dilute phases, while anharmonic elastic effects dominate in surfactant-rich phases. Non-symmetric quadratic and non-quadratic energy densities were also investigated by Leitao et al.21 , as well as by Lade and Krawietz31. While the former model focuses mainly on the scaling of the water/oil interfacial tension (see below), the latter model studies the microstructures and the phase behaviour of microemulsion systems in the vicinity of the three-phase body. Still another approach by Berk41, Teubner51 and Pieruschka and Marcelia61 suggests to treat the bicontinuous phase and its amphiphilic film as level surfaces of Gauss-type random fields. This approach is useful and predictive when the Gauss model of random interfaces is related to statistical mechanics. In doing so, Pieruschka and Safran71 were able to relate the correlation length £TS and the periodicity d TS (see [5.3.10] and [5.3.11 ] of the bicontinuous structure to the mean bending modulus according to 10V3^ fc1(0_____-fcTT S
[5.5.5]
Most recently, the gradual formation of fluctuating, connected microemulsion networks from disconnected globules via cylinders was considered by Safran et al.89 10 ' " quantitatively treating the structural energy , as well as entropy, of endcaps and junctions. This model provides analytic scaling laws, which describe the experimentally observed scaling of the trajectory of the three-phase body, the closed lobes of two-phase coexistence below and above Tj and Tu , respectively (see sec. 5.2i), as well as the water/oil interfacial tension. II
H. Wennerstrom, U. Olsson, Langmuir 9 (1993) 365. H. Leitao, A.M. Somoza, MM. Telo da Gama, T. Sottmann, and R. Strey, J. Chem. Phys. 105 (1996)2875. 31 O. Lade, A. Krawietz, J. Chem. Phys. 115 (2001) 10986. 41 N.F. Berk, Phys. Rev. Lett. 58 (1987) 2718. 51 M. Teubner, Europhys. Lett. 14 (1991) 403. 61 P. Pieruschka, S. Marcelia, J. Phys. II France 2 (1992) 235. 71 P. Pieruschka, S.A. Safran, Europhys. Lett. 31 (1995) 207. 81 R. Menes, S.A. Safran, and R. Strey. Phys. Rev. Lett. 74, 3399 (1995). 91 T. Tlusty, S.A. Safran, R. Menes, and R. Strey, Phys. Rev. Lett. 78 (13) (1997) 2616. 101 T. Tlusty, S.A. Safran, and R. Strey, Phys. Rev. Lett. 84 (6) (2000) 1244. III A.G. Zilman, S.A. Safran, Phys. Rev. E66 (5) (2002) 051107. 21
5.70
MICROEMULSIONS
(i) Bending constants The bending Helmholtz energy per unit area F® depends for all membrane theories parametrically on the unknown coefficients in the expansion, namely the bending constants fcj and k2, as well as the spontaneous curvature Jo . Thus, the determination of these coefficients is of substantial importance. Let us first consider the spontaneous curvature. For spherical structures (K = J 2 /4), which exist over a wide range of spontaneous curvatures J below and above the three-phase region, J can be obtained by minimizing the bending Helmholtz energy (dFJ^/dJ = 0). If higher order entropic and fluctuation terms are neglected, the spontaneous curvature J is given by J
o=J(1 + ^ )
[5 5 61
- '
i.e. by the experimentally observable mean curvature (see sec. 5.3h) and the bending constants. As the latter are parameters, which characterize the bending properties of microscopic interfaces, they are in general difficult to determine. In the last two decades, several methods have been developed (see sec. III. 1.15). They can be extracted, for example, from the relaxation times of the deformation modes of microemulsion droplets and their polydispersity. These properties are measured by a combination of neutron spin-echo spectroscopy (NSES) and small angle neutron scattering (SANS)1 2) or dynamic light scattering (DLS) measurements31, respectively. Here, we will show that the bending constants can be determined directly by analyzing the experimentally easy-to-study phase behaviour, microstructure and water/oil interfacial tension in terms of the above-mentioned membrane theories. According to the model of thermally fluctuating amphiphilic films, the lamellar structure melts into a bicontinuous phase when the renormalized saddle splay modulus fc2(£) approaches zero. Since the instability of the bicontinuous phase runs parallel to the instability of the lamellar phase, the bare saddle splay modulus k20 can be calculated directly from the position of the (optimum) X-point and [5.5.4], By identifying the topological persistence length £fc with the maximum length scale | = 1.75<5/(Z>g (see [5.3.21] with the prefactor a O 2 =7.O), the bare saddle splay modulus k2 0 follows as ,20=_^lnI^ ZM
e,TT
[5 .5.7,
d>°
Furthermore, using the Gauss model of random interfaces (which are thermally fluctuating) and having determined the correlation length £TS and the periodicity d TS
11
J.S. Huang, S.T. Milner, B. Farago, and D. Richter, Phys. Rev. Lett. 59 (1987) 2600. B. Farago, D. Richter, J.S. Huang, S.A. Safran, and S.T. Milner, Phys. Rev. Lett. 65 (1990) 3348. 31 T. Hellweg, D. Langevin, Phys. Ren. E57 (1998) 6825. 21
MICROEMULSIONS
5.71 Figure 5.34. Bending moduli for the systems H2O -n-alkane- C n E x as a function of log £ . The bare saddlesplay modulus fc20 l s calculated from the interfacial volume fraction of surfactant
of the bicontinuous mlcrostructure, the renormalized bending rigidity fcj(£) follows from [5.5.5]. In a second step, the bare bending rigidity fc10 can be calculated by converting [5.5.3]. In doing so, we obtain the bending constants for the systems H2O n-alkane-CnEx as a function of the maximum length scale t, for each system (fig. 5.34). The value of the renormalized saddle splay modulus Jc2(£) (open squares) at the X-point is by definition zero, while the bare saddle splay modulus k20 (filled squares) becomes increasingly negative with increasing £ , i.e. with increasing efficiency of the system. The left-most data point belongs to the short chain C 6 E 2 -system and the right-most one to the long chain C 12 E 5 -system. Having chosen a logarithmic axis for the length scale, the data points fall on a straight line according to [5.5.3]. Looking at the variation of the renormalized mean bending modulus k^g) (open spheres), it can be seen that the value is of the order of 0.4 kT, increasing somewhat with increasing 4 • However, the bare rigidity /c10 (filled spheres) increases significantly as the microemulsion system displays higher efficiency. Also, kl0 varies, according to [5.5.3], linearly with £, . Calculating the ratio k2Q/klQ, one obtains ^ Q V ^ I O ~ ~0.65, which becomes increasingly negative with increasing efficiency, as predicted by Golubovic1' and Morse21. Alternatively, the bending constants can be determined by considering the measurement of the interfacial tension
on a microscopic level31, as has already been discussed to some extent in sec. III. 1.15. Increasing the angular velocity in a spinning drop experiment, in which the tem11
L. Golubovic, Phys. Rev. E5O (1994) 2419. D.C. Morse, Phys. Rev. E50 (1994) 2423. 31 R. Strey, Colloid Polym. Sci. 272 (1994) 1005. 21
5.72
MICROEMULSIONS
Figure 5.35. Schematic illustration of the interfacial tension experiment on an o/w-microemulsion (a) coexisting with an excess oil phase, (b) upon an increase of the bulk interfacial area A by AA via tilting the container, the area is filled in by the amphiphilic film supplied by the microemulsion droplets, which have to unbend. As a result, interfacial tension experiments, at least in part, measure "bending energy".
perature T, volume V and overall composition nt are kept constant, an elongation and, hence, an area increase of an injected oil-rich drop in the surrounding water-rich phase is observed. If either the water-rich or oil-rich phase is a microemulsion, the area is filled-in by surfactant molecules from the microscopic amphiphilic film of the microemulsion. Figure 5.35 shows such an experiment schematically. The initial state is pictured by the upright container on the left-hand side in which an oil-in-water droplet microemulsion (a) is in contact with an oil-excess phase (b). Tilting the container (constant volume, right-hand side) the macroscopic, surfactant-loaded area will increase by an amount AA. Since the system will tend to prevent direct water/oil contact and the monomeric solubility of the surfactant molecules remains unchanged (closed system at constant temperature), the microemulsion droplets have to provide surfactant molecules to shield the w/o-contact. At the same time, their oil content becomes part of the oil-excess phase. Accordingly, the amphiphilic film, which covers the interface, is flat on average, while it was bent before shaping the microemulsion droplet interface. Part of the Helmholtz energy price to be paid is, therefore, the difference in bending energy before and after the interfacial tension experiment. Using only Helfrich's bending Helmholtz energy per unit area, neglecting all other contribution mentioned above, the water/oil interfacial tension is given by yab = -^-k1(j2-2JJ0)-k2c1c2
[5.5.9]
where the energy of the flat macroscopic interface F® (°°) ( Cj = c 2 = J = 0 ) amounts to F a ° H = ^ f c i J § • Eliminating J Q between [5.5.6] and [5.5.9], which may be done in the
MICROEMULSIONS
5.73
regime of o/w- and w/o- droplet microemulsions, one obtains11 yah =-J2[kl+k2)-k2clc2
[5.5.10]
This simple parabolic dependence of the water/oil interfacial tension on J is in good agreement with experiments. However, despite the simplicity of this derivation and the good agreement with the experimental observations, [5.5.10] is inconsistent for bicontinuous structures since it requires a positive value of the saddle splay modulus k2, whereas Helfrich's expression on which it is based requires k2 to be negative. Leitao et al.2) overcame this inconsistency by using a quadratic but non-symmetric energy density. Assuming additionally that the variation of the mean and principal curvatures with the parameter applied (here the temperature) is almost linear (see [5.3.26 and 27]) they found f&=fab(_2kl+k2T2+1)
[5.5.11]
with 2(T-T ) T = -± — T u
_
T i
and
k 7 a b = -=§-
[5.5.12a,b]
^2
Recalling the scaling behaviour of the three-phase bodies, length scales and curvatures as well as water/oil interfacial tensions (figs. 5.11, 5.27 and 5.33), one realizes that T is equivalent to the reduced temperature. Furthermore, [5.5.12b] resembles Volmer's31 stability argument ([5.4.5]) and enables the straightforward calculation of the saddle splay modulus k2 if both the minimum water/oil interfacial tension fah and the maximum characteristic length scale E, are determined. Then the bending modulus kx is the sole free parameter in [5.5.11], which can be obtained by adjusting the opening angles of the V-shaped yab{T) curves. The perfect fit of the solid lines in figs. 5.31 and 5.32 supports this procedure. The values of the bending moduli /Cj and k2 obtained are plotted in fig. 5.36 versus the logarithms of the maximum length scale E, for the systems studied. As can be seen, kx increases with increasing £, i.e. as the microemulsion system displays a higher efficiency, from values of about 0.6 kT to 1.1 kT. According to [5.5.12b], the value of the saddle splay modulus k is negative and ranges between -0.3 kT to -0.7 kT. It stays almost constant for medium- and longchain surfactant systems, while it becomes slightly more negative for the inefficient, short-chain surfactant systems. Collecting the average value for all systems examined, one finds
11
R. Strey, loc. cit. H. Leitao, A.M. Somoza, M.M. Telo da Gama, T. Sottmann, and R. Strcy, J. Chem. Phys. 105 (1996) 2875. 31 M. Volmer, Z. Phys. Chem. 125 (1927) 151. 21
5.74
MICROEMULSIONS
Figure 5.36. Bending moduli fcj and fc2 f° r the systems H2O -nalkane- C n E x as a function of log E, . Alternatively to fig. 5.34, the moduli obtained by the detailed analysis of the interfacial tension experiment in terms of a quadratic, but non-symmetric bending energy, while thermal fluctuations are neglected. Note that with increasing maximum length scale £,, i.e. efficiency of each system, fcj increases slightly, while the absolute value of fc2 decreases slightly.
5.4e, [5.4.5]) does not only apply with respect to the order of magnitude, but also semiquantltlvely within a factor of two. Having determined the bending moduli by two different approaches, one might want to compare the values obtained. However, before comparing these quantitatively one has to recall that two totally different membrane models are used to analyze the experimental data. While the former (fig. 5.34) incorporated the effect of thermal fluctuations and, therefore, the renormalization of the bending elastic constants, the latter added a symmetry-breaking term to Helfrich's bending energy" (fig. 5.36). Here, the question arises what the nature is of the bending modulus deduced. Does one obtain the bare or renormalized values, i.e. length scale-dependent constants? In spite of the differences between the two models, the values of all bending constants obtained are in the order of kT. Furthermore, the values of the bending elasticity are, in general, positive, whereas that of the saddle splay modulus are negative. Comparing the average ratio between the two constants (k20/kl0) =-0.65 and (k2/kl) = -0.56 , a nearly quantitative agreement is found. Interestingly, both ratios are close to those for simple liquid interfaces, derived from the elementary relation kx = -2k2 2). All of this is an extension of sec. III. 1.15. (II) Scaling of Interfacial tensions II The quantitative description of the temperature variation of the water/oil interfacial tension yab(T) with temperature by [5.5.11] (which is based on a quadratic but nonsymmetric bending energy) suggests refining the scaling of the yab(T) curves shown in sec. 5.4e (fig. 5.33). As reduced parameters , we define
11 21
W. Helfrich, Z. Naturforsch. 28c (1973) 693. C. Varea, R. A, Physica A220 (1995) 33.
MICROEMULSIONS
5.75
Figure 5.37. Refined scaling of the yab{T) curves for 19 systems of the type H2O-n-alkaneC n E x . Using a quadratic, but non-symmetric, bending energy to describe the spinning drop experiment, a reduced water/oil interfacial tension y" and a reduced temperature r* can be defined according to [5.5.13a,b]. Note that within this y* versus z* representation, all data points fall on top of each other. The full line is calculated from [5.5.14]. yab J-* = 4-TT yab
and
I 2k, l + k0 T* = T - ^ y k2
[5.5.13a,b]
where t* only differs from the above-used, reduced temperature r (see fig. 5.33 and [5.5.12a]) by a factor determined by a combination of the bending constants. In fig. 5.37 the reduced interfacial tension y* is plotted versus the reduced temperature r* for all 19 H2O -n-alkane-CnEx systems studied. As can be seen, applying this reduction of the scales, all interfacial tensions collapse onto a single curve. The full line is calculated using [5.5.11 ] which, in terms of the reduced variables, reads y* = {r*f+\
[5.5.14]
Comparing the different types of scaling shown in figs. 5.33 and 37, one realizes that the scaling of the yab{T) curves can be improved by using a phenomenological theory for the bending energy of the amphiphilic film. In this section we have reviewed three different approaches to the theoretical description of microemulsions. Although microscopic models and Landau theories are only very briefly considered, the membrane models were discussed in more detail. We learned to appreciate the predictive power of the latter model as it allows one to under-
5.76
MICROEMULSIONS
stand many properties of microemulsions through the curvature and bending properties of the underlying amphiphilic film. The microscopic differences between different surfactants are mainly reflected in different spontaneous curvatures and bending constants of the membranes. Numerical values of the constants are obtained by considering different membrane theories. It is found that the bending rigidity, in general, is positive and increases with the surfactant chain length, while the saddle splay modulus comes out negative and, remarkably, the ratio k2 /fcj of about -0.5 resembles the value expected for simple liquid interfaces. 5.6 Applications As we have stated in the beginning of this chapter, the general features of microemulsions are best understood by studying simple ternary non-ionic mixtures of the H2O n-alkane-n-alkyl polyglycol ether (C n E x )-type. Now we are in a position to apply this knowledge to more complex systems. We found that the basic features of microemulsions correlate with the variation of the spontaneous curvature of the microscopic amphiphilic film. One observes that the spontaneous curvature changes monotonically with temperature from positive (o/w-structures, 2 (Winsor I)) to negative (w/o-structures, 2 (Winsor II)), passing through zero for bicontinuous microemulsions (3 (Winsor III)) where these contain equal amounts of water and oil. Concomitantly, the water/oil interfacial tension passes through a characteristic minimum, while at the same time the length scale reaches its maximum. The maximum, in turn, occurs because the spontaneous curvature passes through zero. That these features are, in fact, general expresses itself in the corresponding state type representation of phase diagrams, length scales and interfacial tension. The identified relevant parameters 0g, Tm and AT especially underline the crucial importance of the three-phase body and the X-point. Because of their special properties, such as ultra-low water/oil interfacial tension, excellent solubility for both hydrophilic and hydrophobic substances, and an extremely large specific internal interface, microemulsions gain an increasing interest in technical applications. Examples are tertiary oil recovery, cosmetics, pharmacy, as well as washing and extraction processes. Furthermore, microemulsions are used as reaction media for organic and enzymatic syntheses1 2) polymerizations341 and the production of nanoparticles5 6). With respect to its importance for technical applications one has to be aware that
1)
M.-J. Schwuger, K. Stickdorn, and R. Schomacker, Chem. Rev. 95 (1995) 849. C. Solans, H. Kunieda, Industrial applications of microemulsions. Marcel Dekker (1997). 31 F. Candau, in Polymerization in organized media, C. M. Paleos, Ed., Gordon & Breach, (1992) pp. 215. 41 M. Antonietti, R. Basten, and S. Lohmann, Macromol. Chem. Phys. 196 (1995) 441. 51 M.P. Pileni, J. Phys. Chem. 97 (1993) 6961. 61 J. Eastoe, B. Warne, Curr. Opin. Colloid Interface Sci. 1 (1996) 800.
21
MICROEMULSIONS
5.77
changing the curvature of the amphiphilic film is not only realized by temperature variations as demonstrated before. In fact the curvature can also be affected by adding a fourth and fifth component. The addition of, for example, salt, alcohol or another surfactant to ternary mixtures provides an additional degree of freedom to adjust the spontaneous curvature of the amphiphilic film. Especially in technical applications It Is often necessary to obtain a microemulsion system with specified properties of the amphiphilic film, i.e. mlcrostructure, conductivity or viscosity in a given range of compositions or temperature with a given oil or surfactant. In the following, we will discuss some mlcroemulsions, which are relevant in applications as systems containing technical-grade, non-ionic surfactants, n-alkylpolyglucosides, ionic surfactants, mixtures of non-Ionic and ionic surfactants and amphiphilic blockcopolymers. We will show that the Insights gained by studying the temperature dependence of ternary nonIonic microemulsions can easily be adapted to these more complex mixtures. 5.6a Technical-grade mixtures of non-ionic surfactants Applications of microemulsions generally involve technical-grade surfactants that are usually mixtures of homologues and/or isomers. The most widely used types of non-ionic surfactants in industrial applications are ethoxylated alcohols or ethoxylated alkyl phenols. In contrast to the monodisperse C n E x surfactants, discussed above, the technical-grade surfactants have a broad distribution of the degree of ethoxylation and contain a residual amount of non-reacted alcohol. The chain length of the alcohol, however, is rather narrowly distributed. Although those surfactants are produced in quantities of some million tons per year, only a few studies have been conducted on microemulsion formulations with technical-grade surfactants 123 ' 41 . In these studies it has been proven useful to treat the blend as a (pseudo) single component. Then the phase behaviour of such a system can be represented in an upright Gibbs phase prism (see fig. 5.3). Having studied the phase behaviour of simple ternary microemulsions in detail before (see sec. 5.2), we know that even systems with a complex phase behaviour can easily be characterized by isoplethal T{wc) sections through the phase prism (temperature versus surfactant mass fraction, see fig. 5.4). Here, the optimum microemulsion Is also given by the X-point of a T(wc) section at an oil-to-water-plusoil volume fraction (j> = 0.50 . In fig. 5.38 the T{wc) sections of the systems H2O-n-octane-C12E6 and the technical grade analogue DA-6 (dodecylalcohol 6) are shown for comparison at
A. Graciaa, J. Lachaise, J.G. Sayous, P. Grenier, S. Yiv, R.S. Schechter, and W.H. Wade, J. Colloid Interface Sci. 93 (1983) 474. 21 H. Kunieda, K. Shinoda, J. Colloid Interface Sci. 107 (1985) 107. 31 H. Kunieda, N. Ishikawa, J. Colloid Interface Sci. 1985 (1985) 122. 41 T. Sottmann, M. Lade, M. Stolz, and R. Schomacker, Tenside. Surf act., Deterg. 39 (2002) 20. 51 B. Jakobs, T. Sottmann, and R. Strey, Tenside Surf. Det. 37 (2000) 357.
5.78
MICROEMULSIONS
Figure 5.38. Isoplcthal T(wc) section through the phase prism of the systems H2O-n-octane-C^Eg and the technical grade analogue DA-6 at an oil-to-water plus oil volume fraction
a horizontal three-phase region, which touches the horizontal one-phase region at the X-point. On the other hand, the phase boundaries of the technical grade (DA-6) system are strongly distorted, especially at lower surfactant mass fractions. Despite this distortion, the two systems behave similarly. Both systems are rather efficient, whereby the X-point of the technical grade (DA-6) system is located at a slightly lower wc, i.e. higher efficiency and a somewhat higher temperature T . Furthermore, both systems show an extended lamellar phase within the one-phase region. The reason for the distortion of the phase boundaries in the DA-6 system can be rationalized considering the broad distribution of ethoxylation in the technical grade surfactant. In addition, the monomeric solubility of every specific homologue in water WQm o n a , and particularly in oil ix>cm o n b , is different. Taking into account only w r m™ u (because wn mnn „ « wr mnn . for non-ionic surfactants (see sec. 5.2f)), the less ethoxylated, more hydrophobic homologous of the surfactant DA-6 are more-likely expected to dissolve in the oil-excess or subphase. Thus, the remaining surfactant mixture in the amphiphilic film is effectively more hydrophilic than the base surfactant DA-6. By decreasing wc by adding water and oil, increasing amounts of the more hydrophobic fractions of the surfactant DA-6 are extracted from the internal interfacial film, which accordingly becomes increasingly hydrophilic. Thus, following the properties of ternary non-ionic microemulsions (see sec. 5.2e, fig. 5.8), the phase behaviour shifts to higher temperatures with decreasing wc , explaining the large distortion of the fish-type phase diagram, i.e. an increasing phase inversion temperature (p.i.t.) with decreasing ix>c . On the microscopic level, this distortion can be discussed considering the spontaneous curvature of the amphiphilic film composed of different surfactant homologous. Upon decreasing wc , the fraction of surfactant molecules with large head
MICROEMULSIONS
5.79
groups increases within the film. Thus, although the temperature is kept constant, the amphiphilic film will be increasingly curved around the oil. As for pure surfactants, this increase in curvature can be compensated by an increase in temperature, which leads to the weakening of the hydrogen bonds. Accordingly, within the technical grade surfactant systems the spontaneous curvature depends not only on the temperature, as found for the ternary model systems, but also on the composition of the mixed amphiphilic film. The effect of the partitioning of the surfactant homologous between the mixed amphiphilic film and both the excess phases and the microdomains of water and oil has been studied quantitatively for systems containing mixtures of two different monodisperse surfactants1-21, called surfactant and co-surfactant, respectively34 . There it turned out that the variation of the spontaneous curvature as a function of the composition of the mixed amphiphilic film is most suitably studied by keeping the temperature, i.e. the other tuning parameter, constant. In the following, we will present how a quarternary system can be driven through zero spontaneous curvature, i.e. through phase inversion, by varying the composition of the mixed amphiphilic film. As an example, we have chosen microemulsions with n-alkylpolyglucoside, which are known to exhibit a rather weak temperature sensitivity, so that tuning the spontaneous curvature by temperature is not an option. 5.6b Alkylpolyglucoside microemulsions Alkylpolyglucosides are ideally suited for formulating non-ionic, non-toxic, biodegradable microemulsions that are temperature-insensitive . These sugar surfactants (C m G x ) have m numbers of carbons within the hydrophobic chain and x numbers of glucose units in the hydrophilic head group. Due to the fact that one glucose group contains 6 hydroxyl groups, the CmGx surfactant is in general rather hydrophilic. Furthermore, the weak temperature dependence of the strong hydration of the hydroxyl groups causes the rather weak temperature sensitivity of CmGx-microemulsions. Thus, temperature is an inappropriate parameter to tune the spontaneous curvature of the amphiphilic film. Instead, the mixing of two surfactants of different hydrophilicity is the appropriate means (see sec. 5.6a). To mimic the effect of increasing temperature discussed at length above, obviously the rather hydrophilic alkylpolyglucosides have to be mixed with a hydrophobic amphiphile, like an alcohol7-8-9'10' or a C n E x -surfactant21 11
S. Yamaguchi, H. Kunieda, Langmuir 13 (1997) 6995. L.D. Ryan, K.-V. Schubert, and E.W. Kaler, Langmuir 13 (1997) 1510. 31 M. Kahlweit, R. Strey, and G. Busse, J. Phys. Chem. 95 (1991) 5344. 41 R. Strey, M. Jonstromer, J. Phys. Chem. 96 (1992) 4537. 51 M.H.G.M. Penders, R. Strey, J. Phys. Chem. 99 (1995) 10313. 61 C. Stubenrauch, Curr. Opin. Colloid Interface Sci. 6 (2001) 160. 71 K. Fukuda, O. Soderman, B. Lindman, and K. Shinoda, Langmuir 9 (1993) 2921. 81 M. Kahlweit, G. Busse, and B. Faulhaber, Langmuir 11 (1995) 3382. 91 C. Stubenrauch, B. Paeplow, and G. H. Findenegg, Langmuir 13 (1997) 3652. 101 T. Sottmann, K. Kluge, R. Strey, J. Reimer, and O. Soderman, Langmuir 18 (2002) 3058. 21
5.80
MICROEMULSIONS
Figure 5.39. (a): Schematic phase tetrahedron of a quaternary water-oil-surfactant-cosurfactant system. The section at constant oil-to-water plus oil volume fraction tp — 0.50 shows phase boundaries resembling the shape of a fish. The dashed curve, ranging from the critical endpoint on the water-side to that of the oil-side of the tetrahedron, displays the trajectory of the middle-phase microemulsion. (b): corresponding section through the phase tetrahedron for the quaternary system D2O -n-octane-n-octyl-fl-D-glucopyranoside ( CgGj )-l-octanol ( C g E 0 ) at a constant temperature of T = 25°C. Note, that the system is driven through phase inversion by the addition of the hydrophobic amphiphile C 8 E 0 . to drive the system through zero spontaneous curvature, i.e. the phase inversion. The phase behaviour of quaternary systems at constant temperature, in general, has to be represented within a phase tetrahedron (see fig. 5.39a). As for the ternary temperature-sensitive microemulsions (see fig. 5.2), insight into the phase behaviour of a quaternary system can be facilitated by considering the phase diagrams of the systems representing the faces of the phase tetrahedron. Accordingly, here one has to consider the diagrams of four ternary systems. Studying as an example the quaternary D2O -n-octane-n-octyl-P-D-glucopyranoside (CgGj )-l-octanol (C g E 0 ) system, it turned out that all ternary systems show extensive miscibility gaps at T= 25°C1). Here, the phase behaviour of the two-side systems D2O -n-octane-CgGj and D2O-rt-octaneC g E 0 are of major interest. Within the former system, the C8Gj molecules prefer the water phase, i.e. a 2 miscibility gap is formed, whereas the latter system shows 2 behaviour, i.e. the C 8 E 0 molecules reside mainly in the oil phase. Furthermore, since there is the demixlng tendency of the third ternary side-system D2O - C g E 0 - CgGj , the formation of a three-phase region is induced inside the phase tetrahedron. Figure 5.39a depicts this situation schematically by means of a section through the phase tetrahedron at a constant oil-to-water -plus-oil volume fraction 0 = 0.5. As can be seen, a typical fish-type phase diagram is found as the ratio of cosurfactant (D) to surfactant (C)-plus-cosurfactant (D)
11
T. Sottmann et al., loc. dt.
MICROEMULSIONS e
Jv =
5.81
—
[5.6.1]
is increased. Here, the overall volume fraction surfactant, i.e. surfactant plus cosurfactant is given by
°
[5.6.2]
To experimentally determine the sections through the tetrahedron, a sample containing the desired amounts of water, oil and surfactant have to be titrated with the cosurfactant. In fig. 5.39b the result of such experiments is shown for the D2O -noctane-n-octyl-p-D-glucopyranoside (CgGj)-1-octanol (C g E 0 ) system at 0 = 0.50 and T = 25°C11. As expected, at low mass fractions of CgGj the phase sequence 2-3-2 is found with increasing 1-octanol content. At higher mass fractions of CgGj , the 2-1-2 sequence is observed. For still higher mass fractions, a lamellar phase appears. Comparing the constant
Va B
[5.6.3]
v°+v°
which dictates in particular the spontaneous curvature and the bending constants, i.e. the efficiency of the system. V^and Vg correspond respectively to the volume of surfactant and cosurfactant molecules residing in the mixed amphiphilic film. To determine Sy in the system under consideration, the monomeric solubilities (see sec. 5.2f) of 1-octanol in n-octane and CgGj in D2O have to be known, while the solubilities of 1-octanol in D2O and CgGj in n-octane can be neglected. The former can be determined individually from the phase behaviour applying the analysis of Kunieda and coworkers2 3). Considering the variation of the composition of the amphiphilic film 8° , one can quantitatively understand the phase behaviour of the system. Starting from the ternary system D2O-n-octane-CgGj at 0 = 0.50 and as an example at wc =0.10 an oil-inwater (o/w) - microemulsion forms that coexists with an excess oil phase. As one adds the 1-octanol it partitions between the bulk oil phase and the amphiphilic film. Thus, on the one hand, the alcohol acts as a cosolvent making the oil more hydrophilic. On
11 21 3)
T. Sottmann et al., loc. cit. H. Kunieda, M. Yamagata, Langmuir 9 (1993) 3345. S. Yamaguchi, H. Kunieda, Langmuir 13 (1997) 6995.
5.82
MICROEMULSIONS
Figure 5.40. Representations of the normalized trajectories of the middle phase for the temperature-dependent H2O-n-octane-CnEx and the quaternary H2O-n-octane-CgGj-CgGg system, (a): volume fraction of the surfactant or surfactant-cosurfactant mixture in the interfacial film, normalized with respect to the maximum of each trajectory, (b): Reduced temperature T and composition of the interfacial film <5y(red) versus 0 . Note, that r and <5^(red) are equivalent parameters so that all trajectories collapse onto a single curve.
the other hand, the alcohol mixes into the amphiphilic film making it increasingly hydrophobic. Although the spontaneous curvature of the amphiphilic film is lowered by both effects, the latter is predominant, since the OH-group of the alcohol is small compared with the large head groups of the sugar surfactant. Further increasing the concentration of 1-octanol, the film is enriched in 1-octanol and decreases its curvature until it inverts to form a water-in-oil (w/o) microemulsion. Accordingly, the tuning parameter S% in quaternary temperature-insensitive n-alkylglycoside systems plays the same role as the temperature in the ternary water-oil- C n E x systems. That this is indeed the case can be shown by scaling the trajectory of the middle phase (X-points) of the quaternary system onto the trajectories of the temperaturesensitive ternary non-ionic microemulsions11. In fig. 5.40a, the reduced trajectory
11
T. Sottmann, K. Kluge, R. Strey, J. Reimer, and O. Soderman, Langmuir 18 (2002) 3058.
MICROEMULSIONS
5.83
l
°V,u °V,eJ
can be defined where the essential parameters are again the lower limit Sv e , the upper limit Sv u and the mean Sv m (Sv m = (<5^ j + Sy u ) / 2 ) of the three-phase body. Plotting the reduced z<0) trajectory of the ternary C n E x systems (see fig. 5.11b), together with the Sy(red)(
K. Kluge, C. Stubenrauch, T. Sottmann, and R. Strey, Tenside Surf. Det. 38 (2001) 30. J.H. Schulman, T.P. Hoar, Nature 152 (1943) 102. 31 J.H. Schulman, W. Stoeckenius, and L.M. Prince, J. Phys. Chem. 63 (1959) 1677. 41 P.A. Winsor, Solvent Properties of Amphiphilic Compounds, Butterworth & Co., London, (1954). 51 A.M. Bellocq, J. Biais, B. Clin, A. Gelot, P. Lalanne, and B. Lemanceau, J. Colloid Interface Sci. 74(1980)311. 61 D. Langevin, Mol. Cryst. Liq. Cryst. 138 (1986) 259. 71 J.F. Billman, E.W. Kaler, Langmuir 6 (1990) 611. 81 H.N.W. Lekkerkerker, W.K. Kegel, and J.Th.G. Overbeek, Ber. Bunsenges. Phys. Chem. 100 (1996)206. 91 H.F. Eicke, in Microemulsions, I. D. Robb, Ed., Plenum Press (1982) p. 17. 101 M. Kahlweit, R. Strey, R. Schomacker, and D. Haase, Langmuir 5 (1989) 305. III S.J. Chen, D.F. Evans, and B.W. Ninham, J. Phys. Chem. 88 (1984) 1631. 21
5.84
MICROEMULSIONS
structure they are nearly balanced amphiphiles, so that no cosurfactant is necessary to reach a zero spontaneous curvature. In the following we will consider the quaternary (pseudo-ternary) system D2O /NaCl-n-decane-AOT to show the main features of ionic microemulsion systems. Subsequently, we will apply the knowledge gained for quarternary non-ionic microemulsions to also understand the complex five-component ionic mixtures. (i) Quaternary AOT microemulsions Due to the fact that only traces of an inert electrolyte have to be added to shift the phase inversion to ambient temperatures, AOT has become the most widely studied amphiphile for the purpose of formulating ionic microemulsions. In general, an ionic amphiphile, and in particular AOT, can be made to change from hydrophobic to hydrophilic by increasing the temperature. This is because upon increasing temperature the effective degree of dissociation of the counterions will increase. On the other hand, addition of a salt has exactly the opposite effect. It will make AOT less hydrophilic because the salt ions will compete with the counterions and screen the ionic head groups. This combination of adding salt and increasing the temperature can be used to tune the spontaneous curvature of the amphiphilic film in a quaternary AOT microemulsion system with precision11. That this is really the case can be shown considering the phase behaviour of the system D2O /NaCl-n-decane-AOT as an example. Using the mixture of D2O and NaCl (often referred to as brine) as a pseudo-component by keeping the ratio between the two constant at e = 0.006 (see [5.3.3]), the phase behaviour of the system can, as a first approximation, be represented in an upright Gibbs phase prism. As for the systems discussed above, we consider the corresponding binary systems first to understand the phase behaviour of the pseudo-ternary ionic microemulsion. Brine ( D2O/NaCl) (A) and n-decane (B) are of course practically immiscible over the experimentally accessible temperature range. The binary system n-decane (B)-AOT (D) shows a lower miscibility gap, which lies below the melting point of n-decane. Hence, complete miscibility of n-decane and AOT exists between the melting point and boiling point of n-decane. However, the situation changes upon adding traces of water. Then an upper miscibility gap appears21. The decisive role in the phase behaviour of the AOT-microemulsion again plays the pseudo-binary system (D2O/NaCl) (A)-AOT (D). Considering first the binary system D2O -AOT, the lower miscibility gap with an upper critical point is covered by a lamellar phase, which extends to low amphiphile concentrations. Switching to the pseudo-binary system by adding NaCl, the situation changes dramatically. The lamellar phase is pushed back while the miscibility gap is simultaneously enlarged towards higher temperatures. As for ternary non-ionic microemulsions, the phase behaviour of the pseudo-ternary 11 21
S. H. Chen, S.-L. Chang, and R. Strey, J. Chem. Phys. 93 (1990) 1907. M. Kahlweit, R. Strey, R. Schomacker, and D. Haase, Langmuir 5 (1989) 305.
MICROEMULSIONS
5.85
ionic mixture can be inferred from considering the interplay of the three binary systems. At low temperatures, the ionic surfactant is preferentially soluble in oil, at high temperatures in brine. Thus, having added the adequate amount of salt, an increase in temperature turns the ionic surfactant from hydrophobic at low into hydrophilic at high temperatures. This temperature dependence is, importantly, the reverse of that for non-ionic microemulsions (see sec. 5.2a and c). Alternatively, the spontaneous curvature of the amphiphilic film can also be considered to understand the behaviour of ionic microemulsions. Let us start with a film, which is curved around the water. Obviously, this situation can be obtained if additional ions (salt) are added to screen the repulsive interaction between the ionic head groups. Increasing the temperature, the degree of dissociation of the counterions increases, i.e. an increasingly repulsive interaction results between the ionic head groups on the water side of the amphiphilic film. Accordingly, the curvature of the film inverts from curved around the water ( J o < 0 ) to curved around the oil ( J o > 0 ) as the temperature is increased. In fig. 5.41 the phase behaviour of the system D O/NaCl-n-decane-AOT at e = 0.006 is shown by means of a T(ix>c) section at an oil-to-water-plus-oil volume fraction tp = 0.60 . As one can see, a three-phase region is obtained at ambient temperatures. At lower temperatures a water-in-oil (w/o) microemulsion coexists with a water excess phase (2 ), whereas at higher temperatures an o/w microemulsion coexists with an oil excess phase (2). Thus, a phase sequence of 2-3-2 is observed in the ionic system, Figure 5.41. T[wc) section of the ionic system D2O /NaCl-n-decanesodium-bis-ethylhexylsulfosuccinate (AOT) at fixed constant
M. Kahlweit, et al., loc. cit.
5.86
MICROEMULSIONS
which is just the inverse of the 2-3-2 sequence found in non-Ionic microemulsions. Apart from this inverse temperature dependence, the overall behaviour of the two types of systems is rather similar. The phase boundaries of the ionic system also resemble the shape of a fish. The three-phase region touches the one-phase region (the X -point for the ionic system) at f = 38.1°C and G5C = 0.059 , implying a rather good efficiency of the double chain ionic surfactant AOT in solubllizlng brine and rt-decane. The analogy between the properties of pseudo-ternary Ionic and ternary microemulsions Is most suitably demonstrated considering the trajectory of the composition of the middle phase {X -points). Figure 5.42 shows the trajectory of the X -points projected onto the
Figure 5.42. Trajectories of the surfactant-rich middle-phase, i.e. the X -point for the systems D2O/NaCl-n-decane-AOT at s = 0.00611. (a): projection onto the 0 C (0 plane of the phase prism. Note that the volume fraction of surfactant tpc passes through a maximum at <j> = 0.5 . (b): projection onto the T(tp) plane of the phase prism. In this projection the trajectory exhibits a sigmoidal shape. Importantly, opposite to the non-ionic systems (see fig. 5.10) it extends from the lower critical endpoint (cep a ) at Tj on the oil-rich side to the upper one ( cepp ) at Tu on the water-rich side. 11
M. Kahlweit, R. Strey, R. Schomacker, and D. Haase, loc.clt.
MICROEMULSIONS
5.87
(ii) Quinary NaDS microemulsions We now turn to microemulsion systems containing ionic surfactants with only one alkyl chain. In general, these surfactants are extremely hydrophilic so that strongly curved, and thus almost solubilisate-less, micelles are formed in ternary water-oil-ionic surfactant mixtures. Also, the addition of an electrolyte does not suffice to curve the amphiphilic film around the water. Thus, a rather hydrophobic cosurfactant has to be added to invert the structure from o/w to w/o1 2). In order to study these complex quinary mixtures of water/electrolyte (brine)-oil-ionic surfactant-non-ionic cosurfactant, we use, as above, brine as a pseudo-component. Then, equivalently to the quaternary sugar surfactant microemulsions, (see fig. 5.39), the phase behaviour of the pseudoquarternary ionic system can, as a first approximation, be represented in a phase tetrahedron if one keeps the temperature constant. In the following, we will concentrate, as an example, on the system H2O/NaCl-ndecane-sodiumdodecylsulphate (NaDS)-l-butanol (C4OH) at the rather large mass fraction of s= 0.10 of NaCl in the mixture of water and NaCl and T= 20°C. As for all systems considered before, the phase behaviour of the side systems have to be considered first. Of major interest are the two side systems H2O/NaCl-n-decane-NaDS and H2O/NaCl-n-decane-C4OH . Both systems show miscibility gaps. While the hydrophilic NaDS molecules prefer the water phase, the C4OH molecules reside mainly in the oil phase. Thus, together with the demixing tendency of the third ternary side system H2O/NaCl-C4OH-NaDS, a three-phase region is formed inside the phase tetrahedron. In fig. 5.43 a section through the tetrahedron of the pseudo-quaternary system H2O/NaCl-n-decane-NaDS-C4OH at
A.M. Bellocq, J. Biais, B. Clin, A. Gelot, P. Lalanne, and B. Lemanceau, J. Colloid Interface Scl. 74 (1980) 311. 21 H.N.W. Lekkerkerker, W.K. Kegel, and J.Th.G. Overbeek, Ber. Bunsenges. Phys. Chem. 100 (1996) 206.
5.88
MICROEMULSIONS
Figure 5.43. Section through the phase tetrahedron for the pseudo-quaternary system H2O / NaCl-n-decane-sodium dodecyl sulphate (NaDS)-l-butanol (C4OH) at a 0 = 0.58, £ = 0.10 and a constant temperature of T = 20°C. Note that as shown for the quaternary sugar surfactant systems (see fig. 5.39), the pseudo-quaternary ionic system can be driven through phase inversion by the addition of the rather hydrophobic amphiphile C4OH
further enriching the film with 1-butanol, the spontanous curvature inverts, which manifests itself in phase inversion. Interestingly, having tuned the phase behaviour from 2 to 2 using the tuning parameter Sy, the system can be tuned back to 2 by keeping S% constant and increasing the temperature. 5.6d Microemulsions with non-ionic and ionic surfactants In the above we used the most frequently applied combination of a short-chain alcohol as non-ionic cosurfactant with a single-tailed ionic surfactant to tune the quinary ionic microemulsion system through the phase inversion. Let us now consider the change in phase behaviour if, instead of the short-chain alcohol, an ordinary long-chain non-ionic surfactant is used. From the discussion of the phase behaviour of ionic microemulsions (see sec. 5.6c) we know that the temperature dependence of their phase behaviour is in almost every detail the inverse of that of non-ionic microemulsions. This evidently raises the question of how the phase behaviour changes from the non-ionic to the ionic behaviour if one varies the ratio S (corresponding to the mass fraction of Sv, see [5.6.1]) between ionic and non-ionic surfactant. One can expect that at a certain value of S the inverse temperature trend tends to compensate, so that a temperature-insensitive microemulsion forms. Needless to emphasize that this property is extremely relevant in technical applications where mixtures of nonIonic and ionic surfactants are often used. To locate the composition where most of the properties of complex quinary (pseudoquaternary) mixtures are expected to be temperature-insensitive, different, time-con-
MICROEMULSIONS
5.89
Figure 5.44. Isothermal trajectories of X points within a £() representation for the quinary system H2O/NaCl-ndecane-C12E4-AOT at a constant 0= 0.6011. Note that one observes some kind of isosbestic point, a point where all isotherms intersect. There, the phase behaviour is temperature-insensitive. All mixtures having a composition, which lies to the left side of the point of intersection, show the phase behaviour of non-ionic microemulsions, whereas the others behave as ionic ones.
suming approaches are possible. One way is to establish sections through the tetrahedron of the pseudo-quaternary system at a constant 0, but at different e and temperature T. Each section yields the ionic to non-ionic plus ionic surfactant mass fraction ~S at the X-point. Plotting the determined X -points within a e{5) representation, isothermal trajectories are obtained. These isothermal trajectories are shown in fig. 5.44 for the system H2O/NaCl-n-decane-C12E4 -AOT and a constant <j>= 0.60. As one can see, all isotherms intersect at a particular composition. This means that at this so-called isosbestic point the phase behaviour is temperature-insensitive. Thereby, all compositions that lie to the left side of the isosbestic point show the phase behaviour of non-ionic microemulsions, whereas the others behave as ionic ones. Figure 5.45 shows the temperature insensitivity of the phase behaviour in terms of a T(wc) section of the quinary system H2O/NaCl-n-decane-C12E4 -AOT. Here, the composition of the system is chosen near the isosbestic point at ift = 0.50, S= 0.60 and a salt mass fraction in water of e = 0.006. Thereby, only the phase boundaries of the one phase region are determined experimentally, whereas the three-phase region is shown schematically. As one can see, the phase boundaries around the X-point are particularly steep, which indicates temperature-insensitivity. Thus, preparing a mixture of H2O/NaCl-n-decane-C12E4-AOT at <j> = 0.50, 5= 0.60, e= 0.006 and an overall surfactant mass fraction of wc = 0.08, a one-phase microemulsion is obtained between 0 and 75°C. This property is extremely relevant in technical formulations.
11
M. Kahlweit, R. Strey, J. Phys. Chem. 92 (1988) 1557.
5.90
MICROEMULSIONS
Figure 5.45. T(wc) section of the quinary H2O/NaCl-ndecanc-Cj2E4 -AOT microemulsion at a constant
5.6e Amphiphilic block copolymers as efficiency booster One of the main factors hindering an even more general use of microemulsions in technical application is the relatively high amount of surfactant needed. The disadvantages are clear; apart from the high costs of those formulations, problems such as higher environmental risks together with more difficult disposal arise. In cosmetic or pharmaceutical products, but also in household cleansers, washing powders etc., the risk of skin irritation or other health problems increases with larger surfactant contents. One way to decrease the amount of surfactant needed to form a one-phase microemulsion is first to increase the hydrophobic chain length of the surfactant used. However, as we have shown already in sec. 5.2d (see figs. 5.6 and 5.7), this increase in efficiency is accompanied by a strong stabilization of the lamellar phase, which in applications could cause problems due to high viscosity. Thus, if the hydrophobic chain length of the surfactant molecule exceeds 12 to 14 carbon atoms, extended regions of the phase diagram around the phase inversion temperature are occupied by the lamellar phase. Furthermore, below the p.i.t. other even more hindering mesophases of hexagonal and cubic symmetry form. Since these phases are often highly viscous and tend to complicate the handling of water-oil-surfactant systems, it is clear that in many applications it is important to try to avoid the formation of mesophases. Recently we have discovered how to increase efficiency while suppressing
MICROEMULSIONS
5.91
Figure 5.46. Samples of H2O-n-decane-C10E4 systems at 0= 0.50 and wc = 0.03 in Hellma flat quartz cells immersed in a thermostatted water bath at T = 30.3 o c". From left to right, the content of the amphiphilic block copolymer PEP5-PEO5 increases as indicated. Note the increasing phase volume of the middle phase due to polymer addition. The increasing darkness of the middle phase is a consequence of the increasing length scale and the resulting stronger light scattering.
the formation of mesophases1 2). In the following we will illustrate this phenomenon. Starting from the well-known phase behaviour of the ternary microemulsion system H2O -n-decane-C10E4 , we observed an enormous efficiency increase by adding an amphiphilic block copolymer. The molecular structure of these (ethylenepropylene)-copoly(ethyleneoxide) copolymers is similar to the C n E x surfactants, differing from these by the branched nature of the hydrophobic block and, of course, by the larger overall molar mass. In the literature, these polymers are abbreviated as PEPx-PEOy, where x and y are the approximate molar masses of the blocks in kg/mol. While mixtures of two surfactants of comparable chain length show small synergistic effects in microemulsions, adding an amphiphilic block copolymer to a conventional microemulsion system leads to a large efficiency increase by traces of polymer. To see this, consider fig. 5.46. A sample containing equal volumes of water and n-decane (
11 B. Jakobs, T. Sottmann, R. Strey, J. Allgaier, L. Willner, and D. Richter, Langmuir 15 (1999) 6707. 21 T. Sottmann, Curr. Opln. Colloid Interface Sci. 7 (2002) 57.
5.92
MICROEMULSIONS
polymer mass fractions. As can be seen by the rightmost cell, even for the highest polymer content the polymer mass fraction amounts only to wcS = 0.004. From this simple experiment a number of immediate observations can be made. The oil and water excess phases visible in the left test tubes are progressively swallowed by the surfactant-rich middle phase, which thereby increases in volume. The accompanying effect of increasing opalescence leads to the increasingly darker appearance of the middle phases in transmitted light. This effect is obviously connected to the increasing length scale in these systems, which leads to stronger scattering. The experiment can be, and actually was, performed at constant temperature so that the spontaneous curvature is apparently not affected. This large increase in efficiency can be discussed more quantitatively considering fig. 5.47. Here, the variation of the phase behaviour as a function of S is shown for the system H2O -n-decane-C10E4 -PEP5-PEO5 in the form of T(wc) sections at > = 0.5 (i.e. the mixture of C 10 E 4 and PEP5-PEO5 is considered as a (one) pseudo component). For comparison and to elucidate the relative magnitude of this effect, the T(wc) sections of the systems H2O-n-decane-C10E5 and H2O-n-decane-C12E4 are also shown. Consider first the influence of the molecular parameters of the surfactants. As we know already from the above, the X-point shifts to higher temperature and to larger surfactant mass fraction if the head group size is increased by one oxyethylene unit (C 10 E 4 —> C 10 E 5 ). On the other hand, keeping the head group size constant and increasing the chain length of the hydrophobic alkyl tail from C 10 E 4 to C 12 E 4 , one observes a shift of the X-point to lower temperatures and a significantly lower wc , implying a large increase in efficiency. One can also see that in the "fish-tail" a lamellar phase appears (for comparison, see also figs. 5.7 and 5.8). The striking phenomenon of adding the polymer PEP5-PEO5 is demonstrated by the filled circles in fig. 5.47. As can be seen, replacing only a mass fraction of 0.015 of the surfactant by the amphiphilic block copolymer shifts the X-point to a significantly
Figure 5.47. T(wc) sections through the phase prism of the system H2O n-decane- C 10 E 4 -PEP5-PEO5 at equal volumes of water and n-decane (0= 0.5). The well-known fish-type phase boundaries are shown for the system H2O-n-decane-Cj0E4 as open circles. The effect of increasing surfactant head group size (C 1Q E 5 ) and tail size (C 1 2 E 4 ) is demonstrated. Note the associated temperature shifts. Adding traces of polymer PEP5-PEO5, i.e. increasing S, leads to an enormous efficiency increase (filled circles) at constant temperature.
MICROEMULSIONS
5.93
lower overall surfactant-plus polymer-mass fraction 03 c + D . Proceeding over 8 = 0.05 to 8 = 0.119, the minimum amount of surfactant plus polymer to form a one-phase microemulsion drops to wc+D = 0.035 . Importantly, the phase inversion temperature is not affected by the addition of the amphiphilic block copolymer. This aspect might be an interesting aspect for applications. Furthermore and on the contrary to the similarly efficient C 12 E 4 system, the addition of PEP5-PEO5 does not lead to the formation of a lamellar phase in the one-phase regions presented in fig. 5.47. In order to understand the mechanism behind the boosting of the efficiency of amphiphilic block copolymers, one has to keep in mind that the behaviour of microemulsions is controlled by the properties of the amphiphilic film. Thus, the influence of the polymers with respect to the microstructure1 21, on the one hand, and concerning the bending elastic moduli31 as well as the spontaneous curvature41, on the other hand, has to be considered. Summarizing the results from the phase behaviour and neutron scattering experiments it is found that the adsorption of the amphiphilic block copolymer into the surfactant membrane influences the membrane properties significantly. Theoretically, the effect of polymers anchored to membranes has been calculated recently for Gauss (ideal) chains51. In the mushroom regime, the effective bending moduli are found to be
and h,ef{=k2-—cr{^+^)
[5.6.6]
where kj and fc2a r e the bending and saddle-splay moduli, respectively, of the pure surfactant membrane, a is the number density of the block copolymer within the membrane and ra and rb are the end-to-end distances of the hydrophilic and hydrophobic block, respectively. Furthermore, the prediction of the dependence of the surfactant volume fraction 0g at the X-point on the saddle-splay modulus k 2 0 ([5.5.7]) and that for the dependence of the saddle-splay modulus k2 eff o n the polymer number density were combined to predict the dependence of the X-point on a and the polymer size61 according to
11
H. Endo, J. Allgaier, G. Gompper, B. Jakobs, M. Monkcnbusch, D. Richter, T. Sottmann, and R. Strey, Phys. Rev. Lett. 85 (2000) 102. 21 H. Endo, M. Mihailescu, M. Monkenbusch, J. Allgaier, G. Gompper, D. Richter, B. Jakobs, T. Sottmann, R. Strey, and I. Grillo, J. Chem. Phys. 115 (2001) 580. 31 G. Gompper, H. Endo, M. Mihailescu, J. Allgaier, M. Monkenbusch, D. Richter, B. Jakobs, T. Sottmann, and R. Strey, Europhys. Lett. 56 (2001) 683. 41 G. Gompper, D. Richter, and R. Strey, J. Phys. Condens. Matter 33 (2001) 9055. 51 C. Hiergeist, R. Lipowsky, J. Phys. II France 6 (1996) 1465. 61 H. Endo, J. Allgaier, G. Gompper, B. Jakobs, M. Monkenbusch, D. Richter, T. Sottmann, and R. Strey, Phys. Rev. Lett. 85 (2000) 102.
MICROEMULSIONS
5.94
Figure 5.48. Logarithm of the volume fraction of the surfactant CJQE 4 in the am-
phiphilic film f#g at the X m -point plotted versus the polymer area
l n ^ + D = In^g = l n ^ ' ° -5Vr(ra2 +r2)
[5.6.7]
Here, 0Q'° is the surfactant volume fraction of the fish-tail point in the system without polymer and £'=-0.628. To prove the predicted dependence, in fig. 5.48 the surfactant volume fraction at the X-point (note, that large portions of the polymer molecules extend into the water and oil sub-domains: thus the fraction of the molecules which resides directly in the amphiphilic film is negligible!) for a series of different systems is plotted as a function of ofr.2 + r 2 ). It was found that all data collapse into one single line, which nicely confirms [5.6.7]. However, the observed slope of E = -1.51 is twice as large as the predicted one. The difference was discussed in terms of the pronounced difference between self-avoiding and ideal chains. Finally, the influence of the amphiphilic block copolymer on the "effective" spontaneous curvature was predicted1 2 to be
W T » = c om-^Jf^Mr a -r b ) *
VD
K
15.6.8]
\ ,eff
Obviously, for a symmetric block copolymer with ra = r b , the spontaneous curvature contribution of the polymer vanishes as observed in fig. 5.47. Furthermore, the contribution of the polymer to c 0eff has to become negligible for strongly curved structures, such as oil-swollen micelles.
11 21
G. Gompper, D. Richter, and R. Strey, J. Phys. Condens. Matter 33 (2001) 9055. C. Hiergeist, R. Lipowsky, J. Phys. II France 6 (1996) 1465.
MICROEMULSIONS
5.95
5.7 General references Structure and Dynamics of Strongly Interacting Colloids and Supramolecular Aggregates in Solutions, S.-H. Chen, J.S. Huang, and P. Tartaglia, Eds., Kluwer (1992). (This proceeding of a NATO school (Italy, 1991) present results on the microstructure of soft condensed matter obtained via different scattering techniques.) H. Endo, M. Mihailescu, M. Monkenbusch, J. Allgaier, G. Gompper, and D. Richter; B. Jakobs, T. Sottmann R. Strey and I. Grillo, Effect of Amphiphilic Block Copolymers on the Structure and Phase Behavior of Oil-Water-Surfactant Mixtures. J. Chem. Phys. 115 (2001) 580. (This article reviews the efficiency boosting effect of amphiphilic block copolymers demonstrating that polymers which are anchored to the interfacial film increase the membrane curvature elasticity.) G. Gompper, M. Schick, Self-assembling Amphiphilic Systems, Vol. 16 of Phase Transitions and Critical Phenomena, Academic Press (1994). (This book focuses on the three principal approaches to describe self-assembling amphiphilic systems theoretically: through microscopic-, Ginzburg-Landau- and membrane models.) M. Kahlweit, R. Strey, Phase Behavior of Ternary Systems of the Type H2O -OilNon-ionic Amphiphile (Microemulsions). Angew. Chem. Int. Ed. 24 (1985) 654. (This review shows that simple ternary mixtures of the type water-n-alkane-non-ionic amphiphile mimics the properties of complex multicomponent microemulsions.) M. Kahlweit, R. Strey, D. Haase, H. Kunieda, T. Schmeling, B. Faulhaber, M. Borkovec, H.-F. Eicke, G. Busse, F. Eggers, Th. Funck, H. Richmann, L. Magid, O. Soderman, P. Stilbs, J. Winkler, A. Dittrich, and W. Jahn, How to study Microemulsions. J. Colloid Interface Sc(. 118 (1987) 436. (Paper with a review character which summarizes the various experimental techniques for studying the properties of microemulsions.) Microemulsions, M. Kahlweit, R. Lipowski, Eds. Ber. Bunsenges. Phys. Chem. 100 (1996) 181-393. (This issue presents the state of the art of microemulsions in the year 1995 with respect to experimental facts, theoretical modelling and experimental techniques.) Handbook of Microemulsion: Science and Technology, P. Kumar, K.L. Mittal, Eds., Marcel Dekker (1999). (This book link the various scientific aspect of microemulsions with their applications as reaction media, Pharmaceuticals, cosmetics, foods, etc.)
5.96
MICROEMULSIONS
Microemulsions, I.D. Robb, Ed., Plenum Press (1982). (Proceedings of a conference on the physical chemistry of microemulsions, motivated by the tertiary oil recovery. First results on phase behaviour, interfacial tension and microstructures of microemulsions systems are presented.) Microemulsion Systems, H.L. Rosano, M. Clausse, Eds. Vol. 24 (1987). Surfactant Science Series, Marcel Dekker. (Proceedings of the 59th Colloid and Surface Science Symposium in 1985; it includes some first studies of the microstructure of microemulsions via SANS and NMR-self-diffusion measurements.) S.A. Safran, Statistical Thermodynamics of Surfaces, Interfaces, and Membranes, Vol. 90, Frontiers in Physics, Westview Press (1994). (Presented as a set of lecture notes, the book presents the statistical mechanics that underlie the macroscopic, thermodynamic properties such as interfacial tension, wetting, membrane elasticity of soft matter.) R. Strey, Microemulsion Microstructure and Interfacial Curvature, Colloid and Polymer Set 272 (1994) 1005. (The author shows the intimate connection between phase behaviour, microstructure and interfacial tensions and their relation to curvature energy for ternary microemulsions.)
6
THIN LiguiD FILMS
Dimo Platikanov and Dotchi Exerowa 6.1 6.2
Introduction
6.1
Experimental methods
6.4
6.2a
The 'film thickness' issue
6.4
6.2b
Macroscopic flat foam films
6.2c
Microscopic foam or emulsion films
6.10
6.2d
Determination of the disjoining pressure
6.11
6.2e
Determination of the contact angle and the film tension
6.13
6.2f
Measurement of foam film elasticity
6.16
6.2g
Measurement of the lateral electrical conductivity of foam films
6.2h
6.5
6.6
6.21
6.3a
Description of the simplified film model
6.22
6.3b
Mechanical equilibrium of a thin liquid film
6.23
6.3c
Fundamental thermodynamic equations of a thin liquid film
6.26
6.3d
Thermodynamic approach using a film model with two Gibbs dividing planes
6.27
Contact between thin liquid film and the adjacent meniscus
6.31
Non-equilibrium properties of thin liquid films
6.36
6.4a
Kinetics of thinning of plane-parallel liquid films
6.36
6.4b
Deviations from the plane-parallel shape during film thinning
6.38
6.4c
Kinetics of rupture of thin liquid films
6.42
6.4d
Jump-like formation of a black film in a thin liquid film
6.45
6.4e
Film thickness upon extraction
6.49
6.5a
DLVO forces
6.50
6.5b
Potential of the diffuse double layer at the aqueous solution-air interface
6.55
6.5c
Steric surface forces
6.57
6.5d
Oscillatory disjoining pressure
6.60
Black foam films and emulsion films
6.62
6.6a
Disjoining pressure in black films
6.64
6.6b
Main properties of the two types of black films and the CBF-NBF transition
6.8
6.48
Surface forces in symmetric thin liquid films
6.6c 6.7
6.19
Thermodynamics of thin liquid films
6.3e 6.4
6.17
Measurement of the coefficient of gas permeability through foam films
6.3
6.7
6.69
Stability and rupture of bilayer black films
6.73
Diffusion processes in symmetric thin liquid films
6.78
6.7a
Gas permeability of foam films including CBFs
6.78
6.7b
Gas permeability of Newton black films
6.82
6.7c
Lateral diffusion in black phospholipid films
6.84
Thin liquid films: a biomedical application
6.85
6.8a 6.8b 6.9
Model study of the lung surfactant system through black foam films
6.85
Black foam film clinical test
6.87
General references
6.90
6 THIN LIQUID FILMS DIMO PLATIKANOV AND DOTCHI EXEROWA
6.1 Introduction
Thin liquid films form automatically in all colloidal systems in liquid dispersion media, where they arise when two particles of the disperse phase (solid particles, liquid drops, or gas bubbles) come close to each other. The liquid film is symmetrical when two particles are identical, i.e., in the case of homo-interaction. Samples of such thin liquid films are foam films, emulsion films, and films between identical particles in a sol. In more complicated cases, when two particles with substances of different composition approach each other, an asymmetric liquid film is formed. These are all cases of hetero-interaction. The most important asymmetric films are the wetting films: thin liquid films separating a solid from a gas phase. Foam films are symmetric liquid films between two gas phases, i.e., GLG films; emulsion films are L'LL' films, L1 being a liquid immiscible with the liquid L. One can distinguish two types of emulsion films: OWO films and WOW films, which correspond to the two emulsion types. Foam and emulsion films are both thin liquid films with fluid interfaces. Only these types of films are examined in the present chapter. Symmetric SLS films have already been considered in chapters 1.4 and IV.3, and asymmetric SLG films (wetting films) in sec. III.5.3. In this chapter we shall use the terms foamjilm and emulsion film. Obviously, the properties and the behaviour of thin liquid films determine the stability or instability of the corresponding disperse system. That is why thin liquid films are among the most important objects of colloid and interface science. Furthermore, liquid films play a central role in many technological processes, such as the flotation of minerals, coatings, oil recovery, detergency and washing. This has contributed to the vast expansion of fundamental and applied research in the domain of thin liquid films, especially in the second half of the 20th century. The first observations of the properties and behaviour of thin liquid films date back to a few centuries ago. Attention was drawn to foam films because their colours can easily be observed by the unaided eye. Mark Twain wrote in 1880, 'A soap bubble is the most beautiful thing and the most exquisite, in nature I wonder how much it would take to buy a soap bubble if there was only one in the world? One would buy a Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
6.2
THIN LIQUID FILMS
hat full of Koh-i-noors with the same money, no doubt.' However, as we shall show, from a scientific point of view, films that are so thin that they exhibit no interference colours are at least as challenging. The prominent scientists Hooke and Newton in the 17th century described the process of thinning of foam films and the formation of black spots and black films in a thicker film. This is why the current IUPAC nomenclature refers to the thinnest foam or emulsion films as Newton black films. In the middle of the 19th century, Plateau, a Belgian, performed extensive investigations on the properties of foam films. He established the important fact that thin liquid films are always connected either to a bulk liquid phase or to a solid wall through a liquid body with concave surfaces (see fig. 1.11. in Vol. I). We now call this liquid body a Plateau border. Marangoni studied the role of the expansion or compression of surfactant adsorption layers for the stability of the films, known now as the Marangoni effect; see sec. III.6e. The great Gibbs contributed significantly to the understanding of liquid film stability, introducing a phenomenon which is now called, Gibbs elasticity. The remarkable book, Soap bubbles by Boys summarized the knowledge of foam films, towards the end of the 19th century, and illustrated not only thin liquid film properties, but also capillary phenomena in general. The reader can find illustrations in figs. 1.1.410. During the 20th century studies of thin liquid films expanded at accelerated rates. In the first decades, Perrin studied extensively the different types of black foam films as well as the interesting phenomenon of stratification (multilayer formation in films prepared from concentrated solutions). In 1936 Deryagin introduced the notion of disjoining pressure, the most important thermodynamic parameter which characterizes a thin liquid film (see sec. 1.4.2). The great success of the DLVO theory for the stability of lyophobic colloids underlined the decisive role of thin liquid films for understanding the stability of all types of disperse systems in liquid media. Significant contributions to the quantitative description of thin liquid films have been made by the scientific schools of Deryagin (Russia), Overbeek (Netherlands), Mysels (U.S.A.), Sheludko (Bulgaria), and many others1 . What exactly should we call a thin liquid film? Let us consider a symmetric liquid film with parallel plane interfaces (fig. 6.1). The Cartesian coordinate system is oriented so that the x- and y- axes lie in the film plane, while the z-axis is normal to the film interfaces. Obviously the film's dimension along z is much smaller than the dimensions along x and y which determine the film area. However, the question arises: what do we mean by, 'much smaller'? The liquid phase from which the film is formed is denoted in fig. 6.1 by a, and the adjacent phases (gas, or another liquid or solid) by p. As already shown in sees. 1.2.5 and 1.2.22.a the a/p interface is not a mathematical plane but an interfacial layer of 11 Note that the Slavic names Deryagin and Sheludko are often transcribed as Derjaguin and Scheludko, respectively.
THIN LIQUID FILMS
6.3
Figure 6.1. Scheme of, (a), a thick- and, (b), a thin liquid film formed from liquid phase, a , between two equal phases, p . finite thickness. Most of the physical quantities inside this interfacial layer vary along the z-axls, owing to the mutual influence of the phases a and p. When the distance between the two interfaces Is rather large (fig. 6.1a), the two interfacial layers are far away from each other. There is liquid between them with the same properties as in the bulk liquid phase, a. Such a GLG system may be considered as a system of three bulk phases, separated by two liquid-fluid interfaces. When the thickness Is a few micrometers or more, the film can be treated by bulk thermodynamics. However, In the case when the film thickness is so small that both interfacial layers overlap (fig. 6.1b), Inside the film there is no liquid with bulk properties. Just as in a single interfacial layer, most of the physical quantities vary along the z -axis over the entire liquid film, from the bulk of the one phase, P, up to the other phase p. As we shall see later, this is the physical ground for the rise of the disjoining pressure in the film, as well as for the change of the surface tension of the film as a function of thickness. The intensive properties of the film become dependent on its thickness. Such a film Is referred to as a thin liquid film. It should be noted that this is a thermodynamlc definition of a thin liquid film and it reflects the peculiar properties of the film and their dependence on thickness. However, when kinetic properties (e.g., film drainage, rheology, dynamic elasticity) are considered, a liquid film Is often called, 'thin' simply because of its very small thickness, irrespective of the question whether it is 'thin' in the thermodynamlc sense. For example, Glbbs has considered foam film elasticity by assuming the surface tension of the film surfaces to be the same as that of the bulk liquid. The thermodynamic definition of thin liquid films shows that the most important factor, which determines its properties, is the interaction between the two film surfaces, i.e., the pair interactions between the two adjacent phases across the liquid film
6.4
THIN LIQUID FILMS
(see chapters 1.4 and IV.3). Hence, single liquid films of all types are very useful models for the investigation of pair interactions (disjoining pressure, interaction energy, stability or instability, etc.). Thus, thin liquid films have attained their own significance because they are powerful tools (both experimental and theoretical) in colloid and interface science. Moreover, as has already been noted, the properties of thin liquid films are decisive for the properties of the corresponding disperse systems of which the films are a part. Certainly, a disperse system is not simply the sum of many liquid films: it is much more complicated. Nevertheless, knowledge of film properties contributes to the elucidation of problems of real disperse systems. Thus, all the information about foam and emulsion films which is considered in the present chapter will be used in chapters 7 and 8 when examining foams and emulsions, respectively. 6.2 Experimental methods This section is devoted to experimental methods involved in thin liquid film research. Most specific properties of thin liquid films cannot be studied using the well known physicochemical methods for investigating the properties of bulk phases and single interfaces. Original and unique methods had to be developed for liquid film research, and our knowledge about thin liquid films is based, to a great extent, on these effective experimental techniques. As already noted in the introduction, this chapter focuses on symmetric thin liquid films with fluid interfaces, i.e., foam films and emulsions films. Accordingly, experimental methods for such films (mainly foam films), will be emphasized. 6.2a The 'Jiltn thickness' issue The film thickness h is the most important quantity to be determined experimentally. It not only characterizes the film, but most film properties also depend on it. Basically, h is defined as the distance between the two film interfaces. This quantity would be defined exactly had the film interfaces been mathematical planes. However, in reality this is not the case, not even for SGS films, because solid surfaces usually have significant roughness. Only in the special case of freshly-cleaved mica sheets could one consider the surface as being approximately, 'ideally' flat. This is why mica is often used in the surface-force apparatus as the solid in SGS films. See sec. IV.3.12b. Defining h for symmetric films with fluid interfaces is much more difficult. The transition from the liquid film to the adjacent gas (or liquid) phase is not abrupt. There is an interfacial layer with finite thickness, even for one-component liquid films, as in fig. III.2.1. However, foam and emulsion films contain more components, at least there is a surfactant which forms stabilizing adsorption layers at the film surfaces. How can one define h for such complicated structures? One possibility is the mechanical definition: the two film boundaries are considered as two plane-parallel mathematical
THIN LIQUID FILMS
6.5
faces, each of which is subject to a uniform, isotropic tension. These faces are the surfaces of tension (see sees. 1.2.23 and V.6.3). The mechanical film thickness can then be defined as the distance h between the two surfaces of tension1 . The thermodynamic description of an interface is based on the concept of the Gibbs dividing plane (sec. 1.2.5). Thin liquid film thermodynamics involves two Gibbs dividing planes, see sec. 6.3. The positions of these planes are defined with respect to the zero surface excesses of a major component, which we can choose (see sec. 1.2.22a). Most appropriate for thin liquid films is to accept a zero interfacial excess of the solvent in the bulk phase from which the film is formed. By this convention, the positions of both Gibbs dividing planes of the film are defined, and the thermodynamic film thickness is the distance between these two planes2'3'. This thermodynamic film thickness should be used when data for the quantities characterizing the film are treated thermodynamically. The main problem is that neither the mechanical, nor the thermodynamic, film thickness coincides exactly with the thickness which is determined experimentally using different physical methods. Most of these are of an optical nature and the measurement and the interpretation thereof are based on the assumed abrupt change of such properties as refractive index, electron density, etc., at the film boundary. With electric methods for film thickness measurement an abrupt change at a given film boundary of quantities, such as the specific conductivity or dielectric permittivity, is required. The question now is whether, by optical or electrical methods, the same, or a similar, h is measured as is defined thermodynamically. Let us consider the most widely used optical method, viz., interferometry, which is based on the measurement of the intensity of visible light reflected from the film. The simplest case of a foam film will be considered first. Classical optics provides relationships that link the thickness of the film to its optical characteristics, assuming that the optical properties vary discretely with z. If foam films are observed in reflected white light, it can be seen that their colour changes periodically during thinning. Initially, when the films are fresh and thick, drainage runs rather rapidly, but later it gradually slows down. The interference can be registered as a photocurrent/time curve in which the extrema correspond to interference maxima and minima, i.e., when film thicknesses are multiples of XI An (where X is the wavelength and n is the refractive index). Assuming the film to consist of pure water, and knowing the order of interference, k, it is easy to determine the thickness at these points. Film thicknesses between a maximum and a minimum are calculated from the ratio between the intensities measured of the reflected monochromatic light I, corresponding to a certain thickness, and Imax corresponding to the interference maximum, according to the
11
J.C. Eriksson, B.V. Toshev, Colloids Surf. 5 (1982) 241. A. Sheludko, Annualre Univ. Sofia, Fac. Chimie 62 (1970) 47. 31 B.V. Toshev, I.B. Ivanov, Colloid Polymer Set 253 (1975) 558, 593. 21
THIN LIQUID FILMS
6.6
formula1 -2)
h = — k/r±arcsin 2nn
l
_H™^ ^l + [(^-l)/ 2 n] 2 (l-//I max )J
[6.2.1]
Here, h w is called the equivalent water thickness, I.e., the thickness of a foam film with the uniform refractive index of the bulk solution, n. The accuracy of thickness measurements with this micro-interferometric technique is ± 0.2 nm.
Figure 6.2. A three layer model of a very thin foam or emulsion film: 1, hydrocarbon layer; 2, aqueous layer; 3, bulk or gas or oil phase. Schematic.
It is clear that hw is not equal to the real physical film's thickness and it does not necessarily coincide with the thermodynamic thickness h. The difference between hw and h can be neglected for thicknesses larger than 30 nm. However, for thinner foam films it is necessary also to account for the film structure. A three layer film model with an aqueous core of thickness h2 and refractive index n 2 , and two homogeneous layers of hydrocarbon chains of the adsorbed surfactant of thickness hj each, and refractive index rij is often used (fig. 6.2). The thickness of the aqueous core, h , is most often determined according to the Duyvis3 formula nf-1 h
2=hw-2hi75-7
|6 2 21
- '
Since, however, each model involves some assumptions, the calculation of h is always somewhat arbitrary. The most important problem in the three layer model concerns the positioning of the plane separating the hydrophobic and hydrophilic parts of the adsorbed surfactant molecule. In some cases it seems reasonable to have this plane passing through the centres of the hydrophilic heads of the molecules, in others the head group is not assumed to enter into the aqueous core. A more detailed model 11
A. Scheludko, D. Platikanov, Kolloid-Z. 175 (1961) 150. A. Scheludko, Proc. Koninkl. Nederl. Akad. Wetenschap B65 (1962) 76. 3I E.M. Duyvis, Ph.D. Thesis, Univ. of Utrecht (1962); see also C.J. Vasicek, Optics of Thin Films, North Holland Publishing Company (1960) 185. 21
THIN LIQUID FILMS
6.7
is the five layer model in which the aqueous layer 2 (fig. 6.2) is divided into three layers: the inner core containing aqueous solution only, and two layers, containing the hydrophilic head groups of the adsorbed molecules; each of these layers is considered to be a separate entity optically, with its own refractive index. The calculations of the film thickness based on different film models do not solve the general thickness problem. However, they provide possibilities for reasonable interpretation of the experimental results. 6.2b Macroscopic flat foam, films The oldest and most widely used method of formation of macroscopic flat foam films involves the drawing out of a film (with an area of the order of square centimeters) in a frame, from a surfactant solution (fig. 6.3)1'. A second frame made of a very thin wire (for example, 12.5 (im diameter) is attached to the glass frame, so that during their withdrawal a vertical foam film is formed. Owing to the small thickness of the wire the film obtained has a thin Plateau border which makes the film thinning better controlled. The film can be drawn out at a known rate (usually very slowly) by a mechanical device. Figure 6.3b presents the profile of a vertical foam film. The films' thicknesses are usually measured in the section which is at a distance H from the solution surface where the pressure differs from the atmospheric pressure by Ap = -pgH . In its lower part, the film contacts the solution by its Plateau border, and at heights greater than K the film is flat ( K is the capillary length, see [III. 1.3.3]). Various physical methods are employed in the study of macroscopic foam films obtained with a frame. Film thicknesses can be determined from the intensity of the
Figure 6.3. Frame for drawing out a foam film: (a) profile of a vertical film and its contact with the solution: (b) 1, glass frame; 2, thin wire frame. 2 is much thinner than 1.
11
J. Lyklema, P.C. Scholten and K.J. Mysels, J. Phys. Chem. 69 (1965) 116.
6.8
THIN LIQUID FILMS
reflected light (sec. 6.2a). This determination can be combined with the measurement of Brewster's angle, 8B , i.e., the angle at which the reflected light is completely planepolarized (fig. 6.4) see sec. 1.7.10a. The change in #B provides additional information about the refractive index of the thin films that have three- or two-layer structures.
Figure 6.4. Scheme of the apparatus for simultaneous measurement of Brewster's angle and the intensity of the light reflected from the film: 1, Nicol prism; 2, frame with the film; 3, diaphragm; 4, mirror; 5, photo-multiplier; 6, cell; A, beam; B, reflected beam. (Redrawn from Corkill etal. 1 '.
In order to determine the concentration of the substances in the foam film, and their change with time, the surfactant or the added electrolyte can be radioactively labeled; the film obtained is scanned with a Geiger counter through diaphragms with narrow gaps2'. Figure 6.5 shows the experimental arrangement for simultaneous ellipsometry and reflectometry3'. A foam film is formed on a vertical frame, in a closed cell C. There is one spot, S, on the film which is illuminated by both horizontal beams which enter through the semi-transparent mirror, BS. The two beams strike the film at different angles of incidence. In this arrangement it is possible simultaneously to gather
Figure 6.5. Scheme for simultaneous ellipsometry and reflectometry of liquid films: L, He-Ne laser; F, attenuation filter; BS, beam splitter; Ml, M2, mirrors; P, polarizer (GlanThomson prism); K, compensator; C, measuring cell; S, observation spot; Al, A2, analyzers; Dl, D2, D3, light detectors. (Redrawn from den Engelsen and Frens, loc. cit.).
11
J.M. Corkill, J.F. Goodman, C.P. Ogden and J.R. Tate, Proc. Roy. Soc. A273 (1963) 84. 2 I. Clunie, J. Goodman, and B. Ingram, in Surface and Colloid Science, Vol. 3, E. Matijevic, Ed., Wiley (1971) p. 167. 31 D. den Engelsen, G. Frens, J. Chem. Soc. Faraday Trans. 170 (1974) 237.
THIN LIQUID FILMS
6.9
ellipsometric data at a large angle of incidence (for optimum sensitivity) and monitor the thickness of the film by reflectometry at small angle of incidence. An optical three layer model has proved superior to a one-layer model for the Interpretation of the ellipsometric data. The refractive indices of the film and surface layers were determined, and it was found that the index for the surface is higher than that for the film core.
Figure 6.6. Simple cell for FTIR measurements of black foam films in a frame. (Redrawn from Umemura et al.l'.)
The measurement of the absorption of infrared light provides information about the water content in the films. This option is of major significance for black films. In order to obtain measurable absorption values, the Infrared light has to pass through a series of vertical films (up to 10), formed for example, in a cylindrical tube acting as the frame2'. This method gained significant improvement with the introduction of the contemporary Fourier transform infrared technique (FTIR), permitting measurable absorption values of the infrared light to be obtained, even from single black films. The thickness of the aqueous core can be derived from the absorption at 3400 cm"1 which is related to the OH stretching vibration of the water molecules. The cell used to form films of ca. 2 cm2 area is Illustrated in fig. 6.6. By fitting the calculated curves of polarized FTIR spectra to the respective experimentally obtained dependence, the complex refractive indices are found, as well as the thickness of the aqueous core. The FTIR technique also allows study of the orientation and phase-state of hydrocarbon chains, supplying additional information concerning the film structure. X-ray reflectivity under small angles has been used in the study of vertical foam films in a rectangular frame and a horizontal X-ray diffractometer3'. Later, this technique was improved in order to reduce the limit of the angle of incident beam to about 10 mrad4'. Horizontal foam films in a frame from porous sintered glass, and a vertical X-ray diffractometer have also been used5'. Models of the film structure were 11
J. Umemura, M. Matsumoto, T. Kawai and T. Takenaka, Can. J. Chem. 63 (1985) 1713. J. Corkill, J. Goodman, C. Ogden, and J. Tate, Proc. Roy. Soc. A273 (1963) 84. 31 J. Clunie, J. Corkill, and J. Goodman, Discuss. Faraday Soc. 42 (1966) 34. 41 O. Belorgey, J.J. Benattar, Phys. Rev. Lett. 66 (1991) 313. 51 D. Platikanov, H. Graf, and A. Weiss, Colloid Polymer Sci. 268 (1990) 760; 271 (1993) 106. 21
6.10
THIN LIQUID FILMS
employed In the interpretation of curves of the reflectivity vs. angle, as well as a means of modelling the distribution of the electron density in a direction perpendicular to the film. Fitting of the experimental points to the curves calculated according to a model, yielded the appropriate model structure which corresponded best to the majority of the experimental results. Data related to the film thickness were also derived. 6.2c Microscopic foam or emulsion films Small circular foam or emulsion films, with a radius within the range of 10500 \xm , are considered as microscopic films. The experimental technique for their study has proved very successful, and Is continually being improved. It permits the measurement of thermodynamic quantities, to follow the kinetic behaviour, the formation of black films, the realization of metastable states, etc. An advantage over macroscopic films is the possibility of working at very low surfactant concentrations.
Figure 6.7. Scheme of a measuring cell for the study of microscopic foam or emulsion films; (a) in a glass tube; (b) with a reservoir of surfactant solution, d'; (c) in a porous plate; a, glass-tube film-holder; b, biconcave drop; c, microscopic liquid film; d, glass capillary; e, surfactant solution; f, optically flat glass; g, porous plate.
The measuring cell of Scheludko and Exerowa has proved to be a versatile and reliable tool for the formation of microscopic horizontal foam or emulsion films1'. It is presented in fig. 6.7, with variants (a), (b) and (c). The foam film c is formed in the middle of a biconcave drop b, situated in a glass tube a of radius R, by withdrawing liquid from it (variants (a) and (b)) or the film is in the hole of porous plate g, (variant (c)). Photographs showing the formation of black foam film via black spots taken under a microscope are presented in fig. 6.8. The suitable range of tube radius R is 0.2-0.6 cm in variants (a) and (b), whereas the film radius ranges from 100 to 500 (im . In variant (c), the radius of the hole can be considerably smaller, for example 120 um , and consequently, the film radius is 10 |im . The film is situated in the closed space of the cell, saturated with the solution vapour. The periphery of the film is in contact with 11
A. Scheludko, D. Exerowa, Comm. Dept. Chem. Bulg. Acad. Set 7 (1959) 123.
THIN LIQUID FILMS
6.11
Figure 6.8. Photographs taken under a microscope of a microscopic film showing the creation and growth of a black film.
the bulk solution from which the film is formed. The film holders, the tube, or the porous plate, are welded to capillary d. The inner part of the tube a carrying the biconcave drop is finely 'furrowed' with vertical lines, situated close to one another, which improves wetting1'. A constant capillary pressure acts on the film formed in variants (a), (b) and (c) in fig. 6.7: it is determined by the radius of curvature of the meniscus. Porous plates of various pore radii can be used in variant (c)2>, fig. 6.7. If the meniscus penetrates into the pores, their size(s) determine the radius of curvature; a small pore size allows increase in the capillary pressure until the gas phase can enter them. The capillary pressure can be increased to above 105 Pa , depending on the pore size and the surface tension of the solution. 6.2d Determination of the disjoining pressure The disjoining pressure 77(h), due to the surface forces acting in the film is the most important thermodynamic quantity which determines the film stability. The experimental techniques described in sec. 6.2b and 6.2c and those using macroscopic films, are usually used for determination of /7(h) as well. At equilibrium, rj(h) is equal to the outside pressure difference Ap, applied to the film. In the case of flat vertical foam films on a frame (fig. 6.3), Ap = -pgH is the hydrostatic pressure, H 11
D. Exerowa, M. Zacharieva, R. Cohen, and D. Platikanov, Colloid Polym. Sci. 257 (1979) 1089. 21 D. Exerowa, A. Scheludko, Compt. Rend. Acad. Bulg. Sci. 24 (1971) 47.
6.12
THIN LIQUID FILMS
being the height of the spot where the thickness is measured above the horizontal liquid reservoir. Vertical flat foam films on a frame are also used in a completely different method for studying long range surface forces, and the interaction Gibbs energy in thin liquid film, namely light scattering. When a vertical macroscopic foam film is illuminated, part of the light is scattered by thermal fluctuational microwaves in the film surfaces. Measuring the intensity of the scattered light enables one to calculate the film tension, and the energy of the long range interactions in the film1 . Surface light scattering has been introduced in sec. III. 1.10. For microscopic horizontal circular foam or emulsion films (fig. 6.7) Ap is the capillary pressure of the concave meniscus around the film, which is determined by the meniscus curvature and the interfacial tension of the bulk solution. The quantity Ap is accessible experimentally and, at equilibrium, /7(h) = Ap can be determined. Let us consider this case more closely, since it is the basis of the successful and widely used experimental 'thin liquid film -pressure balance technique'2).
Figure 6.9. Block scheme of the thin liquid film pressure balance technique: 1, recorder; 2, DC amplifier; 3, photomultiplier; 4, high voltage supply; 5, reflected light microscope; 6, foam or emulsion film; 7, measuring cell; 8, thermostatting device; 9, standard manometer; 10, pressure transducer; 11, membrane pump to apply a gradual pressure change. A block scheme of the apparatus is shown in fig. 6.9. The films are formed in the porous plate cell (fig. 6.7, variant (c)). The hydrodynamic resistance in the porous plate is sufficiently small. The maximum capillary pressure which can be applied to the film is determined by the pore size and the surface tension y of the solution. When the maximum pore size is 0.5 |im , the capillary pressure is ~ 3 x 10 5 Pa at y = 70 mN/m.
11 21
A. Vrij, Adv. Colloid Interface Sci. 2 (1968) 40. D. Exerowa, T. Kolarov, and Khr. Khristov, Colloids Surf. 22 (1987) 171.
THIN LIQUID FILMS
6.13
The cell Is placed in a thermostattlng device, mounted on a microscopic table. Thus, the film can be monitored and measured photometrically in reflected light. The apparatus sits on a special antivibration table in a thermostatted room. In order to eliminate any parasitic stray light, special diaphragms are employed for the incident and reflected light. The reflected light enters the photomultiplier, 3; its signal is amplified and registered by the recorder, 1. The regulation of the capillary pressure is achieved by a special membrane pump, 11, which allows a gradual and reversible change in the gas pressure p g in the closed cell. The pump and the manometer are placed close to the measuring cell and are connected to it by thick-wall tubing, thus ensuring good thermostatting. The thin liquid film pressure balance technique has been used by a number of researchers, who have introduced several technical improvements. For example, values of 77(h) less than 100 Pa prove difficult to measure, but there should be an entire conformity with the equation giving the balance of pressures acting in the film and the geometry of the measuring cell1' IJ[h) = pg-pr+^—Apghc
[6.2.3]
where p r is the external reference pressure; r is the radius of the capillary tube; A/7 is the density difference between the gas and aqueous surfactant solution; and h c is the height of the solution in the capillary tube above the film.
6.2e Determination of the contact angle and the film tension Microscopic foam or emulsion film techniques also allow the determination of two other important thermodynamic characteristics of the films: the contact angle a, appearing at the contact of the film with the bulk solution from which it is formed, and the film tension, yf, related to it. See below, [6.3.47]. For the wetting film equivalent we have already derived [III.5.3.7]. Two methods for the of-measurement have been developed: a topographic method, and a film expansion method2'3'. (i) The topographic method. This is most suitable for small contact angles, and is used in the study of black films of thickness 6-8 nm. The technique is based on the measurement of the radii of the interference Newton rings when the film is observed in reflected monochromatic light (fig. 6.10). By knowing the film thickness, the contact angle can be calculated according to tan2a=B2-4A(C-h/2) where
11
V. Bergeron, C.J. Radke, Langmuir 8 (1992) 3020. A. Scheludko, B. Radoev, and T. Kolarov, Trans. Faraday Soc. 64 (1968) 2213. 31 T. Kolarov, A. Scheludko, and D. Exerowa, Trans. Faraday Soc. 64 (1968) 2864. 21
[6.2.4]
THIN LIQUID FILMS
6.14
(a)
(b)
Figure 6.10. (a) A common black film; (b) and a segment of its meniscus; surfactant: NaDS ( 2.77 x 1(T 4 mol dm" 3 + NaCl 0.2 mol d m ' 3 ); magnification; ~ 300 times (for (b) it is much more); Xj and x2 are the distances between neighbouring interference rings. A_
I 2JX-I Xry
2
*i~*2. .XQ — X-l
^
I A.X-1 Xn
x\-1xx .Xrj
c=l_
X-l
i= £
_X_ 471
Here, Xj is the distance between the first and the second Newton rings; x2 is the distance between the first and the third Newton rings; X is the wavelength of the light; n is the refractive index of the solution, and h , the equilibrium thickness of the film. The applicability of the topographic method can be extended by using an approximate solution of the Laplace equation. In this case, it is necessary to measure the radius rk of the Jc-th Newton ring of the meniscus surrounding the film, and the thickness at which this ring emerges. (ii) The expansion method. This procedure allows accurate determination of larger contact angles, and is suitable for the study of Newton black films of thickness about 5 nm. The method is based on the ratio between the parameters of the thicker film (rx and ay) and the black film, which results from the thicker film at constant volume of the meniscus. An equilibrium black film of radius r2 and a contact angle a is formed from the initial non-equilibrium thicker film, having parameters Tj and a = 0 during its 'expansion' at constant V The value of a can be determined from the experimentally measured values of r t , r2, a2 and R (the tube radius) according to V = — E-RuE(R + 2/3u) + RuF(R + l/3u)--J(r2
-U2)(R2 -r2)
[6.2.5]
THIN LIQUID FILMS
6.15
where E{q>, k2) and F(cp, k2) are elliptic integrals of the second and first order \R2 -r2 w = arcsin, —-x =-, \R2-u2
,
R2-u2 fc= =— R2
, Rsina-r and u = R-rsina
The contact angles can also be determined by the technique of a floating lens1', which is widely used in the study of black emulsion films of hydrocarbons in aqueous media. The, 'diminishing bubble method' is also used for determination of both static and dynamic contact angles of microscopic foam films (see subsec. 6.2h). The contact angle of a macroscopic, flat foam film on a vertical frame can be determined from data on the refraction of a monochromatic light beam passing through the contact zone between the film and the bulk solution from which the film is drawn2 . There is a simple relationship between a and yf at equilibrium (see sec. 6.3e) yf =2ycosa
[6.2.6]
Hence, the film tension is often determined from experimental data for a and y for a given foam or emulsion film. However, yf of foam films can also be measured directly. To this end, the following method is used, both under static or dynamic conditions3'. A hemispherical foam film can be obtained by blowing a bubble from a vertical capillary tube (fig. 6.11). The capillary tube is placed in a vessel with the surfactant solution, so that its upper orifice is close to the solution's surface. When a gas (air) of a given defined pressure is introduced into the tube over the solution surface a foam film is formed, acquiring a hemispherical shape.
Figure 6.11. Scheme of a device for the measurement of the film tension of a hemispherical foam film: A, capillary tube; C, vessel containing the solution; D, cover with small hole, H; M, foam film; L, manometer; P, precision pump. (Redrawn from Platikanov, et al., loc. cit.)
11
D.A. Haydon, J.T. Taylor, Nature 217 (1968) 739. H.M. Princen, S. Frankel, J. Colloid Interface Sci. 35 (1971) 386. 31 D. Platikanov, M. Nedyalkov, and N. Rangelova, Colloid Polymer Sci. 269 (1991) 272. 21
6.16
THIN LIQUID FILMS
The lower part of the capillary Is placed in a closed space filled with a gas, whose pressure can be measured accurately and which can be varied at different rates by means of a special pump. The manometer registers the difference from the atmospheric pressure Ap, which is equal to the capillary pressure of the spherical foam film, Ap = pgAH = 2y{ /Rf
[6.2.7]
The radius of curvature of the film, Rf , is determined from photographs taken with a horizontal microscope. Equilibrium values of yi belong to fixed pressure differences Ap, while the dynamic film tension, / f , is obtained by monitoring the film as a function of Ap , which Is varied at a certain rate. 6.2/" Measurement of foam, film
elasticity
The elasticity of liquid films (Gibbs elasticity under quasi-static equilibrium conditions, or non-equilibrium Marangoni elasticity) plays an important role in the stability of some foam films, and is characterized quantitatively by the elasticity modulus dyf Ef = A^— [6.2.8] dA where y{ is the film tension; A Is the film area. For a common thin film, yf = 2y. The elasticity can be measured both with macroscopic vertical and horizontal foam films and with individual foam bubbles. An apparatus consists of three glass frames one is the main one, and two are auxiliary1'. The main frame is immobile and is connected to a dynamometer which measures the tension of the film after its formation. The film is illuminated with monochromatic light, and the change in film thickness during its thinning or expansion is determined from the horizontal interference fringes. The thickness measurement begins after a black film is formed in the upper part of this frame which is, in fact, the initiation of the elasticity determination. Drawing out the second frame increases the Interface by about 80%. The third frame, which moves in a direction opposite to the second one, serves to ensure a constant level of the solution in the cuvette. The Increase in the surface tension as a result of surface enlargement leads to overflowing of some parts of the film to the lower Plateau border of the main frame, and to stretching the upper part of the film. The relative deformation AA/ A can be determined from the change In the profile of film thickness, and the elasticity modulus can be calculated from [6.2.8]. Another technique employs two frames, with films moved in opposite directions by a reversible motor while the total surface area remains constant2'. When starting the experiment, the frames are submerged in the solution. The solution level is then lowered, allowing films to form on the frames. When black spots are formed in the upper
11 21
K.J. Mysels, M. Cox, and J. Skewls, J. Phys. Chem. 6 5 (1961) 1107. A. Prins, C. Arcuri, and M. van den Tempel, J. Colloid Interface Scl. 24 (1967) 84.
THIN LigUID FILMS
6.17
part of the main frame, the motor brings the frames into periodical motion, leading to changes in the film areas. The average variation in thickness along the profile is determined from a series of photographs of the pattern of interference fringes. The variation is caused by film thinning owing to gravity, and thus the change in film thickness owing to elastic strains of film elements is found. The data obtained serve to determine Ay and AA , which are necessary for the calculation of elasticity. The main disadvantages of the techniques involving vertical films are that the deformation can only develop after black spots are formed which excludes the investigation of low surfactant concentrations; also, the high rate of thinning makes it difficult to distinguish between the effects of drainage and film deformation. In an apparatus with a horizontal film1', the frame is made of four platinum wires of diameter 40 (xm and length 120 mm. The frame area and the film area can vary from zero to 140 cm2 . The film is isolated from the bulk solution, and drops of the solution are introduced by a special device. In order to obtain a film with uniform initial thickness, it is stirred by an intense horizontal air stream. The elasticity modulus is evaluated from the measured film tension and local film thickness using, ^
= 2^-*)/^
[6 2 91
' -
where the subscripts 1 and 2 refer to the initial- and deformed- states of a film, respectively. The film thickness is determined interferometrically. This technique allows one to perform quasi-static equilibrium deformations and to study solutions with low surfactant concentrations (up to the c.m.c). 6.2g Measurement of the lateral electrical conductivity of foam films Figure 6.12 presents the scheme of a device for the measurement of the lateral electrical conductivity of foam films2 . The horizontal film is formed between electrodes 3 and 4; the surfactant solution enters from vessel 1 into cell 2. The thickness of the ring-like film depends on the rate of lowering the liquid level. A portion of the solution is supplied to the cell by means of compressed air, and then re-enters vessel 1. The rate of this process is controlled by the rate at which the mercury shifts from the vessel 8, to tube 7, when air enters the system through capillary 6. The electrical resistance between the electrodes is measured by the bridge, 5. The film thickness can be calculated from the resistance Re]el =
In(r V 29 /r,) . U 2nKhh
[6.2.10]
11 A. Rusanov, V.V. Krotov, in Progress in Surface and Membrane Set Vol. 13, D.A. Cadenhead and J.F. Danielli, Eds., Academic Press (1979) p. 415. 21 A. Scheludko, Adv. Colloid Interface Set 1(1967) 391.
6.18
THIN LIQUID FILMS
Figure 6.12. Scheme of a device for the conductometric study of foam films. 1, a vessel containing the surfactant solution; 2, measuring cell; 3 and 4, cylindrical electrodes; 5, measuring bridge; 6, capillary; 7, tube; 8, vessel containing mercury; 9, microscope. (Redrawn from Scheludko, loc. cit.)
where Rel is the electrical resistance of a thinning (or equilibrium) film of thickness h ; fj and r2 are the inner and outer radius of the ring-like film (the radii of the cylindrical electrodes), respectively. The quantity KL is the specific conductivity of the solution. Another type of measuring cell1' can be used, not only for measurement of the electrical conductivity, but also for determining transference numbers of ions in black films. Two cylindrical hollow electrodes (fig. 6.13), made of silver, 1, are placed coaxially one above the other. The lower electrode is placed in a Teflon vessel 3, containing the solution. The upper electrode can be moved vertically by a precise micrometric system. Ring 2 is made of a porous glass, and is placed on each electrode. When the procedure involves measurement of the electrical conductivity of the film, vessel 3 is filled with the solution. The rings are wetted by the solution and, when
Figure 6.13. Scheme of a measuring cell for the determination of the lateral electrical conductivity and the ionic transference numbers in black foam films. 1, silver electrodes; 2, rings made of porous sintered glass; 3, Teflon vessel. (Redrawn from Platikanov and Rangelova, loc. cit.)
11 D. Platikanov, N. Rangelova, in Research in Surface Forces, B.V. Derjaguin, Ed., Vol. 4, Consultants Bureau, New York (1972) p. 246.
THIN LIQUID FILMS
6.19
drawn apart, an almost cylindrical foam film of a certain thickness forms between them. The film drains until it reaches its equilibrium thickness, after which the electrical resistance Re] is measured by an AC bridge. The lateral electrical conductivity of a film of 1 cm in length and width Kf (Q"1) is calculated from K f =—!— 7tRela
[6.2.11]
where I is the extent of the film, measured by a microscope with an ocular micrometer; a is the minimum film radius at a distance 1/2 from the porous ring; the value ! / a Is calculated from the ratio r I a = cosh((/a); ris the radius of the porous ring. The measurement of the transference numbers of the ions in a film, t£ and t£, is performed when the lower porous ring 2 is wetted with the solution studied, and ring 1 of the upper electrode is wetted by a solution in which the concentration of NaCl is twice as high. A concentration cell is produced, whose voltage is measured by an electrometer with a vibro-transformer. The transference numbers in the film are calculated using [1.5.5.17].
6.2h Measurement of the coefficient of gas permeability through foam films The gas permeability coefficient, Kg (in ms" 1 ), is defined by1' ^=-KAAc dt 8
[6.2.12] 8
where dJV is the number of gas moles passing across the foam film area, A, during a time dt; Ac is the difference in gas concentration between the two sides of the foam film. The application of a concentration difference Ac at constant temperature means that there is a difference Ap between the pressures on the two sides of the film. As a result, the film is curved and the capillary pressure due to this curvature exactly equals the difference Ap. Because of their small thickness, the foam films have a negligible mass, hence the shape of the curved film is virtually spherical. Such a film can be formed as the top of a foam bubble floating freely on the solution surface during the process of film thinning. A typical size of such a bubble Is O(mm). Under gravity, the part of such a bubble below the surface of the gas/bulk solution is non-spherical. The bubble can be photographed from the side in transmitted light. On this basis a method has been developed for determining the gas permeability through foam films2'. The size of the film is measured from the photographs (fig. 6.14) and, accounting for the deviation from the spherical shape in the lower part of the bubble, the decrease in bubblevolume and area with time can be calculated, and hence the gas permeability coefficient.
" A.C. Brown, W.C. Thuman, and J.W. McBain, J. Colloid Sci. 8 (1953) 508. H.M. Princen, S.G. Mason, J. Colloid Sci. 20 (1965) 353.
21
THIN LIQUID FILMS
6.20
Figure 6.14. Gas bubble floating on the surface of a surfactant solution.
A similar bubble which, Instead of floating freely, remains fixed on a well-wetted porous plate, is used in the so-called 'stationary bubble method'11. The foam film is formed on the porous plate, acquiring the shape of a hemisphere. The radius of curvature R is practically equal to the radius of the perimeter at the base of the hemispherical bubble. Because of the gas passing from the bubble through the foam film into the atmosphere R decreases with time t. The values of r are measured and Kg is calculated from Kg = (p/4yit)(R%-R?) + (2/3t)(r0-rt)
[6.2.13]
where i?0 Is the radius at the beginning of the measurement, R^ the same at time t, p is the atmospheric pressure and y* is the film tension. Bubbles having sizes of the order of tens of micrometres, floating on the surfactant solution surface, deviate only a little from sphericity. This fact has been used in the 'diminishing bubble method'2', which allows the measurement of contact angles of black films, the line tension of the contact line film/meniscus, and the coefficient of gas permeability through the film. Figure 6.15 presents a possible set-up. The solution studied Is poured Into the Teflon vessel. The bubble is observed and photographed in
Figure 6.15. Sketch of a device for the measurement of the gas permeability of films by the 'diminishing bubble method': 1, camera, 2, microscopes; 3, thermostat; 4, teflon vessel; 5, plane glass plate. (Redrawn from Nedyalkov et al., loc. clt.}
11
R. Krustev, D. Platikanov, and M. Nedyalkov, Colloids Surf. 79 (1993) 129. M. Nedyalkov, R. Krustev, D. Kashchiev, D. Platikanov, and D. Exerowa, Colloid Polymer Set 266(1988) 291. 21
THIN LIQUID FILMS
6.21
light, transmitted through the flat transparent bottom by means of the microscope 2 and camera 1. Simultaneously, the film at the top part of the bubble is observed and photographed by means of the microscope 4 and camera 5 in reflected light. The bubble radius R and the radius of the black film r are measured from the photographs as a function of the time during which the bubble shrinks as a result of the gas diffusion through the film. The diminishing bubble method is used for the estimation of Kg according to
K g =[(p/2 r )(i?4_ R 4) +( 8 / 9 ) ( R 3_ R 3)j J r 2 d J
[6 .2.i4]
where Ro, Rt p, and t have the same meanings as in [6.2.13]; y is the surface tension. The diminishing bubble method can also be used for the determination of the contact angles of Newton black films, especially under dynamic conditions where a = a(t). The gas pressure p of the bulk gas phase in the closed cell (fig. 6.15) can be varied as a function of time, by using a special device. Accordingly, the bubble varies its size. The experimentally measured bubble radii R(t) and film radii r{t) are used for the calculation of the a(t) values using
a(t) = l L c s i n ^ - ^ ^ l - a r c s i n ^ ^ l 2[ \R 3yr ) Zyr J
[6.2.15]
p being the solution density, and y its surface tension. Certainly, the great variety of experimental techniques cannot have justice done to them in this short overview. Besides many variants of the methods reviewed above, there are also other original methods for studying specific aspects of the behaviour and properties of thin liquid films, including foam film studies with or-particles, the method for determination of surface diffusion coefficients in microscopic foam films by fluorescence recovery after photobleaching (FRAP, see sec. 1.7.15), the formation and studying of microscopic foam films stabilized by insoluble monolayers, either in a Langmuir surface balance, or in the Scheludko-Exerowa cell, by adsorption of the insoluble surfactant from the gas phase. Many more references to the literature can be found in the book by Exerowa and Kruglyakov, mentioned in sec. 6.9b. 6.3 Thermodynamics of thin liquid films A thin liquid film is defined as a liquid phase which is so thin that there is no liquid with bulk properties inside the film. Here, the 'bulk' is the liquid phase, a , from which the film has been formed (fig. 6.1b). Hence, a thin liquid film is typically small thermodynamic phase. Most of the physical properties vary along the z-axis normal to the film. According to Gibbs, the thermodynamic description of such a system is based
6.22
THIN LIQUID FILMS
on a simplified model in which the differences between the values of physical quantities of the model and of the real system are introduced as excess quantities (see also sec. 1.2.5). In this section, a thermodynamic analysis of a symmetric, plane-parallel, horizontal thin liquid film is presented. This is the simplest case for a film with fluid interfaces. More complicated analyses (beyond our scope) involve the influence of gravity on nonhorizontal films, as well as the properties of films curved owing to a pressure difference across them. Different simplified film models can be considered: a model with one Gibbs dividing plane, a model with two Gibbs dividing planes, a model with two dividing planes according to Guggenheim, etc. Here, the thermodynamic analysis is developed using only the model with two Gibbs dividing planes. It seems that this thermodynamic approach is most convenient for interpretation of the experimental results. Moreover, it makes a connection with previous sections (1.2.5 and 22)1'. 6.3a Description of the simplified film model The entire thermodynamic system (fig. 6.16) consists of: (i), the liquid phase a, in general a multi-component solution; (ii), the phase (3 , either a gas (in the case of a foam film) or liquid, immiscible with the liquid a (in the case of an emulsion film); (iii), the small phase / , a thin liquid film formed from and connected to the liquid phase a. The thin liquid film is drawn out in a solid frame; the solid material of the vessel walls and the frame is not soluble in phases a and (3. The following simplifications are accepted in the film model: a. The Plateau borders, specifically the menisci, at the film-contact with the solid walls are neglected. The film/meniscus/bulk liquid transition will be considered in sec. 6.3e. b. Both interfaces between the film and phase (i are replaced by parallel, horizontal,
Figure 6.16. Thin film in contact with phase a . Explanation in the text.
11 J.A. de Feijter, Thermodynamics of Thin Liquid Films, in Thin Liquid Films, I. Ivanov, Ed., Marcel Dekker (1988) Chapter 1.
THIN LIQUID FILMS
6.23
Gibbs dividing planes. The whole space outside the dividing planes consists of fluid with the bulk properties of phase (3 . c. The space between the dividing planes is filled by liquid with the bulk properties of the reference phase a . d. The distance, h , between the dividing planes is defined as the thermodynamic thickness, h , of the thin liquid film. e. The film has a macroscopic area. 6.3b Mechanical equilibrium of thin liquid film Initially, a thick liquid film is drawn by the frame from the liquid phase, a . It can drain only if pa < pP; the pressure difference, Ap = pP -p™ , is the driving force for the film thinning. We may ask why an equilibrium thin liquid film can be obtained at the end of the process of thinning if Ap = constant? As shown in sec. III.2.3, the pressure p In an isotroplc bulk phase (such as phases a and (5) is the same everywhere (Pascal's law). This means that the following components of the pressure tensor are identical: I6'31'
P»=Py,=Pa=P
This does not apply to a real interfacial layer and neither to a real thin liquid film (fig. 6.1). There, only the component pzz of the pressure tensor normal to the film (or single interface) remains constant, while the components parallel to the film (or single interface) are functions of z: Pzz = Pn = constant
[6.3.2]
I63'3!
P«=Pyy=PXz)
As a result, an Interfacial tension y arises at each isolated interface between two semiinfinite bulk phases, its mechanical definition being de Bakker's formula [III.2.3.5]
7= J [Pn-Pt< z ']
dz
= J [p P -P t (z)] dz
[6.3.41
A thin liquid film, just like an interface, is anisotropic along the z-axis (fig. 6.1b) and, by analogy, the film tension y* (fig. 6.16) can be defined by the same expression:
/ = J [P p -P t (zl] dz
[6.3.5]
Here, the integration is carried out from the bulk of phase f$ , over the entire film up to the bulk of phase (3 on the other film side. The film tension is a force per unit length, acting tangential to a surface, which is called the surface of tension of the thin liquid film. In the case of a symmetric, plane-parallel film, the surface of tension coincides
THIN LIQUID FILMS
6.24
Figure 6.17. Tension profile across a thin liquid film according to Eriksson and Toshev (toe. cit.); s.o.t. stands for surface of tension.
with the midplane of the film. The mechanical meaning of the film tension of a thin liquid film is that an external force per unit length, yi, is needed to extend the equilibrium film by an infinitesimal amount (see also fig. 1.2.1). Unless the films are thinner than a few nm, the major contributions to yf arise in the regions at the film interfaces (fig. 6.17), and a better correspondence with the real film can be attained if yf is divided into two equal parts, yf / 2 , acting in two surfaces of tension located at z = ±ht / 2 . The distance ht between the two surfaces of tension can be considered as the mechanical film thickness, ht, and it is determined by1' h
t=4- J [pp - p t ( z »] zdz
t6 3 61
--
On the basis of this simplified film model it is also possible to introduce a surface (interfacial) tension of the film, y° . Its mechanical definition is given by2 h/2
Y3 = J [pa -p t (z)]dz+ J [pP -pt{z)\ 0
dz
[6.3.7]
h/2
In [6.3.7], h is the thermodynamic film thickness, i.e., the distance between the two Gibbs dividing planes. Hence, y° depends on the location of the dividing planes, in contrast to the surface (interfacial) tension y between two bulk phases, which is independent of the location of the single dividing plane. The quantities y° and y{ II both act approximately in the same surface of tension. Both y° and y^ determine the mechanical equilibrium of the film in a tangential direction. We shall now consider the mechanical equilibrium of the film in the normal 11 21
J.C. Eriksson, B.V. Toshev, Colloids Surf. 5 (1982) 241. B.V. Toshev, I.B. Ivanov, Colloid Polymer Set 253 (1975) 558, 593.
THIN LIQUID FILMS
6.25
direction. In the simplified film model, the space between the two dividing planes is filled by liquid with pressure pa of the reference phase a (fig. 6.16). The film is thinning because of the pressure difference, Ap = pP - pa . Obviously, at equilibrium, Ap must be balanced by some additional pressure. This pressure, due to the action of surface forces in the film (or interactions between the adjacent phases (3 across the film), has been introduced as the disjoining pressure !7{h). Its mechanical definition is given by 1 ' n(h) = pn-pa
[6.3.8]
i.e., the disjoining pressure is the difference between the normal component of the film's pressure tensor (which is constant) and the pressure in the reference phase a . The thermodynamic definition of the disjoining pressure is considered in sec. 1.4.2. Taking into account that Ap = p P - p < \
pV = Pn
[6.3.9]
it follows that for symmetric, plane-parallel, horizontal thin liquid films /7(h) = Ap
[6.3.10]
Equation [6.3.10] is the condition for mechanical equilibrium of the film in a normal direction. A positive value for 77(h) means repulsion between the two film interfaces, i.e., both interfaces are 'disjoined' by the interactions due to surface forces. The resultant of all interactions, per unit film area, is actually fl(h). It balances the applied external pressure difference, Ap. For a plane-parallel, horizontal, real thin liquid film, which contacts the frame wall through a Plateau border, Ap is the capillary pressure at the curved surface of the corresponding meniscus. Equation [6.3.5] can be transformed into h/2
/ =2J"[pP-pt(z)]dz = 2 J [pP-pt(z)]dz + 2 J [pP-pt(z)]dz = = 2 J (pP-pa)dz + 2 J [pa-pt(z)]dz+2 J [pP-pt(z)]dz 0
0
h/2
Combination of [6.3.11] with [6.3.7], [6.3.9], and [6.3.10] gives yt=2y°+n{h)h
[6.3.12] 2
{
This is a very important relationship ' between the film tension y the surface tension of the film y° the disjoining pressure /7(/i), and the film thickness h, i.e., the most important thermodynamic characteristics of a thin liquid film. 11 21
B.V. Deryagin, N.V. Churaev, Kofi. Zhur. 38 (1976) 402. A.I. Rusanov, Roll. Zhur. 28 (1966) 583.
6.26
THIN LigUID FILMS
6.3c Fundamental thermodynamic equations of a thin liquid film. The fundamental thermodynamic equation for the internal energy t/f of a thin liquid film follows directly from the First and Second Law dUf =TdS f -pPdV f + y{dAt +^ y u j dn 1 f
[6.3.13]
i
which is an extension of [1.2.8.5]. Here, and subsequently, all superscripts denote the phase to which the respective quantity belongs. The first term to the right reflects the heat transferred reversibly to the film from outside, Sf being the entropy of the film. The second term is the volume work done on the film by its surroundings, Vi being the film volume. Note that this work is done by the external pressure pP onto the film. The work done for a reversible increase of the film area Af is given by the third term to the right (sec. 1.2.3 and fig. 2.1 there). The last term gives the chemical work done when an infinitesimal amount dn[ moles of component i in the film are added. The temperature T and the chemical potential /i{ of each component, i, are considered constant in all three phases, f, a and p (fig. 6.16). Using the common Legendre transformations we can obtain the fundamental thermodynamic equations for the Helmholtz energy F f and for the Gibbs energy Gf of the thin liquid film dF f =-S f dT-pPdV f + yfdAf +^// i dnf
[6.3.14]
i
dGf =-S f dT + Vfdp(3 + y f dA f +^// i dnf
[6.3.15]
i
Integration of [6.3.13] at constant T, pP, y{, and //j gives the total internal energy Uf of the thin liquid film U! =TSf-pSiVf +y!Af +^M,nl
[6.3.16]
I
From [6.3.16], the total Helmholtz energy F f and the total Gibbs energy Gf of a thin liquid film can be obtained easily F f = U f - T S f = -p^V{ + / A{ +^^1
[6.3.17]
i
Gf =U{ -TSf +pPvf = yiAi + ^ ^ 4
[6.3.18]
t
Differentiation of [6.3.16] gives dU{ =TdS f +S f dT-pl 3 dV f -V f dpP + 7fdAf +Atdyt+ i
i
[63 l g ]
THIN LIQUID FILMS
6.27
Combination of [6.3.19] with [6.3.13] leads to the very important Gibbs-Duhem relationship Jor a thin liquid film S f d 7 - V f d / + A f d / + £rcifd//i =0
[6.3.20]
This relationship, expressed per unit film area reads d / =-s!,dT + hdpP -^T/fd/ij
[6.3.21]
i
with
^-=h Af
^=sf Af
'
a
?L=ii Af
'
'
All the above equations, including the Gibbs-Duhem relationship, constitute the basic thermodynamics for a thin liquid film. They do not depend on the film model chosen, since we have not yet specified the extensive film properties. They are valid for symmetric, horizontal, plane-parallel films with fluid interfaces. In this theory, the Plateau borders, or menisci, do not occur, but they do so implicitly. Since the film and bulk phase a are in equilibrium, the border between them must also be at equilibrium with both. So, in principle, it must be possible to extract information about the film from the properties of the Plateau border, and conversely. 6.3d Thermodynamic approach using a film model with two Gibbs dividing planes According to this model, a thin liquid film consists of two parallel mathematical planes (Gibbs dividing planes) with the space between them filled by liquid with the properties of the bulk reference phase a; the space beyond the dividing planes is filled by fluid with the bulk properties of phase (3 (fig. 6.16). The differences between the extensive thermodynamic properties of the model and those of the real film are accounted for as surface excess quantities. So, the surface excess internal energy U^ of the film per unit area of each film surface is presented as h/2
U
f = J [Uv(z)-U«]dz+ J [Uv(z)-L/P]dz 0
[6.3.22]
h/2
In [6.3.22], U® and L/P are the volume densities of the internal energy in the homogeneous bulk phases, a and p, respectively; Vv(z) is the volume density of the internal energy in the inhomogeneous real film. The surface excess Helmholtz energy F^f of the film, per unit area of each film surface, is given by h/2 F
* = J [> v (z)-F v a ]dz+ J [Fv(z)-FvP]dz 0
h/2
[6.3.23]
6.28
THIN LIQUID FILMS
F^* and F$ being the volume densities of the Helmholtz energy in the homogeneous bulk phases, a and (3, respectively, and Fv(z) the volume density of the Helmholtz energy in the inhomogeneous real film. The excess surface concentration r ? f of each component i in the film, per unit area of each film surface, is given by h/2
T f = J [n vi (z)-<]dz+ J [n^zl-n^Jdz 0
[6.3.24]
h/2
n^ and n^ being the volume densities of each component i in the homogeneous bulk phases, a and (3 , respectively, and n^iz) is the volume density of each component i in the inhomogeneous real film. The surface excess quantities as defined by [6.3.22-24], as well as other extensive surface excess properties of the film, depend on the location of the Gibbs dividing planes. Just as in the case of a single interface, different reference conditions for the dividing plane location can be considered. The best option is to choose the same condition for locating both a single dividing plane between the phases, a and (5, and the two film dividing planes. Then, the surface excess properties of the film and the single interface can be compared directly. Most often, the condition Ff{ = 0 is used, i.e., the surface excess of component 1 in the film is zero, component 1 being the solvent in phase a (mostly water). According to this thermodynamic model, the total thin liquid film consists of two Gibbs dividing planes and a volume part filled by liquid with the bulk properties of the reference phase a. Hence, the values of the extensive properties of the film can be obtained as the sum of twice the surface excess plus the corresponding amount in the volume of the film. So, the total internal energy L7f of a film is Uf =2Ua! +Uaf = A1 (2Uf + hU%)
[6.3.25]
where Ua{ is the surface excess internal energy for each dividing plane, and Uaf the internal energy of the volume part of the total film. The total Helmholtz energy F f of the total film is given by Ff =
2F
[6.3.26]
F o f being the surface excess Helmholtz energy for each dividing plane, and F a f the Helmholtz energy of the volume of the total film. By the same token, the total amount n[ of each component i in the film is expressed by n[ = 2nf + nf = Af [2Pf + hn^)
[6.3.27]
with nf1 the surface excess of each component i in each dividing plane of the film, and n"f the amount of component i in the volume of the total film.
THIN LIQUID FILMS
6.29
Given the selected location of the Gibbs dividing planes, we obtain from [6.3.27] an equation which determines the thermodynamic film thickness h
n{=A{hn^
(/ff = o)
[6.3.28]
-1
h= A
[6.3.29]
"vl
According to [6.3.29] the thermodynamic film thickness h is equal to the thickness of a layer of the solution in phase a , which contains the same amount of component 1 as the film. The real films are usually stabilized by dense surfactant adsorption layers at both interfaces. That means that the thermodynamic film thickness is closer to the thickness of the inner liquid layer of the film, rather than to the whole film thickness. However, the stabilizing surfactant molecules can be very different and so are the film structures. Hence the comparison of the thermodynamic film thickness with the real film structure should be done for each system separately. In general, the physical film thickness is usually larger than the thermodynamic thickness determined by condition [6.3.28]. After differentiation, [6.3.25] can be transformed into 2dUof = dUf - dU af
[6.3.30]
As in [I.A4.1], the differential of the internal energy Uai of the volume part of the thin liquid film reads d[/ a f = TdS af - p a d V f +^// i dn 1 a f
[6.3.31]
i
Combination of [6.3.30] with [6.3.13] and [6.3.31], taking into account [6.3.9-10], [6.3.27], and the relationship Sf = 2 8 ^ +S a f , gives 2dti of = 2TdS<7f - /7(h)dVf + / d A f + 2^^dnf
[6.3.32]
i
This is the fundamental thermodynamic equation for the surface excess of the internal energy of the film. Following the usual procedure, the corresponding Gibbs-Duhem relation of the dividing planes of the film can be obtained as 2S of dT-V f d77(h) + Afd yf + 2 ^ n f d ^
=0
[6.3.33]
i
Expressed per unit film area d/ =-2SfdT + hdn{h)-2£r?{dfJi
[6.3.34]
i
which may be compared with [6.3.21]. This is nothing else than the Gibbs adsorption
6.30
THIN LIQUID FILMS
law for the film. Substitution of [6.3.12] into [6.3.32] leads to another form of the fundamental thermodynamic equation for the surface excess of the internal energy of the film 2dUof = 2TdSat - 77(h)Afdh + 2y°dA{ + 2^^dnf
[6.3.35]
i
As a consequence, the Gibbs-Duhem relation of the dividing planes of the film takes the alternative form 2SafdT+/7(h)Afdh + 2A f d7 o + 2^rr? f d^ i =0
[6.3.36]
i
or, per unit film area 2dy° =-2S^ f dT-77(h)dh-2^/] 5 f d / u i
[6.3.37]
i
On the basis of both the Gibbs-Duhem equations [6.3.37] and [6.3.34] very useful relationships between the main thermodynamic characteristics (h, /7(h), y, yf, y°) of thin liquid films can be obtained. From [6.3.37] it follows that 2 \^]—\
=-IJ(h)
[6.3.38]
which, after integration at constant temperature and all chemical potentials (for thick film h = °o , n(h) = 0 , ys = y), gives h
2(y° - y) = - J n(h)dh = AF(h) or 2ya=2y+AF{h)
[6.3.39]
The quantity AF(h) is usually called the Interaction Helmholtz energy of the film. Obviously this is the isothermal, reversible work per unit film area, done against the disjoining pressure, when thinning a thick film down to a small equilibrium thickness h. Equation [6.3.38] shows that when a non-zero disjoining pressure /7(h) appears, owing to interactions across thin liquid film, the surface tension of the film y° depends on the film thickness and differs from the surface tension y of the bulk liquid phase. This point will be discussed further in sec. 6.3e. Another relationship that can be obtained from [6.3.34] is —^—
=h
[6.3.40]
which, after integration at constant temperature and all chemical potentials, gives
THIN LIQUID FILMS
6.31
h
yf =2y-i
n(h)dh + n{h)h
or yf = 2y + II(h)h + AF(h)
[6.3.41]
Equation [6.3.41] can be used to determine the interaction Helmholtz energy of the film AF(h) from experimental data for y and yf (the last two quantities can be measured directly). For the so-called blackfilms, i.e., the case when 2y and y{ noticeably differ, the product 17{h)h is usually very small since h is extremely small. Then, I7(h)h is negligible and AF(h) = /
-2y.
From the fundamental thermodynamic equations for the surface excess of the internal energy of the film, [6.3.35] and [6.3.32], and using the Legendre transformations, i.e., changing variables, we can obtain the corresponding fundamental thermodynamic equations for the surface excess of the Helmholtz energy of the film 2dF(rf = ^S^dT
- n{h)dVf + yiAA{ + 2^// i dnf f
[6.3.42]
i
2dF af = -2SafdT-mh)A{dh
+ 2yadA{ +2^Mi
[6.3.43]
i
An alternative thermodynamic definition of the film tension y{ follows from [6.3.42] y1 =2 ^ r I 3 A )T, vf, nf
[6.3.44]
By the same token, thermodynamic definitions for the surface tension of the film y° and the disjoining pressure I7[h) can be derived from [6.3.43] ra = \^-r\
nih)= ~(^-)
[6.3.45]
[6.3.46]
Equations [6.3.44-46] show that all three quantities characterizing the film, y{ , ya , and, /7(h) are defined in terms of the surface excess of the Helmholtz energy, a major feature of the thermodynamic approach with two Gibbs dividing planes. 6.3e. Contact between thin liquid film and the adjacent meniscus A thin liquid film is always connected with a bulk liquid phase, or with a solid wall, through a meniscus called the Plateau border, a liquid body with concave interfaces. The meniscus contains the bulk liquid a from which the film has been formed. The contact film/meniscus is schematically presented in fig. 6.18. On the left side there is a
6.32
THIN LIQUID FILMS
Figure 6.18. Definition of the contact angle(s) and other parameters determining the transition zone between film and Plateau border. To the right of the figure, the liquid has bulk properties.
part of the horizontal, plane-parallel, symmetric, thin liquid film, and on the right side a part of the meniscus. Between them, a transition zone film/bulk always exists 1 . At equilibrium, the disjoining pressure J7(h) acts in the film, and the surface tension of the film y® differs from the surface tension y of the bulk liquid phase. Both /7(h) and yc depend on the surface forces in the thin liquid film. These interactions operate in the transition zone film/bulk as well, but they decrease with increasing thickness of the transition across the meniscus. This means that both IJ{h) and ( y- y° ) gradually decrease from the film to the bulk liquid, where they become zero. In a real system, the curved meniscus interface gradually becomes a flat film interface, and y gradually turns into y° . The macroscopic thermodynamic description does not consider the transition zone but assumes an abrupt change of y (bulk liquid) into y*5 (film). This approach requires the introduction of two other thermodynamic quantities: the contact angle film/bulk and the line tension of the contact line. For bulk liquid, the surface tension is y, and the shape of its concave interface is determined by the Laplace equation. If we consider y to be constant also in the transition zone, and extrapolate the meniscus interface according to the Laplace equation towards the film (the dashed curve in fig. 6.18), it intersects the midplane of the film (the horizontal dashed straight line). The line of intersection is the contact line of the film, and the angle between both dotted lines is the thermodynamic contact angle23) a0 (fig. 6.18). If we consider a cylindrical meniscus, i.e., a straight contact line, the balance of forces in the tangential direction, taking [6.3.41 ] into account, is given by
11
T. Kolarov, Z. Zorin, and D. Platikanov, Colloids Surf. 51 (1990) 37. H.M. Princen, S.G. Mason, J. Colloid Set 20 (1965) 156. 31 B.V. Deryagin, G.A. Marrynov, and Yu.V. Gutop, Roll Zhur. 27 (1965) 298. 21
THIN LIQUID FILMS
yf =2ycosa0
6.33
= 2y + I7(h)h + AF{h)
[6.3.47]
In the case of a curved contact line, (e.g., a circular film in a spherical meniscus) the balance offerees reads 1 ' 2ycosa 0 = / f +
-^r
[6.3.48]
where r is the line tension, and Rf the radius of curvature of the contact line. The term t/R1
is actually a two-dimensional capillary pressure, i.e., a surface pressure
difference on both sides of the curved contact line. However, this term is usually very small because of the very small value of r . Only for extremely low R{ should it be taken into account. Another definition of the contact angle film/bulk is in terms of the angle ah between the extrapolated Laplace meniscus of the film (the dashed curve in fig. 6.18) and the extrapolated surface of the film (the upper horizontal dashed straight line). Obviously, the contact angles a h and ao are different but this difference is usually very small. For the ah case, the balance offerees in the tangential direction (straight contact line), taking into account [6.3.39], is given by y° = ycosah,
2ycosah=2y+AF{h)
The contact angles, ah
[6.3.49]
and ao, can be measured experimentally using [6.2.4] in
which h = h for ah and h = 0 for a0 . From these experimental values combined with a measured value of y one can calculate values of the interaction Helmholtz energy AFlh) of the film AF(h) = -2y(l- cos ah)
[6.3.50]
AF(h) = - 2 y ( l - c o s « )-77(h)h
[6.3.51]
As already noted in the previous section, the term fl{h)h is usually very small, hence ah ~ ao . Equations [6.3.47 and 49] are actually extensions of the laws of Young and Neumann, see further sec. III.5.1. According to the thermodynamic approach developed so far, the surface tension of the film y° depends on the film thickness and it differs from the surface tension y of the bulk liquid phase because of the surface forces interacting in a thin liquid film. One might wonder whether the surface tension should not, at equilibrium, be the same everywhere, because otherwise a flow in the liquid surface layer would arise owing to the surface tension gradient. It can be shown that this contradiction is only apparent: it is a consequence of the simplifications in the thermodynamic model. Let us to that end consider another simplified mechanical model of the system, thin liquid film/transition zone/bulk liquid meniscus (fig. 6.19). A symmetrical, plane-
11
V.S. Veselovskij, V.N. Pertsov, Zhur. Fiz. Khim. 8 (1936) 245.
THIN LIQUID FILMS
6.34
Figure 6.19. Definition of parameters for describing the mechanical model of the film, transition zone-plateau border (bulk system). (Redrawn from B.V. Toshev and D. Platikanov, Adv. Colloid Interface Sci. 40 (1992) 157.) parallel, thin liquid film is in contact with a cylindrical meniscus. The midplane of symmetry of the film coincides with the basic film surface of tension. Let us assume that the surface tension is constant everywhere at the two outer surfaces of tension, and that it is equal to y of the bulk liquid phase in the meniscus. Deviations from the value of y owing to the surface forces are introduced as excess tensions, for one surface only, Ay{ (in the film) and Ay(z) (in the transition zone, where z is counted normal to the plane of the film). Note that in this model, while the constant y acts in the two outer surfaces of tension, Ay{ and Ay[z) act in the basic film surface of tension (the midplane of the film). The mechanical equilibrium in the normal direction is given by the condition 77(z) + Ap(z) = 77(z) + -£• = I7(z) R 1 - = R
dcos# dsin# =— dz dx
dcos0 r
= Ap(h)
[6.3.52]
dz [6.3.52a]
In [6.3.52], 77(z) and Ap(z) are the disjoining pressure, and the capillary pressure in the transition zone, respectively. Both pressures act normal to the outer surfaces of tension of the transition zone. They vary along the x-axis, but their sum is everywhere constant at equilibrium, and equal to the capillary pressure Ap(h) of the bulk liquid meniscus. The quantity 1/R is the curvature of the surface in the transition zone. Integration of [6.3.52], taking into account the fact that, at z = h / 2 , 0= 0 gives
THIN LIQUID FILMS
6.35
z
J n(z)dz-y(cos8-l) = Ap(h)(z-h/2)
[6.3.53]
h/2
Mechanical equilibrium in a tangential direction, in the symmetry plane of the film requires a constancy of the overall tension at any x 2[y+Ay(z) + Ap{h)N]cos8 = const.
[6.3.54]
z = JVcosfl where JV is the distance, along the normal to the surface, between the symmetry midplane of the film and the outer surface of tension of the transition zone. After elimination of JV , [6.3.54] transforms into 2[y+ Ay{z)]cosd + 2Ap(h)z = 2y+ 2Ayf + 2Ap(h)h/2 = /
[6.3.55]
The left side of [6.3.55] is the overall tension at any x in the transition zone, whereas the right side represents the overall tension in the film. The force quantities, 2Ayf and 2A}'cos#, are applied only in the basic surface of tension (the midplane) of the film. There are two other Aysin0 force components related to the two surfaces of the transition zone. These components counterbalance each other at any point on the basic surface of tension. The disjoining pressure, and the capillary pressure, act always and everywhere normally to the two surfaces of the transition zone, identified as surfaces of tension. The simultaneous validity of the conditions [6.3.53] and [6.3.55], requires z
Ay(z)cos8-Ay{ =- j /7(z)dz
[6.3.56]
h/2
or, written in derivative form dAy(z)cosg dz The results obtained on the basis of this simplified mechanical model can be compared with those obtained previously, if we use the following relationships y+Ay(z)=f(z),
y+Ayf = y°
[6.3.58]
With [6.3.58] equations [6.3.55] and [6.3.57] transform into 27o(z)cos6»+2Ap(h)z = 2 / a + 77(h)h = yf
[6.3.59]
dy° 2-^— = -n(h) dh
[6.3.60]
6.36
THIN LIQUID FILMS
Obviously, [6.3.59] coincides with [6.3.12], and [6.3.60] follows from [6.3.45] and [6.3.46]. This means that both approaches, the one assuming variable surface tension, and the other based on constant surface tension, are completely equivalent. The apparent differences are caused by the simplifying model- assumptions. 6.4 Non-equilibrium properties of thin liquid films When a thin liquid film is formed from a bulk liquid phase a in a surrounding fluid p it is initially a thick film. However, it becomes a thin liquid film as a result of the process of thinning under the driving force, Ap = pP - pa . This pressure difference can be: the capillary pressure of the meniscus, a hydrostatic pressure difference, or an externally created pressure difference. As shown in sec. 6.3b, an equilibrium thin liquid film is obtained only if Ap Is counterbalanced by a positive disjoining pressure n(h). However, when 17{h) remains lower than Ap, or if it is negative, no equilibrium can be established. The process of thinning leads either to rupture of the film or to a jump-like formation of a much thinner blackjilm stabilized by a compensating positive
n(h). The processes of film thinning, film rupture, and jump-like formation of black film, comprise the most important non-equilibrium behaviour of thin liquid films. These processes mainly have a hydrodynamic nature, and depend on the pressure acting and the size and shape of the films. In this section, the non-equilibrium properties are emphasized for microscopic, circular, horizontal films, surrounded by a double-concave meniscus. Such films are obtained in the cylindrical holder of the ScheludkoExerowa cell (fig. 6.7) and they obviously have a cylindrical symmetry. Microscopic films offer certain advantages with respect to treatment and investigation under strictly defined conditions. However, in sec. 6.4e some experiments with macroscopic, vertical films will be discussed.
6.4a Kinetics of thinning of plane-parallel liquid films An important factor determining the kinetics of thinning of thin liquid films is their anisodimenslonality, i.e., the fact that their radii r are much larger than their thickness h. Two quantities are introduced for the quantitative description of the kinetics of thinning: the lifetime T of thinning, and the rate v qfjilm thinning, defined according to the relationships r =h [ * J
v
v =- ^
[6.4.1]
at
ho
where h 0 is the initial thickness, h cr is the critical thickness at which the film ruptures and t is time. Let us first consider the thinning kinetics of plane-parallel, circular films with tangentially immobile interfaces, i.e., surfaces with a very high modulus (sec. III.3.6).
THIN LIQUID FILMS
6.37
Mostly, this is the result of the presence of surfactant adsorption layers. Under such conditions the hydrodynamics of the film thinning is determined only by the volume flow in the interior of the film, which is very well described by the Reynolds lubrication theory1', leading to the well known Stephan-Reynolds relation? for the rate uRe of thinning/thickening of a liquid layer of thickness h and viscosity r] between two solid circular plates of radius r under a pressure drop Ap
Equation [6.4.2] was first applied to the kinetics of thinning of circular, horizontal, microscopic liquid films by Scheludko3'. In his interpretation, the pressure drop Ap = Apc - IJ{h), represents the difference between the capillary pressure Apc of the meniscus and the disjoining pressure I7[h) in the film. Thus, ^
dt
^ 3^r 2
16.4.3,
Since the interfaces of the foam or emulsion films considered here are often not rigid, and in some cases (e.g., large films, or films from very low concentration surfactant solution) can deviate considerably from the requirement of high modulus, the applicability of [6.4.3] to the process of thinning imposes some restrictions: 1) The film interfaces should remain plane-parallel (should not deform); 2) the interfaces should be tangentially immobile (zero surface velocity); 3) the viscosity r/ should not depend on film thickness; 4) the rate of evaporation should be negligible compared to the rate of drainage; 5) the capillary pressure Apc of the meniscus should not be affected by film thinning. Experience with model systems of small circular foam films allows us to state that, usually, evaporation contributes negligibly to thinning of foam films. Obviously, films from highly volatile liquids require special control against evaporation. There is neither theoretical nor experimental evidence indicating that there is a dependence of bulk viscosity on film thickness h within the range of 0.01-1 um 4 '. The tangential mobility of film interfaces is affected considerably by the presence of a surfactant through the Marangoni effect (sec. III.3.6e). When the liquid drains from a film stabilized by a surfactant, a gradient of surface tension is created at its surfaces, which counterbalances the viscous stresses. This gradient, coupled to the surfactant adsorption gradient, causes surfactant mass transfer diffusion and flow from the bulk to the film surface, and surface flow in the direction of the adsorption gradient. The
11
O. Reynolds, Phil. Trans. Royal Soc. A177 (1886) 157. J.S. Stephan, Math. Natur. Akad. Wiss. 69 (1884) 713. 31 A. Scheludko, Kolloid.-Z. 155 (1957) 39. 41 Proved for large films by J. Lyklema, P.C. Scholten and K.J. Mysels, J. Phys. Chem. 69 (1965) 116, see sec. 6.4e. 21
6.38
THIN LIQUID FILMS
tangential velocity of the film interfaces is determined by the surface flow, which is always different from zero, i.e., condition 2), of zero surface velocity, is rarely satisfied1'. Deviations from [6.4.2-3] resulting from the Marangoni effect, have been observed experimentally and theoretically predicted2' in the expression
jL^LaWir]]
3D7
[6.4.4]
where D and Da are, respectively, the bulk and surface diffusion coefficients of the surfactant, y is the surface tension. The dimensionless quantity Ma = \r[dy/dcs)\/Dri is a variant of the Marangoni number^ and plays an important role in all transport processes at phase interfaces4'. Other factors (e.g., surface viscosity) affecting the surface mobility of films and, hence, the kinetics of their thinning, have also been analyzed5'. Equation [6.4.4] describes correctly the film hydrodynamics only for large Ma, equivalent to a small deviation from zero surface velocity. The general analysis of the role of surface mobility on the rate of thinning indicates that, at high surface mobility (Ma < 1), the hydrodynamics of film thinning changes strongly, thus leading to a deviation from approximation [6.4.2]. The outcome is that the rate of thinning v increases strongly compared to that of films with immobile surfaces, uRe 6>. In the limiting case of completely free surfaces (films without a surfactant), the resistance of inertia prevails over the viscous one. 6.4b Deviations from the plane-parallel shape during film thinning The Reynolds relationship [6.4.2] requires liquid drainage from the film to follow strictly the axial symmetry between parallel walls. Rigid surfaces ensure such drainage through their non-deformability, while non-equilibrium foam or emulsion films are, in fact, never ideally plane-parallel. The shape is determined by the balance between hydrodynamic and capillary pressure. Experimental studies have shown that only microscopic films with radii less than 100 |im retain their quasi-parallel surfaces during thinning, which makes them particularly suitable for model studies. Films with larger radii exhibit significant deviations, which affects both the kinetics of thinning and their stability, unless special precautions are taken (for example, by making the films rigid).
11
A. Scheludko, Adv. Colloid Interface Set 1 (1967) 391. B. Radoev, E. Manev, and I. Ivanov, Kolloid.-Z. 234 (1969) 1037. 31 Note that there are various definitions for Ma. In [8.1.4] a more general definition will be given, explicating the surface dilational modulus. In the present definition, the assumption is that there is only one surfactant. 41 A. Avramov, K. Dimitrov, and B. Radoev, Langmuir 11 (1995) 1507. 51 A. Barber, S. Hartland, Can. J. Chem. Eng. 54 (1976) 279. 61 E. Mileva, B. Radoev, Colloid Polym. Set 264 (1986) 965. 21
THIN LIQUID FILMS
6.39
Not-very-large films (r > 100 urn) retain axial symmetry, but lose their planeparallel shape. In their centres there forms a typical thickening, known as a dimple (thicker lens-like patch) which is surrounded by a thinner, almost plane-parallel, barrier ring. The dimple arises spontaneously as a result of the hydrodynamic resistance to thinning in the periphery of the circular liquid film1 2 . With further increase in radius, the thinning films lose their axial symmetry, and disintegrate into individual sub-domains. It was first believed that the dimple in foam films decreases, and even disappears, at small film thickness. Later, experimental investigations proved that the rate of thinning is practically equal in both the centre of the dimple and the barrier ring, implying that the difference of thickness between the thicker and the thinner domains does not decrease down to the critical thickness of rupture. This leads to a relative increase in the non-uniformity. On the other hand, non-uniformity of thickness increases with the increase in film size (film radius). These results can be very useful in the interpretation of experimental data about the dependence of the rate of film thinning and film lifetime on film radius. Compared with [6.4.3] and [6.4.4], the experiments indicate a weaker dependence of r on r (fig. 6.20) than predicted theoretically.
Figure 6.20. Film lifetime as a function of film radius for aqueous films of isovalerianic acid in KC1 solution according to [6.4.3] or [6.4.4]. (Redrawn from Exerova and Kolarov3'.)
A very Interesting, phenomenon, which is different in principle, has been observed for the thinning of microscopic, circular O/W emulsion films: spontaneous cyclic formation of a dimple was found to occur in an emulsion film from an aqueous solution of Tween 20 (a non-ionic surfactant) between two oil droplets. This phenomenon was described as a diffusion dimple formation in contrast to the dimple caused by the hydrodynamic resistance to thinning in liquid films. The dimple shifted from the centre to the periphery and periodically regenerated. During the dimple growth the thickness at the circular, almost plane-parallel, area of the film between the dimple and the meniscus remains approximately constant (no change of the reflected light intensity). Photographs of the different periods of a dimple growth are shown in fig. 6.21, and the
11
S. Frankel, K. Mysels, J. Phys. Chem. 66 (1962) 190. D. Platikanov, J. Phys. Chem. 68 (1964) 3619. 31 D. Exerova, T. Kolarov, Ann. Univ. Sofia, Fac. Chem. 59 (1964/65) 207. 21
6.40
THIN LigUID FILMS
Figure 6.21. Four consecutive pictures, (a) —» (d), of spontaneous dimple growth in an aqueous emulsion film, diameter 330 |im , stabilized by Tween 20. (Photographs from Velev et al.11.)
process Is presented schematically In fig. 6.22. An explanation has been suggested in terms of the Marangoni effect of continuous redistribution of surfactant molecules from the bulk to the surface until film equilibrium is reached. This phenomenon would probably yield new knowledge on the mechanism of instability of newly formed emulsion films and emulsions. During drainage of larger circular, horizontal films (r > 200 urn ), more complex phenomena have been observed (fig. 6.23). One, or more, thick domains form in such films, dividing them into parts: we shall call them channels. During film thinning the channels move, and sometimes separate from one end and disappear into the meniscus. The axial symmetry of drainage, assumed in the Reynolds model, is perturbed. At the same time there emerge 'centres of thinning' in the film. Such a film drains faster than expected for a homogenous symmetrically draining film of the same size. The 11
O. Velev, T. Gurkov and R. Borwankar, J. Colloid Interface Sci. 159 (1993) 497.
THIN LIQUID FILMS
6.41
Figure 6.22. Scheme of the main stages of spontaneous cyclic processes. Cross-section normal to the film. The dimple and film size are not to scale. (Redrawn from Velev et al., toe. cit.)
Figure 6.23. Photographs of microscopic foam films ( h ~ 1 0 0 n m ) prepared from sodium oleate in aqueous KC1 solutions, (a) and (b), small films with radius ~ 100 um (homogeneous); panels (c) through (f|, large films exhibiting the formation of 'channels'. (Photographs taken from A. Scheludko, Adv. Colloid Interface Set 1 (1967) 391.)
6.42
THIN LIQUID FILMS
complex structure of large films appears spontaneously and, therefore, it corresponds to a hydrodynamically more convenient regime of thinning. It seems that the deviations from the plane-parallel shape during film thinning provide the main reason for the considerable differences between the time of thinning measured, and calculated from [6.4.3]. At small radii (r < 200 urn ) and poor mobility of film surfaces, the experimental and calculated rates of film thinning were very close, while at large radii (r > 200 |j.m ) the experimentally determined values exceed the calculated ones by a factor of more than 10. This extremely large difference in the rates and time of thinning cannot be attributed solely to tangential mobility of film interfaces (estimated to be 5-80%). Moreover, such a mobility would not depend on film radius and would not change the functionality of the rate of thinning/radius. The theoretical model assuming non-deforming film surfaces, requires a linear dependence of the rate of thinning on the inverse square of the film radius. In all experiments performed, the u(r) relationship was weaker. These experimental results clearly indicate that the Reynolds relationship can be applied only to sufficiently small films (r < 100 |j.m ), i.e., to films of uniform thickness. Obviously, the main reasons for strong deviations from the theoretical relationship should be attributed to the deformations of the film interfaces. For the thinning dynamics of large, circular, horizontal films which were nonhomogeneous because they developed subdomains In them, the following equation has been derived1
677 \ 4}^r 4
This equation differs strongly from Reynolds' relationships [6.4.2] and [6.4.3], and has been checked with respect to the rate of thinning as a function of the film's radius (v vs. r -4/5 ) -pjjg experimental results are in good agreement with the theoretical prediction. Equation [6.4.5] holds for large films containing strongly expressed non-homogeneities of thickness. When the film's radius decreases, its profile becomes plane-parallel and its rate of thinning asymptotically approaches Reynolds' relationship. The theory also predicts that the transition between the two regimes of drainage should occur at a radius r* = Ayjyh I Ap . A typical value in the model experiments is r* = 50 um which agrees well with the experimental observations2'. 6.4c Kinetics of rupture of thin liquid films3} The study of processes leading to rupture of thin liquid films is useful for under-
11
E. Manev, R. Tsekov, and B. Radoev, J. Dispers. Sci. Technol. 18 (1997) 769. B. Radoev, A. Scheludko, and E. Manev, J. Colloid Interface Sci. 95 (1983) 254. 31 In sec. 8.3e the equivalent issue of emulsion films (coalescence) will be discussed in some detail. 21
THIN LIQUID FILMS
6.43
standing the reasons for their stability. By their very nature, thin liquid films are metastable with respect to bulk liquid. So, the simplest driving force for film rupture involves reaching a deviation from this metastable state. Typical examples of such unstable systems are symmetric thin liquid films in which the Van der Waals contribution to the disjoining pressure is not compensated. This contribution obeys Hamaker's relationship (sec. 1.4.6a) /7 h
( 'vdw=- K vdw/ f l 3
I6-4-6!
where K Vdw = Al&n, and A is the Hamaker constant. To this category belong films from some aqueous surfactant solutions containing sufficient amounts of electrolyte to suppress the electrostatic component of the disjoining pressure, as well as films from non-aqueous solutions (see also sec. 6.5). During thinning, the thermodynamically unstable films retain their shapes over a large thickness range, until a rather small thickness is approached, at which the film ruptures. This thickness is the critical thickness of rupture h cr , mentioned before. Therefore, the thermodynamic instability is a necessary condition for film instability, but mechanistic arguments are required to describe the rupture process. Two steps in the process of film rupture can be recognized: (i) film thinning with retention of film shape and, (ii) film rupture. Which of these is more important, requires analysis of the various mechanisms of film rupture. Contemporary understanding of liquid-film rupture is based on the existence of thermal fluctuations on liquid surfaces1 . These fluctuations are inherent in the equilibrium of liquid interfaces, see sec. III. 1.10. For thick films, the fluctuations in the two film surfaces are independent of each other, but upon thinning they start to correlate. This interference under conditions become destructive, i.e., waves appear, whose amplitudes increase with time. Rupture occurs at the moment when the amplitude Ah , or rather the root mean square value ((Ah)2)1/2of a certain unstable wave grows until it becomes of the order of the critical film thickness <(Ah)2>I/2 = O(hcr)
[6.4.7]
The basis of this model, i.e., the condition of increase in the amplitude of a spontaneous wave, is equivalent to the condition of increase in local pressure 23 ', §(Apc + n(h)) > 0
[6.4.8]
In this expression, <SApc = -yk2Ah is the capillary pressure corresponding to a wavelength X = 2n/k, and amplitude Ah; <5/7(h) = (d/7(h)/dh)Ah is the perturbation of the disjoining pressure /7(h). Hence, equation [6.4.8] yields 11
L. Mandelstam, Ann. Physik 41 (1913) 609. A. Scheludko, Adv. Colloid Interface Set 1 (1967)391. 31 A. Scheludko, Proc. Konlnkl. Nederl. Akad. Wetenschap B65 (1962) 87. 21
6.44
THIN LIQUID FILMS
k2 < [d/7(h)/dh]/;r = kc2r
[6.4.9]
The upper limit of the range of unstable wave spectra kCT is also known as the Sheludko wave number. In thick films (h > 0.5 |im), only capillary forces act against surface deformations; <5Apc » SI7{h)), and fluctuation waves are practically stable over the entire range of wavelength. Moreover, the steady state amplitudes of the capillary waves, determined from
the equipartition 1
50 mN m " ; kT = 4 • 10"
law 21
((Ah) 2 ) 1/2 = *JkT/ y
under typical conditions
(y~
J ) have values of the order of <(Ah)2}1/2 < 1 nm , i.e., thick
films are not only stable but remain practically unaffected by thermal fluctuations. In the process of film thinning the interactions due to surface forces in the film become stronger, and negative values of the disjoining pressure have a destabilizing effect (increase of amplitude). When attractive Van der Waals forces prevail across the film, [6.4.9] and [6.4.6] give 1 ' for k cr Jo \r
Once formed, the unstable waves grow until one of them (the fastest) conforms to [6.4.7] and then the film ruptures. During this time, the film thins additionally, at a rate depending on the conditions under which it is produced. This kinetic part of the theory of the critical thickness of rupture has been formulated independently and partially solved by Vrij 2 '. An important feature of the kinetics of film rupture is the random character of the process. Here, the question is about the correct description of the effect of fluctuations on the evolution of single waves. The further development of the theory gives an expression which seems suitable from the experimental point of view3' h
=:
:
Vdw
(64111
Experimental data are in good agreement with [6.4.11]. The K v d w value recalculated from the experimental data for ha is very close to that calculated theoretically according to the Lifshits theory. Some corrections have been introduced later in the theory of film rupture. Another approach to the rupture of thin liquid films is based on stochastic modeling of this critical transition 4 '. Autocorrelation functions for steady state and for thinning liquid films are formulated. A method for the calculation of the lifetime r , and the critical thickness hcr of films is introduced. It accounts for the effect of the spatial 11
A. Scheludko (1962), loc.cit. A. Vrij, Faraday Discussion Chem. Soc. 42 (1966) 23. 31 B. Radoev, A. Scheludko, and E. Manev, loc. cit. 41 R. Tsekov, B. Radoev, Adv. Coll. Interface Sci. 38 (1992) 353. 21
THIN LigUID FILMS
6.45
correlation of waves. The existence of non-correlated sub-domains leads to a decrease in r, and increase in h cr , as a result of the increasing possibility for film rupture. The coupling of dynamics of surface waves, and the rate of drainage v leading to stabilization of thinning films, is also accounted for1'. The mechanism of film rupture proposed by Scheludko and Vrij has stimulated experimental work for the determination of ficr . Most successful are studies which employ the method of microscopic circular thin liquid films with radii r ~ 100 [im (sec. 6.2c). Since rupture is a process with a stochastic character, reliable measurements of h cr are possible only with microscopic films, from which non-thermal fluctuations are more easily suppressed. The probability character of rupture is illustrated by the curves in fig. 6.24. It can be seen that the most probable critical thickness of film rupture increases with the increase in film radius, its value being ca. 30 nm at r = 0.1 mm.
Figure 6.24. Distribution of curves of the critical rupture thickness h cr for foam films made of isovalerianic acid + aqueous KC1 solution; each curve is for a film with different radius: curve 1, r = 98 [im ; curve 2, r = 138 ^m ; curve 3, r — 197 |im ; curve 4, r = 295 |xm ; curve 5, r = 394 |^m ; curve 6, r = 492 urn ; N is the total number of critical thicknesses measured; AJV is the number of ruptured films with thickness between h and h + Aft . (Redrawn from E. Manev, A. Scheludko and D. Exerowa, Colloid Polymer Set, 252 (1974) 586.)
6.4d Jump-like formation of a black film in a thin liquid film It was shown in the previous section that during thinning, thin liquid films reach a certain critical thickness h cr at which they lose their stability. There are two possibilities: either the film ruptures and disappears completely, or a jump-like local thinning occurs in the film. When it has a very small thickness a foam or emulsion film looks black because it does not visibly reflect light, hence, the local patches formed in a jump-like process are called black spots. Black spots are very small, round areas of black films, which are much thinner than the surrounding area of the unstable thin liquid film. Figure 6.25 shows photographs of black spots at different stages of their development. 11
A. Sharma, E. Ruckenstein, Langmuir 3 (1987) 760.
6.46
THIN LIQUID FILMS
Figure 6.25. Stages of formation and growth of black spots in an isolated gray film: (a) schematic presentation of a black spot in the thicker film; (b), (c), photographs showing the growth of black spots; (d) the whole film area is black.
It appears that the Scheludko-Vrij theory for the rupture of thin liquid films, and the further concepts introduced on that basis (see the previous section), are applicable not only to the process of rupture by local thinning, but also to the formation of black spots. Hence, black spots can be used to monitor the mechanism of local flexion in the film, which allows one to estimate roughly the fluctuation wavelength (A/2 is ca. 1 |im ). Such a general treatment of instability, including the formation of black spots, can be employed as an additional tool to verify the theory of rupture. Another important result is that the rupture of unstable films and the formation of black spots occur at the same critical thickness. The value of h cr is ca. 30 nm for films from aqueous surfactant solutions. Foam and emulsion films look gray at this thickness. Figure 6.26 shows the dependence of the most probable values of h cr (of rupture) and h c r M at which black spots appear in the gray film. The scatter of the h cr
THIN LIQUID FILMS
6.47
values is + 0.2 nm. Thus, the most probable value of h cr or h c r b l can be determined with an accuracy of ca. 0.5 nm. The experimental points drawn in fig. 6.26 reflect the smooth transition from the region of rupture to the region of black spot formation, as well as the independence from h cr on the final state of rupture or black spot formation.
Figure 6.26. Dependence of the most probable critical thickness hCT corresponding to (I), film rupture or, (II), to the formation of black spots upon the surfactant concentration c , in the solution. Foam films from NP20 + aqueous KC1 solution: curve 1, r = 5 5 u m ; curve 2, r = 104p.m; curve 3, r = 200 (im . (Redrawn from E. Manev, A. Scheludko and D. Exerowa, Colloid Polymer Set, 252 (1974)586.)
What will happen in a thinning liquid film when it loses its stability, film rupture at hcr , or black spot formation, depends on the kind of surfactant and its concentration in the solution. On that basis, a new parameter has been introduced1' for quantitative characterization of the surfactants as foam or emulsion stabilizers. This is the minimum bulk surfactant concentration at which black spots begin to form in the unstable, thin liquid film. It is called the concentration of black spots formation c b ,. This concentration c bl is a very important characteristic of foaming agents and emulsifiers. It can be determined microscopically in reflected light. Films are prepared from surfactant solutions with different, gradually increasing concentrations, keeping all other parameters constant. The lowest surfactant concentration at which black spots first appear is cb) . The concentrations c bl at which various kinds of surfactants ensure the formation of films that are stable towards rupture can depend strongly on the nature of the surfactant molecule. Within homologous series, the c bl values decrease with increaseing chain length of the molecule. Furthermore, the value of c bl is a function of the temperature, electrolyte content, pH (for ionic surfactants), and the presence of admixtures in the solution. For emulsion films the nature of the second liquid phase also influences c b l . That is why the measurement of c bl should be standardized. The comparison of the concentrations c bl corresponding to the formation of black spots in emulsion and foam films, obtained from solutions of one and the same surfactant (which acts both as a foaming agent and as an emulsifier), indicate that c bl for " D. Exerowa, A. Scheludko, in Chemistry, Physics and Application of Surface Active Substances, Vol. 2, J.Th.G. Overbeek, Ed., Gordon & Breach, London (1964) p. 1097.
6.48
THIN LIQUID FILMS
foam films is considerably lower than c b | for emulsion films. This means that stable foam films (usually black films) can be obtained at lower surfactant concentrations than stable emulsion films, even from a non-polar organic phase. With increasing polarity of the molecules of the organic phase c bl for emulsion aqueous films increases1', which is analogous to the increase in c bl for hydrocarbon emulsion films. In fact, in real emulsions, black films are rare or absent. After the jump-like formation of the black spots, they expand, merge, and finally the entire area is occupied by black film (fig. 6.25). The kinetics of the expansion of the black spots in the gray film have also been studied. The process of expansion in an emulsion film is quite similar to that in a foam film. At low electrolyte concentrations the black spots in both types of films expand slowly, at high electrolyte concentrations the process is very fast (within a second or less), and ends with the formation of a black film with a large contact angle film/bulk liquid phase (meniscus). In the process of transformation of the black spots into a black film, the emulsion film is very sensitive to any external effects (vibrations, temperature variations, etc.) this is in contrast to the black foam film. The characteristic concentration c bl which is specific for each surfactant parameter, is also important for foam, and emulsion, stability (see chapters 7 and 8, respectively). 6.4e Film thickness upon extraction Thin films can be extracted from bulk liquids, for instance by submerging a thin frame and pulling that slowly out of the solution. This case is rather typical for macroscopic films. Suppose the film is vertical, i.e. it is extracted against gravity. When the rate of pull-out of the frame uf is constant, a stationary state develops, in which the thickness h is a compromise between the amount entrained by the frame and the amount that flew down by gravity and the suction of the plateau border between film and solution. The issue is of practical relevance. In the first place, it plays a role in the process of marginal regeneration, discovered by Mysels2' which plays a dominant role in the thinning of macroscopic films, and which hardly occurs in the microscopic films discussed before in this section. Except for rigid films, having a very high surface viscosity, thinning takes place by the suction of thicker film parts into the border, in exchange for thinner parts. Most macroscopic films are mobile, as can be verified by the changing colour patterns and hence display marginal regeneration at the Plateau border, visible as turbulent motions. To quantify this process one should know both the rate of extraction and the rate of suction. Because of this argument, Frankel21 set
11
H. Sonntag, J.Netzel, and H. Klare, Kolloid-Z. Z. Polym. 211 (1966) 121. K.J. Mysels, K. Shinoda, and S. Frankel, Soap Films: Studies of their Thinning, Pergamon Press (1959). 21
THIN LIQUID FILMS
6.49
out to compute h(uf) and found 2/3 2/3
h = 1.88
I yl/bql/2 pl/2
[6.4.11]
where r\ is the liquid viscosity and where y, g and p enter the equation because they account for the influences of gravity and capillary suction. It is of historical interest that, independent of Frankel, a similar equation has been derived by Deryagin and Lev!1', which did not immediately draw interest in the West because it was published in a Russian Journal. Deryagin and Levi addressed another practical application, viz. that of the thickness of a film entrained on a (wetting) drum, rotating in the surface of a fluid, out of which a thin film is to be made, say a photographic film. By a later comparison of the two theories it became clear that they were based on the same fluid mechanics tools. Equation [6.4.11] can only be studied for rigid films. It was found to apply exemplarily2'. Hence it could be inferred that r\ retains its bulk value for the fluid in the heart of the film except for a layer on the two surfactant layers on the film surfaces that is thinner than a few cross-sections of a water molecule. This finding supports using of the bulk viscosity in the Reynolds equation [6.4.2 or 3] and is in line with the present-day insight into the thickness of the stagnant layer in electrokinetics. 6.5 Surface forces in symmetric thin liquid films According to the thermodynamics of the thin liquid films, the most important factor which determines their properties is the interaction between the two film interfaces, i.e., the interaction between the two adjacent phases across the liquid film (see chapters 1.4 and IV.3). A very important thermodynamic parameter is the disjoining pressure, which is the sum result of the interactions arising from different types of surface forces acting across thin liquid films. The dependence of the disjoining pressure on the film thickness, the so-called disjoining pressure isotherm fl{h) is closely related to the more general problem of the stability of a disperse system. It is not surprising, therefore, that the knowledge of the forces in thin liquid films has developed parallel to the theories for colloid stability, the first of these being the DLVO theory. Major efforts have been directed to understanding the nature and origin of surface forces acting in thin liquid films, including bilayer films. Moreover, explanations are sought for several cases in which classical DLVO theory is unable to account for all experimental findings. Two categories of surface forces are usually distinguished: DLVO, and non-DLVO surface forces. The Van der Waals molecular interactions or dis11 B.V. Deryagin, S.M. Levi, Izv. Akad. Nauk. SSR (1959); B.V. Deryagin, Zhur. Eksp. Teor. Fiz. 15 (1995) 9. 2) J. Lyklema, P.C. Scholten, and K.J. Mysels, J. Phys. Chem. 69 (1965) 116.
6.50
THIN LIQUID FILMS
persion molecular forces, as well as the electrostatic or double layer forces are called DLVO-forces (both are long range forces), the balance of which provides the foundation of the DLVO theory. Non-DLVO forces are of a different nature. Most important for symmetric thin liquid films with fluid interfaces are the steric surface forces, which are long range for macromolecules, but shorter range for smaller molecules such as surfactants. Theoretical, as well as experimental consideration has been given for the existence of solvent structure-mediated forces, as expounded in sec. III.5.3a. In the extended DLVO theory, DLVOE, these are explicitly incorporated, see sec. IV.3.9. In specific cases, other forces could also operate. The precise direct measurement of surface forces is a subject of current interest. It should provide a reliable distinction between the forces, along with elucidation of their mutual influence, and their dependence on h in systems of different composition, temperature, etc. This enables a more critical evaluation of the familiar theories, but also stimulates theoretical analysis. Surface forces measurements in foam and emulsion films and particularly, in microscopic films stabilized with amphiphilic molecules (surfactants, phospholipids, polymers) enable the study of these forces from large h (long range surface forces) down to those at bilayer contact. The binding energy of an amphiphilic molecule in the bilayer is dominated by interactions between the nearest-neighbour molecules in the lateral and normal directions of the film, i.e., by short range forces. A new approach will be put forward in the evaluation of the binding energy (sec. 6.6). Measurements with microscopic foam films have also permitted a detailed study of the transition between long and short range interactions and their reversibility. 6.5a DLVO forces The dependence of the disjoining pressure on film thickness, the /7(h) isotherm, is at the root of the DLVO theory of stability of lyophobic colloids. According to this theory, the disjoining pressure in thin liquid films is considered as the sum of an electrostatic component, 77el (see chapter IV.3), and a Van der Waals component, / 7 v d w (see chapter 1.4) /7(h) = /7el(h) + /7 vdw (h)
[6.5.1]
A complete analysis of the theory can be found in several monographs, e.g. 123 '. The theories for both components of the disjoining pressure are considered in detail in the two FICS chapters cited. Here, we repeat only a few expressions which are used for the interpretation of experimental results. The electrostatic contribution to the disjoining
11
E.J.W. Verwey, J.Th.G. Overbeek, Theory of Stability of Lyophobic Colloids, Elsevier (1948). B.V. Derjaguin, N.V. Churaev, and V.M. Muller, Surface Forces, Consultants Bureau N.Y. (1987). 31 J.N. Israelachvili, Intermolecular and Surface Forces, Academic Press, 2nd ed. (1991). 2)
THIN LIQUID FILMS
6.51
pressure is given by [IV.3.3.21] /7el(h) = 2cRT(coshzy m -l) with m = h/2
[6.5.2]
and ym = Fy/m / RT . Here, c is the concentration of a symmetric
electrolyte of valence z, F is Faraday's constant; \//m is the midway potential, i.e. the potential at a distance hi2 which is related to the potential y/^ of the diffuse part of the electric double layer at the film/adjacent phase interface through the relationship [IV.3.3.6]
_ ^ 2
U= =
f
^
[6.5.3]
y=y d^2[cosh(zy)-cosh(zy
... Fy/ with y =——, RT
Fy/1 d y a =—-— RT
m
)]
and
K = ^JF2z2c/e0eRT
[6.5.4]
where e is the relative dielectric permittivity, and eQ = 8.854xlO" 12 C 2 V" 1 m" 1 is the dielectric permittivity of free space (vacuum). An approximated expression (see [IV.3.3.27]) of the electrostatic component of the disjoining pressure can be derived from [6.5.2] at small values of y/m nel(h) = ^ ^ I [ t a n h ( z y d /4)] 2 e~Kh
[6.5.5]
Equation [6.5.5] is quite adequate for rapid calculations of 77e]. However, for a wide range of yfi and c values, it is necessary to use the more general equations [6.5.2] and [6.5.3] in combination. In figs. IV.3.7 and 9 we indicated how close the approximate solution is to the exact one. These figures apply to plots of the Gibbs energy per unit area, from which FI(h) is found as the slope. Equation [6.5.5] applies for the case of constant y/1. Appendix IV.2 contains a collection of more rigorous equations. Two theories, macroscopic and microscopic, are involved in the calculation of the Van der Waals component /7 V d w of the disjoining pressure in thin liquid films (chapter 1.4). According to both theories (sec. 1.4.6), the total interaction force in a flat gap between two semi-infinite phases decreases with distance much more slowly than does the interaction force between two individual molecules. The following general expression is obtained for the Van der Waals component of the disjoining pressure in a symmetric film bordering gas- or condensed phases /7
Vdw = - K v d w / h n
I6-5-6!
where K v d w is the (Van der Waals-) Hamaker constant (the name was introduced by
6.52
THIN LIQUID FILMS
A. Scheludko), Kvdw{h) = A(h)/6n, with A{h) being the Hamaker constant. It has mostly a weak dependence on film thickness. For small ft , n = 3 ; however, at larger ft a correction for electromagnetic retardation has to be taken into account, ultimately leading to n = 4, according to Casimir and Polder (sec. 1.4.6). A general formula for the calculation of the dispersion force between any pair of condensed phases has been derived in the macroscopic theory (sec. 1.4.7). According to this theory, the attraction between two macrobodies results from the overlap of the fluctuating electromagnetic fields of the two substances. If this field is known for a thin liquid film, it is possible to determine the disjoining pressure in it (see [1.4.7.1][1.4.7.9]). The stricter macroscopic theory avoids the approximations assumed in the microscopic theory, viz., the additivity of pair interactions; the integration of pair interactions instead of summation, and the extrapolation of interactions between individual molecules in a gas to interactions in condensed phases. The macroscopic theory has been used successfully for the calculation of the dispersion interactions, i.e., of /7 vdw (h) curves. Several approximations have been derived, since the exact equations are too complicated for everyday application. A good approximation for symmetric flat films is (see [1.4.7.11]) DO
n
h
/
\2
+x+1 x da
y^ >-^h!lT
with x = 2cohey2 ; ff
}- ^
16571
--
and £ are the relative dielectric permittivities (functions of the
frequency co) of the film and the adjacent phases, respectively; h is Planck's constant divided by 2n. For small film thickness, a simpler approximation can be derived
Another approximation for a foam film having a relatively large thickness reads „
he
nVAYu=—. vdw
4
K2 ;
(\-e(O]f
( \ \ Of
„_
m
[6.5.9
h 240Vi(0)ll + £(0)J ^U(O)J
where e(0) is the static value of the dielectric permittivity of the film, and c is the speed of light in a vacuum. Here, /7 V d w is inversely proportional to h4, as in the Casimir and Polder formula which accounts for the electromagnetic retardation of the dispersion forces1'. For practical use of the macroscopic theory, equations with empirical constants 2 ' are also used; sometimes for example
11 21
H.B.G. Casimir, D. Polder, Phys. Rev. 73 (1948) 360. W.A. Dormers, Ph.D Thesis, University of Utrecht (1976).
THIN LIQUID FILMS
6.53
where a, b, c, d, and e are empirical constants. Quantitative experimental verification of the DVLO theory has been performed with macroscopic and microscopic foam films. As shown above, the disjoining pressure is given as the sum of /7el and /7 V d w , i.e., they are considered to be additive. In a symmetric foam or emulsion film, /7el is always positive, while / 7 v d w is always negative. In an equilibrium film, whose state is determined by the action of positive disjoining pressure, /7(h) can be calculated from the balance of the forces obtained experimentally. When /7(h) is negative it is possible to use the dynamic method according to which IJ(h) is calculated using the hydrodynamic equation for the rate of film thinning (sec. 6.4). This method enables studying /7(h) isotherms over a large thickness range. It also provides the possibility of verifying the different theories of disjoining pressure, including the cases in which I7(h) is always negative and no equilibrium films are formed. Unique measurements of disjoining pressure have also been performed with nonaqueous films1'2'. In such films the disjoining pressure is caused by Van der Waals forces only, and these can be studied without interference from electrostatic repulsion or repulsion forces due to short range interactions. Direct experimental evidence for the existence of an attractive Van der Waals disjoining pressure has been obtained from studies of aniline, benzene, and chlorobenzene films. For these substances, Hamaker constants of the order of 10~ 19 J have been evaluated, and the electromagnetic retardation of dispersion forces was also established. The length of London's wave has also been calculated; for benzene it is 64 nm, for chlorobenzene, 81 nm, and for aniline, 80 nm. The first quantitative proof of the DLVO theory for thin films has been conducted with a model system a microscopic foam film, using the micro-interferometric technique (sec. 6.2). Independent studies have been performed of n^(h) and /7 vdw (h), as well as of their joint action at various concentrations c of electrolyte. At very low c, equilibrium films of large thickness were formed, in which the electrostatic interaction prevails so that their behaviour could be described completely by this interaction. At such a film thickness, / 7 v d w is still very low, so the equilibrium film state was reached from the balance between the electrostatic disjoining pressure and the capillary pressure, /7e|(h) = Ap . Non-equilibrium thinning films were formed at high electrolyte concentration, with a prevailing contribution of the negative nVA^j(h) . At intermediate electrolyte concentrations (10~2- 10~3 mol dm" 3 ), /7el(h) and /7Vdw(h) are competitive. At large film thickness, accelerated drainage was observed initially, after which 11
A. Scheludko, D. Platikanov, and E. Manev, Faraday Discuss. Chem. Soc. 40 (1965) 253. B.V. Derjaguin, A.S. Titijevskaya, Proc. 2nd Intern. Congr. Surface Activity, Vol. 1, J.H. Schulman, Ed., Butterworths (1957) p. 211. 21
6.54
THIN LIQUID FILMS
Figure 6.27. Disjoining pressure vs. film thickness; 5xlCP 4 wt % saponin + 10~ 2 m o l dm" 3 aqueous KC1 solution; solid lines calculated from [6.5.5] and [6.5.6] at (/*=90mV and T = 25°C; curve 1, K v d w = 3.3xlO~ 21 J ; curve 2, K v d w =5.1xlCT 2 1 J ; curve 3, K v d w = 8 . 5 x l O " 2 1 J ; the point marked with ' • ' corresponds to the equilibrium film thickness. (Redrawn from A. Scheludko and D. Exerowa, Kolloid Z., 168 (1960) 24.)
the process was delayed until equilibrium was established. This delay was due to the rise of /7el(h) with decreasing thickness. Figure 6.27 plots the /7(h) isotherm, obtained by the dynamic method, for aqueous saponin films. The experimental point marked by a filled circle1' corresponds to an equilibrium film at /7(h) = Ap = 73 Pa . Despite the considerable scatter, a clearly expressed minimum can be seen, similar to the dependence calculated from [6.5.1], [6.5.5], and [6.5.6] with ( / i = 9 0 m V
and
n = 3 . The value of the Van der Waals-Hamaker constant has been determined to be K v d w = 4xlO~ 21 J ; however, one must note the dependence of K v d w on the film thickness, predicted by Lifshits theory and calculated by some authors using approximate methods. DLVO surface forces have also been studied using vertical macroscopic equilibrium foam films (sec. 6.2) from aqueous sodium octylsulphate containing KC1. The film thickness covered a large range from 80 to 8 nm, at electrolyte concentrations from 10"3 to 1 mol dm" 1 (fig. 6.28). A smooth fall of the equilibrium thickness is seen clearly as the counterion concentration is increased. By studying films of small thickness the authors considered, for the first time, the film's structure, i.e., the presence h, nm 100 -
^
\ °^s.
Figure 6.28. o*»o Ovo ^^s«»o°
I
i10" u
|
„ 4
I
Q 3
i10" u
I
r> 2 110" U
i ID" u
|
, 1
C, mol dm"
11
L1
Equilibrium thickness of a
macroscopic film vs. electrolyte concentration; sodium octylsulphate + aqueous LiCl solution; solid line, calculated from [6.5.5] and [6.5.6] at y/d = 65 mV and Q.
Kirju; = 3.21 x 10~ zl J . [Redrawn from J. Lyklema and K.J. Mysels, J. Am. Chem. Soc, 87 (1965) 2539.])
A. Scheludko, D. Exerowa, Kolloid. Z. 165 (1959) 148.
THIN LIQUID FILMS
6.55
of an aqueous core and adsorption layers (sec. 6.6). This is important for the treatment of the results from optical thickness measurements, and also for the estimation of /7el(h) and /7 v d w (h). On the basis of the above experimental data, and other experimental facts, one can conclude that deviations from DLVO theory commence at foam film thicknesses below 20 nm (sec. 6.6). 6.5b Potential of the diffuse double layer at the aqueous solution-air interface Under certain conditions, aqueous electrolyte solutions form equilibrium thin liquid films. Their equilibrium thickness is determined by the positive electrostatic component of the disjoining pressure 17 (h) which depends on the potential y/^ of the diffuse double layer at the liquid film-air (gas) interface. The relationship between the film thickness h and electrolyte concentration c obeys DLVO theory. The potential y/1 at the solution-air interface can be calculated from the measured equilibrium film thickness h and the known c value. This is a method1' for determining yfi and hence, for studying the electric properties at such surfaces. The measurements extend those described in sec. III.4.4b. The method is valuable, since an equilibrium potential and diffuse charge density cr"1 can be evaluated, avoiding all the complications that occur with electrokinetic measurements. The equilibrium film method also allows the determination of yA at electrolyte solution-air interfaces without any surfactant, finding the origin of the surface charge in this case, finding the isoelectric points at these interfaces, studying the effect of the concentration of various kinds of surfactants, the nature of the ions, and the (/*(h) dependence, etc. 2 ' 3 ' Detailed studies with the equilibrium film method for the h(c) - and h[pH) dependence in the absence of a surfactant, as well as h{cs) at very low surfactant concentrations c s , gave yf1 ~ 30 mV at the interface of aqueous electrolyte (HC1 + KC1) solution-air. Without further information, the sign of yA cannot be established. At low electrolyte concentrations, yA increases with increasing c s , from its value for an aqueous electrolyte solution without surfactant, to a constant value typical for each kind of surfactant4'. The increase of yA at low surfactant concentrations c s is connected with its adsorption at the solution-air interface. The diffuse charge density (f* reaches a constant value when, or just before, the adsorption layer becomes saturated. Values for yA at the aqueous surfactant solution-air interface can be used to calculate the diffuse charge density cfi according to [IV.3.3.13] 11
D. Exerowa, Kolloid.-Z. 232 (1969) 703. D. Exerowa, loc. cit. 31 D. Exerowa, T. Kolarov, and Khr. Khristov, Colloids Surf. 22 (1987) 171. 41 D. Exerowa, M. Zacharieva, R. Cohen, and D. Platikanov, Colloid Polym. Set 257 (1979) 1089. 21
6.56
THIN LIQUID FILMS
( -2
£ £CRTS^^
1/2
(cosh y d - cosh y m ) [6.5.11] n ) ' Although DLVO calculations do not provide the signs of yA and &1, experiments may indicate which ions are adsorbed at the interface, and so it is possible also to determine the signs. For non-ionic surfactants, XJA and cA are considerably lower than for ionic surfactants, but they are non-zero, and slightly higher than yA at the aqueous electrolyte solution-air interface. For example, t/A = 39 mV for decyl methyl sulphoxide. For ionic surfactants such as Na-dodecylsulphate, yA is higher (yA = -82 mV ). The i/A "values are in good agreement with the data for the f-potential obtained through electrophoresis of bubbles1'. The very weak i/A(h} dependence has indicated that these yA values can be accepted for the surface of bulk liquids as well. At the aqueous electrolyte solution-air interface, without surfactant, t/A has been measured very precisely with low electrolyte concentrations and extremely pure solutions and vessels. Figure 6.29 presents the yA[pH) dependence for aqueous solutions at constant ionic strength (HC1 + KC1). It can be seen that at pH > 5.5, the potential becomes nearly constant around -30 mV. At pH < 5.5, the potential decreases sharply and becomes zero at pH = 4.5, i.e., an isoelectric state is reached at the solution surface. This observation indicates that the charge at the surface of the aqueous solutions is caused by preferential adsorption of OH~ ions over H+ ions. Experiments for non-ionic surfactants show that yA also depends on pH, and that there also exists an isoelectric point which depends on the surfactant concentration. The i.e.p.'s obtained (accuracy ±0.1) are within the acidic range of pH and differ between the various non-ionic surfactants. Hence, these i.e.p.'s can be used as parameters characterizing the specific ion binding properties of all kinds of surfactants and small ions at the solution-air interface2 . It has been shown that, in general, interfaces acquire charges that depend on the composition of the aqueous solution. Both in pure water, and in the presence of indif-
Figure 6.29. Dependence on pH of yA at the surface of aqueous HC1 + KC1 solution-gas phase; ionic strength of the solution 10~4 mol dm"^ (Redrawn from D. Exerowa, Kolloid Z., 232 (1969) 703). 11 2)
R.W. Huddelston, A.L. Smith, in Foams, A.J. Akers, Ed., Academic Press (1976) p. 163. Th. van den Boomgaard, J. Lyklema, Langmuir 5 (1989) 245.
THIN LIQUID FILMS
6.57
ferent electrolytes, the air-water interface is negatively charged. The same applies to aqueous solutions of non-ionic surfactants, while in the case of cationic and anionic surfactants the interface exhibits positive and negative charges, respectively. A charge reversal at the air-water interface, owing to increased adsorption of a cationic surfactant with increasing surfactant concentration, has also been observed1'. 6.5c Steric surf ace forces The surface forces considered in the previous two sections (Van der Waals and electrostatic forces across diffuse double layers at fixed y/1) are the DLVO surfaceforces. However, foams and emulsions practically always contain surfactants as the stabilizing agents. Hence, for thin liquid films with fluid interfaces (foam or emulsion films) the most important additional forces are the, steric interactions short range for small molecules (to be discussed in sec. 6.6), and long range for macromolecules, whose action in thin films will now be discussed. For an extensive discussion of the principles and elaborations, see chapter 1. Microscopic foam films have been used to study the steric interaction between two liquid-gas interfaces. Two ABA triblock copolymers have been employed: P85 (PEO27PPO39PEO27, M = 4600 Da) and F108 (PEO122PPO56PEO122, M = 14000 Da). Blocks A are hydrophilic polyethylene oxide (PEO) chains, while block B is a hydrophobic polypropylene oxide (PPO) chain. The film thickness has been measured by the micro-interferometric method (sec. 6.2). Figure 6.30 presents the dependence of the equivalent film thickness h w on the NaCl concentration c for aqueous solutions of the two copolymers. The h w values initially decrease with increasing c, and then level off to a constant value. The plateau starts at a critical electrolyte concentration c cr = 3xlO" 3 mol dm~3 NaCl for F108, and c cr = 3xlO~ 2 mol dm" 3 NaCl for P85, similar
Figure 6.30. Equivalent film thickness of films stabilized by non-ionic block copolymers P85 and F108 (described in the text), as a function of the NaCl concentration; aqueous solutions: •, 7x lCr 5 moldm~ 3 P85 and o, 7 x l ( r 7 m o l d m - 3 F108; T = 23°C; the arrows show c c r . (Redrawn from D. Exerowa, R. Sedev, R. Ivanova, T. Kolarov, and Th.F. Tadros, Colloids and Surfaces, 123 (1997) 277.)
11 T. Kolarov, R. Yankov, N.E. Esipova, D. Exerowa, and Z.M. Zorin, Colloid Polym. Sci. 271 (1993)519.
6.58
THIN LIQUID FILMS
to low molecular mass surfactants. The bulk properties of aqueous solutions of PEOPPO-PEO copolymers are essentially similar to these of non-ionic surfactants. It is very likely that the interaction behaviour is determined by the hydrophilic PEO chains protruding into the solution. Although the stabilizing blocks are non-ionic, the left branch of the h w (c)- dependence is dominated by electrostatic repulsion. Charge creation must be attributed to preferential anion adsorption, for example of OH~ ions, at the air-water interface (sec. 6.5b). The trends are similar for both copolymers. In contrast, the plateau values for the two copolymers are very different, at 39 + 3 nm and 17 ± 1 nm, for F108 and P85, respectively. Since the copolymer having the higher M gives thicker films, a surface force component of steric origin may be inferred. However, the thickness hw
is an
effective parameter which is too crude for quantitative interpretations. As a reasonable compromise between physical relevance and tractability, the three layer model is adopted: the foam film is viewed as having a symmetric sandwich structure two adsorption layers symmetrically confining an aqueous core (fig. 6.2). The equivalent film thickness h w is larger than the total film thickness h since the refractive index of the adsorption layer is higher than that of water (r^ > n 2 ) . Allegedly, the electrolyte concentration affects only the aqueous core thickness, h2, which can be calculated using [6.2.2]. The thickness
h^ of the adsorption layer is determined by the
conformation of the macromolecules at the interface. The radius of gyration a
of a
1
flexible neutral chain in a good solvent is given by the Flory relationship ' a
=aN3/5
[6.5.12]
where a (— 0.2 nm) is an effective monomer size, and the number JV = JVpo + 2iVEO . If a macromolecule adsorbs at the interface as a separate coil, it should occupy an area of the order of the projected area of the molecule in the bulk solution, i.e., a 2 . The figures obtained (ca. 2 nm 2 ) are much lower and therefore the PEO chains are crowding the solution-air interface and are stretched, i.e., they form a brush. Its thickness, hj, can be calculated from the simple brush model2' h1=aJVEO(a2/am)1/3
[6.5.13]
a m being the area per molecule. For a detailed treatment of brushes, see sec. 1.11. The most significant finding is that the plateau thickness values at high electrolyte concentration are larger than twice the adsorption layer thickness: i.e., h > 2h1 (Table 6.1). This is rather unexpected, since above c cr electrostatic repulsion is suppressed, and only steric interaction is expected to stabilize the film. If so, a total thickness close to the double brush thickness, i.e. h ~2h , would be expected. Both h and h{ are different between the two polymers. However, the thinnest films from both copolymers
11 21
P. Flory, Principles of Polymer Chemistry, Cornell University Press (1956). P.G. de Gennes, Macromolecules 13 (1980) 1069.
THIN LIQUID FILMS
6.59
have the same structure two brush layers, and an aqueous core of thickness ~ 3aff . The equilibrium thickness of the thinnest films is probably due to the same type of steric stabilization, which is different from the brush to brush repulsion. Table 6.1. Thickness of the layers in the three layer film model. Copolymer
h, nm
2hj , nm
h 2 , nm
h
P85
15.1 37.9
6.4
8.5
21.2
16.7
-3.4 -3.1
F108
2/ag
Rather similar results have been reported for foam films from aqueous solutions of polyvinyl alcohol1'. Their equilibrium thickness at low pressure is much larger than twice the adsorption layer thickness measured by ellipsometry. The core thickness is again several times the radius of gyration of the polymer molecules in bulk solution. The explanation given by the authors is that the upper limit of interaction is governed by a few, essentially isolated, polymer tails. Similar arguments can also be given for the case of an ABA block copolymer. In other words, the film thickness at high electrolyte concentration is governed by the longest, and not by the average, PEO chains. A chain longer than the average brush thickness will behave as a brush only up to hy, but rather as a mushroom (swollen coil) further into the solution. Thus, the following qualitative picture is arrived at. Above c cr , at lower pressure, the mushroom tails of the longer chains interact rather softly; at higher pressure, a brush to brush contact is realized and steeper repulsion is expected. This behaviour is confirmed experimentally. At lower c, electrostatic repulsion dominates, and decreases with c until steric repulsion (which is independent of c ) becomes operative at c = c cr . The transition from electrostatic to steric repulsion occurs at a value of c cr given by "cr =\
%
[6.5.14]
where KC^ may be interpreted as the critical Debye length. The most detailed information about the steric surface forces can be obtained from the disjoining pressure vs. thickness isotherm for films from aqueous polymer solutions with c > c c r . A disjoining pressure range encompassing four orders of magnitude (up to 5xlO 4 Pa) has been monitored by two complementary techniques: the dynamic method, and the thin liquid film pressure balance technique (sec. 6.2). Figure 6.31 shows the [7{h) isotherm for F108 solutions. The two cases discussed above can be clearly distinguished: (i) at low pressures, FI(h) < 102 Pa, the 'soft' steric repulsion determines quite large film thicknesses of about 40-50 nm; (ii) at high pressures, /7(h) > 102 Pa, a strong steric repulsion, owing to brush-to-brush contacts, determines considerably smaller thicknesses. 11
J. Lyklcma, T. van Vlict, Faraday Discuss. Chem. Soc. 65 (1978) 25.
THIN LIQUID FILMS
6.60
Figure 6.31. Disjoining pressure /7(hw) isotherms for; aqueous copolymer solutions, T = 23°C; porous plate method: D , 7 X 10" 6 mol dm" 3 F108 + 0.05 mol dm" 3 NaCl; O, 1.43xlO~5 moldm" 3 F108 + 0.05 mol dm~ 3 NaCl; dynamic method: . , 10- 5 mol dm" 3 F108 + 0.1 mol dm~ 3 NaCl; the solid line is the best fit of [6.5.15]. (Redrawn from R. Sedev and D. Exerowa, Adv. Colloid Interface Set, 83 (1999) 111.)
Presumably, the PEO brushes extending from the two surfaces come into contact and repel each other. Under these conditions the de Gennes scaling theory for interaction between two surfaces carrying polymer brushes applies. Accordingly the steric component of disjoining pressure is given by
" D3|Ul.J
(21,,) f
where h 11h^ is the dimensionless film thickness (hj is the brush layer thickness at infinite separation) and D is the distance between two grafted sites. The first term is the osmotic pressure arising from the increased polymer concentration in the two compressed layers. The second one is an elastic restoring force (polymer molecules always tend to coil, hence the negative sign of this term). Such a mechanism is operative at h < 2hj = 21.2 nm , i.e., h w < 28.0 nm . The solid line in fig. 6.31 is the best fit to [6.5.15]. The Van der Waals component has no significant influence on the numerical procedure, and the electrostatic component is eliminated because the experiments are carried out at high c. The fitted value, h j = l l . l n m , is in good agreement with the value of 10.6 nm obtained in the three layer model (table 6.1). Thus, de Gennes' theory gives a satisfactory description of the steric surface forces at pressures and film thickness where brush to brush contact is realized. 6.5d Oscillatory disjoining pressure
In sec. IV.3.9 it was shown that, even in the absence of steric interactions, DLVO theory accounts for only part of the electrostatics, i.e., only for diffuse double layers at fixed potential yfi that is. The approximations involved neglect of Stern layers, ignored interaction at constant cfi , and regulation was not considered. Also, solvent structure mediated interactions were not considered. In sec. IV.3.9 these features were added to
THIN LIQUID FILMS
6.61
those of DLVO, leading to the extended DLVO theory, called DLVOE. At the level discussed so far, DLVO is satisfactory as far as the electrostatics is concerned. However, the solvent structure-mediated contributions, /7 so i vstr (W. deserve special attention because they can provoke the oscillatory interactions which have been observed for pure liquids (fig. II.2.3) and for silica sols between solid plates1'. These forces result essentially from the repulsive parts of the molecule-molecule or particle-particle interaction Gibbs energies, respectively. The advantage of thin film studies is that such oscillations can be made beautifully visible. Consecutive stepwise film thinning has been observed for thin liquid films with fluid interfaces, formed from solutions with surfactant concentrations much higher than the c.m.c. This phenomenon is known as stratification. During this process, the initially formed films thin stepwise, becoming black in most cases, and sometimes even thin down to bilayers. The stratification phenomenon in foam films was already described by the beginning of the 20th century2'3'. Later, the stepwise thinning was rediscovered by Bruil and Lyklema4'5' and later extended by several other authors, e.g.6'7'8'. The phenomenon was to be universal: it can also be observed in emulsion films, asymmetric films of the air-water-oil type, films made of latex suspensions, and liquid crystalline films. Stratification is connected with the layering of latex particles or micelles inside the film. In turn, this layer-like arrangement is caused by confinement. During drainage, the film thicknesses adjust themselves so as to accommodate an integer number of layers, hence the step-wise nature of the thinning. Very precise measurements of IJ(h) isotherms have been performed with foam films from aqueous solutions with high concentrations of Na-dodecylsulphate, employing both the pressure-balance technique and the dynamic method. Disjoining pressure isotherms were established down to pressures of 10 Pa, with specially constructed film holders and careful pressure control. The typical h{t) curve shows a stepwise film thinning with a pitch of 10 nm. The stepwise thinning is in line with the oscillatory form of the I7[h) isotherm. Bergeron and Radke introduced the term oscillatory disjoining pressure (fig. 6.32). As seen from the figure, the results obtained by the dynamic and the equilibrium methods agree well. The former gives data in the negative range of disjoining pressures, 11
D. Atkins, P. Kekicheff, and D. Spalla, J. Colloid Interface Scl. 188 (1997) 234. E.S. Johnonnott, Phil. Mag. 11 (1906) 746. 31 J. Perrin, Ann. Phys. Paris 10 (1918) 160. 41 H.G. Bruil, Ph.D Thesis. Wageningen Agricultural University (1970). 51 H.G. Bruil, J. Lyklema, Nature Phys. Set 233 (1971) 19. 61 A. Nikolov, D. Wasan, Langmuir 8 (1992) 2985. 71 E. Perez, J.E. Proust, and L. Ter-Minassian-Saraga, in Thin Liquid Films, I.B. Ivanov, Ed., Marcel Dekker (1988) p. 291. 81 J.W. Keuskamp, J. Lyklema, in Adsorption at Interfaces, ACS Symposium Series 8, K.L. Mittal, Ed., ACS Washington DC (1975) p. 191. 21
6.62
THIN LIQUID FILMS
which is an advantage of this method. It is worth noting that the length scale of the oscillations was large, about 10 nm, and even reached about 50 nm. Theoretical anbalyses are based on the statistical thermodynamics of concentrated colloids, as treated in chapter IV.5, but now extended to 2D confined layers1'2'. So, essentially these systems behave as 2D supermolecular fluids. The interesting point is that the ordering can be observed on two levels, the molecular, and the colloidal one, the driving force (essentially the repulsion between pairs of molecules or colloids) being identical. The general conclusion of this section is that thin film studies, and in particular those directed at disjoining pressure isotherms, continue to contribute to our understanding of colloidal forces in the DLVO range and beyond.
Figure 6.32. Oscillatory disjoining pressure Ue^{h) vs. film thickness isotherm; aged aqueous solution 0.1 mol dm~3 Na-dodecylsulphate; o, dynamic; ", equilibrium; T = 24°C. (Redrawn from V. Bergeron and C.J. Radke, Langmuir 8 (1992) 3020.)
6.6 Black foam films and emulsion films As shown in sec. 6.4d, upon thinning, thin liquid films lose their stability at a critical thickness hcr . Depending on the type of surfactant and its concentration, a jump-like formation of black spots can occur in the foam or emulsion films. The black spots expand, merge, and the entire film area is occupied by a black film at the end (fig. 6.25). These films can reach extremely small thicknesses. When observed under a microscope they reflect very little light and appear black when their thickness is below 20 nm. Therefore, they could also be called nanofilms. The IUPAC nomenclature distinguishes two equilibrium states of black films: the common black film (CBF) and Newton black film (NBF). There is a pronounced transition between them, i.e. CBF can transform into NBF or the reverse. The CBF, just as the common thin liquid films, can be described by the three layer or sandwich model, a liquid core between two adsorption layers of surfactant mole"A.D. Nikolov, P.A. Kralchevsky, I.B. Ivanov, and D.T. Wasan, J. Colloid Interface Set 138 (1989) 13. 21 P.A. Kralchevsky, A.D. Nikolov, I.B. Ivanov, and D.T. Wasan, Langmuir 6 (1990) 1480.
THIN LIQUID FILMS
6.63
cules (fig. 6.2). The NBFs, however, are bilayer formations without a free liquid core between the two layers of surfactant molecules. Thus, the contact between droplets, or bubbles, in an emulsion or foam, can consist of a bilayer from amphiphilic molecules. In the behaviour of the latter, the short range molecular interactions prove to be of major importance. The definition, 'liquid film' is hardly valid for bilayers. They possess a higher degree of ordering, similar to that of liquid crystals. The disjoining pressure vs. thickness dependence for relatively large h values of common thin liquid films, stabilized by surfactants, is consistent with the DLVO theory. However, black films exhibit a deviation from DLVO theory which is expressed in the specific course of the /7(h) isotherm. Fig. 6.33 depicts a /7(h) isotherm (arbitrary scale) of an aqueous film from a surfactant solution containing an electrolyte. The two states of black films, CBF and NBF, are clearly distinguished. Such a presentation of the n(h) isotherm can explain the thermodynamic state of the two types of black films, stabilized by long- and short range surface forces, respectively. See also the similar graph for wetting films in fig. III.5.12.
Figure 6.33. General schematic presentation of the disjoining pressure isotherm of a symmetric thin liquid film; 1, region referring to CBF; 2, region referring to NBF.
On the right-hand side of the isotherm (fig. 6.33) the curve possesses a shallow minimum after which, upon decreasing h , the disjoining pressure becomes positive and increases to a maximum. In this range, common thin liquid films exist, their equilibrium being described by DLVO theory. If h < hCT, the film is a common black film CBF presented schematically in the figure. Such a film forms through black spots and, at the equilibrium film thickness h,, the disjoining pressure equals the external (capillary) pressure IJ{h) = Ap. The equilibrium thickness of CBF is also described by DLVO theory, although discrepancies between the theoretically and experimentally obtained /7(h) isotherms are established1' at film thicknesses below 20 nm. The pressure difference, /7(h)max - Ap , is the barrier which inhibits the transition to a film of smaller thickness. 11
J. de Feijter, A. Vrij, J. Colloid Interface Set 70 (1979) 456.
6.64
THIN LIQUID FILMS
According to DLVO theory, upon further reduction of h after /7(h) max , the disjoining pressure should reduce indefinitely. However, experimental results 1 ' 2 show the existence of a second minimum in the FI(h) isotherm, after which the disjoining pressure ascends sharply. Another equilibrium is established on the rising left-hand side of the isotherm, again under the condition 17{h) = Ap. This black film, with a smaller equilibrium thickness h 2 , is actually the bilayer Newton black film (NBF), also presented schematically in the figure. The left branch of the fl(h) curve, like the preceding minimum, is not described by DLVO theory. Obviously, at this extremely small film thickness, short range non-DLVO surface forces (for example, steric interactions) determine this part of the FT(h) isotherm. This is a typical illustration of the E in DLVOE interactions, as introduced in sec. IV.3.9. 6.6a Disjoining pressure in black films The direct experimental measurement of the disjoining pressure TJ{h) as a function of the film thickness h (the /7(h) isotherm) is of major importance for understanding the nature of surface forces which determine the equilibrium state of both types of black films, and establish the CBF-NBF transition. Such isotherms have been measured with macroscopic and microscopic films. In particular, the thin liquid film pressure balance technique, employing the porous plate measuring cell of ExerowaScheludko (sec. 6.2d), provides equilibrium values over a wide range of pressure and thickness values. This technique has been applied successfully by many authors for plotting the /7(h) isotherms of black films from various surfactant solutions. The various kinds of surfactants (emulsion or foam stabilizers) exhibit their own idiosyncracies. ionic surfactants. Black films from Na-dodecylsulphate (NaDS) will be used here as our paradigm. This typical representative of ionic surfactants is a very good stabilizer of emulsions and foams, and has been studied extensively3'4'5', so most of the film parameters are well known. Figure 6.34 presents a summary of the experimental results obtained by three different teams at different times. The figure indicates good agreement between all the data of on the right-hand side of the curve, i.e., for the long range interactions. Mysels and Jones gave no data for the left-hand side of the isotherms, i.e., within the NBF region and the CBF-NBF transition, since their porous ring measuring cell did not allow such measurements; Bergeron and Radke used the porous plate measuring cell. Comparison of fig. 6.34 with the FI(h) isotherm drawn on an arbitrary scale in fig. 6.33, shows that only positive 17(h) values have been measured, since equilibrium
11
T. Kolarov, A. Scheludko, and D. Exerowa, Trans. Faraday Soc. 64 (1968) 2864. F. Huisman, K.J. Mysels, J. Phys. Chem. 73 (1969) 489. 31 K.J. Mysels, M. Jones, Discuss. Faraday Soc. 42 (1966) 42. 41 D. Exerowa, T. Kolarov, and Khr. Khristov, Colloids Surfaces 22 (1987) 171. 51 V. Bergeron, C.J. Radke, Langmuir 8 (1992) 3020. 21
THIN LIQUID FILMS
6.65
Figure 6.34. Disjoining pressure isotherm for l m m o l d m " 3 NaDS + 0.18 mo] dm" 3 NaCl solution; T = 23-24°C; O, Mysels and Jones (1966); A, Exerowa, Kolarov and Khristov (1987); a , Bergeron and Radke (1992).
black films only form when the positive Il(h) = Ap. It is clear that when (d/7(h)/dh) > 0, films are thermodynamically unstable, so that their thickness cannot be measured. The right-hand branch of the J7{h) curve refers to CBFs, and indicates that their thickness h decreases with an increase in Ap. Below a thickness of 7.1 nm at /7(h)max = 105 Pa , the CBF transforms through a jump-like transition into an NBF of h = 4.3 ± 0.2 nm. The increase in capillary pressure up to 1.2xlO5 Pa does not alter the NBF thickness. Nor does the NBF change its thickness with a reduction in pressure, down to Ap = O.lxlO 5 Pa . Around this point, however, again with a jump, it transforms back into a CBF with the corresponding thickness. In the electrolyte concentration range c = 0.165 - 0.18 mol dm"3 , a pressure region of fluctuating NBF-CBF transitions has been found. At higher electrolyte concentrations, when a certain pressure value is reached, a jump-like transition from CBF to NBF occurs. The film thickness of such a transition does not depend on c, which means that there is a fluctuation zone where the energy which is needed to overcome the barrier /7(h) max , in the 77(h) isotherm (fig. 6.33) is of the order of kT. Films that satisfy this energy requirement are metastable. Upon increasing c, the pressure at which the transition occurs is reduced as a result of the lowering of the barrier 77(h)max. At c> 0.2 mol dm" 3 there is no longer a NBF to CBF transition, which means that the left minimum in the /7(h) isotherm lies very deeply in the region of negative 17{h) values. Thus, an electrolyte concentration range from 0.165 to 2.0 mol dm" 3 confines the isotherm region in which both CBF-NBF and NBF-CBF transitions can take place. The NBF thickness measured (within experimental accuracy) does not depend on capillary pressure, and its average value is 4.3 nm. The situation below this value will be discussed further. The depth of the left minimum in the n(h) isotherm, within the small film thickness range, is strongly affected by the electrolyte concentration. Similar 77(h) isotherms have been observed for other anionic surfactants1'. Com11
I.J. Black, R.M. Herrington, J. Chem. Soc. Faraday Trans. 91 (1995) 4251.
6.66
THIN LIQUID FILMS
pared with the DLVO theory, the experimental data indicate a disagreement between theory and experiment. For small film thicknesses, the theory seems to need improvement for the calculation of the electrostatic component /7el and Van der Waals component /7 v d w of the disjoining pressure. Also, to properly describe the right branch (CBF) of the /7(h) curve in fig. 6.34, one of the DLVO equations for high if* is needed (app. IV.2), or Stern theory has to be incorporated. The left branch (NBF) is not foreseen by DLVO theory; it rather belongs to the E of DLVOE. Non-ionic surfactants. Two types of IJ{h) isotherms have been established1 for black films from aqueous solutions of non-ionic surfactants. The first can be illustrated by solutions of decyl 4-oxyethylene, (C 10 E 4 ). The surfactant and electrolyte concentrations were chosen so that equilibrium films are obtained within a large thickness range, including the CBF-NBF transition region. The change in film thickness has been achieved by gradually and isothermally increasing the capillary pressure. The initial thickness of the C 10 E 4 film decreases to 11 nm and black spots appear, gradually invading the whole film and turning it black. Its thickness falls to 6 nm and changes no more upon further increase in capillary pressure. This result indicates that the transition to NBF occurs via a jump-like overcoming of the (/7(h)max - Ap) barrier, i.e., the /7(h) isotherm has the same shape as that in fig. 6.34. The other type of FI(h) isotherm for non-ionics can be illustrated by black films of 20-oxyethyl nonylphenol (N0E2O). The curve obtained experimentally has another shape. With the increase in /7(h) the film thickness decreases gradually with no jumplike transition, and an equilibrium thickness of 9 nm is reached, which remains constant upon further increase of pressure. Since i//^ is almost equal for both surfactants, there should be an additional repulsion force in the N^E20 films. Knowing that the thickness of these films does not depend on c but is a function of 17{h), such an assumption seems reasonable. It Is then interesting to see how the experimental isotherms conform with the DLVO theory. Of course, the fact that, even for non-ionic surfactants, /7(h) has an electrostatic contribution, indicates that these interfaces acquire a charge by some specific ion uptake. The experimental /7(h) curve for films from C 10 E 4 lies between the two theoretical curves obtained at o d = const and yfi = const. Therefore, it may be supposed that, in this case, the DLVO theory plus regulation describe well the electrostatic and van der Waals interactions in the common black films. The experimental /7(h) isotherms of NPE20 films are in agreement with the theory only on the right-hand side of the isotherm, that is, at relatively large thickness. The curve follows a smooth course which probably reflects the existence of an additional, gradually increasing, repulsive force. Since this force appears in films from surfactants with long oxyethylene chains, it can be considered as a steric interaction (sec. 6.5c). An interesting option for studying FI(h) isotherms for non-ionic surfactants is to 11
T. Kolarov, R. Cohen, and D. Exerowa, Colloids Surfaces 42 (1989) 49.
THIN LIQUID FILMS
6.67
plot the disjoining pressure vs. pH at a constant ionic strength of the aqueous solution. This possibility was suggested after the finding of a strong pH effect on the y/A potential, including the occurrence of an isoelectric point (sec. 6.5b). It has also been shown that the CBF-NBF transition can be realized when the pH of the solution is altered1'. This is also of particular interest for finding the conditions for studying NBFs in the absence of /7el at not-too-high ionic strengths. The fl{h) isotherms of black films from N0E2O solutions, measured at pH = 5.7, 6.1 and 4.0, show clearly the influence of OH" ions, and the isotherms are in very good agreement with the h(pH) dependence determined at Ap = const using DLVO theory. At pH values close to the i.e.p., the isotherm has a different shape, and the films rupture at lower pressures than those at neutral pH. Detailed quantitative analyses of the disjoining pressure measurements, performed with black films from non-ionic sugar-based surfactant2' at various c s , pH, and ionic strength have also shown the adsorption of OH" ions to be responsible for the surface charge at the film-air interface. Phospholipids. The formation of thin aqueous films from phospholipids is experimentally difficult since many of them are insoluble in water. Sonicated dispersions of insoluble phospholipids can be u s e d , and the first direct measurements of interaction forces in films stabilized by neutral phospholipids were made with films from suspensions of unilamellar DMPC vesicles . The 17(h) isotherm indicates a barrier transition to NBF of thickness 7.6 nm, which remains constant upon further increase in pressure. This isotherm is similar to that obtained for non-ionic surfactants, such as C 10 E 4 . The right-hand side of the isotherm can be interpreted in the same way: with DLVO forces in the film. Metastable films exist in the c range from 1.5xlO~2 to 2 x 10~2 mol dm" 3 of NaCl, in which Newton black spots appear in the CBF. The NBF do not change their thickness of hw = 8.2 nm, up to c = 0.5 mol dm" 3 . Interesting results have been obtained for black films stabilized by soluble zwitterionic phospholipids51: lysophosphatidylcholine (lyso PC) and lysophosphatidylethanolamine (lyso PE) in the presence of Na+ or Ca2+ . Besides the /7(h) isotherm another type of isotherm proves to be very informative, i.e., the dependence of film thickness h on electrolyte concentration c at c s = const, Ap = const, and T = const. This h(c) dependence allows one to distinguish clearly the action of electrostatic disjoining pressure, and to find the electrolyte concentration at which the CBF-NBF transition occurs. The effect of Ca2+ on the equilibrium thickness of films from aqueous lyso PC solution is shown in fig. 6.35. Unlike the results obtained with the monovalent Na+ ions, black films of constant thickness h = 7.6 nm are always formed at low Ca2+ 1) T. Kolarov, R. Cohen, and D. Exerowa, Colloids Surfaces A129-130 (1997) 257. V. Bergeron, A. Waltermo, and P. Claesson, Langmuir 12 (1996) 1336. 31 T. Yamanaka, M. Hayashi, and R. Matuura, J. Colloid Interface Sci. 88 (1982) 458. 41 R. Cohen, R. Koynova, B. Tenchov, and D. Exerowa, Eur. Biophys. J. 20 (1991) 203. 51 R. Cohen. D. Exerowa, and T. Yamanaka, Langmuir 12 (1996) 5419. 21
6.68
THIN LIQUID FILMS
h, nm Figure 6.35. Equivalent thickness of lyso PC foam films as a function of CaCl2 concentration; Ap = 35 Pa; T = 30°C; pH = 5.5. (Redrawn from R. Cohen. D. Exerowa, T. Kolarov, T. Yamanaka and V.M. Muller, Colloids Surfaces, 65 (1992) 201.)
concentrations. At 10 3 mol dm 3 CaCl2 , however, a dramatic increase in film thickness to about 35 nm has been observed, i.e. a transition to comparatively thick films. Further increase in c led to a monotonous reduction in film thickness until c reached 2 x 10"2 mol dm" 3 CaCl2 , where gradually expanding black spots appeared, until the whole film became black. The thickness of these black films (obviously CBF) continues to decrease with increase in c, up to 0.2 mol dm" 3 CaCl2 when black films of the same thickness as in the low concentration range are again obtained. This is indicated in fig. 6.35 by a small thickness jump. Further increase in c, up to 0.5 mol dm" 3 , has no effect on the thickness of these black films. The IJ(h) isotherm at 2xlO~ 3 mol dm" 3 CaCl2 is also similar to that for C 10 E 4 ; relatively thick films are formed at low pressures, and their thickness decreases with increasing FI(h}. The jump-like transition, CBF-NBF, occurs in the pressure interval, 5xlO" 3 to 6xlO" 3 Pa. The NBF thickness does not change with further increase in pressure. Measurements by various techniques have shown that monovalent ions do not bind to zwltterionic phospholipids. In contrast, the observed thickness transition in fig. 6.35 demonstrates that the divalent Ca2+ ions do have a specific effect on the properties of the black films studied. Most probably this is caused by the specific binding of Ca2+ ions to the adsorbed lyso PC1'. The binding of the positive Ca2+ ions in the low concentration range can lead to low \jA values and, consequently, to a reduction in the electrostatic repulsion, and to formation of NBF (the left-hand side of fig. 6.35). The transition from NBF to thicker films, observed at c = 10~3 mol dm" 3 , can be related to further Ca2+ ion-binding at higher c, which induces higher y/ values. If the surface density of phopholipid molecules is known, the fraction of bound Ca2+ ions can be calculated from y/ . In this way a ratio of 2:1 for lyso PC-Ca2+ has been derived from the experiments '. The resulting /7e] is evidently sufficiently strong to allow the D T. Yamanaka, T. Tano, O. Kamegaya, D. Exerowa, and R. Cohen, Langmuir 10 (1994) 1871.
THIN LIQUID FILMS
6.69
formation of thicker films. Further increase in c leads to thinner films as a result of the competitive action of further Ca2+ binding (higher y/ ) and higher ionic strength of the solution (lower /7 el ). 6.6b Main properties of the two types of black films and the CBF-NBF transition The two types of black films, the common black film (CBF), and Newton black film (NBF), are two different equilibrium states of a black emulsion or foam film, and show rather different properties, owing to their different structures. This difference also determines the CBF-NBF transition. A number of thermodynamic parameters determine whether CBF or NBF is the equilibrium state dictated by the temperature T, electrolyte concentration c , surfactant concentration c s , pH and capillary pressure Ap. Besides the values of these parameters the nature of the surfactant (including amphiphilic polymers) is also decisive for the existence of the black films. The values of the parameters at which CBF or a NBF is stable are within definite (sometimes narrow) intervals and only the combination of all parameters determines the type of black film observed. Table 6.2. Capillary (disjoining) pressure and film thickness at rupture or at the CBFNBF transition c c h h Ap 10"4 Ap 10"4 NaCl NaCl 3 3 nm nm Pa Pa mol dm" mol dm" Film rupture 4
lxlO" lxlO" 3 0.1
0.15
2.5 -9.8 2.0-8.3 10.0- 13.0 12.0- 14.0
CBF-NBF transition 19.0 - 11.0 16.0-9.0 6.4-6.0 5.8-5.3
0.165 0.180 0.20 0.25 0.31
10.0 + 0.5 9.0 ±0.5 5.5 ±0.5 2.0 ±0.2 0.015 ±0.001
7.1 7.1 7.1 7.1 7.1
+0.2 ±0.2 ±0.2 ±0.2 ±0.2
The following example illustrates this. The good foam or emulsion stabilizer, Nadodecylsulphate, provides the formation of stable black films only if c s > c bl (sec. 6.4d). However, the c and Ap values are also decisive. Table 6.2 presents1' the values of capillary (disjoining) pressure /7(h) = Ap and film thickness h at which films from 1 mmol dm" 3 NaDS aqueous solutions (lower than the c.m.c.) either rupture or where a CBF-NBF transition occurs. In the NaCl concentration range from 10~4 to 0.15 mmol dm 3 , the films rupture in a certain pressure interval which becomes narrower with rising c. The experiments indicate that film rupture at pressures lower than /7(h)max is a random phenomenon. The existence of a critical pressure Apcr of film
11
D. Exerowa, T. Kolarov, Khr. Khristov, loc. cit.
6.70
THIN LIQUID FILMS
rupture was observed with various types of films (common thin films, CBF and NBF) stabilized with various kinds of surfactants but a satisfactory theoretical explanation of this effect has not yet been proposed. The temperature is another very important parameter: systematic studies of black films from aqueous NaDS solutions have shown that, at given values of c , c s and Ap, the change of temperature influences the CBF-NBF transition: at high temperatures the equilibrium black films are CBF, but at low temperatures they are NBF. In fig. 6.36, the line divides the diagram between an area of CBF stability (left upper part) and the area of NBF stability (right lower part).
T/°C
•
^o'*
30 -
y
yx 20 5/° . /( 3 „ Q 1
0 2
Q 3 c
salt /
m o 1
Q 4
dm"
3
Figure 6.36. Dependence of the temperature of the CBF-NBF transition on the electrolyte concentration in the initial solution: x , data from M. Jones, K. Mysels and P. Scholten, Trans. Faraday Soc, 62 (1966) 1336; O, data from D. Platikanov and M. Nedyalkov, Ann. Univ. Sofia, Fac. Chem., 64 (1969/70) 353.
The influence of temperature is well demonstrated by measurements of the longitudinal specific electrical conductivity Kf (sec. 6,2g) of black films from 1 mmol dm" 3 NaDS + 0.3 mol dm" 3 NaCl solution, and the specific electrical conductivity KL of the initial bulk solution as a function of temperature. Comparison of the results (fig. 6.37) allows one to find the temperature at which the CBF-NBF transition occurs. Line 1 presents the temperature dependence of the longitudinal specific electrical conductivity K of CBF. It has the same slope as line 3 which refers to the temperature-dependence of the specific electrical conductivity KL of the bulk phase. The slope of line 2 which applies to NBF is much steeper. The cross-point of lines 1 and 2 at 31.6°C indicates the temperature of the CBF-NBF transition at the given concentrations of the initial solution. The slope of the lines in Arrhenius co-ordinates in fig. 6.37 is proportional to the activation energy of the ionic movement. The near-equality of the slopes of lines 1 and 3 indicates that the liquid core of CBF contains practically the same solution as the bulk. On the other hand, the fact that the slope of line 2 is 2.5 times larger proves that the NBF contains no liquid solution. Another very informative fact is that the transport numbers of the counterions in NBF from ionic surfactants are close to 1, i.e., the electric current is due to the surfactant counterions only: the film contains no measurable amounts of other ions, although there is a high electrolyte concentration in the bulk solution. The transport numbers of ions in CBF are almost the same as in the bulk.
THIN LIQUID FILMS
6.71
Figure 6.37. Temperature dependence of the longitudinal specific electrical conductivity, Kf , of CBF (line 1), NBF (line 2) and the specific electrical conductivity, KL , of bulk solution (line 3); films made from NaDS + 0.3 mol dm" 3 NaCl solution (Redrawn from D. Platikanov and N. Rangelova in: Research in Surface Forces (Ed. B.V. Deryagin), Vol. 4, Consultants Bureau, New York (1972) p.246).
Thermodynamic quantities which show rather different values for CBF and NBF are the contact angle a and film tension y{ (sec. 6.3e), which are related through equation [6.3.47]. Systematic measurements of a have been performed with black films from NaDS aqueous solutions. Figure 6.38 presents results obtained by three different teams. A very sharp jump in the a(c) curve is observed at the CBF-NBF transition. The critical electrolyte concentration c cr at which the transition occurs from one type of black film to the other can be determined from this jump. For example, for microscopic black films from (NaDS+NaCl) solutions, c cr = 0.334 mol dm" 3 . At lower c values CBF with a < 1° are stable, while at higher c values NBF with a of the order of 10° are stable. The data obtained for macroscopic flat films in a frame indicate somewhat lower values, i.e. c cr = 0.2 mol dm"3 . Hence, a range of NaCl concentration (0.2-0.334 mol dm" 3 ) is distinguished in which the a{c) curves follow a different course. This results from the metastable state of CBF: in this concentration range, after a certain time, the microscopic CBFs transform into NBFs. In the metastable region, indicated by the dashed lines in fig. 6.38, both NBFs and CBFs are observed, exhibiting, respectively, large or small values of the contact angles. The a values obtained with microscopic NBFs are close to those obtained with macroscopic NBFs. When there is no external influence NBFs are only formed at electrolyte concentrations where the maximum in the /7(h) isotherm is overcome /7(h)max < Ap. However, if the film area is large, or when the film is subjected to an external disturbance, NBFs can be obtained at lower c when n(h)max > Ap . This is the reason for the difference in c cr values derived from microscopic and macroscopic films (fig. 6.38). It is established that c cr decreases monotonously with increasing film size. At the smallest film radii r , c cr remains constant, which allows one to determine it at r —> 0 . Obviously, c cr = 0.334 mol dm" 3 applies only to a definite film radius, r = 250 um .
6.72
THIN LIQUID FILMS
Figure 6.38. Contact angle vs. NaCl concentration for black films from aqueous NaDS solutions; T = 22°C; x , data from T. Kolarov, A. Scheludko and D. Exerowa, Trans. Faraday Soc, 64 (1968) 2864; • , data from F. Huisman and K. J. Mysels, J. Phys. Chem., 73 (1969) 489; O, data from D. Exerowa, Khr. Khristov and M. Zacharieva, in: Pouerkhnostnye sily v tonkikh plenok, Nauka, Moscow (1979) p. 186 (in Russian).
By analogy, the film tension y f , being related through [6.3.47] to the contact angle a , shows similar behaviour. Its value for CBF (a < 1°) is close to 1y of the bulk solution, while for NBF (a of the order of 10°) it is essentially different. Accordingly, the values of the interaction Helmholtz energy, AF{h), of the film (sec. 6.3d) are rather different between CBF and NBF. The two types of black films from NaDS solutions exhibit completely different behaviour in another way. In contrast to CBFs, the NBFs do not change their thickness with Ap and c . However, their properties depend on the composition of the bulk surfactant solution, e.g. a depends on c . The thickness of NBFs, determined from the h(c) dependence, is approximately equal to twice the thickness of the adsorption layer as assumed by Perrin. This is confirmed for NBFs obtained from other surfactants. However, it has been proved by infrared spectroscopy1' and electrical conductivity measurements2 , that there is a little water in the NBF. It is most probable that the adsorption layers do contain minor amounts of water, but that these layers are not separated by a macroscopic aqueous core. This conclusion is also supported by ellipsometric measurements3'. The data for the film thicknesses of NBF from NaDS solutions, obtained by several authors with different techniques, are rather scattered between 3.3 and 4.5 nm. Precise X-ray reflectivity measurements with CBF and NBF films from NaDS + NaCl aqueous solutions provide more details about their structure4'. The thickness of the respective layers, which detail the film structure is: hydrocarbon chains layer 1.08 nm, polar groups layer 0.38 nm, water layer 0.38 nm, total film thickness 3.29 nm. Although these data indicate an aqueous core, this water is most probably hydration water, and not bulk liquid, so the NBF can be considered as bilayers.
11 2
J. Corkill, J. Goodman, C. Orgden, and J. Tate, Proc. Roy. Soc. A273 (1963) 84.
D. Platikanov, N. Rangelova, in Researchin Surface Forces, B.V. Deryagin, Ed., Vol. 4, Consultants Bureau, New York (1972) p. 246. 31 D. den Engelsen, G. Frens, J. Chem. Soc. Faraday Trans. I 70 (1974) 237. 41 O. Belorgey, J.J. Benattar, Phys. Rev. Lett. 66 (1991) 313.
6.73
THIN LIQUID FILMS
6.6c Stability and rupture ofbilayer black films Newton black foam films, Newton black emulsion films, and bilayer lipid membranes (BLM), are bilayers of amphiphile molecules. Their stability with respect to rupture, and their permeability, can be considered from the same point of view. A possible mechanism for these processes is based on the appearance and growth of holes in the bilayer. For thicker films, i.e. not for bilayers, such a mechanism was first proposed by de Vries1'. A theory for the hole rupture of NBF (bilayers) was developed by Deryagin et al. 2 ' 3 '. According to these authors the mean time for rupture is not a function of any experimentally controllable parameter. In contradistinction, the theory to be presented below4'5 considers the bilayer film at equilibrium with the surfactant solutions, of which the concentration can control the process of the nucleation of holes. Figure 6.39 represents schematically a generally accepted molecular model for such bilayers. The description of the formation of microscopically small holes responsible for the bilayer stability and permeability can be based on both thermodynamic and molecular models. Figure 6.39. Molecular model of; (a), foam or O-W emulsion and, (b), lipid membrane or W-O emulsion bilayer; open circles denote vacancies and holes; u and UQ are, the energies of the lateral and normal bonds, respectively between the nearest-neighbour amphiphile molecules in the bilayer.
The driving force or supersaturation A[i for bilayer rupture is a thermodynamic quantity, since it is defined by Afi = fib-fis
[6.6.1]
where jub and jus are the chemical potentials of the amphiphile molecules in the bilayer, and in the solution, respectively. For sufficiently dilute solutions //s depends logarithmically on the concentration c s of the monomer amphiphile molecules in the solution. Equation [6.6.1] can be approximated by Afj{cs) = kTln(ce/cs)
[6.6.2]
where ce is a concentration of monomer amphiphile molecules in the solution called the bilayer equilibrium concentration. 11
A. de Vries, Proc. 3rd Intern. Congr. Detergents, Cologne, Vol. 2 (1960) p. 566. B.V. Deryagin, Yu.V. Gutop, Kolloid. Zhur 24 (1962) 431. 31 B.V. Derjaguin, A.V. Prokhorov, J. Colloid Interface Sci. 81 (1981) 108. 41 D. Exerowa, D. Kashchiev, J. Colloid Interface Sci. 77 (1980) 501. 51 D. Exerowa, D. Kashchiev, Contemp. Phys. 27 (1986) 429 21
6.74
THIN LIQUID FILMS
Thermodynamics can also be used for determining the work w{ for formation by fluctuation of an i-sized hole, i = 1,2,3,... being the number of amphiphilic molecules that would fill the hole. Since a hole appears as a result of transferring i molecules from the bilayer into the solution; the work associated with this process is -i Afi . On the other hand, work equal to the total peripheral Gibbs energy of the hole has to be done, which is proportional to the length of the hole's perimeter. For circular holes of bilayer depth, this is (2/ra m i) 1/2 . Hence, wi is given by wi =-(A// + r L (2/ra m i) 1 / 2
[6.6.3]
where rL ( J m" 1 ) is the hole-specific edge Gibbs energy which, in principle, can be a function of i; am is the area per molecule. When Aju > 0 (for example, when c s < ce ), the competition between the two energy terms in [6.6.3] causes wi to pass through a maximum w * at i = i* . The hole of size i * is the so-called hole nucleus, and w' is the nucleation work. While the subnucleus holes tend to decay (w i decreases with decreasing i
supernucleus holes can
grow spontaneously (u>i decreases with increasing i> i*). For this reason the bilayer can rupture only after the appearance, by a fluctuation, of at least one nucleus hole per unit time and, accordingly, w* is the energy barrier for bilayer rupture. From [6.6.4] and the extremum condition, dwi/di
at l = i* , and provided rL is independent of i, it
follows that for A// > 0 (i.e., for c s < c e ) t* = ; r a " 1 ^ L 2A/J2
[6.6.4]
w* = n
[6.6.5]
Equations [6.6.4] and [6.6.5] show that both the nucleus size and nucleation work become infinitely large if A//= 0 , i.e., if c s = c e . Physically, this means that the formation of holes in the bilayer is then impossible. Accordingly, the bilayer is then truly stable (and not metastable) with respect to rupture by this mechanism. It must be emphasized that the bilayer also retains this true (or indefinite) stability for A^ < 0, i.e., for cs> ce, since both terms in [6.6.3] are then positive, and wi can only increase with increasing i. The bilayer cannot rupture despite the presence of a population of a certain number of holes. We shall now turn to results obtained by using the molecular model of an amphiphile bilayer illustrated in fig. 6.39. The basic idea in the theoretical description is to regard the bilayer as consisting of two mutually adsorbed monolayers of amphiphilic molecules. Each of the monolayers can be filled to a maximum of JVm = l / a m ~ 1018 molecules m" 2 , but the thermal motion of the molecules reduces their density to below JV . This means that vacancies of amphiphilic molecules (i.e., molecule-free sites) exist in the bilayer, their density being nx (m~2 ). At high values of
THIN LIQUID FILMS
6.75
rij, the vacancies cluster together to form holes. If these holes are sufficient in number and size, they can make the bilayer permeable to molecular species. When A/u > 0 (for example, when c s < ce ) a nucleus hole can come into being and, by irreversible growth, cause the rupture of the bilayer. In other words, rupture occurs as a result of a two dimensional (2D) first-order phase transition of the, 'gas' of amphiphile vacancies in the bilayer into a, 'condensed' phase of such vacancies which is equivalent to a ruptured bilayer. The model of the bilayer (fig. 6.39) can also be used to express ce in terms of intermolecular bond energies, the result being ce=c0exp(-U/2kT)
[6.6.6]
Here, cQ is a reference concentration, and U > 0 with U given by U = zu + zQu0
[6.6.7]
is the binding energy of an amphiphile molecule in the bilayer; u and u 0 (positive for attraction, negative for repulsion) are respectively, the energies of the lateral and normal bonds (owing to short range surface forces) between the nearest-neighbour amphiphile molecules in the bilayer, z and zQ are the lateral and normal co-ordination numbers of these molecules (for example, z = 6 , z 0 = 1 or 3 for hexagonal packing). The molecular model given for amphiphilic bilayers can also be used for describing the process of hole-nucleation. Using the classical nucleation scheme, this process results from a series of, 'bimolecular reactions' characterized by the nucleation rate, which is the frequency with which the i * -sized nuclei become supernuclei holes of size i*+l. For steady state nucleation, the nucleation theory derives an explicit r(c s )dependence of the bilayer mean lifetime r on the bulk concentration c s as r(cs) = , 4 e x p - ^ r ln c ( e/cs) with A = \/ZcoNQA
and B= 7CAQT1 12{kT)2
[6.6.8]
[6.6.9]
where Z is the so-called Zel'dovich factor (a number having values from ca. 0.01 to 1); A is the bilayer area ftjfs"1) is a frequency of a vacancy joining the nucleus hole, NQ (m~2) is the density of available lattice sites on which a vacancy can be formed. In some cases r can be so short that experimental observation of the bilayer after its formation is possible only with a certain probability W which depends on the resolution time t r of the particular equipment used. In the direct, visual observation of bilayer rupture, for example tr ~ 0.5 s, which is the reaction time of the eye. Since observation of the bilayer is only possible if the bilayer has ruptured during the time t > tT , this nucleation theory yields W as a function of c s < ce
6.76
THIN LIQUID FILMS
W(cs) = exp{-t r /^exp[5/ln(c e /c s )]}
[6.6.10]
Equations [6.6.8] and [6.6.10] show that both r and W increase sharply with the bulk surfactant concentration c s over a relatively narrow range. For this reason a critical concentration cc for bilayer rupturing in less than rc seconds, can be defined by the condition r(cc) = r . From [6.6.8] we obtain c c =c e exp[-5/ln(r c /A)]
[6.6.11]
The theoretical t(cs) dependence [6.6.8] can be checked easily, since both x and c s are measurable quantities. The bulk concentration c s of the amphiphilic molecules is determined when preparing the surfactant solution. The mean lifetime r is measured as the time elapsing from the moment of formation of a bilayer with a given radius until the moment of its rupture. Microscopic emulsion or foam bilayer films have been investigated1' using the experimental cell and device described in sees. 6.2c and 6.2d. Owing to the stochastic character of the film rupture, the film lifetime r has to be determined by averaging over a large number of measurements.
Figure 6.40. Dependence of the mean lifetime on the surfactant concentration for foam (O) and emulsion (•) bilayer films from ^12^22 + a t l u e o u s 0-3 mol dm~ 3 KC1 solutions; r = 100 (im . (Redrawn from H.J. Miiller, B. Balinov and D. Exerowa, Colloid Polym. Sci., 266 (1988) 921.)
A comparative investigation of the rupture of microscopic foam and emulsion bilayers, obtained from the same aqueous solutions of dodecyl 22-oxyethyl (C12E22) non-ionic surfactant, at electrolyte concentrations higher than c cr , has been carried out. The emulsion bilayer was formed between two oil phases of nonane. Figure 6.40 shows the r(cs) dependence for foam (open circles) and emulsion (black circles) bilayer films. It is seen that, in both cases, x depends strongly on the surfactant concentration. The solid curve for the foam bilayer represents the best fit of [6.6.8] with A = 0.145 s, B = 25 and ce = 3.3 x 10~5 mol dm" 3 . Using these values for B and ce in the theoretical equations leads to i* between 1 and 4, W* = (5.6/9)/cT, and rL = 1.1 x 10" u J m" 1 . Obviously, in this case rL is only an effective quantity because of the very small size of the nucleus hole. Since the data for the emulsion bilayers are
11
D. Exerowa, B. Balinov, and D. Kashchiev, J. Colloid Interface Sci. 94 (1983) 45.
THIN LIQUID FILMS
6.77
rather scattered, only ce and r could be estimated; their values are ce =(0.5/3)x 1CT3 mol dm" 3 and rL = 6xlO" 12 J m" 1 . It is also seen in fig. 6.40 that the stability of the foam bilayers exceeds that of the emulsion bilayers, and that ce is much lower for the foam bilayers. The same experiment can also be used to check [6.6.10], W being the ratio between the number of bilayer films observed, and the number of all films studied. Study of the W(cs) dependence is possible, and is particularly convenient at low cs values when the mean bilayer lifetime r is comparable to tT. According to [6.6.10], a characteristic feature of W is its sensitivity to changes of cs over a very narrow range only. The W(cs) dependence obtained fits [6.6.10] well, and its shape is different for different types of surfactants1 .
Figure 6.41. Temperature dependency of the critical rupture concentration of DMPC phospholipid bilayer films. Drawn line according to [6.6.11]. (Redrawn from D. Exerowa, Adv. Colloid Interface Sci. 96 (2002) 75.) A very steep W(cs) dependence has been measured for dimyristoyl phosphatidyl choline (DMPC) foam bilayers. The W values jump from 0 to 1 within a very narrow cg range. The critical concentration, c c , for bilayer rupturing is determined with accuracy of 5% from this steep curve at a given temperature. The temperature dependence of the DMPC critical concentration c c is shown in fig. 6.41. Experimental results are denoted by circles and the solid lines are least squares fits. The break in the lnc c vs. 1/T dependence corresponds to the temperature of a film thickness transition at 23°C. It is clear that the temperature dependence of c c is very sensitive to the occurrence of some phase transitions in the bilayer films, and may therefore be used for their detection. The transitions in the phospholipid bilayer films described above, were observed when the thickness hw of DMPC foam bilayers was measured 2 ' micro-interferometrically over the temperature range from 10°C to 30°C. Three temperature ranges could be 11 21
D. Exerowa, B. Balinov, A. Nikolova, and D. Kashchiev, J. Colloid Interface Sci. 95 (1983) 289. D. Exerowa, A. Nikolova, Langmuir 8 (1992) 3102.
6.78
THIN LIQUID FILMS
distinguished with respect to the formation of bilayers with constant thickness: 1012°C (h w = 7.5 nm), 13-20°C (h w = 6.9 nm) and 24-30°C (h w = 6.2 nm). The two thickness transitions may be regarded as reliable, because the accuracy of the microinterferometric method is ±0.2 nm and there is good reproducibility of the experimental results. An important result is the coincidence of the temperature of the main phase transition determined for the DMPC water-ethanol dispersion by DSC, with the temperature of the jump-like change in the bilayer thickness (23°C). At the same temperature, a break is observed in the temperature-dependence of cc (fig. 6.41). These facts show that, both in the bulk phase and in the foam bilayer, a chain melting phase transition occurs, which was found to be related to a sharp shift in the number of gauche conformations of carbon-carbon bonds1'. Hence, a gel state of the foam bilayer at T < 23°C, and a liquid-crystalline state at T > 23°C, are likely. It follows from the theory we have described that cc = ce in the case of a missing metastable region, when only thermodynamically stable foam bilayers are formed. The high stability of DMPC foam bilayers gives us reason to assume that cc =ce, thus permitting the computation of the binding energy U from the experimental dependence of cc on temperature (fig. 6.41). The values of U obtained from the best fit of [6.6.6] to the experimental data are (1.93±0.04)xl0- 19 J at T < 23°C, and (8.03±0.19)x 10" 20 J at T > 23°C. The higher value of U for the gel state is expected, as it refers to a state of higher degree of order. The good fit of the experimental results to the theoretical dependence, and A high stability of the DMPC foam bilayers, show that the assumption cc = ce is probably accurate in this case. The good agreement between theory and experiment permits the determination of several molecular quantities in the theory: w*, I*, rL, cc, ce, etc. From the hole edge energy rL the bond energy u between two nearest-neighbour surfactant molecules in the bilayer plane may be estimated, since rL originates from short range molecular interactions in the bilayer. This important parameter for the bilayer stability and permeability has now been determined for a number of surfactants. 6.7 Diffusion processes in symmetric thin liquid films Like interfaces, thin liquid films separate two phases. Under appropriate conditions, the transport of molecules from one phase to the other occurs by diffusion across the film. On the other hand, lateral diffusion of some molecules can take place along the film. In this section, diffusion processes in symmetric thin liquid films will be illustrated for three cases only: gas permeation through foam films including CBFs, gas permeation through bilayer films, and lateral diffusion in black films. 6.7a Gas permeability of foam films including CBFs The process of gas transfer through a liquid film, towards a lower partial gas 11
N. Yillin, I.W. Levin, BiochemistrylG (1977) 642.
THIN LIQUID FILMS
6.79
pressure, consists of three stages: gas dissolving from the side of the phase with higher partial pressure, diffusion through the film, and evolution from the opposite side of the film. Because of the difference Ac between gas concentrations in the liquids on either side of the film, molecular diffusion occurs across the liquid film. According to Fick's first law for unidimensional flow at steady state diffusion (see table 1.6.1 and sec. 1.6.5a), the mass, dm , transferred through an area of cross-section A, during a time dt, is proportional to the concentration radient Ac/h of the substance across film thickness dm = _DAAc dt h where D is the coefficient of molecular diffusion and K is the permeability coefficient, in this case, K = D/h . At equilibrium, the gas concentration at the surface liquid layer is expressed by Henry's law, c = KHpi, see [1.2.20.3], where KH is Henry's constant and p ( is the partial gas pressure. Thus, the rate of gas transfer through a flat liquid film can be given by dm
— = dt
DAKuAD:
^-^- = -KAAp, h
[6.7.2]
where Ap{ is the difference between the partial gas pressures in the phases separated by the film; in this case K = DKH I h. If the gas in a foam bubble and in its surroundings is the same, the diffusion transfer occurs under the action of the capillary pressure Ap = Ay I a ; solving [6.7.1] yields, 9
o 16y, 3po °
aZ-a2+—/-(aril -a)
°
, 8RTDKHy 8RTK.Y = ^-t = -t ' poh p0
6.7.3
where aQ is the initial bubble radius, a the bubble radius at time t, y is the surface tension, R is the gas constant, p 0 is the pressure in the large gas phase. At p 0 » Ay I a the second correction term on the left side of [6.7.3] can be neglected. Equation [6.7.3] suggests that the rate of diffusion is much lower than the rate of gas dissolution and gas evolution from the two film surfaces, and that the adsorption surfactant layers do not affect gas transfer. However, it is known that monomolecular layers from some insoluble surfactants considerably reduce the rate of evaporation of the water surface1'. Hence, the effect of adsorption layers on the rate of the gas-transfer should also be taken into account. An analysis of the gas permeation through thin liquid films can be made on the basis of the three layer film model (fig. 6.42). The gas flow through the film's inner liquid layer, of thickness h w , equals 11
A.W. Adamson, Physical Chemistry of Surfaces, 4lh ed. Wiley (1982).
THIN LIQUID FILMS
6.80
Figure 6.42. Calculation of the gas permeability through a foam film; hw, thickness of the liquid core; S, thickness of the adsorption layer; h *, thickness of the aqueous layer, equivalent in diffusion resistance to the monolayer. (Redrawn from H.M. Princen and S.G. Mason, J. Colloid Sci., 20 (1965) 353].)
dm DA(c\ -c'o) 1 — = | i _ 2 - = KA(c 1-c'2) where c\ = KHcx,
[6.7.4]
c'2 = KHc2 are, respectively, the equilibrium gas concentrations in
the liquid phase at both film surfaces (fig. 6.42); KH is again Henry's constant. At steady state, this gas flow is equal to the flow through both surfactant monolayers. In the general case, c\ < Cj and c'2 > c 2 , since the gas passes through the monolayer (thickness S) at a limited rate. In a similar way gas flow through the adsorption layer is determined from [6.7.4] ^ f = ' c ML A (ci-c' 1 ) =
fcMLA(c'2-^)
[6.7.5]
where lcML = £>ML / h * is the permeability of the surfactant monolayer with thickness 8; D ML *s *he a n a l ° g u e °f the diffusion coefficient of a monolayer; h * is the thickness of an aqueous layer, equivalent in diffusion resistance to the surfactant monolayer. From [6.7.4] and [6.7.5] it follows that C
''- C '2
=
h
K + 2D/k
16 7 61
^ - ^ )
- '
The total permeability K of the film, accounting for the monolayers, is K=
DKy. H hw+2D/fcML
=
DKy. H— hw + 2h*
[6.7.7]
In the limiting case of relatively thick liquid films, when h w »2D/kML,
the per-
meability is determined by the thickness of the liquid core K =^ i In the other limiting case of Newton black films when h w « 2D/kML is determined by the diffusion resistance of the surfactant monolayers
[6.7.8] , the permeability
THIN LIQUID FILMS
K
_
K
H k ML
2
6.81
[6.7.9]
Therefore, the effect of the monolayer reduces to an additional resistance, equivalent to that of an aqueous layer of thickness h * . For many surfactants h * is within the range of 7 to 12 nm. This means that the permeability of thick films is determined by the rate of molecular diffusion in the liquid film core, while for black films (NBFs, or very thin CBFs with h < 10 nm) the permeability is determined by the properties of the surfactant monolayers. Electrolytes do not significantly affect the gas permeability of monolayers. It was considered that gas diffusion through the monolayer occurred as the result of the creation of microscopic vacancies between the surfactant molecules. This model was called, the 'energy barrier model'. However, this model later proved to be unsatisfactory1'. Experimental data for the gas permeability coefficient K of CBFs indicate that in such particularly thin foam films the permeation is mainly determined by the surfactant monolayers21. The monolayer permeability is lower at small thickness, since the normal interactions between the surfactant molecules make the monolayers better ordered and less permeable. This mutual influence weakens with increasing film thickness and fcML tends to a constant value. Thus, two opposite tendencies act: kML increases with increasing h, but the inner liquid layer's permeability simultaneously decreases. The monolayer permeability, and consequently the film permeability, depend on the specific adsorption of the counterions. Ions which are more strongly adsorbed in the monolayer make it more compact and ordered, and hence less permeable.
Figure 6.43. Film permeability K vs. NaCl concentration; the dashed line indicates the CBFNBF transition according to literature data. (Redrawn from R. Krustev, D. Platikanov and M. Nedyalkov, Colloids Surfaces A79 (1993) 129.)
11 21
H.M. Princen, J.Th. Overbeek, and S.G. Mason, J. Colloid Sci. 24 (1967) 125. R. Krustev, D. Platikanov, and M. Nedyalkov, Colloids Surfaces A123-124 (1997) 383.
6.82
THIN LIQUID FILMS
The reduction in film thickness h is achieved by increasing the electrolyte concentration in the solution. An experimental K(cNaC|) dependence over a large concentration range is plotted in fig. 6.43. The jump-like change in K occurs at a cNaC1 value equal to that of the CBF-NBF transition. This is a remarkable experimental result, since CBF is about three times more permeable than the much thinner NBF. This again confirms that the properties of the adsorption layers, and not the film thickness, control the permeability of the extremely thin black foam films. In a bilayer film the coupled monolayers are better ordered, and the film is less permeable although it is much thinner than CBF. 6.7b Gas permeability of Newton black films Gas permeation through NBFs is of particular interest, not only because of its role in foams, but also because Newton black films represents bilayers which are model for biological membranes. According to the theory of bilayer stability and permeability, discussed in sec. 6.6c, the holes formed by fluctuations in a bilayer consist of a different number, i = 1,2,3
of vacancies of surfactant molecules. It can be assumed
that gas transfer through the bilayer occurs simultaneously through the hole-free homogeneous film area (permeability coefficient Ko representative of the background permeability of the bilayer) and through the holes of different size I (permeability coefficient, Ki). In this case, the permeability coefficient of the bilayer film is given by1' K =K0+^Kj
[6.7.10]
i=l
According to the theory for hole formation in sec. 6.6c, K can be expressed as K = K0+Yiai[\/cs)i
[6.7.11]
i=l
where v eii
m /—
[6.7.12]
cJexpGu/fcT In [6.7.12], Aeff is the effective area of one vacancy in the hole (e.g., Aeff = a m II for a hole of a bilayer thickness); a m is the area of a surfactant molecule in the monolayer plane; GLi Is the total linear Gibbs energy of a hole of size i; ce is the equilibrium surfactant concentration, i.e., the concentration at which the hole cannot enlarge to a size resulting in film rupture. Experimental K{cs) curves of the gas permeability coefficient as a function of the bulk surfactant concentration, obtained by the 'diminishing bubble' method (sec. 6.2h)
11 M. Nedyalkov, R. Krustev, D. Kashchiev, D. Platikanov, and D. Exerowa, Colloid Polym. Sci. 266 (1988) 291.
THIN LIQUID FILMS
6.83
Figure 6.44. Dependence of the permeability coefficient on surfactant concentration for NBFs from NaDS + 0.5 mol dm" 3 aqueous NaCl solution at four temperatures; solid lines, calculated using [6.7.11]. (Redrawn from M. Nedyalkov, R. Krustev, A. Stankova and D. Platikanov, Langmuir, 8 (1992) 3142.) are presented in fig. 6.44 for NBFs from aqueous NaDS solutions, at four different temperatures. Over a large concentration range the permeability coefficient remains constant, while at lower concentrations it increases sharply. This is in agreement with the theory and indicates that the permeability through holes in the bilayer increases considerably over the background permeability at concentrations below
4.3 x
10~4 mol dm" 3 and T = 25°C. Since this concentration is lower than the c.m.c, it can be considered that the K value determined at c s >4.3xlO" 4 mol dm" 3 is the coefficient of background permeability Ko across the homogeneous hole-free film part. Statistical treatment of these results, has been carried out over the entire range of surfactant concentrations, using [6.7.11] with all possible combinations of the other terms in [6.7.12], the summation limit being i = 4 . This treatment indicates that, under the given conditions, the holes consisting of three vacancies dominate the gas permeability, i.e., K = KQ + a 3 ( l / c s ) 3 . In such a case, the diffusion coefficient is D 3 = 7 . 2 x l O " 5 cm 2 s" 1 . This value is higher than the typical diffusion coefficient of air in water, but it is considerably lower than the diffusion coefficient in a gas. Hence, gas-transfer through three-vacancy holes is more difficult than in air. The gas diffusion coefficient through the hole-free homogeneous film part, DQ =1.4xlO~ 8 c m 2 s ~ 1 , is much lower than that for the typical aqueous films, and corresponds to a liquid film with viscosity of about 2.5 N s m~ 2 . This is completely understandable, bearing in mind that the bilayer structure of the NBF consists mainly of ordered surfactant molecules. The results on the temperature-dependence of K for CBFs and NBFs, although insufficient for definite conclusions, favour of the theory for an energy barrier for
6.84
THIN LIQUID FILMS
monolayer permeability. This theory is also suitable for the interpretation of the gas permeation through hole-free bilayer films (background permeability). The sharp increase in gas permeability of NBFs which is found at low surfactant concentrations, at which the film stability also decreases strongly, is evidence for the existence of spontaneous hole formation in the film. Hence, bilayer stability and permeability can be described by similar models: bilayers have the highest permeability and the lowest stability (to rupture) within a definite narrow range of surfactant concentrations. 6.7c Lateral diffusion in black phospholipid films Measurements of the lateral diffusion in microscopic phospholipid foam films-including black films-using fluorescence recovery after photobleaching (FRAP, see sec. 1.7.15), are of particular interest, because they provide an alternative model system for the study of lateral molecular mobility in biological membranes.
Figure 6.45. Dependence of the lateral diffusion coefficient D of adsorbed fluorophore molecules on film thickness h for microscopic thin liquid films from aqueous lecithin solutions; T = 24°C. (Redrawn from Z. Lalchev, R. Todorov, H. Ishida and H. Nakazawa, Eur. Biophys. J., 23 (1995) 433.) Systematic measurements of the lateral diffusion coefficients1'2' D of
fluorophore
molecules in the phospholipid adsorption layers of various types of foam films have been performed with microscopic films from aqueous lecithin solutions. The experimental D values are plotted as a function of film thickness h in fig. 6.45. For common thin liquid films a significant decrease in D with decreasing h was established for the range between 100 nm and 30 nm . A further decrease in D was observed for CBFs (D = 5 x 10" 8 cm 2 s" 1 ) and for NBFs ( D = 2.2 x 10" 8 cm 2 s" 1 ). The CBF has an equivalent water thickness of approximately 13 nm and consists of a liquid layer between the two adsorbed phospholipid layers according to the three layer film model (fig. 6.2). The value of the lateral diffusion coefficient in the thinnest bilayer NBF, approximately 7 to 8 nm thick in this case, and containing no free liquid core, is half of that in the CBF (fig. 6.45). Since the decrease of the film thickness reflects the decrease of the inner liquid layer thickness (at constant temperature) these results support the 11 21
Z. Lalchev, P. Wilde, and D. Clark, J. Colloid Interface Set 167 (1994) 80. Z. Lalchev, P. Wilde, A. Mackie, and D. Clark, J. Colloid Interface Sci. 174 (1995) 283.
THIN LIQUID FILMS
6.85
conclusion that reduction in the thickness of the liquid core of the phospholipid films slows down the lateral molecular diffusion. The results presented in fig. 6.45 are in agreement with the experimental data for the gas permeability coefficient of CBFs (sec. 6.7a), which show that the monolayer permeability is lower at smaller film thickness. The reason is that the normal interactions between the surfactant molecules make the monolayers better ordered and less permeable, so that their molecules are less mobile. Investigations of the influence of temperature on the lateral diffusion have allowed identification of the temperature at which a transition occurs in the mobility of the absorbed phospholipid at the film surface. The position of this transition from a largely immobile adsorbed layer to a state consistent with free molecular self diffusion depends on the chemical structure of the phospholipid. Comparison with the phase transition observed in the bulk phospholipid/water system allows one to conclude that the lower temperature threshold of measurable diffusion correlates with the onset of the formation of a lamellar liquid crystalline phase of the phospholipid. Therefore, the data support a correlation between the surface and bulk phase transitions. 6.8 Thin liquid films: a biomedical application An example will now be presented to demonstrate how the fundamental knowledge of thin liquid films can be used for solving practical problems, particularly in medicine. As is well known, a bilayer of phospholipid molecules between aqueous electrolyte solution phases (BLM) is widely used as a model for biomembranes (fig. 6.39b). As a model for the lung surfactant system, the bilayer Newton black film (fig. 6.39a) will be considered here, where the NBF is obtained from amniotic fluid, lung surfactant extract, or single phospholipids and their mixtures. 6.8a Model study of the lung surfactant system through black, foam films The lung is a unique biostructure. It is important that the alveolar surface is in contact with atmospheric air. This phase interface has its specific features. A fundamental question is to understand the particular structure, consisting predominantly of dipalmitoyl-phosphatidylcholine (DPPC), phosphatidylglycerol (PG), and specific proteins, which play very important roles in lowering the interfacial Gibbs energy of the alveolar surface and imparting the required rheological properties (sec. III.3.9). Contemporary science allows for physicochemical modeling of this structure and understanding the basic reasons for its formation, rheology and stability. In order to clarify this in vitro model of the alveolar surface we shall consider the scheme in fig. 6.46, which also presents the link 'in vitro model - in vivo structure'. A horizontal microscopic foam film under the ambient conditions of lung alveoli (capillary pressure Ap film radius r electrolyte concentration c and temperature T )
6.86
THIN LIQUID FILMS
Figure 6.46. Schematic presentation of an in vitro model and an in vivo structure.
has been proposed as a model of the alveolar surface1'2'. It is known that the alveoli radii are in the range of 30 to 200 (im, the surface tension amounts to 25 to 30 mN/m, and the capillary pressure is in the range of 103 Nm~2 to 102 Nm~2. At a baby's birth, Ap is between 0 and 300 Nm~2 (or 3 cm H2O column) which lies within the same range. These data suggest that the microscopic foam film is a very suitable model for the study of the alveolar surface and for the alveolar stability. The formation and stability of foam films is determined by molecular interaction forces which can lead to long range surface forces. The microscopic foam film under the conditions of lung alveoli is a bilayer film or NBF (fig. 6.39b), connected to the bulk liquid phase. A theory for the formation and stability of bilayer films has been developed, based on the short range interaction forces between nearest neighbour amphiphilic (e.g., phospholipid) molecules. The energies of lateral and normal bonds owing to these interactions, u and u 0 , respectively, determine U, the binding energy of a phospholipid molecule in the bilayer (sec. 6.6c). When the bulk surfactant concentration c s i.e. the phospholipid concentration, is changed, the stability of the bilayer film is also affected. This can be expressed in terms of the film lifetime r or by the probability W for observing a bilayer film (the probability for film formation), which are given by [6.6.8] and [6.6.10) (sec. 6.6c). According to 11 21
D. Exerowa, Z. Lalchev, Langmuir 12 (1996) 1846. D. Exerowa, Z. Lalchev, and D. Kashchiev, Colloids Surfaces 10 (1984) 113.
THIN LIQUID FILMS
6.87
these equations the dependence r(cs) or W(cs) is very steep. This permits one to define very precisely the initial concentration, called the critical concentration, cc , of bilayer film formation, as well as the concentration at which the probability for their formation is 100%, named the threshold concentration ct (fig. 6.47). The c c and ct values are specific for the different phospholipids mixtures and can be used as sensitive characteristics for the formation of bilayer films. Thus c c and ct are successfully used to characterize bilayer films from amniotic fluids1', or lung surfactant solutions, as models for studying the alveolar surface and stability. Previously, the basic in vitro physicochemical model used was the monomolecular one.
Figure 6.47. Dependence of the probability W for observing a foam bilayer at bulk phospholipid concentration cs (schematic presentation). A study of microscopic black films has been carried out in the measuring cell of Sheludko-Exerowa (fig. 6.7, variants A and C), in which a microscopic foam film with radius of about 100 |xm is formed in the middle of a biconcave drop (sec. 6.2c). The main advantage of these measuring cells is that one can work at very low surfactant concentrations, when the formation and stability of the microscopic film can be studied simultaneously. This is very important for the in vitro model of the alveolus. 6.8b Black foam film clinical test The development of fetal lung maturity tests is based on the use of parameters related to the suggested physicochemical in vitro model of the alveolar surface. The application of the parameters characterizing the formation and stability of foam bilayers or NBFs has proved to be most efficient and useful. As shown in sec. 6.8a, c c and c t are sensitive characteristics for the formation of bilayer NBFs. For amniotic fluids (AF) it is convenient to use the threshold dilution data d,. instead of the threshold concentrations ct since there is a linear dependence between ct and dj and between the critical values c c and dc . Figure 6.48 depicts the dependence d (c Dppc ) of an AF sample for a lung at the 32nd gestation week in the presence of a physiological solution of NaCl and 47% ethanol. The arrows indicate (a), the threshold dilution d and (b), the threshold concentration c t characterizing the stability of the AF bilayer. This is the very parameter which relates to the respiratory status of the neonates, i.e. it is a decisive indicator of whether they have the respiratory 11
A. Nikolova, D. Exerowa, Langmuir 12 (1996) 1846.
6.88
THIN LIQUID FILMS
Figure 6.48. Probability W for observing AF foam bilayers: (a) experimental dependence on dilution, d ; (b) experimental dependence on the phosphatidylcholine concentration, c s . distress syndrome (RDS) or not and hence this diagnostic tool has promoted the development of a new method for the assessment of fetal lung maturity. It has been found that dt = 3.1 is the threshold value that separates mature from immature samples, thus providing a precise and early estimation of lung maturity. In brief, the principle of the diagnostic test 1 ' is based on the relationship which exists between the rupture of the foam films (W = 0) and development of RDS in the newborn, as opposed to the 100% formation of bilayer NBFs, (i.e., above the threshold concentration c t and stable black films) for mature lungs. Statistics have been collected on a great number of AF samples at various gestation ages. Figure 6.49 presents the threshold concentration c t or d^ of AF along with a picture (in the microscope eye-piece or via a VCR) of a gray foam film, (a), which ruptures at lower concentrations, and a black bilayer film, (c), formed via black spots (b) at higher concentrations. It is clearly seen that c t (or dt , see also fig. 6.48) separates the mature from the immature AF samples. In the last case, RDS is developed in the newborn. It is worth noting that the comparison of the black foam film (BFF) clinical tests for the prediction of RDS with the largely used lecitin/svingomielin (L/S) method 2 ' shows good correlation. Furthermore, in the range of L/S ratio from 1.5 to 2.0 the BFF test gives better diagnostic prediction. The black foam film (BFF) test can also be used successfully for testing new therapeutic surfactants, such as infasurf, exosurf, and survanta. The advantages 3 ' of the BFF method are: high reliability (about 95 %), speed (= 30 min), the small quantity of AF sample needed (= 1 ml), ease of performance, and that the object (the model, in vitro) of the method represents the system in vivo.
11
D. Exerowa, Z. Lalchev, B. Marinov, and K. Ognyanov, Langmuir 2 (1986) 664. J. Clements, A. Platzker, D. Tierney, C. Hobel, and R. Creasy, JV. Engl. J. Med. 286 (1972) 1077. 31 M. Cordova, A.J. Mautone, and E. Scarpelli, Pediatr. Pulmon 21 (1996) 373. 21
THIN LIQUID FILMS
6.89
Figure 6.49. Schematic presentation of the relationship between threshold concentration ct and surfactant insufficiency/sufficiency, or lung immaturity/maturity. Photographs of microscopic foam films: a) unstable thin liquid film, b) black spots in a thin liquid film, c) bilayer Newton black film.
The response of the in vivo situation gives reason for the following conclusions: (i) The threshold dilution d^, or threshold concentration c t , (a measure of fetal lung maturity) corresponds to about 40 \±g cm"3 DPPC concentration in AF, which is close to the values of, 'critical DPPC concentration', known in the literature. It is also known that DPPC plays a main role in lung maturation and ensures its normal function. (ii) The curve of threshold dilution, c^, vs. the gestation period in weeks indicates an increase in DPPC, which corresponds to the increase in DPPC during fetus growth. (iii) The structure of the alveolar surface is not a monolayer at the hypophase/air interface but an ordered structure - most probably a bilayer structure, where lateral and normal short range interactions act between the adjacent amphiphilic molecules. These considerations undoubtedly indicate the value of model studies of the lung surfactant system and give reason for further development of the in vitro model so that a clearer description of the, in vivo structure can be achieved. It is interesting to relate these findings to the Theological ones on expansion-compression studies, alluded to in sec. III.3.9.
6.90
THIN LIQUID FILMS
6.9 General references 6.9a IUPAC Recommendations Definitions, Terminology and Symbols in Colloid and Surface Chemistry, D.H. Everett, Ed. Part I, Pure & Appl. Chem. 31 (1972) 579. Thin Films Including Layers: Terminology in Relation to Their Preparation and Characterization, L. Ter Minassian-Saraga, Ed. Pure & Appl. Chem., 66 (1994) 1667. 6.9b Books and reviews (entire or partly) on thin liquid films A.W. Adamson, A.P. Gast, Physical Chemistry of Surfaces, Wiley (1997). (A textbook which includes chapters concerning thin liquid films.) Foams, R.J. Akers, Ed., Academic Press (1976). (Proceedings of a conference on foam and foam films.) J.J. Bikerman, Foams, Springer-Verlag (1973). (Textbook.) V.C. Boys, Soap Bubbles, Dover (1959). (One of the first books on foam films reflecting the knowledge at the end of 19th century.) I.S. Clunie, J.F. Goodman and B.T. Ingram, Thin Liquid Films, in Surface and Colloid Science, Vol. 3, E. Matijevic Ed., Wiley (1971). (Review.) B.V. Derjaguin, Theory of Stability of Colloids and Thin Films, Consultants Bureau, New York (1989). (A monograph mostly on disjoining pressure and surface forces in thin liquid films.) S.S. Dukhin, N.N. Rulev and D.S. Dimitrov, Koagulyatsiya i dinamika tonkikh plenok, Naukovaya dumka, Kiev (1986) (in Russian). (A monograph mostly on kinetic properties and dynamics of thin liquid films.) D. Exerowa, D. Kashchiev and D. Platikanov, Stability and Permeability of Bilayers, Adv. Colloid Interface Set, 40 (1992) 201. (Review.) D. Exerowa, P.M. Kruglyakov, Foam and Foam Films, Elsevier (1998). (Monograph which extensively covers almost all themes of this chapter plus much about foams.) J.W. Gibbs, The Scientific Papers of J. Willard Gibbs, Dover (1961); J.W. Gibbs,
Collected Works, Longmans Green, (1928). (A book of greatest importance for the thermodynamics of thin liquid films.) The Modern Theory of Capillarity, F.C. Goodrich, A.I. Rusanov, Eds. AkademieVerlag Berlin (1981). (Collection of papers on interfacial phenomena including liquid films dedicated to the Centennial of Gibbs' Theory of Capillarity.)
THIN LIQUID FILMS
6.91
Thin Liquid Films, I. Ivanov, Ed., Marcel Dekker (1988). (15 chapters by different authors cover different aspects of thin liquid films.) P.M. Kruglaykov, Yu.G. Rovln, Fizikokhimiya chernykh uglevodorodnykh plenok, Nauka, Moscow (1978) (in Russian). (A monograph on specific properties of black hydrocarbon liquid films.) K. Mysels, K. Shinoda and S. Frankel, Soap Films, Pergamon Press (1959). (A monograph on foam films summarizing the knowledge in the middle of 20th century.) A.I. Rusanov, Phasengleichgewichte und Grenzjldchenerscheinungen. AkademieVerlag Berlin (1978). (Monograph, significant parts are devoted to thin liquid films.) A. Scheludko, Colloid Chemistry, Elsevier (1966). (A textbook which includes chapters concerning thin liquid films.) A. Scheludko, Thin Liquid Films. Adv. Colloid Interface Set, (Review.)
1 (1967) 391.
H.T. Tien, Bilayer Lipid Membranes (BLM), Marcel Dekker (1974). (Monograph.)
This Page is Intentionally Left Blank
7
FOAMS
Vance Bergeron and Pieter Walstra 7.1 Definition 7.1 7.1a Foam structure 7.1 7. lb Wet and dry foam: spherical bubbles versus polyhedral structures 7.3 7.1c Foaming agents 7.6 7.2 Foam formation 7.7 7.2a Principles 7.8 7.2b Various methods 7.11 7.3 Foam stability 7.14 7.3a Drainage 7.14 7.3b Coalescence 7.17 7.3c Ostwald ripening 7.18 7.4 Foam characterization 7.21 7.5 Foam properties 7.24 7.5a Rheology 7.24 7.5b Optical properties 7.27 7.6 Antifoam and defoaming 7.28 7.6a Strategies 7.29 7.6b Defoaming and antifoam for formulations 7.31 7.6c Form destruction by defoaming and antifoaming formulations 7.32 7.6d Antifoam performance lifetimes 7.34 7.7 Applications 7.34 7.7a Industrial applications 7.34 7.7b Food 7.36 7.7c Detergents 7.36 7.7d Cosmetics 7.37 7.7e Miscellaneous 7.37 7.8 General references 7.37
This Page is Intentionally Left Blank
7 FOAMS VANCE BERGERON AND PIETER WALSTRA
7.1 Definition Foam, in the most general sense, constitutes any material whose structure consists of gas cells dispersed in a continuous phase of liquid, solid or a gel. However, in the realm of colloid science, foam is most often considered to be a dispersion of gas in a liquid phase, and that is the system to be considered here. Although the dispersed gas bubbles in foam are typically macroscopic (i.e. > 10 (im ) and considered as beyond the colloidal range, the thin liquid films that are situated between adjacent gas bubbles in a foam can attain thicknesses of some nanometers. Moreover, their existence, as well as that of the foam, relies primarily on colloidal and surface forces. In other words, foam is a macroscopic gas dispersion, whose existence and properties are controlled by colloidal and surface forces and interactions across the films that separate Individual gas bubbles. Furthermore, it should be noted that foam is a non-equilibrium dispersion. Given long enough time, complete separation of the gas and liquid phases will ensue. As such, understanding 'foam stability' is fundamental to the understanding of foam properties and applications. We have dedicated chapter 6 to the study of isolated 'thin liquid films' and their stability. In the present chapter we focus on the macroscopic aspects of foams, i.e. including assemblies of thin films, and show how these aspects are tied to the colloidal and surface forces that create this unique type of dispersion. Unless explicitly noted, the continuous phase is aqueous.
7.2 a Foam structure Observed from a distance, foam produced from a transparent or coloured liquid appears homogeneous and white. However, upon closer inspection the intricate structure formed by the close packing of individual bubbles becomes apparent. Figure 7.1 illustrates several features of the microstructure common to most liquid phase foams under the influence of gravity. This photograph portrays foam generated by bubbles rising through a surfactant solution. Near the top of the foam, most of the liquid has drained away under the force of gravity, leaving a so-called dry Joam consisting of nearly polyhedral gas bubbles separated by thin liquid films of relatively uniform thickness. At the bottom of the foam, we find bubbles that are more or less
Fundamentals of Interface and Colloid Science, Volume V J. Lyklema (Editor)
© 2005 Elsevier Ltd. All rights reserved
7.2
FOAMS
spherical in shape and the foam is considered wet due to the larger water content present. Sometimes these two types of foam are called old and young, respectively, or distinction is made between polyhedral and bubble foam (or Kugelschaum, a term of German origin). Even further down in the photograph we see individual bubbles rising through the solution in what can be considered a simple gas dispersion. See also fig. 1.1.7.
Figure 7.1 Photograph and schematic image of an aqueous foam. The arrows depict light waves that are scattered (reflected and refracted) by the bubble interfaces, which create the white image of foam. Smaller bubbles generate more scattering and the foam appears whiter. Accompanying the photograph in fig. 7.1 is a scheme illustrating why foam appears white at a distance. Light from a multiple wavelength broad spectrum source (so-called white light) is shown as a series of rays labelled as I o that are impinging on the foam. Once the rays enter the foam they are partly reflected by the bubble interfaces and thin films that comprise the foam. These multiple reflections in all directions scatter the incident light efficiently, providing for the white image. We note that when the bubble sizes are small, more interfaces are present, thus providing more reflected light and a whiter image of the foam. This is why shaving cream foam, with micron-sized bubbles, appears much whiter than our dishwashing foam. The non-equilibrium nature of foams is also revealed in fig. 7.1. Initially all the bubbles to generate the foam were of about uniform size. However, we clearly see a wide size distribution of the bubbles in the foam. Moreover, the larger bubbles tend to be present near the top of the foam. This distribution is due to gradual formation of larger bubbles by coalescence and their subsequent segregation under the influence of gravity. Liquid drains downward while the buoyancy force of large bubbles pushes them to the top. Bubble growth is controlled by two effects: film rupture, leading to coalescence, see sec. 7.3b, and diffusion of gas molecules through the liquid from
FOAMS
7.3
small bubbles to larger ones. The latter effect is called Ostwald ripening, introduced in sec. 1.2.23c, and which will be discussed in sec. 7.3c. The rupture of individual films has been treated in chapter 6. No matter which of these processes dominates, the final outcome is the same. At equilibrium there is no foam, only one region of consolidated liquid and another of gas. The physical chemistry of the interfaces and the foam structure are the primary factors controlling the relative rates of the ageing mechanisms. In practice we find a broad spectrum of foam lifetimes, ranging from seconds, as with champagne foams, to hours or even days with foam produced by carefully formulated surfactant solutions. To obtain an impression of some physical properties, two foams and two types of food emulsion are compared in table 7.1. Table 7.1 Orders of magnitude of quantities in foams compared with food emulsion,
at room temperature a) Property
Foam
Foam B
Emulsion w/o
Emulsion
A
Drop/bubble diameter
103
102
3
0.5
um
Volume fraction Drop/bubble number
0.9
0.9
0.1
0.1
-
109
10 12
10 16
10 18
m"3
Interfacial area
0.005
0.05
0.2
1.2
m 2 ml" 1
Interfacial tension
40
40
6
10
mN m" 1
Laplace pressure Solubility D in C Density difference D-C Viscosity ratio D/C Typical time scale a)
10
2
10
b)
3
10
b)
4
10
2.1 -103 lO- 4
2.1 -103 lO- 4
0.15
3
4
5
io-
lO-
lO- 2 10-
5
c)
o
-102
102
Units
o/w
Pa % v/v kgm~ 3
102
io- 6
s
Oil phase is a triacylglycerol oil; gas phase is air D = disperse phase, C = continuous phase. If the gas phase is CO 2 , the solubility is about 100% v/v, but strongly dependent on composition of the aqueous phase, especially pH and salt composition. Oils often contain minor components that are somewhat soluble in water, but the solubility of the triacylglycerols is generally negligible. Time needed for separate events during formation, e.g. the deformation time of a drop or bubble.
7.1b Wet and dry foam: spherical bubbles versus polyhedral structures Foam structure is characterized primarily by the ratio of liquid to gas present in the foam. This ratio is typically quantified using the gas volume fraction cp . As already mentioned, foams with a low gas volume fraction are considered to be wet foams. Once the gas fraction in foam becomes high enough, bubbles are forced into contact and form polyhedral structures, leading to dry foam. There is no formal criterion that distinguishes wet from dry foam. However, it is generally accepted that a system having a gas fraction near 0.63 (the limit for randomly close-packed spherical objects) is a
7.4
FOAMS
very wet foam, while gas fractions exceeding 0.8 constitute dry foam. In certain systems, extremely dry conditions prevail where gas fractions can surpass 0.95. The dryness of the foam has a major influence on its mechanical and Theological properties, and this will be discussed in detail later. In the 19th century, the Belgian Joseph Plateau extensively studied the structure and form of bubbles and films that comprise foam. Plateau's meticulous observations have led to what are now referred to as the first laws of foam geometry or Plateau rules. These rules are a consequence of the fact that the liquid films organize themselves to minimize their surface area for a given gas volume. This area minimization is driven by the surface tension of the air-liquid interfaces and results in a lowering of the Gibbs energy of the system. The three basic rules developed by Plateau are: (i) three smooth surfaces of a soap film intersect along a line; (ii) the intersection formed by three bubbles coming into contact will produce borders that have an angle between them of 120°; (iii) the vertices generated when four borders come together forms a tetrahedral angle of 109.5°.
Figure 7.2 Polyhedral foam structure, (a) Cross-section of three bubbles in contact. Three films join at a Plateau border. The angle between two films is 120°. (b) Enlargement of the cross-section near the Plateau border. R is the radius of curvature of the latter. The pressures p at various positions are also indicated, (c) The vertex where four Plateau borders meet. The angle between borders is ideally 109.5°. Figure 7.2 provides schemes of these geometrical conditions. In honour of his contributions, the transition zone formed when three films are in contact is defined as the Plateau border. Plateau borders constitute the structural skeleton of the foam. Plateau's rules lead to the following guidelines for the polyhedral structure of the bubbles in a foam: only sets of four bubbles can be in mutual contact, all four bubbles share a common vertex, each of the four combinations of three bubbles share a common Plateau border, each of the six combinations of two bubbles share a common film, and the angles between pairs of films and of borders are 120° and 109.5°, respectively. According to these local structural rules, it is impossible to construct a
FOAMS
7.5
regular, periodic, polyhedral foam from a single polyhedron. No known polyhedral shape has been found that can be packed into a space-filling structure while simultaneously satisfying the intersection rules required of films and borders. As such, there is no simple model structure that can serve as a convenient mathematical idealization of polyhedral foam. The closest polyhedral structural element that nearly satisfies all the mechanical constraints is Lord Kelvin's minimal tetrakaidecahedron cell, known as the Kelvin cell, which is pictured in fig. 7.3.
Figure 7.3 Two Kelvin cells in contact, providing a polyhedral structure that minimizes the surface area for a given gas volume.
In a real foam, the system has further degrees of freedom for establishing local mechanical equilibrium, notably the bending propensity of films and Plateau borders. The curvature in such cases is due to a pressure difference Ap between adjacent bubbles, as described by Laplace's law, which depends on the film tension y{ and the principal radii of curvature Rx and R2 Ap = r f ( R f 1 + R 2 1 )
[7.1.1]
The pressure inside a given bubble must be the same everywhere; therefore, at the facets of the bubble where the surface is nearly flat, a force balance is maintained by the disjoining pressure (see below), which must balance the Laplace pressure in the regions of high curvature; see sec. 6.5. Although a pressure difference can exist between adjacent bubbles and between the gas and liquid phases, the pressure throughout the continuous liquid phase (i.e. films, borders and vertices) must be constant at constant height in the foam. When there is no mechanical equilibrium, liquid flow will occur until it is reached. The Laplace pressure in the Plateau borders is negative, as shown in fig. 7.2b, and is given by ylR^; it concerns a cylindrical surface, hence R2 = °° • The pressure is balanced by the pressure due to gravity, which leads to Rx=ylpgH
[7.1.2]
where p is density of the liquid, g the acceleration due to gravity and H the height in
7.6
FOAMS
the foam above the bulk liquid. This means that the radius of curvature of, and thereby the amount of liquid in, a Plateau border will be smaller at a greater height in the foam. If the bubbles are not very small, nearly all of the liquid in the foam will be in the Plateau borders. The density of the foam relative to that of the liquid will then be given by 1 ' prel=O.5(y/pgHq)2
[7.1.3]
where q is the length of the Plateau border, roughly corresponding to the bubble size. This implies that for bubbles of 1 mm, and at a height of 3 cm, the relative density of the foam would be on the order of 0.01 , i.e. very small. It will take a long time, however, before the foam will have drained to reach mechanical equilibrium. 7.2 c Foaming agents Foaming agents consist wholly or partly of surfactant(s). Without a surfactant a foam cannot be made or kept. The surfactant has three main Junctions. (i) Its adsorption allows the formation of a surface tension gradient, whereby tangential movement of the surface by an external shear stress is greatly retarded or even arrested; see sec. 8.1c. This is all that makes foam formation possible. First, the gradients formed upon flow enormously slow down drainage and thinning of foam lamellas, as illustrated in fig. 7.4a-b. Second, the Marangoni effect (a surface tension gradient induces flow of the bordering liquid) is essential in the Gibbs mechanism of film stability illustrated in fig. 7.4c; if a thin spot develops rapidly in a film, it causes local increase of the surface tension, which in turn creates a gradient, and the resulting flow 'repairs' the thin spot. The main variable governing these effects is the magnitude of the surface dilational modulus K^ .
Figure 7.4 Foam lamellas. (a) Downflow of water (W) between two air bubbles (A) in the absence of a surfactant, (b) Same, in the presence of a surfactant, which causes a surface tension gradient, (c) Gibbs mechanism of film stability. Surfactant molecules are depicted by short lines. The arrows indicate motion of the surface and the bordering liquid. From P. Walstra, Physical Chemistry of Foods, Dekker (2003). ' This and some other relations for a foam are given by J. Lucassen in E.H. Lucassen-Reynders, Ed. Anionic Surfactants: Physical Chemistry of Surfactant Action, Marcel Dekker (1981) p. 217.
FOAMS
7.7
(ii) The surfactant adsorption layers provide colloidal repulsion between both surfaces of a film, which often is paramount in counteracting bubble coalescence (see sec. 7.3b). (iii) The surfactant lowers the surface tension, and hence reduces the driving force for Ostwald ripening (see sec. 7.3c). The surfactant should have the following properties. It must be well soluble in water, and should not be volatile. It must be quite surface active, i.e. it must give a substantial surface pressure at a low bulk concentration. The magnitude of K^ must already be significant at a relatively low value of the surface excess F, and K^ should not vanish at high surfactant concentration (see fig. 8.4b). It is quite difficult to find surfactants that fulfil all of these requirements. Hence, most foaming agents are surfactant mixtures. Most surfactants used in foams are small-molecule amphiphiles. The hydrophobic part is generally an aliphatic chain. The hydrophilic part can be ionic, which can provide electrostatic repulsion; non-ionics have to provide steric repulsion, generally caused by one or more protruding poly(oxyethylene) chains. These surfactants generally give a surface tension of about 40 mN m" 1 . The concentration needed should be at least sufficient to produce a plateau value of the surface excess on all bubbles. Assuming a foam of
7.8
FOAMS
Unfortunately, fundamental investigations on foam formation, where the effects of several variables on foam properties are systematically studied, are scarce. Because much more is known about emulsion making, and several aspects are comparable for emulsions and foams, the reader Is referred to sec. 8.2 on emulsion formation for much of the basic aspects. Here, points that are typical for foam making and properties will be discussed. First, some general principles are treated, followed by a brief overview of methods applied in practice. 7.2a Principles Three main methods for foam making are used, i.e. supersaturation of the gas phase, by gas bubble Injection, and by agitating gas-liquid mixtures. (i) Supersaturation. Under pressure a gas can be dissolved in a liquid; when the pressure is released, the gas becomes supersaturated and gas bubbles can form. The gas should be well soluble in the liquid, usually water, to obtain a sufficient volume of bubbles. Carbon dioxide and laughing gas (N2O) are suitable. Pressures up to about 8 bar are applied. CO2 can also form in situ by fermentation, as in beer, champagne and bread dough. The bubbles do not form spontaneously by homogeneous nucleation. The critical radius of an embryo gas bubble that can spontaneously form is at most 2 nm . Such a bubble will have a Laplace pressure of about 70 MPa (700 bar). Assuming ideal behaviour, the supersaturation ratio of the gas should thus be about 700 for such a small bubble to survive, an impossibly high value in virtually all practical situations. Hence, some form of 'seeding' is needed. This can occur in a number of ways. First, air bubbles can be entrapped during the transfer of the pressurized liquid to another vessel, say a beer glass. Generally, this leads to quite large bubbles. Second, there can be persistent remnants of air bubbles that have shrunk by Ostwald ripening until the surface is fully covered by solid particles that had adsorbed on the bubble by chance; this is illustrated in fig. 7.5a. Such bubble remnants are found, for instance, in all natural waters, but in small numbers. Third, there may be gas pockets remaining in crevices in the vessel wall, provided that the contact angle is as depicted in fig. 7.5b, in which case the Laplace pressure in the gas pocket is negative. Upon coming into contact with a supersaturated liquid, gas diffuses to the pocket and bubbles are formed, as illustrated. It has also been observed that air pockets remain in textile fibres originating from a cloth used to dry a champagne glass. Since a pocket remains after a bubble has grown and is dislodged, generally due to its buoyancy, one pocket can produce numerous bubbles. Altogether, it is quite difficult to obtain small gas bubbles (< 1 mm) or to control bubble size via supersaturation and pressure release. The amount of foam obtained is determined by the supersaturation of the gas, hence by the pressure in the liquid. (ii) Injection. Gas can be injected in a liquid via narrow capillaries, whereby
FOAMS
7.9
bubbles are directly formed at the end of each capillary. This occurs in a simple laboratory method where gas is injected through a piece of sintered glass; see fig. 7.7b. The principles are the same as those of 'membrane emulsification'; the reader is referred to sec. 8.2a, sub (i), for the conditions allowing bubble formation. In principle, the bubble size is governed by the diameter of the capillaries, but immediate coalescence of newly formed bubbles will occur if: (1} the capillaries are too close to each other; (2) too little surfactant is present; or (3) the gas flow rate through the capillaries is too high. The latter implies that foam making by injection is a slow process. Application of membrane emulsification apparatus for industrial foam formation is presently in a stage of development.
Figure 7.5 Gas pockets in a liquid. G is a gas, W is water, and hatching indicates a solid. For explanation, see text. In (c) the contact angle as measured in the water phase is too small to allow the gas pocket to survive. (Redrawn from P. Walstra, Physical Chemistry of Foods, Marcel Dekker (2003).)
(iii) Agitation. This implies that mechanical energy is applied to the liquid in the presence of a gas phase. It is the most common method. It occurs in various forms, as illustrated in fig. 7.7c-f. In industrial processing, a form of beating by one or more rotating elements is generally applied. Again, the reader is referred to the emulsion chapter, parts of sec. 8.2, for fundamental aspects. The following phenomena or process steps that occur during agitation can be distinguished. All steps occur numerous times and during most of the process all of them occur simultaneously. 1. Bubble formation occurs by entrapment of gas pockets in the liquid. These will soon attain a more or less spherical form; the resulting bubbles tend to be quite large. 2. Transport of surfactant to, and its subsequent adsorption on, the a-w surfaces.
7.10
FOAMS
This lowers the surface tension, which somewhat promotes step (3) and strongly counteracts step (4). The rate of transport Is proportional to surfactant concentration. Bubble formation by agitation is a much faster process than formation by supersaturatlon or Injection, and the time scale needed for surfactant to adsorb Is an important variable. See sec. 8.2b, sub (l)-(il). 3. Breakup of bubbles. The agitation causes external stresses to act on a bubble, causing its deformation; If the external stress Is larger than the Laplace pressure of the bubble, the latter tends to break up Into smaller ones. Commonly, agitation causes intensely turbulent flow. Moreover, the forces acting on the bubbles will be predominantly inertial; see sec. 8.2a sub (iii). The dominant regime will then be TI (turbulent flow, Inertial forces) and the resulting bubble size would be given by [8.2.12]. The power density e, i.e. the rate at which mechanical energy is dissipated In the system, Is the parameter characterizing the intensity of agitation. For most foam making operations e ranges between 104 and 107 W m" 3 . Assuming the surface tension during break-up to be 50 mN m" 1 , the range of bubble diameters obtained would be between 1 mm and 65 um. As an order of magnitude, these values agree with experimental results, but smaller diameters can also be observed. This may be due to the high volume fraction of air bubbles that results after some agitation. The regime then is no longer TI, but is closer to 'bounded flow' (sec. 8.2b). The breakup in such a case is insufficiently understood; see sec. 8.2e for a discussion. 4. Coalescence of newly formed bubbles. This is primarily counteracted by the formation of surface tension gradients; see sec. 8.2c sub (vi). Coalescence will always occur to some extent, but less so at higher molar concentrations of surfactant. This means that small-molecule amphlphiles will be more efficient than polymers at the same mass concentration. Generally, a constant bubble size Is obtained after prolonged agitation, indicating that a steady state of bubble breakup and coalescence is attained. If the agitation intensity allows formation of a larger bubble surface area (large
FOAMS
7.11
Presumably, the rate at which the surfactant can cause a significant surface pressure on a bubble surface is the main variable. Figure 7.6a also shows that above a certain value of the beating speed, the amount of foam may decrease again. The following explanation has been offered11. The motion of the beating rods causes a local pressure decrease, according to Bernoulli's law (p + pv2 II = constant); hence, bubbles can locally expand, according to the gas law ( pVIT = constant) and they are locally pressed close together, whereby a film is formed. This film is enlarged by biaxial extension; hence, the surface excess on the film surfaces and, hence, the surface dilational modulus decreases so the film can rupture and the bubbles coalesce. Such coalescence would then occur more readily for a higher beater speed and a lower surfactant concentration, as observed. Other workers have observed a similar sudden decrease in foam volume above a critical beating rate for other surfactants (including proteins).
Figure 7.6 Examples of foam amount obtained by beating, (a) The amount of foam obtained, V , in litres out of 275 ml of a nonylphenol surfactant (CgiJ)Ej4) solution, as a function of beater speed v , for solutions of various concentration (indicated in nig per ml). (Redrawn from A. Prins, loc. cit.) (b) The overrun
In a closed system, e.g. a flow-through apparatus, the flows of air and liquid can be set beforehand, to obtain the desired overrun. Of course, the agitation intensity and the surfactant concentration must be sufficient to realize formation of a 'stable' foam. The bubble size distribution cannot be controlled in a similar manner. Smaller bubbles are obtained for a more intense agitation; agitation during a longer time (until a plateau value is obtained); and for a higher surfactant concentration (here also a plateau value can be reached). All agitation methods tend to produce a wide size distribution, especially when the surfactant concentration is relatively small. 7.2b Various methods Several methods for making foam have been devised and fig. 7.7 illustrates the 11
A. Prins, in R.J. Akers, Ed., Foams, Academic Press, (1976), pp. 51-60.
7.12
FOAMS
most important ones. All of these can be used on a laboratory scale and some, especially (d) and (e), also on an industrial scale. Several applications, often including small-scale foam making, occur in the household: pouring beer, whipping egg white, applying shaving soap, etc. Especially when studying foam making, stability, or specific applications, it will be useful to do this on a type of foam and in a manner that correspond with the practical situation. Since the methods are widely variable in the foam produced, a brief synopsis is provided to help as a guide for their utilization.
Figure 7.7 Six different methods commonly used for generating foam (i) Release of dissolved gas (see sec. 7.2a, sub (i)). A method common to many of us is the generation of foam through the release of dissolved gas. The most familiar example of this method occurs when pressurized carbonated beverages spout foam out of the top of their containers when opened to atmospheric pressure. Another illustration is the formation of a whipped topping from an aerosol can. This type of foam generation can be a nuisance during fermentation processes; gases produced during fermentation are dissolved under pressure and released into a solution of proteinaceous material that can generate vast amounts of undesired foam. (ii) Gas bubbling (Bikerman test; see sec. 7.2a, sub (ii)). A method that is widely used to generate foam for academic studies is to disperse gas by means of controlled bubbling into the foaming solution. With a porous disk or an injection tube, gas is
FOAMS
7.13
introduced directly into the solution. Bikerman extensively used this method to evaluate the stability of foam, and such methods of this type often bear his name. In some cases authors distinguish between dynamic and static regimes for such methods, depending upon the rate of gas supplied. The static regime corresponds to foam formation under low gas flow rates, whereby the surfactant is given enough time to establish equilibrium adsorption onto the bubble surfaces. In contrast to this, in the dynamic regime the bubble generation rate is too fast and the adsorption equilibrium condition is not generally fulfilled. A major advantage with this type of method is that, when using a single bubble tube in the static regime, one can produce nearly monodisperse foams because each bubble released in solution has the same size. The vessel dimensions (size and shape) and quantity of solution are important, but the gas flow rate and the pore size (or diameter of the injection tube) are the primary control variables. (iii) Shaking. Simple shaking of a surfactant solution in the presences of a gas phase will produce a foam. This is one of the easiest and most evident methods to generate foam. However, in many cases this method is not very practical (particularly for industrial applications) and the type of foam produced is often very polydisperse and hard to reproduce. The later constraint can be greatly improved upon if a mechanical shaking device is used to ensure reproducible conditions. In addition to changing the amplitude and speed of the agitation, one can use different vessels or levels of solution to vary the foaming conditions. (v) Beating. Another simple method to create foam is to blend the gas and foaming solution with a mixing device. This method is well known to any one who has spent time in a kitchen beating foam from egg-whites for a host of different recipes. For industrial applications, this method is more adjustable than mild shaking because of the wide range of mixers available. However, as with weak agitation, the foam produced is generally quite polydisperse and a strict protocol should be used if one wishes to reproduce specific foams. Key parameters are the beating speed, beating time, type of mixer, size of the vessel, quantity of solution and type and concentration of the surfactants. (vi) Co-injection. Foam can also be produced when gas and the foaming solution are co-injected into a porous material or injection tube. Both the relative and the absolute flow rates of the gas and liquid are important. The type and size of packing in the injection tube or porous material also play a major role in the properties of the foam formed. This method is commonly used in the cosmetic industry for generating shaving cream foam. Many devices used to generate fire-fighting foams also employ coinjection strategies. (vi) Pouring (Ross-Miles). One of the best known laboratory methods for generating foam comes from the beverage pouring industry and is known as the Ross-Miles test (for example, the standard test method ASTM Dl 173-53: "Test Method for Foaming Properties of Surface Active Agents"). This test consists of pouring a given amount of foaming solution into an empty vessel and subsequently monitoring the foam generated
7.14
FOAMS
and Its decay rate. There exists a certified procedure to carry out the test, which serves as an Industrial standard, but many variations have emerged and proved useful for different applications. In one well-used version the foaming solution is recycled, letting it continuously fall Into a horizontal vessel. In this case, the pumping rate, tube sizes, height from which the solution falls and the size of the vessel are the main controlling parameters. 7.3 Foam stability One of the most Important requirements for the application of a foam is its stability characteristics. In some cases highly stable foam is desired, such as for whipped cream and shaving foam, while in other instances transient foams are sought, for example In the short-lived foam at the top of freshly poured champagne. In general, there are three primary modes by which foam structure evolves over time resulting in its inevitable destruction: liquid drainage out of the foam; film rupture between adjacent foam bubbles, causing bubble coalescence; and gas diffusion between neighbouring bubbles, i.e. Ostwald ripening (often referred to as foam dlsproportionation), which also implies an increase in bubble size. It is far more difficult to make a 'stable' foam than a 'stable' emulsion. The lifetime of many foams is on the order of an hour, whereas a typical emulsion can be kept for months without appreciable change In droplet size. This difference is for the most part due to the difference In particle size; over 100 um for most foams, about 1 um for many emulsions. In a foam, drainage is nearly always a fast process, at least initially, and it often leads to leakage of liquid from the foam. When the bubbles are relatively small, Ostwald ripening is the most important process causing coarsening, whereas coalescence tends to be the dominant process when they are large. The three changes affect each other. Drainage leads to thinner films, hence to a higher probability of film rupture; Ostwald ripening leads to larger bubbles, which also implies a decreased stability to coalescence; on the other hand, coalescence causes slowing down of Ostwald ripening. 7.3a Drainage The difference in mass density between water and gas is quite large, about 103 kg m~3 ; the density of the gas can generally be neglected. Since, moreover, the gas bubbles are quite large, they will rapidly cream and form a packed layer on top of the liquid whence liquid will drain under gravity and the bubbles will deform each other. Liquid will then flow from the films to the adjacent Plateau borders, which are interconnected and form a network of channels that allows further drainage. This has already been discussed in sec. 7.1; see especially fig. 7.1. For drainage, the following three features are relevant: (i) Equilibrium structure. If drainage proceeds unhindered, a structure will develop
FOAMS
7.15
that is in mechanical equilibrium. The films will drain until the colloidal disjoining pressure equals the gravitational stress, pgH , where H is the height in the foam layer. The equilibrium curvature of the Plateau border is reached when the Laplace pressure in the border equals the gravitational stress; see [7.1.2]. Assuming that the amount of liquid in the films is negligible compared with that in the Plateau borders, the foam density is given by [7.1.3]. It is seen that the density strongly depends on q , hence on bubble size (= q). For bubbles of 1 cm or larger, the (hypothetical) equilibrium foam would hardly contain liquid. For bubbles of 0.1 m m , the foam would contain very little gas; however, [7.1.3] does not apply for a bubble size smaller than about 0.3 mm . Consequently, only for a foam with small bubbles can the initial air content be high enough to allow the formation of an equilibrium structure without leakage of the liquid from the foam. The density distribution over the height of the foam will then roughly be as given by [7.1.3]. The actual structure is somewhat different because the bubbles are likely to be polydisperse. This also causes the average bubble size to be higher near the top than near the bottom of the foam layer. If the initial foam has large bubbles or a low air content, drainage will lead to leakage of liquid from the foam. An aqueous layer forms and the boundary between liquid and foam will move upward until equilibrium is reached. In practice, however, an equilibrium structure is rarely obtained, because film rupture and Ostwald ripening will occur during and after drainage. (ii) Drainage rate. Calculation of the drainage rate in a foam is a formidable task and full solutions have not been given. This depends on three main variables: the viscosity rj of the liquid; the continuously changing geometry of the foam; and the Marangoni number Ma, which determines whether the film surfaces are rigid. A highly simplified case will be discussed first1'. Consider a vertical film at a height H in the foam of thickness h and height and width q . The film surfaces are assumed to be rigid. The pressure difference between the film and the adjacent Plateau borders equals pgH . Under these conditions, the average linear velocity of the liquid flowing out of the film will be given by the StephanReynolds equation [6.4.2], from which (v) = 2pgHh3/3q2r]
[7.3.1]
As (v) = -dh / d t , the time needed for the film to drain to a thickness h can be derived by integration t(h) = 3q2rj/4pgHh2
[7.3.2]
Assuming q = 10" 3 , r] = 10" 3 , p = 1 0 3 , g = 10 , and H = 0.1, all in SI units, drainage to a film thickness of 1 pun would take about a second, and to 20 nm, i.e. a distance where Van der Waals attraction over the film becomes appreciable, about half an hour. 11
Drainage, and related phenomena for isolated foam lamellae have been discussed in some detail in chapter 6.
7.16
FOAMS
A prerequisite for [7.3.2] to hold is that the film surfaces are indeed rigid; if not, drainage proceeds much faster, as discussed in relation to figs. 7.4a and b. The distance over which the liquid has to flow must then be small enough to produce a substantial surface tension gradient. This comes down to q<2nipgh
[7.3.3]
where n is the surface pressure. It is readily calculated that this condition is always fulfilled in a foam. Moreover, the Marangoni number, as defined in [8.1.4], has to be much larger than unity. It is given by Ma = K%h 12qr)(v)
[7.3.4]
where Kg is the surface dilatlonal modulus; in a thin film or narrow Plateau border its value will nearly always be above a few mN m" 1 . Sample calculations show that then the condition Ma » 1 is always fulfilled. However, [7.3.2] is not valid for a real foam. Drainage tends to be much slower, often by an order of magnitude or more. The actual situation is far more complex1'. After a while, most of the liquid may originate from a decrease in the cross section of plateau borders, rather than from the thinning of films. The films are not vertical, but are oriented in any direction, as are the plateau borders. The plateau borders will also cause a considerable resistance to flow, which implies that the pressure difference between film and plateau border will be smaller than in the near equilibrium situation that was implicitly assumed in deriving [7.3.1 ]. The flow resistance is further enhanced by the inhomogeneity of the structure due to the polydispersity of the bubbles. Moreover, bubbles may coalesce and show Ostwald ripening during drainage. Nevertheless, the trends given in [7.3.2] appear to be well obeyed. Drainage time strongly increases with decreasing bubble size, and drainage slows down progressively as film thickness decreases. The time needed for drainage to a given thickness is shorter at a higher position in the foam layer. In practice, some successive drainage regimes often are distinguished. The first regime is the formation of a polyhedral foam. It can proceed quite rapidly, as the films are still quite thick and the condition Ma » 1 may not yet be fulfilled. The second regime Is characterized by a constant drainage rate, dV/dt = constant, where V is the volume of liquid that has drained from the foam. This regime gradually changes into a regime given by V = V0-\/kt
[7.3.5]
where Vo is the volume of liquid originally in the foam and k is a rate constant (in
1 ' An extensive treatment is given by G. Narsimhan, E. Ruckenstein, chapter 2 in R.K. Prud'homme, S.A. Khan, Foams (see General references).
FOAMS
7.17
m~3 s~ ! ). The drainage rate then is proportional to t~2. Eventually, an exponential V(t) relation is observed. Altogether, it takes a long time before something like mechanical equilibrium in the remaining foam is attained. (iii) Counteracting drainage. For many purposes, it is undesirable when a foam shows rapid drainage to a very high value of
7.3b Coalescence Coalescence can be a desired process (when the presence of a foam is a nuisance) or it can be detrimental (when foam is desired). For control of either case, it is very useful to understand the coalescence process. A film between two bubbles can rupture, which will lead to immediate coalescence of these bubbles into a larger one. Likewise, a bubble on top of a foam can coalesce with the air above, thereby disappearing. When these processes take place, the foam will eventually disappear. This will occur faster for larger bubbles. The probability that a film ruptures within a given time is, in most cases, about proportional to its area, hence to bubble radius a squared. Moreover, the number of rupture events needed to destroy a foam will be proportional to the number of bubbles present, hence to a 3 . This implies that the process accelerates and that coarse foams tend to have a short lifetime. Coalescence causes the surface area to decrease, hence leading to a decrease in interfacial Gibbs energy (although not necessarily in proportion to the decrease in area). This provides a driving force for film rupture, implying that the film is thermodynamically unstable. In practice, some films may, however, have a substantial lifetime, which means that an activation Gibbs energy barrier for rupture must exist. In cases where a foam has to be destabilized, the mechanism by which the barrier can be overcome must thus be established. Three situations can be distinguished. (i) Relatively thick films, which generally implies young films, are stable against rupture by the Gibbs mechanism; see fig. 7.4c. A prerequisite is that the film surfaces have a substantial surface dilational modulus. For instance, the ethanol in champagne lowers the surface tension, but it cannot produce a significant modulus. This leads to rapid coalescence. Also see sec. 7.2a sub (iii). (ii) Thin films, meaning films thin enough for colloidal interaction forces acting
7.18
FOAMS
across the film. Especially for large films, the drainage is more complicated than discussed in sec. 7.3a sub (ii). This is treated in detail in sec. 6.4, including the kinetics of film rupture. An important parameter is the disjoining pressure. Broadly speaking, a film tends to be quite stable if the equilibrium film thickness is relatively large. Here it should be realized that the Van der Waals attraction is considerable since the Hamaker constant air-water-air is fairly large, about 13 kT; see app. IV.3, table A 3.1. Strong repulsion can be achieved by a surfactant that causes a significant zeta potential, combined with a low ionic strength. A non-ionic surfactant should provide steric repulsion over a fairly large distance. Several factors affect the probability of film rupture, such as the occurrence of surface waves, the formation of black spots in the film, and film elasticity. These are discussed in chapter 6, especially in sees. 6.5 and 6.6 (and some aspects also in sec. 8.3e sub (iii)). Newton black films can be very stable. Foams of black films have invisible lamellae and only the Plateau borders can be seen. It can further be mentioned that film thinning is most rapid in the upper layers of the foam so that coalescence is likely to occur earlier near the top. Moreover, film thinning can occur by evaporation of water, especially from the film between the upper bubbles to the air above. Such thinning can be relatively fast and readily lead to coalescence at the interface with the air. (iii) The third mechanism of film rupture is due to the presence of hydrophobic particles in the films. Such particles are deliberately used in anti-foaming agents, but similar particles, e.g. emulsion droplets, can also be present as a contaminant in various systems. The rupture mechanism will be discussed in sec. 7.6. 7.3c Ostwald ripening According to Henry's law [I 2.20.3], the solubility of a gas in a liquid is generally proportional to its pressure. The gas in a bubble has a pressure exceeding that of the surroundings by an amount given by Laplace's law, as Ap = 2y/a, where a is the bubble radius. The smaller the radius, the larger Ap and the greater the gas solubility. Consequently, gas will diffuse from a small bubble to a large one. Hence, small bubbles will shrink and finally disappear, whereas large bubbles become larger. This is the Ostwald ripening introduced in sec. I 2.20c, and it results in a coarsening of the foam. To be sure, large bubbles that are at the top of the foam can also disappear since they lose gas to the air above (which has an 'infinite' radius of curvature). Ostwald ripening occurs in several types of dispersions and the change in particle size distribution is generally described by the so-called LSW theory (after Lifshits, Slezov and Wagner); see sec. 8.3b, especially sub (i). However, this theory does not apply if the particles are very close to each other, hence for nearly all foams. More enlightening, though not precisely correct, is the theory by de Vriesv, which considers 11
A.J. de Vries, Rec. Trav. Chim. 77 (1955) 209.
FOAMS
7.19
the shrinkage with time of a small bubble surrounded by large ones. This is illustrated in fig. 7.8. It is assumed that the Laplace pressure of the large bubbles is negligible. The equation derived is a 2 (t) = a 2 -(RTDc sat («.) yl ph)t
[7.3.6]
where aQ is the initial bubble radius, D is the diffusion coefficient of the gas in the liquid (in water about 1.5 1CT9 m 2 s"1 at room temperature), c s a t (~) the solubility (near a flat surface) of the gas in the liquid (in moles per unit volume and per unit pressure) and p is atmospheric pressure (about 105 Pa ). The value of h is an assumed average. When putting a 2 (£) = 0 , t equals the lifetime of the bubble.
Figure 7.8. A small gas bubble surrounded by large ones, as envisaged in the de Vries theory.
Some sample calculations of the lifetime of small bubbles in water may be enlightening. Assume that the gas is either N2 (c sat (°°) = 7 \imol m~3 Pa"1) or CO2 (c sat (oo) = 0.4 mmol m" 3 Pa"1) , y = 50 mN m" 1 , aQ = 50 urn , and h = 20 um . The lifetime of the N2 bubble would then be 3800 s , or about one hour, and 66 s , or about one minute, for the CO2 bubble. Experimental results on single bubbles agree reasonably well with this prediction. Small bubbles can thus disappear fast, especially if the gas is highly soluble, like CO2 . If the bubbles contain a mixture of (two) gases of substantially different solubility, Ostwald ripening leads to a change in gas composition. The poorer soluble gas enriches the smaller bubbles, the more soluble gas enriches the larger ones. This causes a difference in mixing entropy, which results in a counteractive driving force for gas diffusion. In other words, Ostwald ripening is slowed down. The theory is briefly discussed for emulsions in sec. 8.3b sub (iii). Admixture of a poorly soluble gas can be applied in foams to slow down coarsening. Adding small quantities of perfluorohexane vapour has been shown to be quite effective. By simply bubbling the gas to be used in foam making through a small amount of liquid perfluorohexane, enough vapour is introduced to increase the foam lifetime from only a few hours to several days 11 . Addition of some N2 to beer (a few percent of the amount of CO2 present) is common to maintain the head on beer for some time after it has been drawn into a glass. 11
F.G. Gandalfo, H.L. Rosano, J. Colloid Set 194 (1997) 31.
7.20
FOAMS
It has often been observed that the rate of Ostwald ripening may substantially depend on the type of surfactant used. Probably, this is not caused by differences in surface tension, because the range of y values is mostly between 40 and 50 mN m" 1 , which is a relatively small variation. It has also been assumed that the permeability of the adsorption layer for gas may be quite small, thereby slowing down gas transport. However, the adsorption layer is very thin, 2 or 3 nm for small M surfactants, which cannot provide a significant barrier. Moreover, polymeric surfactants may substantially slow down Ostwald ripening, while their adsorption layer is quite porous for small molecules. A third explanation, already offered by Gibbs, is far more likely. Upon shrinkage of a bubble, its adsorption layer becomes compressed, which means that the surface excess F increases, which causes, in turn, a decrease in y. Hence, the driving force will decrease by an amount that depends on the surface dilational modulus Kg . The theory is discussed in sec. 8.3b sub (ii), and the conclusion is that Ostwald ripening will virtually stop if Kg > y. However, bubble shrinkage is not a very fast process, the reciprocal of d In Aid t generally ranging between a minute and some hours. This means that the dilational modulus will relax. For most small M surfactants, Kg will remain quite small, but for polymeric surfactants, it may remain substantial. Some workers do not use a timedependent dilational modulus (discussed in sec. 8.1c), but a strain rate-dependent surface dilational viscosity at the relevant time scale, defined as 77% = Ay/(dlnA/dt)
[7.3.6]
which can be experimentally determined. By combination with [7.3.6], the shrinkage can be calculated (by a numerical method). Some experimental results are given in fig. 7.9, which agree well with the prediction. It is seen that the effect is considerable, but it concerns a highly soluble gas. Ostwald ripening of air bubbles proceeds much slower, as mentioned, and then the surface dilational modulus will generally be too small to have a counteracting effect. Figure 7.9. Shrinkage of a single CO2 bubble. Bubble radius as a function of time for (1) a bubble in water ( ^ = 0) ; (2) in a beer sample with ^ = 0.01 (d In AI d t)~0-9 ; and (3) in another beer sample with 7^ = 0.08 (din AI d t p 0 ' 9 (all expressed in SI units). (Redrawn from P. Walstra, Physical Chemistry of Foods , Dekker, 200311.)
11
After results by A.D. Ronteltap and A. Prins, Colloids Surf. 47 (1990) 285.
FOAMS
7.21
Surfactants that become cross-linked directly after adsorption, e.g. some proteins, can give a very stiff and persistent adsorption layer, which will strongly counteract shrinkage. However, gas diffusion out of the smaller bubbles tends to go on, which may cause bubble deformation and finally buckling. The result is a flattened 'bubble ghost' not containing a gas phase. A bubble with a layer of solid particles, as shown in fig. 7.5a, can remain almost indefinitely stable11. 7.4 Foam characterization First, and foremost, foam is characterized by its gas volume fraction,
[7.4.1]
PF={mG+mL)/VF=(PcVG+pLVL)/VF
[7.4.2]
where VF , VL, and VG are the volumes of the foam, liquid and gas, respectively, and p F , PL and pG are their corresponding densities (by definition, m G and mL are simply the mass of liquid and gas, respectively). Noting that the density of the liquid phase is nearly always approximately 1000 times larger than that of the gas phase, we obtain p F =(p L V L )/V F
[7.4.3]
and hence,
Consequently, a measurement of the foam volume and mass, coupled with a knowledge of the liquid density, provides all the information needed to obtain the average liquid and gas volume fractions in the foam. This method can be useful for dry foams in mechanical equilibrium, but presents a problem when one needs to evaluate the volume fraction for a draining foam. In this case, part of the liquid collects at the bottom of the foam and one needs to consider this amount separately from the foam. In such cases, direct volume measurements of the total amount of liquid used to create the foam and the amount of liquid drained from the foam can be used to determine the average volume fraction as a function of time:
11
These aspects are extensively discussed by Z. Du, M.P. Bilbao-Montoya, B.P. Binks, E. Dickinson, R. Ettelaie and B.S. Murray, Langmuir 19 (2003) 3106.
7.22
fraction of the liquid that is in the foam = Vo - VL (t) / Vo
FOAMS [7.4.5]
where Vo represents the initial volume of the foaming solution and VL (t) the volume of liquid below the foam as a function of time. These methods are widely used to determine average volume fractions within a foam; however, more sophisticated techniques are required to determine the local volume fraction within a foam. Complete characterization of the foam structure implies knowledge of the structure of individual bubbles that comprise the foam. This includes the average bubble size and size distribution, along with their shapes. Measurements to obtain this information are thwarted by problems with visualizing the interior of the foam. Obstacles are created by absorption of light in the liquid and by the substantial difference between the refractive indexes of gas and liquid, causing considerable scattering of light. Therefore, indirect methods are usually employed to study the internal structure of foam.
Figure 7.10. Obtaining the bubble size in a foam can be a delicate procedure. Photographs from the side of a vessel can be used, but deformations and optical aberrations may lead to differences between the measured and actual bubble size.
An optical method that circumvents the problems with multiple light scattering is to estimate the bubble size and distribution from the area of individual bubbles at the side of a transparent window containing the foam; see fig. 7.10. There are many studies reviewing the differences between bulk and surface bubble distributions, and the method can be particularly useful to compare relative differences between different samples, provided that the container walls do not influence foam stability. This criterion requires walls that are well wet by the foaming solution. Similarly, foam samples can be placed in tubes or between glass plates allowing one to convert the threedimensional foam into a two-dimensional one for which Individual bubble sizes can be obtained. This can yield a direct measure of the bubble size distribution, but it is a rather Invasive technique that may locally alter the structure of the foam.
FOAMS
7.23
A direct method that can be used in some cases to study the internal foam structure is cryo-microscopy. In this case, a foam sample is quickly frozen by quenching it in liquid oxygen or nitrogen and the resulting solid structure is cut open and imaged with an optical or electron microscope. Although direct, this method generally suffers from artefacts and is very time consuming. The sample processing tends to perturb the foam structure in an uncontrolled manner during the freezing, and an error in the determination of bubble size is related to the fact that the cutting plane practically never passes through the bubble centres. This effect results in an underestimation of the bubble size. Nevertheless, much information about foam structure has been gleamed from such studies and correction methods are available. Another rather direct method to probe the internal structure of foam is to insert a fibre-optic probe into the foam that can detect the interfaces of the bubbles without causing bubble coalescence. This method requires a probe that is well wet by the continuous phase liquid to prevent film rupture as the fibre-optic probe traverses the films. The probe is inserted into the foam and each time a bubble interface is encountered, this event can be detected by changes in light transmission through the probe. In some cases, direct visual information can be obtained if the probe is used in conjunction with a CCD camera or an equivalent imaging device. Other indirect methods have been developed to determine the local liquid fraction in a foam. For example, in foaming solutions where the dispersed phase has a significant conductivity, measurements of the electrical conductivity can be used to determine the liquid volume fraction. This is carried out by inserting two partly stripped wire electrodes into a tube filled with foam and measuring the local resistance of the medium 11 . One can also increase the number of electrodes through the foam and perform what are called segmented measurements. This type of arrangement is particularly useful for detailed studies of liquid drainage. Electrodes are placed along the outer walls of the vessel containing the foam and the liquid fraction as a function of foam height can be monitored as a function of time. Unfortunately, the foam conductivity and the liquid volume fraction are not linearly related and careful attention must be paid to calibration and interpretation of these measurements. Of the more recent non-imaging techniques developed, one type exploits the strong multiple light scattering in foams. These procedures can provide direct, non-invasive information concerning the foam structure and dynamics. They are called diffuse transmission spectroscopy (DTS) and diffusing wave spectroscopy (DWS). In DTS one measures the probability that an incident photon will be diffusely transmitted through an opaque sample, whereas in DWS fluctuations in the intensity, rather than just the average, are measured. Using the time-averaged transmission of light through the foam gives a measure of the average bubble size, whereas time fluctuations in the scattered light intensity probe the motion of bubbles within the foam.
Recall fig. 6.12, describing a device for the conductometry of an isolated foam film.
7.24
FOAMS
7.5 Foam Properties 7.5a Rheology Foam is exploited in a large number of applications for its striking rheological properties. The characteristics of foam rheology can be classified as both solid-like and liquid-like, depending on its structure and the type and magnitude of the stresses that are applied to it. Foam with a low gas fraction tends to behave like a non-Newton liquid, whereas high gas fraction foams manifest (visco)elastic behaviour. Like an ordinary solid, polyhedral foams have a finite shear modulus and respond elastically to a small shear stress. Beyond a certain stress the foam exhibits viscous flow. Furthermore, due to the relatively large quantities of gas in foam, it can be quite compressible. Thus, Theologically speaking foams defy classification as solid, liquid or vapour; rather, they must be treated as unique substances.
Figure 7.11 A typical stressstrain relationship for foams.
Figure 7.11 depicts a general scheme of the stress versus strain relationship that is typically observed with foam. At low stress the foam behaves as an elastic solid and, upon relinquishing the stress, it recovers its initial shape. The elastic modulus can be determined from the slope of the stress-strain curve in this range. As the stress is increased, part of the deformation becomes permanent and viscoelastic behaviour is observed; the shear modulus is now a complex number, composed of a storage modulus G' and a loss modulus G"; see sec. IV.6.6d. Finally, above a critical yield stress (a ), the foam starts to flow and most of the deformation is permanent. If the strain is increased rather than the stress, the resulting stress decreases after yielding, a phenomenon called stress overshoot; it is also depicted in fig. 7.11. Since foam generally is quite fragile and subject to instability, it is very difficult to obtain reliable rheological data. To gain understanding of the rheological behaviour of foam, many studies are based on results obtained with oil-in-water emulsions of very
FOAMS
7.25
high q>. An essential variable is the Laplace pressure (2y/ a) of the drops or bubbles. As a bubble is deformed, its Laplace pressure is increased. This has two conesquences. First, upon release of the stress, the bubbles will regain their original shapes, which is responsible for the elastic behaviour of the foam. Second, the elastic modulus will be proportional to the Laplace pressure. Another essential variable is the dispersed phase volume fraction, as in all dispersions. Moreover, it turns out that the distribution of the bubble radii has a substantial effect, especially because it determines the value of the critical volume fraction
that for q>c<
G'~{p{(p-
17.5.1]
where
[7.5.2]
Figure 7.12. Rheological parameters, relative to the (extrapolated) value at
11 21
T.G. Mason, J. Bibette, and D.A. Weltz, Phys. Rev. Lett., 75 (1995) 2051. H.M. Princen, A.D. Kiss, J. Colloid Interface Set 112 (1986) 427.
7.26
FOAMS
where d 32 is the volume-surface average diameter of the undeformed drops or bubbles. A result is shown in fig. 7.12, curve (b). The relation found for the yield stress is Oy = P 1 / 3 / M r / d 3 2
[7.5.3]
where f((p) is an empirical function; see fig. 7.12, curve (c). All of these relations are approximate. They will also depend on the width and shape of the bubble size distribution. A problem is that the values of y and d 32 are often difficult to establish. Moreover, the surface dilational modulus will have some effect on the elastic shear modulus of the foam. When the stress applied surpasses the yield stress, the foam will flow. Such systems are often treated by a Bingham model; see sec. IV.6.6.3a. However, the situation is fairly complex. In a relatively wet foam (see e.g. fig. 7.12 r.h.s.), the bubbles can pass each other when the system is subject to a velocity gradient, although they are temporarily deformed during the passage. In a relatively dry foam, however, the bubbles have a fixed mutual orientation. Flow implies that sudden transitions in orientation occur. For instance, a common structure transition in a two-dimensional foam is the so-called Tl transition, as illustrated in fig. 7.13. Nevertheless, for a polydisperse system, a smooth increase of the apparent viscosity with increasing
[7.5.4]
where Vy is the shear rate (velocity gradient) and r]L the viscosity of the liquid phase. It is seen that the system is shear rate-thinning ( r/a will decrease with increasing Vi>).
Figure 7.13 Photograph of a two-dimensional foam undergoing a Tl transition as a result of flow. However, some additional variables must be expected to affect the apparent viscosity, such as the polydispersity, the viscosity of the dispersed phase, and the surface viscosity (both in dilation and in shear). Moreover, the relations will become substan11
H.M. Princen, A.D. Kiss, loc. cit.
FOAMS
7.27
tially different for elongational flows. More detailed models and simulations attempt to relate foam rneology to its structure and individual bubble behaviour. However, the agreement between the predicted behaviour and that observed remains semiquantitative. In part, the poor correspondence between theory and experiment for foam viscosity lies in the difficulty in measuring foam flow properties. To determine Theological parameters, such as the yield stress or the apparent viscosity, commercial rheometers are often employed. These include standard rotational devices and continuous flow tube viscometers. However, obtaining reproducible results independent of the sample geometry is quite challenging and some workers argue that perfection has not yet been achieved. One of the most obvious difficulties is that the rheological properties depend in a sensitive way on the bubble sizes and their distribution, and on the gas volume fraction of the foam, quantities that are hard to control and that change in time and potentially change under shear forces. Furthermore, wall slip often occurs when subjecting foam to a shear stress. In this way a non-uniform shear flow is created during the measurement; see sec. IV.6.7. For example, viscous dissipation in wall slip depends on the thickness of the wetting layer in a very sensitive way and this layer thickness can change as the foam drains and evolves over time. As the extent of plug flow (see also fig. IV. 6.17) depends on the amount of wall slip and the sample geometry, one is faced with a highly complicated system for quantitative interpretation. To characterize the foam-flowing properties for practical situations, it is often useful to construct a foam-flowing device tailored to the specific application for which one intends to use the foam. Using measurements from instruments that do not mimic the conditions found in a particular application can lead to confusion and false leads when attempts are made to engineer foam for specific uses.
7.5b Optical properties Another property of foam, which is often considered in applications, is its level of whiteness. This property plays a key role for sensorial properties of the foam. As mentioned above, foam is capable of scattering light efficiently due to large differences in the refractive indexes of the gas and liquid phases. As a consequence, when more interfaces are present in a foam to reflect the light incident upon it, the result will be a whiter appearance. Thus, foam whiteness is directly related to the sizes of the bubbles that comprise the foam. Smaller bubbles packed into a given volume increase the number of interfaces present, and hence result in whiter looking foam. This explains why typical shaving cream foam, which contains micron-sized bubbles, appears significantly whiter than a dishwashing or bubblebath foam, which is composed of centimetersized bubbles. Another common observation is the iridescent colours observed in foams made from very large bubbles. This phenomenon arises from a property of the thin liquid
7.28
FOAMS
films that form the foam bubbles. Upon exposure to white light (light with multiple wavelengths), reflections occur at both the front- and backsides of the films. The reflected beams having the same wavelength interfere with each other, with some wavelengths being cancelled out and others reinforced. This in turn produces what are called interference colours, with each colour corresponding to a certain film thickness. It Is the distance that the wave travels through the film that determines the phase difference between the light reflected from the two sides of the film. (For more details, see subsecs. 6.2a and 2b.) However, when a multitude of bubbles are present, the different light waves Interfere from different bubbles and different interfaces in a random fashion once again producing white light. Thus, we can observe foam with multicoloured bubbles only when the bubbles are large and not too numerous. 7.6 Antifoam and defoaming In some cases It is desirable to have large quantities of voluminous, long-lasting foam, while in others we need to eliminate or entirely prevent its formation. Rich shampooing lathers are an example of the former, while foam created during distillation or water treatment operations illustrates the later. Unwanted foam can produce product defects, interrupt process operations causing downtime, and create safety hazards, like the malfunctioning of relief valves. Many Industries have combated the difficulties arising from the presence and persistence of foams. These various industries include food and beverage manufacture, the preparation of textiles and dyestuffs, inks and coatings, pulp and paper, lubricants, and even semiconductors. Breaking of foam is also an issue in waste water treatment. Of these, the paper industry is one of the major consumers of foam control agents. In the paper industry, foam can be problematic at nearly any stage; from the pulp and paper fabrication stage through the coating and printing stage, as well as in the reprocessing of waste paper (e.g. de-inking) and the treatment of process wastewater. Although we often associate foam with aqueous solutions, many processes in the petrochemical industry and oil field exploration are faced with organic liquid-based foams. In Its natural state, crude oil contains high levels of dissolved gases, held in by high pressure, and of surface active components. When this live crude Is extracted and passed Into the low pressure environment of a gas-oil separator, the dissolved gases are released and can cause major foaming problems. This can lead to excessive oil carry-over in the downstream gas line and to damage in downstream equipment. One industry that is gaining major importance in this area of foam control Is the biotechnology industry. In biotechnology processing systems the conditions for foam production are always potentially present. If not already intrinsically capable of stabilizing foam, many organisms generate proteins that are released into the surrounding medium, which can act as foam stabilizers. In some cases the medium itself may contain foam-stabilizing agents. The sparging, which is always present, together with the
FOAMS
7.29
presence of proteins, produces the sort of situation in which foam can develop quickly. Foam formation can be a disaster in biological fermenters as it frequently results in loss of production. There are various reasons for this. As the density of foam is so low compared with that of the liquid phase, it can be blown out of the vessel with the exit gas. In addition, any filters designed to prevent the escape of the organisms in the vessel will be blocked by the moisture in the foam as the filter medium swells. This results in pressurisation of the vessel and loss of gas flow or the undesirable opening of overpressure valves. Consequently, the sterility of the whole system will be imperilled once the valves are open to the atmosphere. Sometimes overflow will occur with the foam being discharged into the area surrounding the fermenter. This may well represent a health hazard and will certainly result in a large cleaning task. Moreover, there are occasions when foam builds up so rapidly that within a few seconds the entire liquid contents turn to foam and are lost before any remedial action can be taken. 7.6a Strategies To prevent foam formation, or to eliminate foam once formed, one must first consider the phenomena responsible for the formation and stability of foam. In particular, it is important to realize that foaming depends on the physical processes used to disperse the gas in the liquid phase, together with the accompanying physicochemical phenomena that facilitate these processes and stabilize the foam after it is generated. To illustrate this point, consider the different mechanisms involved in washing machines, which mechanically agitate the gas and liquid phases together, compared with foam in champagnes and carbonated drinks, which are formed by the release of dissolved gas into the liquid phase. As mentioned above, dispersed gas can also be generated as a byproduct in fermentation processes, which can lead to large quantities of unwanted foam at the top of fermentation vessels. In each of these examples, gas-liquid interfaces are created accompanied by the adsorption of material onto these interfaces (i.e. surface active molecules, solid particles e t c . ) , which inhibits the coalescence of adjacent bubbles and eventually leads to formation of the foam. In certain cases, these surface active agents are directly introduced to the solution as in emulsion polymerization that requires surfactants to stabilize the emulsified reactants. In other situations, the surface active material is simply a byproduct or impurity, as in wastewater treatment processes. In fact, a slight process deviation, or contamination by products resulting from corrosion, can create a foaming problem. If there is none or very little surface active material present, any foam that is formed will not be stable and rapid coalescence of the gas bubbles will take place. In these situations, however, the presence of small particles in the interface (fig. 7.5a) can substantially improve stability after reduction of the size of the bubble. Likewise, if we adjust the time-scales for generating the bubbles and bringing them into contact, we can strongly influence the development of foam. These two considerations provide us
7.30
FOAMS
with two scenarios for suppressing foam formation. The first scenario corresponds to a chemical modification to the system by removing the 'foaming agent', while in the second a process adjustment to reduce foaming is realized. In order to remove a foaming agent, one must first establish the nature of the surfactant, often necessitating a first step of thorough chemical analysis of the foaming solution. Specific process modifications include: the elimination of cascades and closing up leaks in pumps, fittings and seals. Simple reductions in gas velocities and the incorporation of expansion tanks can also be tried. These basic foam control strategies can be useful in some cases, but many systems are far too complex for such simple procedures and one must rely on the addition of so-called defoaming and anti-foaming formulations. Accelerating the destruction of relatively unstable foam relies on our ability to increase the drainage of the liquid separating the gas bubbles. The two most common ways to achieve this are to decrease the liquid viscosity and/or the viscoelastic properties of the gas-liquid interface (i.e. surface dilational viscosity and elasticity). Decreasing the liquid viscosity can be effective when the foaming liquid is rather viscous on its own and can often be accomplished by raising the temperature. Unfortunately, many common systems, for which foaming is a problem, already have quite low viscosity and a further decrease has no practical benefits. On the contrary, lowering the viscoelasticity at the bubble surface can have a substantial effect on just about any system. The energy barriers that stabilize long-lasting foam also originate from the material adsorbed at the gas-liquid interface. In general, these barriers are classified as electrostatic or steric, as described in previous chapters. Charged surfactant molecules, like those found in detergent formulations, are a common example of substances that can create strong electrostatic repulsion in soap films. This can be reduced by adding electrolyte to the foaming solution. In certain systems this can have a strong destabilizing effect. In particular, for anionic surfactants (the largest group), adding divalent electrolytes, like calcium ions, often has a dramatic effect as these ions both screen the double layer more effectively and may bind specifically to the charged surfactants, causing a strong reduction in the diffuse layer potential or even causing the surfactants to precipitate out of the solution. These properties also explain why it is often difficult to foam solutions of anionic surfactant in hard water. Sterically or solid particle-stabilized foams are at present the most problematic to break. Robust foams encountered in the pulp and paper industry are typical illustrations. In this case, complex polysaccharide molecules extracted into the wash solution adsorb on the air-solution interface, block bubble coalescence and reduce Ostwald ripening. To destroy these foams, one needs to disrupt or displace the molecules (or particles) from the surface. Certain short-chain alcohols have been used to accomplish this. Other additives include fatty acids, waxes and saturated soaps. However, the most broad-based foam control additives are formulations made from hydrocarbon and
FOAMS
7.31
silicone oils. These formulations have proven effective in both aqueous and organic based foams and are used as multipurpose antifoam additives, regardless of the foam stabilization mechanism. Finally, for some manufacturing processes, foam is not acceptable at any level. Companies faced with this problem have resorted to using chemical defoamers or antifoams. A defoamer is a compound that knocks down a foam already present in the mixture (foam elimination). An antifoam is a compound that prevents the solution from creating a stable liquid-gas interface (foam prevention). Strictly speaking, antifoam is added before the onset of foam, and defoamer after foaming has begun, but the terms and products are frequently used indiscriminately. Defoaming agents/antifoams include soaps (carboxylates) and nitrogenous antifoams such as monoamides, phosphoric acid esters, mineral oil blends, long chain alcohols, fluorosurfactants, and hydrophobed silicon/hydrophilic oil mixtures. In the following sections, we provide a physical understanding of the working mechanisms of defoaming/antifoaming formulations and give guidelines for their utilization.
7.6b Defoaming and antifoam formulations Although a vast range of antifoam substances exist, the most common defoaming and antifoaming formulations are emulsions, generally of the oil-in-water type, with droplet sizes in the 10-50 nm range. They are classified as filled or unfilled. Filled antifoams refer to those in which the droplets contain about 3-10% by weight of finely divided inorganic filler. In most cases, the filler is fumed silica with an average particle size in the range of 0.1-1 um. These particles are often provided with a hydrophobic surface layer to enhance their dispersion in the oil phase, which consists of a nonpolar oil, such as a mineral oil or an unmodified polydimethylsiloxane oil. On their own these oils do have antifoam properties, but the addition of fine particles greatly enhances their performance. Commercially available products of this type are normally sold as an emulsion to promote easy dispersion of the formulation in the foaming solution. Somewhat less universal are unfilled antifoams. These products are highly effective only in specific systems, but they tend to be comparatively less expensive than filled antifoams. Furthermore, they are usually auto-dispersible due to the presence of a hydrophilic group on the molecule. Thus, addition of surfactant or polymers to their formulation is not needed. A large amount of research has been dedicated to the development of antifoaming formulations, and a detailed account of defoaming and antifoaming theory and applications can be found in a relatively recent book . In the following subsection, we summarize how these formulations work to shed light on the utilization and selection of a particular additive or formulation. 11
P.R. Garret, Ed., see the General references in sec. 7.8.
7.32
FOAMS
7.6c Foam destruction by defoaming and antifoaming formulations All antifoaming agents added to foam made of an aqueous solution contain, or form after addition, hydrophobic particles, generally emulsion droplets. Mechanisms by which these particles can cause rupture of a film are illustrated in fig. 7.14. These mechanisms will be discussed below, though not in detail. It goes without saying that film rupture causes coalescence of bubbles, hence eventual destruction of the foam.
Figure 7.14. Possible mechanisms involved in the rupture of aqueous films of thickness h induced by hydrophobic particles of diameter d; the numbers 1-3 indicate subsequent stages; (a) solid particle; (b) and (c) oil droplet; (d) oil droplet or composite particle. Thick arrows indicate spreading of oil or surfactant. A denotes air and W water, a is the contact angle as measured in the water phase; 77 is the viscosity of the water phase (subscript W) or the oil phase (subscript O). S W G '°' is the two-dimensional spreading tension, quantifying the spreading of O at the WG interface (see [III.5.1.1]). (Redrawn from P. Walstra, Physical Chemistry of Foods, M. Dekker, 2003.)
In cases (a) and (b), a particle becomes trapped in a thinning film and then makes contact with both air bubbles; in case (b) the same occurs, possibly after some squeezing deformation of the droplet. Because of the obtuse contact angle a , the curvature of the film surfaces where they reach the particle becomes high, leading to high Laplace pressure. Hence, the water will flow away from the particle, leading to film rupture. This mechanism is designated bridging or film pinch-off. A prerequisite is that the particles are preferentially wet by air, as determined by the contact angle. In the case of a solid, hence possibly nonspherical, particle, the critical value of a depends on particle shape. In cases (c) and (d), contact with one of the film surfaces may suffice. An oil droplet
FOAMS
7.33
reaching the surface will suddenly alter its shape to a flat lens, which induces flow of the water away from the particle; if the film is relatively thin, this may well cause its rupture, as depicted. Whether the formation of a lens will lead to film rupture will also depend on the deformation speed, hence on the viscosity of the oil. If the contact angle equals 180°, the spreading tension s W G ' ° ' will be positive and the oil will spread over the film surface, analogous to the situation in (d); see sec. III.5. la for a discussion on wetting and spreading. In case (d), the particle is generally a composite, from which some substance, be it oil or a surfactant, spreads over the film surface. The mechanism described in this paragraph is designated fluid entrainment; the aqueous liquid is indeed entrained by the moving surfaces (i.e. the Marangoni effect, cf. fig. 8.3), although bridging is the final cause of rupture.
Figure 7.15. Oil drops in antifoaming/defoaming agents and their interaction with an air bubble. A is air, W is aqueous phase, O is oil. (a) Plain oil drop; difficult to achieve wetting of the drop by air. (b) Drop contains solid particles that are oriented in the O-W interface; a protruding particle can pierce the film between O and A. (c) The drop contains crystals that form a solid network; a protruding crystal can pierce the film.
It is well known that oil or fat, even In trace amounts, can be very detrimental to foam stability. For instance, when somebody puts a finger in the foam on a glass of champagne, the foam Is immediately destroyed, presumably due to natural oil from the finger spreading over the film surfaces. Likewise, beer foam is sensitive to lipstick. Nevertheless, plain emulsion drops are not always effective as defoaming agents. The reason is that they often cannot make contact with the air bubble. The emulsion drop tends to be covered by a surfactant layer that causes colloidal repulsion between drops. The same applies to the surface of the air bubbles. Consequently, an air bubble and an emulsion drop will probably repel each other, leaving a thin water film, which may be quite resistant to rupture. This situation is depicted in fig. 7.15a. If the drop contains solid particles that have such an oil-solid-water contact angle that they will preferentially be lodged in the oil-water interface, as depicted in fig. 7.15b, they will likely be able to pierce the film between droplet and bubble due to stress concentration. This will lead, in turn, to wetting of the droplets by air. Then any of the mechanisms depicted in fig. 7.14c-d can be triggered. As mentioned, most commercial antifoaming agents consist of an oil-in-water emulsion with suitable solid particles in the drops. Moreover, drops containing paraffin oil or triglyceride oil often contain
7.34
FOAMS
paraffin or triglyceride crystals, respectively, which tend to form a solid network, as depicted in fig. 7.15c. Such 'fat globules', e.g. those of milk, are efficient foam breakers. 7.6d Antifoam performance lifetimes In addition to losing antifoam through adsorption onto the walls of a vessel and other equipment, there are two additional reasons why antifoams lose their effectiveness over time. First, the surface and interfacial tensions may change with time, and hence all the conditions for spreading and bridging can change. These changes are often brought about by the adsorption of surface active material present in the solution onto the antifoam globules. If this happens, an oil or compound that starts off with favourable conditions for antifoaming may eventually end up stabilizing foams (e.g. the adsorption of surface active material can change the wettability of the antifoam globule). One advantage of silicone-based antifoams is that they are rather resilient to these problems. The second ageing problem, which is more general, involves the size of the antifoaming globule. In a foaming environment the sizes of the globules are often observed to decrease with time. The explanation is not fully clear; perhaps drops that are spreading over an air bubble, to become a thin layer, are subsequently disrupted. Once these droplets fall below 5 \im, the antifoaming efficiency drops off. The lower efficiency may arise because the microscopic globules are so small that they remain in the Plateau borders. Antifoam drop size reduction can be minimized by increasing the oil viscosity and interfacial tension. However, increases in viscosity can have adverse effects on spreading and antifoam dispersion (inhomogeneous spreading of the material). Thus, when applying these criteria, a balance must be sought. Often formulations contain mixtures of oils with different viscosities to provide the best overall benefits. 7.7 Applications As already noted, because of its particular structure, foam has some rather unique properties from which a wide variety of applications have emerged. Owing to its low density (i.e. high gas volume fractions) and unusual rheology, foam has found use in the home, in personal care and in the industrial sector. In many cases, the sensorial properties of foam are exploited to enhance a product's value, while in others it plays a direct role in a given process. Furthermore, just as manufacturing and stability of foams is important in some industrial processes, in others elimination and destruction is crucial. 7.7a Industrial applications (i) Separation and extraction. One of the oldest applications of foam is in the flotation industry. This, and other separation processes, takes advantage of the
FOAMS
7.35
buoyancy and mechanical rigidity of foam. In so-called froth flotation, foam is used to separate the more precious minerals from the waste rock, extracted from a quarry; see sec. III.5.1 lb. This separation method relies on the different wetting properties of the various minerals. Typically, the more valuable minerals tend to be hydrophobic in nature, whereas the waste rock is often preferentially water-wetted. In many cases supplementary chemical additives are used to ensure these conditions. Therefore, once formation of a foam is achieved by bubbling air through a slurry, the waste rock remains in the water, whereas the hydrophobic mineral particles gather at the surfaces of the bubbles and are entrapped in the foam. The foam rises to the top of the vessel where it is skimmed off and the minerals are collected. Foam Jractionatlon is another common foam separation technique that is chemical in origin. This method exploits the preferential adsorption of certain molecules. Foam has the advantage of generating a relatively large surface area for adsorption and collection. Foams can also be used to recover and separate colloidal particles if they have been coated with a surfactant to bring them to the interface. For example, foam fractionation has been used to remove radioactive waste from contaminated water streams. (ii) Oil recovery. Foam has found many uses in the oil recovery industry, from the initial drilling stage in the primary recovery stage, all the way to the tertiary or enhanced oil recovery stage. Again, these applications make use of the unique features of foams, primarily of the large interfacial area and distinct rheological properties of foam. Drilling fluids are used when boring a hole for oil recovery; the fluid protects the walls of the hole and helps to circulate down the waste generated by the drilling process. In many cases, clay or mud slurries are used; however, the density of these fluids can exert excessive pressure on the bore hole walls, which can damage the rock structure and weaken the well. In such cases, an aqueous foam is often used. Foam can be made to have the same rheological features of a standard drilling mud and it can be very effective in cleaning the drilling material from the well. Foam is also used for the direct recovery of oil. The most common form of secondary oil recovery is to inject water into a well in one location to force oil out in another place. This displacement technique can be effective, but it is plagued by rock formations that have high permeability zones. In such cases the water flows along these zones and bypasses the oil. One way to alleviate this problem is to first inject foam, which will tend to plug the high permeability zones, after which subsequent water injections regain their oil recovery efficiency. (iii) Fire-fighting. Foams are particularly useful in combating fire from flammable liquids. Whereas water tends to agitate and further spread a fuel fire and eventually sink to the bottom, foam is less dense than the burning liquid and remains floating at the surface where it can cut off the supply of oxygen, ultimately extinguishing the flames. Foam also allows a more efficient use of the fire-fighting liquid. Foams used for fire-fighting applications are typically produced with a co-injection
7.36
FOAMS
method using a foaming agent diluted in water and air. Typical foaming agents include fluorinated surfactants that can lower the air-water surface tension to below 20 mN m"1 and therefore produce foam that readily spreads over the surface of the burning organic liquid. Instead of considering the volume fraction of the air directly, fire-fighting foams are characterized by their expansion ratio. This ratio is the increase in volume of the liquid after the foam is formed and ranges from 5 :1 to over 1000 : 1 . Ratios of 5 :1 to 20 :1 are considered low-expansion foams, while 21:1 to 200 :1 are medium-expansion foams and ratios greater than 200 :1 are taken as high-expansion foams. Low-expansion foams are used most commonly because they are relatively dense and can be more readily sprayed over larger distances. Their primary disadvantage is the relatively smaller coverage areas (medium- and high-expansion foams are usually too light to be sprayed over any distance and must be used near the flames). However, they provide good coverage and their low density minimizes surface disturbances of the burning liquid. 7.7b Food Foam finds use in many food applications. Examples include whipped cream, meringues and a host of foamed milk products, such as foamed yoghurt and ice cream. In most cases, the foams are stabilized by proteins, the most important being egg-white (egg albumen) and milk proteins. Some of these proteins provide elastic layers at the air-water interface that prevent bubble coalescence and reduce Ostwald ripening. Eggwhite is one of the best foaming agents, primarily due to its high protein content. Moreover, it consists of a mixture of different proteins. These proteins include globulins, which give rise to good foamability, in addition to slow foam drainage by the high viscosity produced by globulin-ovomucoid complexes. Surface complexes of lysozyme and ovomucins also enhance surface viscosity. In milk foam, the active foaming proteins tend to be whole casein and whey protein (of which P -lactoglobulin is the major component). One foam, used in the food industry merely for its aesthetical value, is the foam produced at the top of a glass of freshly poured beer. The foam from beer is formed by the dissolved CO 2 , although some nitrogen is sometimes added to reduce Ostwald ripening. The primary foam stabilizers in beer foam are proteins and polyphenols. Occasionally, other components, such as trace metal ions, iso a- acids and propylene glycol alginate are added to enhance the stability of beer foam. 7.7c Detergents Although nearly all detergent formulations can produce foam, more often than not foam is non-essential, or even detrimental, to their functioning. Instead, foam in these formulations often provides a desirable sensorial attribute for the consumer. One example is the foam generated from hand dishwashing detergent. Mostly a bit of foam is desirable because its presence indicates that the surfactant concentration is above
FOAMS
7.37
its c.m.c. so that it can take up fatty materials. However, consumer panels judging these products invariably rate products that produce high levels of sustainable foam higher, even though no real cleaning benefits are provided. On the other hand, antifoam formulations are added to machine clothes washing and dishwashing formulations to prevent excessive foam that decreases washing efficiency and can cause damage to pumps. One exception is the case where cleaning must be restricted to a particular area and where excessive amounts of water must be avoided. An excellent example of this is foam used for carpet cleaning. Here, foams are often used to clean rugs because they can spread the detergent over the rug surface while avoiding excessive wetting at the base of the carpet.
7.7d
Cosmetics
Two of the foremost applications of foam are found in the cosmetic industry for shampoos and shaving creams. In addition to the aesthetic appeal of foams, they are exploited for their ability to retain different substances and distribute them as required while using a relatively small quantity of fluid. The lather formed by some shampoos is designed for this purpose. Shaving cream foams keep the skin moistened; the foam can be easily used to spread moisturizer or lubricant onto the skin, at the same time still providing enough rigidity to hold the liquid in place. Toothpaste is another common cosmetic product that uses the unique properties of foam to distribute substances while providing a non-Newtonian fluid viscosity that enhances product application throughout the entire oral cavity.
7.7e Miscellaneous In addition to these somewhat common applications of foam, many smaller markets exist where foam is widely used. The following list provides a glimpse: aircraft de-icing, plant frost protection, herbicides, sanitary landfill cover, dust control, law enforcement, military and defence applications. Add to these the many uses for solid foam, which always passes through a liquid-gas state before solidification, and we start to realize the vast industrial importance of foam. 7.8 General references J.J. Bikerman, Foams: Theory and Industrial Applications, Reinhold Publishing, (1953). J.J. Bikerman, Foams, Springer, (1973). C.V. Boys, Soap Bubbles, Dover Publications, (1959). These are the first comprehensive books on foams and provide many empirical insights and historical background.
7.38
FOAMS
D.R. Ekserova, P.M. Kruglyakov, Foam and Foam Films: Theory, Experiment, Application, Studies in Interface Science, 5, Elsevier, (1997). (An in-depth text that covers many details of foam and foam-films from a physico-chemical perspective.) Foams: Theory, Measurements, and Applications, R.K. Prud'Homme, S.A. Khan, Eds. Surfactant Science Series, 57, Dekker, (1995). (This book is a collection of different foam topics ranging from practical engineering applications to basic physicochemical principles.) D. Wealre, S. Hutzler, The Physics of Foams, Clarendon Press, (1999). (Foams from the point of view of physics with a strong underscore on cellular structures and foam drainage.) Foams: Physics, Chemistry, and Structure, A.J. Wilson, Ed., Springer, (1989). (General book that covers many different aspects of foam.) Defoaming -Theory and Industrial applications, P.R. Garrett, Ed., Marcel Dekker, (1993). (A modern and comprehensive collection of defoaming principles and applications.) D. Edwards, H. Brenner and D.T. Wasan, Interfacial Transport Processes and Rheology, Butterworth Heineman (1991). (Contains a complete treatment of surface rheologlcal principles.)
8
EMULSIONS
Pieter Walstra 8.1
8.2
8.3
Characterization
8.2
8.1a
8.2
Description
8.1b
Surfactants as emulsifiers
8.1c
The role of interfacial tension gradients
8.11
8.4
8. Id
Emulsion properties
8.15
8.1e
Determination of drop size distribution
8.20
8. If
Determination of interfacial properties
8.25
8.1g
Determination of colloidal interaction forces
8.27
Emulsion formation
8.30
8.2a
Introduction
8.30
8.2b
Hydrodynamics
8.37
8.2c
Roles of the surfactant
8.46
8.2d
Formation of surface layers
8.59
8.2e
Effect of volume fraction
8.61
Stability
8.62
8.3a
8.63
Overview
8.3b
Ostwald ripening
8.65
8.3c
Aggregation
8.71
8.3d
Sedimentation
8.73
8.3e
Coalescence
8.80
8.4
Case study: Pickering emulsions
8.91
8.5
General references
8.92
This Page is Intentionally Left Blank
8 EMULSIONS PIETER WALSTRA
Emulsions can be formed and used in various situations: - Several organisms produce emulsions, the prime example being the milk excreted by female mammals. Milk contains small oil droplets dispersed in an aqueous liquid. The function of the droplets is to transport substances that do not, or insufficiently, dissolve in water: this includes the transport of a large quantity of edible energy in a limited volume without increasing the osmotic pressure. - Most types of emulsions encountered in daily life are man-made. Their most common function is the transport of water-insoluble substances in a stable, and hence finely dispersed form, but other, more specific functions are also involved. Such products include a range of foodstuffs, Pharmaceuticals, cosmetics, pesticide formulations, paints, lubricants, and finishing agents. - Emulsions can be used in intermediate stages in manufacturing processes. This may concern extraction in a stirred tank or a column. Another example is the formation of latices by emulsion polymerization11. - Emulsions can also be a nuisance, since they can be formed inadvertently during some processes, and then have to be broken. An example is crude oil, which is often obtained as a water in oil emulsion. The fundamentals of interface and colloid science can be applied fruitfully to understand the various properties of emulsions and to predict how these properties can be realized. Interfacial properties and processes are essential in formation and stability of emulsions, and the interfacial area is large. If the drops are not too large, emulsions can be treated as lyophobic colloids. In many respects, they form ideal systems for studying such colloids. The particles are perfect spheres and, for many purposes, they can be considered to be rigid. Several emulsion properties, such as volume fraction, average droplet size, and interfacial composition, can be varied as desired. Polydispersity of the drops may pose a problem, but there are ways of obtaining almost homodisperse emulsions, at least in some systems2'. On the other hand, it is often far from easy to understand what happens in practice, let alone to predict what will happen. This is due to the complexity of the 11
See e.g. the review by J.W. Vanderhoff, Chem. Eng. Set. 48 (1993) 203. J. Bibette, J. Colloid Interface Set. 147 (1991) 474. Fundamentals of Interface and Colloid Science, Volume V © 2005 Elsevier Ltd. J. Lyklema (Editor) All rights reserved
21
8.2
EMULSIONS
phenomena involved. First, several different changes can occur simultaneously, both during emulsification and at rest, and each of these may depend in a different manner on internal and external variables. Second, the conditions during the application and storage of the emulsion may vary widely; for example, chemical changes can occur that affect physical properties. Moreover, several products are not simple emulsions, but contain other structural elements such as gas bubbles or solid particles, or mesomorphic phases. In this chapter, we shall primarily consider simple o/w (oil in water) or w/o (water in oil) emulsions, excluding microemulsions (see chapter 5) and very coarse emulsions. The average droplet diameter is typically of the order of a micrometer. Even such a simple emulsion may contain numerous components, for example, the emulsifiers used in practice are virtually always mixtures. 8.1 Characterization 8.2 a Description An emulsion is a dispersion of drops in another liquid where the two liquids are not miscible in all proportions; it is thus a two phase system. The phases are, for simplicity, called oil and water. In this context 'oil' means a hydrophobic liquid, e.g. benzene, hexadecane, or a triglyceride oil; and 'water' means an aqueous solution. Nearly all emulsions contain at least one other substance, generally called an "emulsifier". It is needed in the formation of an emulsion, and the emulsifier's properties then also determine whether the emulsion will be of the o/w or of the w/o type. The emulsifier is generally also needed to stabilize the drops in the finished emulsion against aggregation and coalescence. All emulsifiers are surfactants that adsorb onto oil water interfaces, lowering the interfacial tension. Since the surfactant must be soluble in at least one of the two phases in order to be active during emulsification, the adsorption layer formed will generally be a Gibbs type monolayer. Emulsions differ from suspensions in that the particle surface is deformable, both in directions perpendicular and parallel to the surface. This permits a relatively simple way of formation of an emulsion from two bulk phases. It also causes the particles to assume a spherical shape. On the other hand, the droplets behave like rigid particles under most conditions. Consider a drop of diameter d = lum, and interfacial tension / = 0.005 mN m" 1 . Its Laplace or capillary pressure 4y/d (see sec. 1.2.23) will then equal 20 kPa. Assume further that the liquid around the drop is stirred, that it has a viscosity of 0.002 Pa s, and that the local shear rate is 5 x 103 s" 1 , which is a very high value. The shear stress acting on the drop then will equal 10 Pa, much smaller than the Laplace pressure. As we will see in sec. 8.2b, sub (i), the stress would lead to a maximum relative deformation of the droplet by 0.1%, a negligible amount. The shear stress can, in principle, also move the droplet's surface in a tangential direction, at least locally. However, the stress will induce an interfacial
EMULSIONS
8.3
tension gradient, which means that a counteracting stress will develop, which can stop the motion. This is discussed further in sec. 8.1c. For very small droplets, say with a radius <0.1|im, the tendency of many surfactants to impose a 'spontaneous curvature' to a monolayer, will become important (see Vol. Ill, sec. 4.7). We then have a regime between that of micro emulsions and macro emulsions. In this chapter we will primarily consider emulsion droplets that are large enough for the spontaneous curvature of the surfactant to have no effect, but small enough for the drop to be rigid under most conditions. In many respects, an o/w emulsion resembles an aqueous foam. However, foam bubbles tend to be one or two orders of magnitude larger than emulsion droplets, and the difference in density between both phases is larger by about one order of magnitude. This means that foam bubbles cream very rapidly, to form a packed layer. Liquid then drains from that layer and the Laplace pressure of the bubbles is small enough to allow their deformation. In this way a polyhedral foam is formed. Such a transition generally does not occur in emulsions, except when they are centrifuged. Important internal variables affecting emulsion properties include: - Emulsion type, i.e., o/w or w/o. The macroscopic properties of an emulsion are generally dominated by those of the continuous phase, especially if the volume fraction of droplets (
Sec e.g., E. Dickinson and D.J. McClements, Advances in Food Colloids, Blackie, (1995), chapter 9, which includes a general discussion of the formation and the various instabilities of multiple emulsions.
8.4
-
EMULSIONS
Volume fraction of droplets. This is often taken to equal the volume fraction of the
disperse phase, but if the droplets are quite small, the adsorption layer may contribute substantially to
Size distribution of the droplets. In most cases, the stability of an emulsion is
greater if the droplets are smaller, except with respect to Ostwald ripening. Hence, the size distribution is an important variable, both the average size and the distribution width. The volume surface average droplet diameter, d 32 , is rarely below 0.3 um , and it may be up to a few times 10 |im. An important variable determined by the size distribution is the total specific surface area of the droplets, given by, Av = 6
[8.1.1]
in units of area per unit of emulsion volume. 8.1b Surfactants as emulsifiers In this section we proceed on treatments given in chapters III.3 and 4, and IV.4. A surfactant used as an emulsifier has two main functions: allowing emulsion formation and providing stability to the emulsion once made. It generally also determines the emulsion type, o/w or w/o. Emulsifiers come in two main types: low molar mass amphiphilic compounds, called 'amphiphiles', here, for convenience (although it may be noted that some authors reserve the words surfactant or emulsifier for 'amphiphile'), and surface active polymers. These differ in several respects, as discussed below. Moreover, properties such as the interfacial tension and the surface excess obtained vary among types of interface: s-w, s-o, a-w and o-w. The differences can be large, and for water soluble surfactants, much more is known about the behaviour at the a w than at the o w interface. For many polymers, nearly all of the literature concerns solid interfaces. Moreover, for a given surfactant, surface activity and interfacial properties may vary substantially with the composition of the aqueous phase and that of the oil phase. For example, benzene, a liquid paraffin, or a triacylglycerol mixture, as the oil phase will result in different y values. It is generally difficult, and often impossible, to find adequate and reliable data. (i) Amphiphiles. The hydrophobic part of the molecule is generally an aliphatic chain, e.g., of palmitic or oleic acid, to which a polar head group is esterified. The head groups, and thereby the amphiphiles, are classified as non-ionic, anionic and cationic. In principle, the adsorbed and the dissolved surfactant are in thermodynamic equilibrium, as given by Gibbs' adsorption law (see 1.2.22). This means that the amphiphile forms a Gibbs monolayer (see III.4). It also means that for an emulsion of
EMULSIONS
8.5
given composition, specific surface area, and temperature, the surface properties [y, r, etc.) are fixed. Several surfactant properties are of importance for application in emulsions. These include:
Solubility in both phases, and also the critical micellization concentration (c.m.c), if it exists. The solubility depends greatly on temperature. Many water soluble amphiphiles have a so called Krafft temperature, below which they form crystals (mostly a-crystals), and leave a very small concentration in solution. Several amphiphiles can also form crystals in an oil phase. For water soluble amphiphiles, a plateau value for the interfacial tension is generally reached at the c.m.c. or at the solubility limit; these concentrations are, for the most part, between 0.01 and 10 mmolar, or between 10 mg and 3 g per litre, the lower values generally applying to non-ionics. Hydrophile Lipophile Balance (HLB) . This is a measure of the preponderance of the hydrophilic over the lipophilic moiety of a surfactant molecule at 25°C. The HLB number changes linearly with the molar Gibbs energy of micellization, and arbitrary scaling constants are introduced in such a way that the minimum HLB number equals about zero, and the number for a molecule that is neither hydrophilic nor lipophilic about seven. Some methods exist to determine HLB numbers; they can also be estimated by summation of numbers assigned to specific groups of the molecule. Tabulated values are available. A surfactant with a low HLB number (2-6) is better soluble in oil than in water and tends to make a w/o emulsion, and for higher HLB numbers (10-18) it is the other way round. This is in agreement with Bancroft's rule (see Vol. Ill, sec. 4.8): the continuous phase of an emulsion is the one in which the surfactant is best soluble. There are, however, exceptions2'. The numbers given refer to non-ionics. Most ionic amphiphiles have quite high HLB numbers and are poorly soluble in oil. A related concept is that of the phase-inversion temperature (p.i.t.) or HLB temperature. It applies to non-ionic amphiphiles containing one or more poly oxyethylene chains. The p.i.t. is defined as a property of the emulsion, since it depends not only on the surfactant but also on the compositions of both phases. At a temperature below the p.i.t., the surfactant tends to make an o/w emulsion. If the temperature is increased, the solubility of the surfactant in water decreases (cf. Vol. Ill, sec. 4.6c), yow decreases to a very low value, and the emulsion becomes very unstable to coalescence. At the p.i.t., y reaches a minimum. Upon further increase in the temperature, the surfactant becomes increasingly soluble in the oil phase, y increases again, and a w/o emulsion forms. For a homologous series of surfactants,
11 K. Shinoda and H. Kunieda, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1, chapter 5, p. 337, Marcel Dekkcr (1983). B.P. Binks, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 1, p. 1. Royal Soc. Chem. (1998).
8.6
EMULSIONS
the p.i.t. increases with increasing number of oxyethylene groups. Broadly speaking, the HLB number and p.i.t. are positively correlated. The interfacial tension caused. In most emulsions, sufficient surfactant is present to reach a plateau value. This is, for the most part, between 2 and 5 inN m"1 at triglyceride oil water interfaces, and often has somewhat higher values at paraffin oilwater interfaces. Oxyethylated surfactants can give much lower values near the p.i.t. Some anionic surfactants give very low y values at a specific ionic strength, e.g. < 1 uN m"1 for Aerosol OT at 0.05 molar NaCl11.
Figure 8.1. Interfacial tension ( yovr) against temperature for cottonseed oil/water systems with 1 monostearin or 1 monopalmitin (broken curve) in the oil at various concentrations (indicated in %). (Redrawn from data by Lutton et al., loc. cit.)
For many non oxyethylated amphiphiles, yow decreases with decreasing temperature. For those with a saturated aliphatic chain, a sharp break in the y(T) curve can occur, as shown in fig. 8.1, if the surface excess is high21. The temperature at which this occurs, Tc , is called the chain crystallization temperature. It is assumed that the aliphatic chains crystallize in the interface, as in an a-crystal of the amphiphile. The molar melting enthalpy in the interface is of the same order as, but somewhat smaller than, that of a-crystals. Tc is lower by 20 to 50 K than the melting temperature of the crystal, the difference being smaller for a higher value of F. The plateau values of F differ greatly, below and above Tc , being, e.g., 6 and 3 |imol m"2 , respectively, for monoglycerides; the respective values of yow are, e.g., 7 and 20 mN m" 1 . The importance of a low y value is that it facilitates the break-up of droplets into smaller ones during emulsification. On the other hand, the stability against coalescence tends to be smaller. Surface layer composition. Often, the surface excess equals, again, about the plateau value, for the most part one to a few mg m~2. The composition of the surface layer is important for the stabilization of emulsion droplets against aggregation and coalescence. Electrostatic repulsion can occur for ionic surfactants, and it depends 11
R. Aveyard, B.P. Binks, S. Clark, and J. Mead, J. Chem. Soc. Faraday Trans. I 82 (1986) 125. 21 E.S. Lutton, C.E. Stauffcr, J.B. Martin, and A.J. Fehl, J. Colloid Interface Set 30 (1969) 283.
EMULSIONS
8.7
mainly on the surfactant type, pH, and ionic strength. Steric repulsion can be provided by surfactants that have flexible chains protruding in the continuous phase, e.g., Tweens for o/w emulsions. The repulsion depends strongly on the solvent quality for the chains, and hence on temperature. Interfacial rheology will be discussed in sec. 8.1c. (ii) Mixtures of amphiphilesl). In practice, surfactant preparations are mixtures. The length and the number of unsaturated bonds, or any branching of the hydrophobic chain, may differ, and so may the composition of the hydrophilic head group. Moreover, mixtures of amphiphiles with different properties are sometimes used intentionally. A special case is the addition of a, 'cosurfactant', i.e., a weakly surface active substance such as a medium chain length aliphatic alcohol. The application of mixtures rather than pure amphiphiles has some important consequences. The plateau value of the interfacial tension may be either higher or lower than for pure amphiphiles. Presumably, this depends on the effect of composition on the maximum packing density in the interface. Mixtures, especially equimolar mixtures, of an anionic and a cationic amphiphile can yield quite low y values (high surface pressures). Something similar can happen for equimolar mixtures of comparable Spans and Tweens (e.g., sorbitan monopalmitate and sorbitan polyoxyethylene monopalmitate). However, some mixtures yield higher y values than each of the pure components. The equilibrium composition of the interfacial monolayer in an emulsion will be different from, and the interfacial tension will be higher than, that at a macroscopic o/w interface. This is because the area to volume ratio will be much larger in emulsions. Relatively small quantities of a surfactant that gives a relatively low y value can eventually dominate in the macroscopic surface layer, but will remain a minor component in the layer on the emulsion droplets. Moreover, it may take a long time, say an hour, to reach a constant y value at a macroscopic interface. Many amphiphiles with long aliphatic chains are quite poorly soluble in water, and possibly even in oil, at low temperature. They mostly form a-crystals upon cooling, leaving too little surfactant in solution to make an emulsion. Mixtures, however, tend to crystallize very slowly, if at all, and emulsions can readily be made. The most important difference may be in the relatively high value of the dilational modulus of the adsorption layer obtained with mixtures, as compared to pure surfactants at high concentration, which will be explained in sec. 8.1c. It may be the main reason why it is difficult to make emulsions (and especially foams) with a pure amphiphile.
Sec, for a general discussion, see E.H. Lucassen Reynders, in E.H. Lucassen Reynders, Ed., Anionic Surfactants
chapter 1, p. 1.
Vol. 1 1 , Physical Chemistry
of Surfactant
Action. Marcel Dckker (1981),
8.8
EMULSIONS
(iii) Polymers. We will first consider synthetic polymers, whose surface properties are discussed in chapters 1 and 2; see also III.3.4i and III.3.8e, f. The simplest are linear homopolymers, such as poly(ethylene), but these are generally unsuitable for emulsions. They tend to be either insoluble in the continuous phase or non adsorbing. Hence, copolymers, usually of two kinds of monomers, are generally applied. An example is the water soluble polyvinyl alcohol (PVA), which is actually a partially hydrolyzed poly(vinyl acetate); the -OCOCH3 groups left, mostly between 2 and 12%, are hydrophobic and tend to become lodged in an o w interface, the other segments assure solubility in water. Two block or three block copolymers can be made, as well as graft polymers (several short chains covalently bound to a long linear chain) and these may have very good emulsion stabilizing properties, but most of the surfactants used in practice are random copolymers. The conformation of the adsorbed polymer then depends on molar mass (degree of polymerization), fraction of hydrophobic groups, and especially on the distribution of these groups over the chain. The conformation, especially the protrusion of long flexible tails into the continuous phase, is an important variable for the strength of the steric repulsion between emulsion drops. In practice, several kinds of copolymers are used, for o/w and w/o emulsions. Some of these are polyelectrolytes, which can also provide electrostatic repulsion.
Figure 8.2. Equilibrium values of the surface pressure, IT, as a function of concentration in the continuous phase, c , and of the surface excess, r, for the surfactants sodium dodecyl sulfate (S), a PVA sample (P), and p- casein (C; molar mass 24 kDa). Approximate results at o-w interfaces, meant to illustrate trends.
Figure 8.2. shows an important difference between a typical adsorbing polymer (PVA) and most other surfactants: the concentration range over which the surface pressure increases from zero to a plateau value is very wide, e.g., a factor of 105 rather than 10-100. This is because of the wide variation in molar mass and in the distribution of acetate groups. Other things being equal, the largest molecules are the most surface active ones. This means that extensive exchange of molecules at the surface occurs, which is a slow process because an adsorbed polymer molecule will
EMULSIONS
8.9
generally be adsorbed with many segments, and all of these must be desorbed simultaneously. Consequently, it can take a very long time, e.g., a day, before equilibrium values of surface pressure, composition, and conformation of the monolayer are obtained. Moreover, the relationship between surface pressure and surface excess will depend on the polymer concentration and specific surface area, since these factors also affect the various concentrations in the continuous phase. Another characteristic of polymeric surfactants is that the maximum attainable surface pressure is generally lower than that reached by most amphiphiles. Hence, very small values of yow cannot be reached, although there is considerable variation among polymers. Nearly all natural polymers that are used as surfactants are proteins; see chapter 3. Because they are not soluble in oils, they can only be used to make o/w emulsions. An important difference between a pure protein and a synthetic polymer preparation is that all the protein molecules are, in principle, identical. Some of the side groups are hydrophobic, others are negatively or positively charged. Virtually all water-soluble proteins adsorb onto o-w interfaces. They are very surface active, as illustrated in fig. 8.2a: the concentration at which a protein starts to significantly increase the surface pressure is always very small, as compared to most amphiphiles. By and large, proteins of larger molar mass are more surface active. The high surface activity is due to the large reduction in Gibbs energy per molecule upon adsorption, e.g., 104 -10 5 times kT. This also means that desorption will be a very slow process. 'Washing away' of a protein from the emulsion drops by dilution with water (or buffer) is virtually impossible, the more so because the very high surface activity precludes the formation of a significant concentration difference. Desorption is even slow upon compression of a monolayer, especially for large molecules11. On the other hand, the plateau value of yow is not very small, of the order of 10 mN m"1 . Figure 8.2b shows two equations of state, and it is seen that at low F values (expressed in units of mass per unit surface area) the value of n is much smaller for the protein than for the amphiphile. This is, for the most part, due to the very small value of F (expressed in moles per unit surface area for proteins), taking into account the fact that n = RTF for small F. Proteins as emulsifiers are conveniently divided into two classes. One of these contains proteins that form more or less random coils in solution. These include the caseins and the gelatins (although gelatin is an exception in being not very surface active: it has only a small proportion of hydrophobic side groups). They adsorb onto an o w interface in a manner comparable to synthetic copolymers, with trains, loops and tails. However, the conformation in the adsorbed state is much more fixed, since it depends on the invariant primary structure of the protein. The conformation of adsorbed (3- casein is well known. The time needed to obtain an equilibrium 11
F. MacRitchie, J. Coll. Interface Set 105 (1985) 119.
8.10
EMULSIONS
conformation is, e.g., 10 s (at the a w interface)11. Most of the other proteins used are globular. Upon adsorption they change conformation, but they generally do not unfold to a considerable extent: the molecular diameters in the interface and in solution tend not to be greatly different, and the adsorbed proteins may be considered as very small deformable particles21. The time scales involved in obtaining an equilibrium conformation vary between 2 and 15 minutes, but some of the conformational changes occur much faster. Many proteins tend to give roughly the same relation between y and concentration at a given interface, but the equation of state and the surface rheological properties can vary considerably3'. It may finally be noted that proteins are not always applied in their native state. Globular proteins that have been heat denatured before adsorption show markedly altered surface properties. The same holds for proteins that have formed aggregates, in which case the surface coat of the droplets is not a monolayer. The surface excess and structure of the surface layer can also depend on its history; for example, the emulsification process itself may cause protein denaturation and aggregation. (iv) Mixtures of polymers and amphiphiles. This combination can give rise to complicated phenomena, depending on the type and the concentration of both compounds4'. This applies to synthetic polymers as well as proteins. For non-ionic amphiphiles, the situation is often relatively simple. They give a lower y value than most polymers (cf. fig. 8.2a) and upon increasing amphiphile concentration, they gradually displace the polymer from the interface. At first, two dimensional phase separation occurs in the interface, and at about the c.m.c. of the amphiphile all polymer is in solution. The rate of desorption tends to be much faster than will occur upon compression of the interface in the absence of amphiphile. Ionic amphiphiles often bind to polymers, even if the latter bear no charges. This means that the activity of the amphiphile in solution is decreased owing to the presence of polymer. This can result in a higher y value at the same total amphiphile concentration, and the c.m.c. of the latter is apparently increased. If the polymer is surface active, the polymer amphiphile complex will also adsorb, giving its own y value. Nevertheless, at high amphiphile concentration, the polymer is likely to be completely displaced from the o-w interface.
H.K.A.I, van Kalsbeek and A. Prins, in E. Dickinson and J.M. Rodriguez Patino, Food Emulsions and Foams, Royal Soc. Chem., (1999) p. 91. 21 J.A. de Feijter and J. Benjamins, J. Coif. Interface Set 90 (1982) 289. Several data for proteins at the interface between triaeylglyeerol or paraffin oil and water are given by J. Benjamins and E.H. Lucassen Reynders, in D. Mobius and R. Miller, Eds., Proteins at Liquid Interfaces, Elsevier (1998) p. 341. B. Lindman, in K. Holmberg, Ed., Handbook of Applied Surface and Colloid Chemistry, Wiley, (2002) Vol. 1, chapter 20, p. 445..
EMULSIONS
3.1 1
Figure 8.3. Interfacial tension gradients Vy in relation to flow of the adjacent liquids, u = linear velocity. Further see text. (Redrawn from P. Walstra, Physical Chemistry of Foods, Marcel Dekker (2002).)
8.1c The role of interfacial tension gradients^ Figure 8.3a shows an interface between water and oil, devoid of surfactant, where the water is caused to flow parallel to the interface. At the interface, the velocity gradient Vt> equals (du x /dz) 0 . There is thus a tangential (shear) stress, T]wVv0, acting on the interface; rjw is the viscosity of the water phase. The interface cannot withstand a tangential stress and the liquid velocity must thus be continuous across the interface: the interface and oil also move. The shear stress is also continuous, but the velocity gradient is not, because generally z?w ^ r\o . In fig. 8.3b the interface contains a surfactant. The flow will now cause the surfactant to be swept down. This implies that an interfacial tension gradient (Vy) is formed, which exerts a tangential stress of magnitude d//dx. If the gradient can be large enough, the stresses will be equal and opposite, hence;
This section is mainly based on Vol. Ill, sec. 3.6, especially 3.6e and 6f, and on the reviews by Lucassen Rcynders (chapter 5) and by Lucasscn (chapter 6) in Lucassen Rcyndcrs (1981),
mentioned in the General References.
.12
EMULSIONS
w
{ dz J z = 0
dx
If so, motion of the interface and the oil phase is arrested; mechanically, the interface acts as a solid wall for tangential stresses. However, the average Vy can never exceed KI tsx , where Ax is the distance over which the shear stress is acting. Moreover, the gradient can readily relax, as will be discussed below. In other words, the interface shows viscoelastic behaviour. In fig. 8.3c, it is seen that a y-gradient, e.g., as generated by local application of surfactant, may cause flow on both sides of the interface. This is called the Marangoni effect (some authors use this term also for the phenomenon depicted in fig. 8.3b); cf. Vol. Ill, sec. 3.6e. If one of the phases has a much lower viscosity than the other (e.g., at an a-w surface), [8.1.2] will hold. It is of considerable interest to know whether the o-w interface of an emulsion droplet resists motion in the tangential direction, since it can affect droplet deformation and break-up, droplet stability against coalescence, and the (bulk) rheological properties of the emulsion. The question is whether Vy is large enough, and lasts long enough, i.e., during the time that an external tangential stress acts on the droplet. In fig. 8.3b the interface is subject to uniaxial dilation, and the parameter of interest then is the (complex) surface dilational modulus, K^. In practice, the stress acting on a droplet may involve surface shear deformation, but the dilational effect is mostly predominant; moreover, for an interface containing amphiphiles, the surface shear modulus tends to be far smaller than the dilational modulus. The surface dilational modulus is given by (see Vol. Ill, sec. 3.6c, h), K^ = dy/dlnA
( = d^r/dlnD
[8.1.3]
The equality between parentheses is only valid for a Langmuir monolayer (i.e., no exchange of surfactants between interface and bulk), where AF is constant and, moreover, the surfactant should not change conformation during dilation. In that case, the modulus is purely elastic. If the deformation of the interface is uniaxial, we have dy/dx = dy/xd\n
A = KJ Ix , since dlnA = dlnx in this case (for isotropic biaxial
dilation din A = 2dlnx ). Hence, the stress caused by Vy can be given by KJ / x, and for x we can use the droplet radius, a. This leads to a Marangoni number for the ratio of the internal over the external stress, given here by K%/a KfJa Ma = -^— = -^— C7ext iVv
[8.1.4]
where the second equality applies for an external shear stress (cf. Vol. Ill, pp. 3.97 98). For Ma » 1, the droplet surface is rigid (as in fig. 8.3b); for Ma « 1, it is fluid (as in fig. 8.3a).
EMULSIONS
8.13
We thus need absolute values for the time-dependent dilational modulus. For a Gibbs monolayer this is given by, d^/dlnr
Note that the numerator equals K^ for a Langmuir monolayer. For an amphiphile that is soluble in one of the phases and that can exchange freely between the interface and this phase, while surfactant transport is by bulk diffusion, we have, dc n \1'2 ^=- ( l D r )
[8.1.5a,
where T is the characteristic time scale for deformation, e.g., 1/Vu if deformation is due to simple-shear flow. The value of cf2 equals the ratio of r over the diffusion time, 2(d/7dc) 2 ID . Generally, the diffusion coefficient D = O (1CT10 m 2 s" 1 ).
Figure 8.4. Surface dilational moduli. In (a) the dimensionless ratio Q - |Kg|/(d^/dlnr) is given versus the dimensionless time, £2 : after results by J. Lucassen and M. van den Tempel, Chem. Eng. Sci. 41 (1972) 1283. In (b) the value of |Kg| is given for an SDS solution of varying concentration (c, millimolar), for a time scale of 0.02 s (after results by J. Lucassen, in Lucassen Reynders (1981), see General References). Some trends are illustrated in fig. 8.4. Panel (a) shows that for small £, i.e., short T and/or low c, the interface behaves like a Langmuir monolayer; it is as if the surfactant is insoluble. For large £ values, the modulus becomes quite small, the surface properties being close to their equilibrium values. Panel (b) gives an example, for SDS, of the dependence of the modulus on the surfactant concentration in solution, for a given value of r . At low c the modulus rises steeply with c, in accordance with the steep increase of n with increasing F, as seen in fig. 8.2b. At high c, the modulus tends to go to zero, because dT/dc goes to zero, and hence E, goes to infinity. In practice, this is generally not observed. The value of the modulus decreases at high c values, but it remains at a significant level, because the 'surfactant' is, in fact,
8.14
EMULSIONS
a mixture of components, in different concentrations and of different surface activities. However, at longer time scales (several minutes or more) the modulus goes to zero. Some sample calculations may be useful. Assume that an emulsion droplet with a = 1 um is subject to a shear stress of 10 Pa, owing to a shear flow of Vv = 5 x 103 s"1; this implies a characteristic time r of 0.2 ms. At such a small r value the dilational modulus will probably be high, but even if it is as low asl raNm"', the Marangoni number will be as high as 100. This means that the surface behaves as if it is rigid. For a sedimenting droplet, the shear stress acting on the drop is of the order of g aAp , where g is the acceleration due to gravity (10m s"2 ) and Ap is the density difference between drop and continuous phase, say 100 kg m~3 . Even for a modulus as low as 1 |j.N m" 1 , we still find Ma = 103 . Also here, the interface will be perfectly rigid. During emulsion formation the surface excess on the drop may be quite low, but since the time scales involved are very short (e.g., 10 us ), the modulus may still be appreciable. These high values of Ma are, to an important extent, due to the small value of a, hence the quite high value that Vy can have, even at a low value of n. However, all of the discussion above applies to amphiphiles, whereas for polymers the trends may be quite different. Equation [8.1.5] cannot be applied, because the polymer can, and generally will, change conformation upon expansion or compression of the interface, which can markedly affect the values of n . Nevertheless, the values of the surface dilational modulus are generally high. For several proteins, at a range of concentrations, values for |Kg| ranging between 20 and 70 mN m"1 were observed at a time scale of about 1 s, and for some PVA samples between 10 and 15 mN m"1 1). This means, again, that the interfaces tend to be quite rigid, for the most part owing to the slowness of any change in surface excess. In other words, the interface behaves more like a Langmuir layer. At a longer time scale, say of an hour, the moduli tend to be smaller, e.g., 5 mN m" 1 . At short time scales, the dilational modulus would thus be given roughly by d ^ / d l n T . Figure 8.2b shows that for (3-casein |Kg| will be quite small at low values of F, and the same holds for other proteins. Such a situation can occur during emulsion formation. At the very short time scales involved, K may even be smaller at low F, because the molecules have not yet had time to expand in the interface and thereby increase the value of n2). Altogether, the modulus will be very small for polymers at quite low r values. We should add that the rigidity of a polymer coated interface may be higher if the interfacial deformation is not truly dilational, but also has a shear component. This has been observed for globular proteins and bituminous substances. Furthermore, the equations and results given for |K^| are only valid for quite small values of the strain. During emulsification, very large strains and very high strain rates can occur, and the
21
J. Benjamins, E.H. Lucassen Reynders, loc. cit. J.A. de Feijtcr, J. Benjamins, J. Colloid Interface Set 40 (1982) 289.
EMULSIONS
8.15
modulus then Is highly non Newtonian. Very little is known about the dilational behaviour under such conditions.
8.Id Emulsion Properties Several characteristics of emulsions have been mentioned or briefly discussed in the previous sections. Some other properties are discussed below. (i) Viscosity1 . In table IV.6.4 a collection was given of semi empirical formulas for the (p dependence of the viscosity of dispersions of solid particles. Of these, we repeat the Rrieger-Dougherty equation [IV.6.9.10], which is also often used for emulsions: ^=^ c (l-^/^ m a x r [ " 1 «'max
[8.1.6]
where
18.1.7]
For a polymeric surfactant, J = 1 0 n m is a reasonable value, and for droplets of a = 0.5 pirn this results in (3eff = 1.06(p . In most emulsions the effect will be smaller. (2) The fluidity of the disperse phase may contribute to (i.e. decrease) the emulsion viscosity, according to the Taylor equation: nn + 0.4??., [^] = 2.5-2 S-
[8.1.8]
However, it is implicitly assumed that the viscous stress is continuous across the drop's interface or, in other words, that Kg = 0, see [8.1.5]. As discussed in sec. 8.1c, under nearly all conditions the value of Kg will be large enough to prevent any lateral movement of the interface, and hence any flow inside the drops. The drops will thus behave as solid spheres. (3) A drop can be deformed by the shear stress acting upon it in the rheometer, which would increase the viscosity. As mentioned in sec. 8.1a, the deformation is negligible under most conditions. See sec. 8.2.6, sub (i), for the factors determining the deformation. (4) Colloidal interactions between the drops will, in principle, affect the value of r;.
11
Primarily based on a review on emulsion rheology by E. Dickinson, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 5, p. 145. Royal Soc. Chem., Cambridge, 1998.
8.16
EMULSIONS
Figure 8.5. {a! Apparent shear viscosity r]a versus shear stress, a , for emulsions (mineral oil in water, d 32 = 0.55 |im , stabilized by Tween 20) of various volume fractions,
As will be discussed in sec. 8.1g, the effect is quite small in most emulsions, unless the net attractive forces are large enough to cause permanent aggregation of drops. Figure 8.5 shows some representative experimental results. Up to
11
E. Dickinson, M.I. Goller, D.J. Wedlock, J. Colloid Interface Set 172 (1995) 192. Sec e.g. the review by H.M. Princen, in R.D. Bee et al., Eds., Food Colloids, chapter 2, p. 14, Royal Soc. Chem. (1989). 21
EMULSIONS
8.17
Semi empirical theory has been developed, in which the numerical constants have been found by fitting to experimental results. For the elastic shear modulus we have G = 0.9p La
the emul-
sion will exhibit flow. Expressions for the yield stress and the apparent (shear rate thinning) viscosity have also been given, but these are less certain. In principle, other variables will contribute to the results. These include the surface dilational modulus of the o-w interface and, for the viscosity, the value of //D . These aspects seem to have been insufficiently studied. Furthermore, the viscosity will be enhanced by aggregation of the drops, and by addition of materials, especially polymers, that increase the viscosity of the continuous phase. In these cases, the emulsion will often show visco elastic behaviour. (ii) Optical properties.
The optical appearance of an emulsion is generally
dominated by the droplets' light scattering. The scattering is also of importance as a tool in the estimation of droplet size distribution (sec. 8.1e), or interactions between droplets (sec. 8.1g). Light scattering by small particles is discussed in 1.7, especially sec. 7.8. Small droplets scatter light according to the Rayleigh-Debye theory, which implies that the shorter the wavelength, the stronger the scattering. For larger drops, say > 1 \im , we approach the domain of anomalous diffraction; the scattering then is dominated by interference between light passing through
and light passing by
a
11
particle . The scattering coefficient is given by, 4
4 sinp + —(1-cosp) p p2V ^ p = 2xd\An\/A. 9 =2
[8.1.10]
where p is called the phase shift parameter; An is the difference in refractive index between droplet and continuous phase, and XQ is the wavelength of the incident light in vacuum. The theory is approximately valid for p > 2 and |An|<0.15 (accuracy better than 10%). 11 H.C. van de Hulst, Light Scattering by Small Particles. Wiley, (1957); there is also a Dover edition, Dover (1981).
8.18
EMULSIONS
The total scattering intensity is given by the turbidity, 3;r^|An|g>
[8.1.11]
which has dimension [L '|. The absorbancy of a dispersion is given by 0.434rf, where C is the optical path length. Accurate values of Q can be calculated by the rigorous Mie theory, which applies for spheres of arbitrary size and refractive index. An example of r/g> versus p is given in fig. 8.6.
Figure 8.6. Turbidity over volume fraction (in um^ 1 ) versus phase shift parameter p for homodisperse spherical particles, calculated for |An| - 0.1.
The appearance of an emulsion is due to diffuse reflection, a complicated multiple scattering phenomenon. Nevertheless, it correlates quite reasonably with the turbidity, provided that (. (i.e., the thickness of the emulsion layer), is not very small (say, over a few mm). For white light, i.e. average XQ = 0.55 |j.m, we may conclude that for particles of various diameters, the appearance of an emulsion will be: d = 0.02 |im grayish, almost transparent d = 0.2 p.m bluish, transmitted light red d =2 |im white d = 20 |im less white, maybe some colour The smallest particles scatter very little light (Rayleigh scattering). Those of about 0.2 |im scatter more, and the intensity depends strongly on the wavelength (RayleighDebye scattering), although the resulting colours will not generally be conspicuous because of the polydispersity. In accordance with fig. 8.6, the maximum whiteness will be observed for about 2 \\m droplets. For still larger drops, the scattering is less and the whiteness even more so, because a substantial part of the scattered light is scattered at very small angles from the incident beam. Furthermore, total scattering is proportional to particle concentration. (For very small drops and high
EMULSIONS
8.19
besides scattering, and the emulsion may show colour. However, for nearly all emulsion droplets n'
[8.1.12]
where Kc is the conductivity of the continuous phase and 'em' refers to the entire emulsion. For o/w emulsions, Kc is high; for w/o emulsions, it is low. The trend is that with increasing co: K increases, E' decreases and e" goes through a maximum; some illustrations are in figs. II.4.37 39. The decrease of e'(co) has been analysed in some detail for dilute sols of charged spherical particles in electrolyte solutions. The presence of electric double layers leads to gigantic values of e' at low frequencies; with increasing co the value of e' relaxes at a frequency that depends, inter alia, on the particle radius. Models studied include latices and microemulsions. For true emulsions, the situation is different, because; (i), they are rarely dilute; (ii), they are generally polydisperse; and, (iii), depending on the type, polarization of the disperse phase must also be considered. Condition (ii) implies a spectrum of relaxation times, hence broadening of the e'(co) decay and the e"(co) maximum. Dielectric relaxation measurement is not a suitable method for determining particle size distributions. 11
T. Hanai, in P. Sherman, Ed., Emulsion Science, chapter 5, p. 353. Academic Press, (1988); and M. Clausse, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1, chapter 9, p. 481, Marcel Dekker (1983).
8.20
EMULSIONS
Condition (iii) is challenging: the permeability of ions, migration across and tangential to the interface of ions and surfactants, slip phenomena, etc., are all reflected in e'(a>) and e"{co). Models to implement this would be needed. Awaiting such general studies, most investigations have a semi empirical nature. For instance, o/w emulsions stabilized by lecithins were studied . The e((a) spectrum could be interpreted by four parameters, including a storage and a loss contribution of the interfacial layer, which enabled the identification of lecithin preparations under some conditions. At high frequencies, only the continuous phase permittivities remain, which differ between oil and water: relaxation in this range can help detect the oil and water contents. This procedure is also in use for oil prospecting in rocks. 8.1e Determination of drop size distribution Virtually all emulsions are polydisperse, so we need to know the size distribution2 . The cumulative distribution of the droplet diameter, d, will be called F{d); it gives the number of droplets of size smaller than d per unit volume, versus d; the dimension is [ L~3 ] (where L stands for length). The size frequency distribution is then defined as, f{d) = dF(d)/dd, of dimension [L~4]. Often, the volume frequency distribution is used, given by (7r/6)d 3 /(d), dimension [ L"1 ]. A useful auxiliary parameter is the n th moment of the distribution, given by Sn = J d n / W ) d d o
[8.1.13]
which has dimension [Ln~3]. See also IV.app. 1. The quantity S o gives the total number of droplets per unit volume, nS2 the specific surface area Av of the droplets, etc. Any type of average droplet diameter can be expressed by; d a b =(S a /S b ) 1 / l a - b )
[8.1.14]
which has dimension [L]. For the same distribution, an average type of higher order (a + b) will have a larger value. The number mean diameter is given by d10 , etc. Often, the average type d 32 (also called the volume-surface diameter d ys , or Sauter mean diameter) is used, in part because of its relationship to the specific surface area; see [8.1.1]. The spread in size is best expressed by a relative standard deviation,
11
R.M. Hill, E.S. Bcckford, R.C. Rowc, C.B. Jones, and L.A. Dissodo, J. Colloid Interface Set 138 (1990) 521. An overview of distribution functions used for emulsions is by C. Orr in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1, chapter 5, p. 369, Marcel Dekker (1983).
EMULSIONS
8.21
'n-f^-lf2 I b n+l
18-1-151
)
For example, c2 is the relative standard deviation of the surface weighted droplet size distribution, and so it corresponds to the average d 32 . If the distribution is not too wide, say, c 2 < 0.5 , the distribution is usually close to log normal and can be characterized with sufficient accuracy by the parameters d 32 and c 2 . If c2 is large, the whole distribution should be given; it may have various kinds of shape, e.g. with a long 'tail' or even bimodal. Although the droplets in an emulsion are generally homogeneous spheres, reliable and precise determination of the size distribution is far from easy. When applying two different methods, it is not exceptional if the values obtained for a given type of average diameter differ by a factor of two; the distribution shapes obtained may also differ substantially. Some of the methods used are listed in table 8.1". The experimental methods vary widely in principle. Simplest is the determination of the diameter of a large number of single droplets by microscopy, both light and electron microscopy, applied in various modes of specimen preparation. Currently, image analysis is often applied. Very useful for emulsions is confocal scanning laser microscopy (CSLM), where the image is obtained directly in digital form; when using a fluorescence microscopy mode, the limit of resolution is quite good. If the specimen is liquid, it is generally advisable to suppress particle motion (sedimentation, Brownian motion, convection currents) by the addition of a transparent gelling agent. Some methods detect individual particles. The highly diluted emulsion flows through a small sensing zone, where individual particles are sensed and sized. Best known are the 'Coulter counter' and comparable instruments, where the particle increases the electrical resistance in a narrow opening; the change in resistance is proportional to particle volume. In another method, the amount of light scattered by each particle is monitored; this needs a calibration curve to convert scattering intensity to particle size21. These methods tend to be relatively accurate. A problem may be the reliability of the correction for coincidence, which is needed because two, or even more, particles may simultaneously pass the sensing zone. In most other methods a macroscopic property is determined, often as a function of some external variable, from which the data are converted into an average size or a size distribution. Most popular are light scattering methods, nearly always of strongly A general book on methods is T. Allen, Particle Size Measurement, 5th ed., Vol. 1. Chapman & Hall, (1997). Specifically for emulsions, C. Orr, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 3, chapter 3, p. 137, Marcel Dekker (1988). 21 See e.g. E.G.M. Pelsers, M.A. Cohen Stuart, and G.J. Fleer, J. Colloid Interface Set 137 (1990) 350.
8.22
EMULSIONS
Table 8.1 Summary of methods for the estimation of droplet size distributions Method
Determines
Droplet size range (|rm)
Dilution needed
Problems/remarks
I Id) fid)
> 0.3 > 0.1
Generally Generally
Electron microscopy
J(d)
> 0.003
No
Tedious, several pitfalls Dcconvolution; immobilizing may be needed Artefacts; 'tomato salad' analysis
Particle sensor Electrical resistance Light scattering
F(d)
fid)
0.5-300 0.2-10
Greatly Greatly
Coincidence correction Tedious; can handle droplet aggregates
Greatly
Microscopy Light microscopy CSLM
not
Static light scattering Forward scattering Spectroturbidimetry
W)
Ttt)
0.4-200 0.2 10
Yes
Matrix inversion Tedious,
Dynamic light scattering Photon correlation spectroscopy
/(At)
<1
Yes
'Average' only
Diffusion time
1-10
No
Rather tedious; no full distribution
Nuclear Magnetic Resonance Pulsed field gradient proton NMR*
Sedimentation Under gravity
d
53
2-100
Generally
Several
Centrifugal
d
53
0.1-5
Generally
Several
Note*: especially for w/o emulsions. Symbols: / = intensity; 8 = scattering angle, r = turbidity; A = wavelength; At = time shift;
diluted emulsions. Often, forward scattering of laser light is measured over a range of scattering angles. By use of Fraunhofer diffraction theory (for large particles) or of Mie scattering theory (valid for spheres of arbitrary size and refractive index), it has been tried to convert the angular spectrum into a size distribution. The available instruments are very easy to use, and they immediately produce a size distribution, as well as some distribution parameters. However, although the software for the methods has gradually improved in quality, tire results can still be insufficiently reliable, especially in the small particle range. Another method is spectroturbidimetry: the turbidity is
EMULSIONS
8.23
measured over a range of wavelengths, and the spectrum obtained is compared to spectra calculated for various size distributions11. The volume fraction and the refractive indices (as a function of wavelength) must be known. The method is somewhat tedious but gives reliable results, especially for droplet sizes of order 1 ^m . Approximate values of d 32 > 3 |im can be obtained from measurements of turbidity over cp at one (relatively short) wavelength. Besides these static scattering methods, dynamic (or quasi elastic) light scattering can be useful21; basic aspects are in sees. 1.7.6-8. Application to very dilute systems will be considered first. Because the particles show Brownian motion, the wavelength of the scattered light is slightly affected, whose shift is detected somehow. Various modes exist, although photon correlation spectroscopy is now the method of choice. Anyway, what is essentially measured is the diffusion coefficient of the particle, which can, via the Stokes-Einstein relation, be converted into an equivalent sphere diameter. In this way, the size of homodisperse small particles (d < 1 um) can be estimated accurately. For polydisperse emulsions, the problem is that the scattering intensity is strongly dependent on size, so the diameter obtained is an average of high and unknown order. Neither can the width of the distribution be established with accuracy. More sophisticated methods, using a range of scattering angles, are in principle more powerful31. Concentrated systems can be investigated by various forms of fibre optics quasi elastic light scattering41, one form being known as, 'diffusing wave spectroscopy'. A problem is that at low
P. Walstra, J. Colloid Interface Set 27 (1968) 493. A clear introduction, including possibilities and limitations of various methods, is given by D.S. Home in E. Dickinson, Ed., New Physico Chemical Techniques for the Characterization of Complex Food Systems, chapter 11, p. 240, Blackic (1995). 31 See e.g. D.S. Home, D.G. Dalglcish, Eur. Blophys. J. 11 (1985) 249. 41 Application to emulsions is discussed by P. van der Meeren, M. Stastny, J. Vanderdeelen, and L. Bacrt, Colloids Surfaces A76 (1993) 125. J.C. van den Enden, D. Waddington, H. van Aalst, C.G. van Kralingen, and K.J. Parker, J. Colloid Interface Sci. 140(1990) 105. 21
8.24
EMULSIONS
ultrasound attenuation, to give immediate results. Figure 8.24 Illustrates some sedimentation profiles. Some other methods are not (yet) very suitable for emulsions. These include, for example, size exclusion or hydrodynamic chromatography (d < 1 urn); field flow fractionation in a centrifugal field (d-cl^im); electro acoustic methods; and ultrasound velocity measurements. Accuracy. Several kinds of uncertainty can arise In the determination of size distributions. Systematic errors can occur readily for Indirect methods. The signal measured can depend on factors other than particle size. The relationship between the magnitude of a signal and the particle size may not be known with sufficient accuracy; often a linear relationship Is assumed, but Is not always true. Several methods underestimate the number of the smallest particles, or do not even notice them. This means that the average size is overestimated, especially for averages Involving So , such as d10 and d 30 ; an estimate of d 32 will then be closer to reality. Even for direct methods, such problems may exist. Several microscopic methods, in fact, see cross sections or thin slices of the material. Then a number of circles is observed, and the problem Is to convert their diameter distribution to that of the original spheres; this is known as the 'tomato salad problem'. Good solutions exist for spheres, which are to be found in texts on stereology11. The droplets are often allowed to sediment (or cream) before they are viewed, and the small ones may then escape notice. Furthermore, it may be difficult to distinguish between separate particles that are close to each other and previously formed aggregates, especially when using an image analysis programme. In indirect methods, the presence of aggregates may completely upset the results, and care should be taken that the droplets are fully deaggregated. Conversion of the raw data to a size distribution especially poses problems for indirect methods. For instance, In scattering methods a range of data (a spectrum) has to be determined, be it a range of wavelengths or a range of scattering angles, to allow derivation of anything more than an average size. If the particle size distribution and the refractive indices are known, it is relatively easy to calculate a spectrum. However, the Inverse problem, calculating the distribution from a spectrum, is far more difficult, especially because the amount of information and Its accuracy are limited. The algorithms involved always involve rounding off, or even shortcuts, and may even be based on a given type of size distribution. All of this may lead to considerable error. Finally, the reproducibillty should be taken into account. Sampling should be representative. Sizing may give rise to random errors. The greatest uncertainty is generally due to random errors in the counting or, more precisely, in establishing the number of particles in a size class. This is largely owing to the Poisson statistics of 11
See e.g. E.R. Weibel, Stereological Methods, Vols. 1 and 2, Academic Press, (1980).
EMULSIONS
8.25
counting. The standard deviation of the number of particles in a certain volume is, for a completely random distribution, equal to the square root of the average. This means that the relative standard deviation of the number of particles in a size class, i, is equal to, or larger than, 1/ JN~ , where JV; is the number actually counted (i.e., before any multiplication with a dilution factor, etc.). Countingjust one particle thus leads to an uncertainty (relative standard deviation) of over 100%. This becomes especially manifest for very large particles, which means that the large particle end of a volume distribution is often subject to considerable error. Especially if the size distribution is wide (high value of c2 ), tens of thousands of particles may have to be counted to obtain reliable results. Another variable of great importance is the total volume fraction [cp = (TI/6)S 3 ] of the droplets. For some methods of estimating size distributions, (p can be derived from the results, but it is better to use this as check, after having obtained the value of cp in another way. Deriving it directly from the quantities of materials used in making the emulsion is often inaccurate, especially when making small volumes. Chemical analysis of the components making up the disperse phase, or of an added marker substance that only dissolves in the disperse phase, is to be preferred. 8. If Determination of interfacial properties This section concerns interfacial tension, interfacial rheology, surface excess concentration and surface layer composition. Nearly always, interfacial properties are determined at a macroscopic interface. The continuous phase as used in making the emulsion, i.e., with surfactant(s) added, and the material making the disperse phase, are brought into contact and measurements are made on the interface. These may concern static or dynamic interfacial tensions, and surface rheology (see Vol. Ill, sees. 3.7e for Langmuir monolayers, and 4.5 for Gibbs monolayers). However, the macroscopic monolayer can be representative of the monolayer around the droplets in only one case, when there is one (pure) non polymeric surfactant, which is insoluble in the disperse phase, and its concentration in the emulsion is in excess of the concentration that provides a plateau value of y. In all other cases, the composition of the macroscopic monolayer will not equal that around a droplet in the emulsion. This is because the surface area to volume ratio is higher in the emulsion than in the macroscopic case by some orders of magnitude, and because some of the surfactant species will be depleted more strongly than others during emulsification. Moreover, it may take a long time to obtain equilibrium. See also sees. 8. lb and c. What can be done, is to make the emulsion and, after allowing some time for equilibration, separate the continuous phase by centrifugation or microfiltration. This liquid and the material of the disperse phase are brought into contact, providing an interface. After keeping the system for a considerable time (some hours often suffice)
8.26
EMULSIONS
the interfacial composition may be virtually equal to that in the emulsion, provided, again, that the surfactants do not dissolve in the disperse phase. Moreover, the separated liquid must be free of droplets. For some polymeric surfactants, notably proteins, the monolayer composition in the emulsion may again be different, because an equilibrium interfacial composition may not have been reached; see sec. 8.2c. In some cases, the interfacial tension can be measured at the surface of a droplet, provided this is larger than, say, 4 \im. Under a microscope, a drop in an emulsion is partly sucked into a very narrow cylindrical capillary, and the pressure difference to achieve this situation is established. This must be equal to the difference in Laplace pressure, given by 2/(1/f^ - 1 / R2) • Here, Rj is the radius of the capillary and R2 that of the protruding part of the drop, and these radii can be measured. In this way, the interfacial tension can be obtained11. It is a very tedious method, but can be used if needed to check results obtained at a macroscopic interface. Altogether, reliable data on interfacial tension and rheology of monolayers on emulsion drops are hardly available. Estimation of the surface excess F is, in principle, not difficult, although it is not so easy to obtain high accuracy2'. One determines in the emulsion the values of (p, d 32 (which yield Av =6) + 6 / > / d 3 2
[8.1.16]
where cc is the concentration in the continuous phase. After separating most droplets from the emulsion by centrifuging or microfiltration, the same values are determined in the 'skimmed' emulsion. We now have two linear equations with the two unknowns, F and c c , which can therefore be calculated. The assumptions underlying this method are that no desorption of surfactant occurs during separation, that the surfactant does not dissolve in the disperse phase, and that F does not depend on the droplet size. Often a simpler method is tried: one determines cc in a skimmed emulsion in which (p is virtually zero, which yields the amount of surfactant adsorbed. However, this is not really possible if the emulsion contains a significant amount of very small drops, say < 0.4 |a.m . The composition of a mixed adsorption layer can be obtained by determining F for each surfactant
present. Some polymers, especially proteins, effectively
give a
Langmuir monolayer. In such a case, one can 'wash away' the dissolved surfactant by repeated dilution and skimming of the emulsion. One may then desorb the adsorbed
11 21
L.W. Phipps, D.M. Temple, J. Dairy Res. 49 (1982) 61. H. Oortwijn, P. Walstra, Neth. Milk Dairy J. 33 (1979) 134.
EMULSIONS
8.27
polymer by adding a suitable amphiphile, and remove the disperse phase material (generally oil). The polymer solution can then be analyzed by a chemical method11. If the emulsion drops are small, and q> is not high, addition of an excess of SDS not only dissolves the protein: the oil is also 'solubilized' in SDS micelles. The resulting liquid can be analyzed directly by SDS gel electrophoresis11. 8.1g Determination of colloidal interaction forces Knowledge of the colloidal interaction Gibbs energy (G), or the corresponding force (F), between emulsion droplets as a function of inter droplet distance, (h), will be very useful, since these are essential parameters in determining whether aggregation and/or coalescence of drops will occur. Pair interactions between hard surfaces are discussed extensively in ch. IV.3, which also gives methods of determination. We shall now discuss the extent to which these methods also work for emulsions. Nearly all measurements are done on macroscopic surfaces, e.g., determining the disjoining pressure in a liquid film in various states of drainage, or measuring directly forces (in the surface force apparatus), as a function of h , between very smooth solid surfaces covered by surfactant monolayers. The results are generally not representative for emulsions. In the first place, the monolayers involved may differ in composition from those on the emulsion drops, because, (i), the adsorbent is generally different (a-w or s-w, rather than o-w) and, (ii), the monolayer will generally not have the same composition, for reasons given in the previous section. Nevertheless, some interesting results have been obtained, which do throw light upon the colloidal interactions between emulsion droplets. Secondly, the time scales involved can be quite different. Measurements on macroscopic surfaces take at least several minutes. The time during which two emulsion drops are close to each other is far shorter. Consider droplets of diameter 1 urn in water. The time, t, needed for a particle to diffuse in a given direction over a distance Al = 2-Jot , where D is the diffusion coefficient. From the Stokes Einstein relationship, we find for an individual droplet, D = 4xlO~ 13 m 2 s"1 . The pair diffusion coefficient will be twice as large, but at close approach the diffusion will be greatly slowed down. When, for example, we consider diffusion from 25 to 5 nm interparticle distance, it is by a factor of about 100 21. Hence, we will have D = 10"14 , and the time needed for diffusion over 20 nm would be about 10 ms. This is shorter by some orders of magnitude than the time needed to obtain an equilibrium composition for a mixture of surfactants, and an equilibrium conformation of adsorbed polymers. Consequently, it would be desirable to make dynamic measurements directly on emulsion droplets. One method, called 'colloidal particle scattering' has been
11
J.A. Hunt, D.G. Dalgleish, Food Hydrocoll. 8 (1994) 175. See e.g. L.A. Spielman, in K.J. Ives, Ed., The Scientific Basis of Flocculation, Sijthoff & Noordhoff (1978) p. 65. 21
8.28
EMULSIONS
Figure 8.7. Principle of the colloidal particle scattering apparatus, (a) Shown are a fixed particle (F) and a mobile one (M) in its original position, as well as the trajectory of the latter in the z-y plane, (b) A cross section through F in the x-z plane {seen in the y direction), v = liquid velocity, a = particle radius, R and 9 are the cylindrical coordinates of the centre of M.
developed by van de Ven and co workers11. The principle is illustrated in fig. 8.7. On a plate is fixed a particle and a dilute dispersion of the same particles flows past it, in simple-shear. The trajectories of several particles that come close to the fixed particle are monitored via a light microscope. These particles will be deflected and show a trajectory roughly as depicted, which is calculated for the case where there is only hard core repulsion between the particles. The trajectories before and after the encounter are considered. The coordinates of the moving particle will have changed in the x and the z directions. When using cylindrical coordinates, as depicted in fig. 8.7b, it is generally observed that the angle 0 changes very little; hence, the deflection can be expressed in the change in the radial coordinate, AR. Let ARQ be the (calculated) deflection if only hard core repulsion and no attraction occurs. When colloidal repulsion acts between the particles, e.g., of electrostatic or steric nature, the result should be that AR > ARO . For net attraction, the opposite is to be expected. In fact, the static particle may capture the moving one if the attraction is strong, but the latter generally escapes after rolling for some distance over the former one, owing to the presence of the supporting plate. Nevertheless, the proportion of encounters that lead to capture can be established. One can try to estimate G(h) by comparing observed trajectories, or merely AR values, with those calculated for various assumed relationships between G and h . This is no simple task. First, the calculation of the trajectories is intricate, time consuming, and not completely exact. Second, there are some important disturbances, since the particles are also deflected by sedimentation (in a single direction) and by Brownian motion (in a random way); the latter means that the translational Peclet number for the particles (shear force over the diffusional force, i.e. d 2 Vu/8D, where ''T.G.M. van de Ven, P. Warszynski, X. Wu, and T. Dabros, Langmuir 10 (1994) 3046. We shall largely follow papers from the group of Dickinson, especially M. Whittle, B.S. Murray, E. Dickinson, and V.J. Pinfield, J. Colloid Interface Sci. 223 (2000) 273; and M. Whittle, B.S. Murray, and E. Dickinson J. Colloid Interface Sci. 225 (2000) 367.
EMULSIONS
8.29
Vu is the shear rate, and D is the particle diffusion coefficient) should be high, say >100. Third, the accuracy needed may be higher than the apparatus can provide; moreover, the particles should be perfect spheres and not too small (say, > 4 nm ) to obtain reliable results. This means that the method is not suitable for routine work. On the other hand, some interesting results have been obtained, although the experiments often involved density matching to minimize sedimentation, and increasing the viscosity to suppress Brownian motion; the inherent compositional changes may therefore have affected the surfactant adsorption. Moreover, emulsion drops are not as easy to handle as polystyrene latex particles, and the latter are often used as a model. Although these may be good as an adsorbent for surfactants such as proteins, where the monolayers mimic those on oil droplets, the van der Waals attraction between these particles is inevitably different from that between oil drops (see e.g. IV.app. 3). For emulsions stabilized by polymeric surfactant(s), it may be interesting to estimate the hydrodynamic thickness of the adsorbed monolayer, since this parameter correlates well with the range over which steric repulsion acts. If the average droplet size is small (say, d < 0.6 |im ), this can be done by determining the droplet size by photon correlation spectroscopy before and after removal of the polymer from the surface by adding a suitable small molecule surfactant11.
Some macroscopic 'solution' properties can also yield information on colloidal interaction forces, provided that no irreversible aggregation of particles occurs. One group of methods employs light scattering (discussed in sec. IV.2.3). Static scattering in the Rayleigh-Debye domain at a range of volume fractions results in a structure factor that depends on colloidal interactions. It is limited, however, to quite small particles (say, d < 0.6), which makes it unsuitable for most emulsions. For small particles, interaction forces can also be derived from dynamic light scattering, but the method is rather unreliable for nearly all emulsions, owing to their moderate to strong polydispersity. Another method involves the determination of specific viscosity as a function of volume fraction. For hard particles, the relationship is often given as; = kl
tlsp=IL^
[8.1.17]
We where r\c is the viscosity of the continuous phase, see sec. IV.6.9a. For hard, noninteracting spheres, /Cj = 2 . 5 . For ^><0.15, the third term is generally not needed. The second term is due to volume exclusion of, and pair Interactions between, the particles. For hard non-interacting spheres, k2 = 6 . The constants are different when colloidal interaction forces act between the particles. A repulsive force primarily leads to a (small) increase in kl, depending on the strength and range of the interaction. In 11
D.G. Dalgleish, Coll. Surf. Bl (1993) 1.
8.30
EMULSIONS
the case of attractive forces, k2 is increased". In a common treatment, where the attractive free energy is depicted as a narrow square well, k2 would equal (5.9+ 1.9/r B ), where rB is the Baxter parameter, related to the second virial coefficient for, say, osmotic pressure by B2 = 4 - l / r B . For B2 = 0, r =0.25, and for stronger attraction its value is smaller. The equation predicts results that agree well with observations, as long as the spheres are not very polydisperse. The treatment is based on the assumption that the translational Peclet number, Pe =
3jid3f7rVu -— 8kT
[8.1.18]
is much smaller than unity, which needs quite small particles and low shear rates. Ideally, T]o should be determined, i.e., as an extrapolated value at zero shear rate. For most emulsions, however, Pe » 1 . This implies that the constants in [8.1.17] will be different, but the theory has neither been well developed, nor checked. Moreover, any effect of colloidal interaction will be relatively weak, since hydrodynamic effects are prevalent. This means, again, that interaction forces cannot be determined readily in most emulsions. Attractive forces leading to irreversible aggregation naturally give rise to aggregates (coagulates or flocculates). This will greatly increase viscosity, owing to the increase in effective volume fraction, and also increase light scattering. This yields possibilities for determining aggregation. However, one may also use a simple light microscope to detect aggregates. 8.2 Emulsion Formation Emulsion formation involves several aspects and can be achieved in several ways. The literature on this subject is vast, and we will primarily follow the contents of three reviews by the present author2', supplemented with newer literature. 8.2a Introduction To make an emulsion, one needs oil, water, surfactant and energy. Generally, mechanical energy is employed, but chemical or electrical energy can also play a part. Energy is needed because the interfacial free energy of the emulsion is higher than that of the original non emulsified mixture, by an amount Avy. The counteracting increase in mixing entropy upon emulsification is negligible. If we assume that we make an emulsion with cp = 0.1 and d 32 =0.3 |im , and that the final value of y = 10 mN m" 1 , the increase in Gibbs energy amounts to 20 kJ m~3. However, the amount of 11 21
See e.g. A.T.J.M. Wouterse and C.G. de Kruif, J. Chem. Phys. 9 4 (1991) 5739. P. Walstra, (1983, 1993), and P. Walstra, P.E.A. Smulders (1998); see General References.
EMULSIONS
8.31
mechanical energy expended is likely to be about 20 MJ m~3 . This huge excess is to a small extent, due to the droplet surface area and the y value being temporarily larger than their final values; moreover, this occurs a number of times (say, 100) during the process. However, by far most of the mechanical energy is dissipated into heat, since it acts on all of the liquid during all of the time of emulsification. The process is thus very energy inefficient. Much research is thus aimed at increasing the efficiency. Moreover, it is often tried to obtain narrow drop size distributions. The larger the droplets, the lower generally is the emulsion stability , but the amount of energy needed is larger for smaller droplets. The optimum is clearly a homodisperse emulsion, and methods have been developed to achieve this, but have not yet met with much success for large-scale manufacture. (i) Emulsification methods. A wide range of apparatus and process-conditions can be applied to make emulsions. The underlying principles can be classified as follows. 1. Nucleation and growthl). In some cases, the material intended to make up the disperse phase is, to some extent, soluble in the continuous phase, e.g., water in benzene. It may then be possible to realize supersaturation, e.g., by altering the temperature or by partial evaporation of the continuous phase. This will then generally cause nucleation, followed by growth, of the disperse phase. Fairly homodisperse drops can often be obtained, and if needed, nucleation may be regulated by adding a limited amount of tiny catalytic impurities. However, coalescence of the newly formed drops becomes appreciable when d and
For a review see B. Vincent, Z. Kiraly, and T.M. Obey, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 3, p. 100, Royal Soc. Chem. (1998). 21 See Ill.fig. 5.47. There is considerable literature on the subject, for the most part of a practical nature. A good overview of all the variables affecting the results is by V. Schroder, Herstellen von Ol-in-WasserEmulsionen mit mikroporosen Membranen, Shaker Verlag, (1999). Some fundamental aspects are discussed by N.C. Christov, D.N. Ganchev, N.D. Vassilcva, N.D. Denkov, K.D. Danov, and P.A. Kralchevsky, Colloids Surf. A209 (2002) 83.
8.32
EMULSIONS
pores. Along the membrane is made to circulate a surfactant solution, the future continuous phase:: the two phases are thus in cross flow. Oil 'drops' emerging from the pores are dislodged from the membrane by viscous forces exerted by the water phase. To allow dislodging, the membrane must be preferentially wetted by the continuous phase. Ideally, all pores should have the same diameter, and be separated from each other by a distance of at least three diameters, to minimize immediate coalescence of emerging drops. If these conditions are fulfilled, and a suitable surfactant is present in sufficient concentration, and the flow in the pores is relatively slow, a roughly homodisperse emulsion is formed, with d ~ 3d m .
Figure 8.8. Membrane emulsification. Cross section through a cylindrical emulsification module, pj — p 2 is the trans membrane pressure. Also, see text.
The trans-membrane pressure should be at least larger than the Laplace pressure in the pore, i.e.,4y/dm, which ranges for the most part between 2 and 100 kPa. To achieve a superficial oil flow rate (volume flow rate per unit ofmembrane surface area) that is large enough for practical purposes, e.g., 10~5 m s" 1 , a much higher pressure has to be applied, e.g., by a factor of five. The result is that the drops are larger and far less homodisperse. Factors leading to a smaller average droplet size are: a smaller pore diameter; a lower trans membrane pressure; a higher viscous stress at the membrane surface (given by r/Wv^ where Vt>0 is the velocity gradient at the surface); a higher concentration, and a more suitable type of surfactant. It appears that large drops often result from immediate coalescence, especially when partial wetting of the membrane by the drops occurs, which depends, in turn, on the type and concentration of surfactant. Moreover, inertial forces may cause formation of larger drops at a higher flow rate through the pores, also if no coalescence occurs. Finally, a longer time lag before the drop is dislodged (for example, owing to a low flow velocity of the
EMULSIONS
8.33
continuous phase) causes an Increase In drop size, possibly owing to enhanced coalescence. The time scales for drop formation are relatively long, for the most part between 0.1 and 10 s. The results also depend significantly on the structure of the membrane. A full explanation of the factors affecting the process has not yet been given. Membrane emulsificatlon offers a good method for making homodisperse emulsions (droplet diameter > 1 u.m) on a laboratory scale. Larger scale production may become possible and profitable for small
See e.g., E.S.R. Gopal, In P. Sherman, Ed., Emulsion Science. Academic Press (1968) p. 1. See e.g., P. Walstra, (1983) loc. cit.
8.34
EMULSIONS
disrupt fluid particles in the liquid. This is by far the most applied method for emulsion formation. The apparatus used varies from simple stirrers or static mixers to more sophisticated machines that can produce very intense agitation; see fig. 8.9. In (a) a rotor stator type mixer is depicted that is commonly used for low viscosity liquids (it is often referred to by a trade name, e.g., Ultra Turrax). In (b), a colloid mill is shown, a rotor stator device for highly viscous liquids. A high pressure homogenizer, (c), is used for low viscosity liquids; the liquid is put under high pressure and forced through a very narrow slit, where potential energy is first converted into kinetic energy which is subsequently dissipated into heat. It results in very intense agitation for a very short time, O(0.1 ms). In the rest of sec. 8.2 we will primarily consider emulsion formation during agitation. It may be added that surface forces arising from interfacial tension gradients can, in some cases, contribute to droplet formation; see, e.g., sec. 8.2c, sub (v).
Figure 8.9. Cross section of the active part of some emulsion making machines, (a) Rotor stator type stirrer. (b) Colloid mill, (c) Valve of a high pressure homogenizer; the zig zag line denotes a heavy spring. The slit width in (b) and (c) is greatly exaggerated. The arrows indicate the flow direction. (After P. Walstra, Physical Chemistry of Foods, Marcel Dekker (2002).)
(ii) Phenomena occurring. Droplet formation owing to hydrodynamic forces can occur by various mechanisms11. We will not discuss the phenomena involved, since they are never critical: it is very easy to make a coarse emulsion. However, not all machines can make emulsions: for example, a high pressure homogenizer can only handle a pre made (coarse) emulsion: when a non emulsified mixture of oil and water passes through the valve, the whole slit would be filled either with oil or with water. The critical step, then, is the break-up of drops into smaller ones. This needs drop deformation, which is counteracted by the Laplace pressure,
+
^^(i i) 11
See E.S.R. Gopal, loc. cit., and P. Walstra, (1983) loc. cit.
[8 2 i]
--
EMULSIONS
8.35
where Rj and R2 are the principal radii of curvature of an interface. Deformation will lead to an increase in p L , as is discussed, e.g. in ch. III. 1. The resistance against deformation is therefore greater if the drop is smaller and y is larger. Drop deformation must be followed by its break-up into smaller ones, which leads to an increase in surface area. Hence, surfactant must be transported to the new surface. Moreover, newly formed droplets will frequently encounter each other and then may coalesce; whether this occurs depends, among other factors, on the surfactant load. The processes are illustrated in fig. 8.10. They all have their own time scales, as discussed below.
Figure 8.10. Various processes occurring during emulsion formation. Drops are depicted by thin lines and surfactant by heavy lines and by dots. Schematic, and not to scale. The dichotomy between fully covered and bare surfaces is also an oversimplification. (After P. Walstra and P.E.A. Smulders, (1998) loc. cit.)
(Hi) Regimes. In agitated systems, the forces acting on the drops are of a hydrodynamic nature. It is useful to distinguish between some emulsification regimes, based on two important criteria: Flow type (see sec. 1.6.4), being laminar (L) or turbulent (T), depending on the Reynolds number. We recall that Re = Lvpl r\, where v is the (average) liquid velocity. The characteristic length, L , depends on the geometry: it equals four times the area over the peripheral length of a cross section of the vessel perpendicular to the flow direction. The critical Re for turbulence to occur depends on geometry, but is often roughly 2300. Type of force. Forces can be frictional, also called viscous (V), or inertial (1). Viscous Jorces act tangentlally to the droplet surface and are given by rjVv , where Vv is
8.36
EMULSIONS
Table 8.2. Various regimes for emulsion formation. See text. Regime TV LV Flow type laminar turbulent viscous Forces viscous Reynolds number: for flow around drop
< 2000 <1
External stress acting on a drop
**,
Resulting d sa
2yWeCI
> 2500 <1
TI turbulent inertial > 2500
£ 2/3 d 2/ 3/ ,l/3
r
Characteristic times b)
T
ID
def ~
ads
671 r p l dmcVv n
TIT}'2
d 2/ 3/ ,l/3
15^1/3 a)
RedT is > 1 if d > rft, I yp. b) Applies only if I]D»TJC. Symbols: Re = Reynolds number; We = Weber number; d = droplet diameter; e - power density (energy dissipation rate); y— interfacial tension; r\= viscosity; Vu = velocity gradient; T= characteristic time scale; F= surface excess; m = [surfactant]; (p= volume fraction of disperse phase; We and e are defined below. Subscripts: D = disperse phase; C = continuous phase; def = deformation of a drop; ads = adsorption of surfactant; enc = encounter with another drop; cr = critical value; pi = plateau value.
the local velocity gradient. Inertial forces can be due to velocity fluctuations that lead to fluctuations in pressure (p), according to Bernoulli's law: p + ^pv2 = constant
[8.2.2]
Inertial forces act perpendicularly to the drop surface. Whether viscous or inertial forces are predominant depends on the drop Reynolds number, Re dr , where L = d and v is the velocity of the drop relative to the liquid. The critical number for inertial forces to occur equals about unity. Table 8.2 gives particulars of the relations in the various regimes. Besides the criteria for the regime, the resulting stress, the approximate droplet size obtained, and characteristic times for the various events occurring during emulsifieation, are given. In practice, intermediate situations occur; hence, the relations given are approximate.
EMULSIONS
8.37
Moreover, most of the equations are scaling laws, although the proportionality factor is generally of order unity. Nevertheless, the equations are quite useful, if only because they make clear what variables are essential. The underlying theory will be discussed further below. In all of the theories underlying the relations in table 8.2 it is assumed that the flow is unbounded, meaning in practice that the drops formed are much smaller than the width of the slit in which break-up occurs. However, in very small machines, especially some laboratory homogenizers, this condition is not fulfilled, since d is comparable to the slit width. This gives rise to a fourth regime, 'bounded flow' (LB) the flow is always laminar (and the term agitation may be inappropriate). It will be briefly discussed below.
8.2b
Hydrodynamics
In this section, the interfacial tension is assumed to be constant. This means either that no surfactant is present, which implies that a single drop is considered or that the surfactant is in excess and conditions are such that r a d s « r def ; see table 8.2. In the first two subsections, unbounded flow is assumed. A brief introduction into flow types is in sec. 1.6.4; see especially fig. 6.7. (i) Laminar Jlows1]. This concerns the Regime LV. Viscous forces are responsible for deformation and break-up of drops. Examples are given in fig. 8.11. The external stress, <7ext, exerted by the flow equals rjSlv , where Vi> is the velocity gradient (strain rate) at the drop surface. For simple-shear flow it equals the shear rate: for true elongational flows, the elongation (extension) rate. Here, T] is the viscosity of the surrounding liquid, but its value depends on the flow type. For example, in elongational flow, rjel = Trrj, where the proportionality constant is called the Trouton number. Tr = 2 for uniaxial two dimensional flow (depicted in fig. 8.1 lb) and Tr = 3 for axisymmetric uniaxial flow; for biaxial flows, the values are 4 and 6, respectively. The quantity r\ is the 'common' shear viscosity, and this value is usually determined and inserted in equations. However, the stresses acting on the drops are larger than in the case of simple-shear flow. The numerical values given are only valid for Newtonian liquids; for visco elastic liquids Tr will be (much) larger. Calculation of drop deformation, under the assumption that the drop shape remains very close to equilibrium during the deformation process, has been applied successfully for various flow types. Use is made of a Weber number (by some authors called the capillary number for laminar flows), the external stress divided by the Laplace pressure, which for viscous forces is defined as We = Vvr]cd/2y 11
[8.2.3]
More information is in P. Walstra (1983), and P. Walstra and P.E.A. Smulders (1998): see the General References. A review on the theory of, and experiments on, drop deformation and break-up is given by H.A. Stone, Ann. Rev. Mech. 26 (1994) 65.
8.38
EMULSIONS
For simple-hear flow the deformation of the drop is given by, D = (L-B)/(L+B) , as illustrated in fig. 8.11a. For We«1 , D = We; incidentally, this implies that the elastic shear modulus of a drop equals 2y/d or half the Laplace pressure. It is seen that the drop turns into a prolate ellipsoid, with the long axis at 45° to the flow direction. The liquid inside the drop rotates, in accordance with simple-shear being a rotational flow. The drop is broken up if We exceeds a critical value. The magnitude of WeCT is given in fig. 8.12 (curve marked a = 0 ) as a function of the viscosity ratio, X = rjD 17]c . It is seen that break-up will not occur for X > 4 , irrespective of how large We is. This is because the characteristic deformation time of the drop, given by r
def = % I °exr
^D » %
[8.2A]
is clearly longer than the time during which the stress acts, r act . In the present case, where crext = /fcVu and r act = 1/Vu, it would imply no break-up if X > 1. (This is an oversimplification, because, e.g., at X = 4 , [8.2.4] does not hold precisely.) At high X values, the elongated drop ( D = 5/4y) starts to rotate in the shear field. At low values of X, where r def =1/Vu, the drop is elongated into a slender thread. The smaller is X, the higher is WeCI; i.e., the longer the thread has to become before it breaks. For X = 10"3 , the draw ratio (L/d) at burst is about 12.
(a )
simple shear flow
Figure 8.11. Two types of two dimensional laminar flow, and the effect on deformation and break-up of drops at increasing velocity gradient {Vv) . (Redrawn from P. Walstra, Physical Chemistry of Foods, Marcel Dekker (2002).)
EMULSIONS
3.39
Figure 8.12. Critical Weber number for break-up of drops in various types of flow, as a function of the viscosity ratio, X . Results (from various sources) of single drop experiments in two dimensional, simple-shear flow ( a = 0 ), hyperbolic flow ( a = 1 ), with intermediate types, as well as a theoretical result for axi symmetrical elongational flow (ASE). The hatched area refers to apparent Wecr values obtained in a colloid mill . (Redrawn from Walstra and Smulders, (1998) loc. cit.)
As an example of elongational flow, two dimensional hyperbolic flow (curve marked 1 in fig. 8.12) may serve. It is seen that WeCT is much smaller than for simple-shear flow. Qualitatively, this is so because Tr = 2 , hence the stress is twice as high, and also because the stress increases as the drop is elongated. Moreover, the drop does not rotate, which means that it has enough time to become deformed and be broken up. A curve for axisymmetrical flow, where Tr = 3 , shows that Wecr is still smaller. Flow types which are intermediate between simple-shear and plane hyperbolic flow can also be realized. Some results are in fig. 8.12. The variable is a parameter a in the velocity gradient tensor, which varies between 0 (simple-shear) and 1 (rotational component zero). It is seen that a fairly small elongational component in the flow already lowers Wecr , especially at high X . Results obtained on (large) single drops under ideal conditions {i.e., very near equilibrium) agree well with theoretical predictions. This would mean that the value of d obtained is inversely proportional to y, proportional to VL> and r], and roughly proportional to Tr ; it is, further, a function of X , the function depending on the flow type. However, between the two main drops resulting from break-up, often some much smaller 'satellite' drops are observed (see, e.g., fig. III. 1.17). These mostly do not form under conditions where drops of order 1 fim result, since the satellite drops would then be < 0.1 |im , implying an excessive Laplace pressure. In other cases, the drops 11
H. Armbruster, Ph.D. thesis, University of Karlsruhe (1990).
8.40
EMULSIONS
attain a pointed shape with two tips, from which very small drops then emanate. It has been shown11 that such 'tip streaming' is due to the presence of a trace of surfactant, which is swept towards the tips (cf., fig. 8.3b). Also, these satellite drops will generally not form in practical emulsification. Assuming the theory to hold under practical conditions, the resulting drops are predicted to have a diameter that is 2~ 1/3 times the diameter corresponding to Wecr . This is, however, not observed in practice, for the following reasons. - The deformation is too fast to allow equilibrium shapes to establish. It has been shown in single-drop experiments that the magnitude of dVu/dt can substantially affect the result. This is especially obvious if dVu/dt is so high that We » Wecr can be reached before break-up occurs. Then a strongly elongated drop results, which subsequently breaks into many (up to 100) small ones, owing to Rayleigh instability21. Such 'capillary break-up' appears to occur frequently in practice. The results depend on the magnitude of r]D I T]c, but it generally implies that the value of d obtained is smaller than predicted by fig. 8.12. - Generally, the conditions encountered by a volume element vary among sites in the apparatus and with time; this applies to the velocity gradient and to the time during which a drop is subjected to a given stress. Hence, a size distribution of substantial width results, and it is mostly unclear what kind of average should be taken when comparing theory with results. - The presence of surfactant can have various effects, depending on the type and the concentration of the surfactant see section 8.2b. Some results obtained with a colloid mill are inserted in fig. 8.12. Care was taken to obtain conditions very close to simple-shear at constant shear rate. The agreement with theory is reasonable, although break-up also occurred at X > 4; presumably, there was still a small elongational component in the flow. (ii) Turbulent Jlows3). Turbulent flow is characterized by the presence of eddies (whorls, vortices), which implies that the local flow velocity, u, generally differs from its time average value, (u). The velocity fluctuates in a chaotic way and the average difference between u and (u) equals zero. The root mean square average velocity, u', is, however, finite and is given by,
u^((u-(u» 2 y / 2
11
[8.2.5]
R.A. de Bruijn, Chem. Eng. Set 48 (1993) 277. The theory is due to J.M.H. Janssen, H.E.H. Meijcr, J. Rheol. 37 (1993) 597. The idea has been further worked out and experimentally confirmed by J.A. Wieringa, F. van Dicren, J.J.M. Janssen, and W.G.M. Agterof, Trans. last. Chem. Eng. 74 A (1996) 554. 31 For general aspects, including droplet break-up, see V.G. Levich (1962), and J.T. Davies (1972), General References. 21
EMULSIONS
8.41
The value of u' generally depends on direction, but at very high Re (say above 50 000, and small length scale), it hardly does so. The local turbulence then is isotropic. Kolmogorov has developed theory for this case, which will be used below. It is largely based on dimensional analysis, which implies that the theory gives scaling laws. Fortunately, most of the unknown constants in the equations are of order unity. Turbulent flow shows a spectrum of eddy sizes, C. The largest eddies have the highest u 1 . They transfer their kinetic energy to smaller eddies; these have a smaller value of u ' , but a higher velocity gradient, u ' / t . Small eddies thus have a higher energy density. They are called 'energy bearing eddies', of length scale Ce . These transfer their energy to still smaller ones, in which the kinetic energy is finally dissipated into heat. The smallest eddies are of length Co, also called the Kolmogorov scale. It is given by io=rf/4p-V2E-l/4
[8 _ 2 .6]
where E is the power density, i.e., the amount of mechanical energy dissipated per unit volume and per unit time. Its magnitude ranges from 104 to 10 12 Wm~3 for most emulsifying machines. The variables e and rj are the most important ones that determine whether isotropic turbulence occurs, and that characterize its properties. Local flow velocities depend on the distance scale, x, considered. For a scale comparable to £e we have, u'(x) = £ 1 / 3 x 1 / 3 p - 1 / 3 ,
x = OUe)
[8.2.7]
Since the velocity gradient in an eddy is given by u'{x)/x , it follows that the gradient increases markedly with decreasing size. The characteristic lifetime of an eddy is given by Ttfe)= Ce/ul(le) = l2J3e-1/3pl/3
[8.2.8]
For water, the time scale is between 0.1 and 100 (is in most emulsifying apparatus. The local flow velocity for the smallest eddies is given by u'{x) = el/2x?]-l/2,
x<
[8.2.9]
In these eddies, the velocity gradient is therefore independent of distance. It may be noticed further that [8.2.7] does not contain the viscosity, although this is a factor that determines the values of Re and CQ, and hence the range over which the equation is valid.
The regime TV will be considered first; here, the diameter of the drop to be broken up is much smaller than Ce . Break-up can occur if r]c is rather high (say, 0.1 Pa s)
8.42
EMULSIONS
and L is not very small (to ensure a high JRe value). These conditions will be further discussed below. Break-up of a drop will occur if We > Wecr . Since the flow between two eddies will always have an elongational component, say a > 0.5 , WeCT will be fairly small and not depend greatly on the viscosity ratio A; see fig. 8.1 1. The stress acting on the drop equals r}cVv . Equation [8.2.9] should be used, because the distance over which the gradient must act, d « Ce , and it yields that Vv = (E/TJ)112 . By using in [8.2.3] for We the maximum size of a drop that can remain unbroken the result is d = 2Wecrye-1/2irl/2
= y£-l/2j]-l/2
[8.2.10]
and the size of the drops formed will be given roughly by the same equation. Its validity has not been checked precisely, to the author's knowledge, but approximate agreement has been obtained11.
The regime TI is the common one in the formation of o-w emulsions (and in making aqueous foams by beating), provided that j]c is kept small. Now a drop to be broken up will be of a size comparable to Ce and it will be subject to pressure fluctuations. From [8.2.7] for u', and the Bernoulli equation [8.2.2], it follows that, Ap(x) = p[u'(x)]2 = f 2/3 x 2/3pl/3
[8.2.11]
If Ap is larger than the Laplace pressure of a drop that is near the eddy, it will be broken up. Combining [8.2.1 and 11], and equating x with the diameter of the largest drop that cannot be broken up, we obtain d==£-2/5^/5p-l/5
[8.2.12]
This equation (actually a scaling relation) is often also used for the average droplet size obtained, e.g., d 3 2 . This generalization is allowed as long as the size distribution is of constant shape when, say, the power density is varied. Essentially, the relative standard deviation c 2 should remain constant. This is often the case for high speed rotor stator stirrers and for high pressure homogenizers, but not always for stirred vessels. In this last case, a more elaborate treatment is needed21. The value of c 2 (see [8.1.15]) often is of order unity. It can be reduced by repeating the treatment in the emulsifying machine, since the various volume elements have not all passed through the zones of highest power density in the machine; d 3 2 is also reduced. See also fig. 8.17.
11
Sec R. Shinnar, J. Fluid Mech., 10 (1961) 259, and especially W.J. Tjabbcringa, A. Boon, and A.K. Chesters, Chem. Eng. Set 48 (1993) 285. 21 See A.W. Pacek, C.C. Man, and A.W. Nicnow, Chem. Eng. Sci. 53 (1998) 2005.
EMULSIONS
8.43
Equation [8.2.12] has been very successful in many cases. For a high speed stirrer, the local value of the power density, i.e., near the stirrer tips, is e ~ pco3X2 where a> = revolution rate (s~') and X= stirrer diameter. Inserting this in [8.2.12] gives d
32 = X- 4 / 5 ftT 6 / 5 / cr 3 / 5 y 3 / 5
[8.2.13]
which is known as the Hinze-Clay relation. The lowest attainable value of d32 is about 1 |im . In the valve of a high pressure homogenizer, e ~ p l l o m / 1 , where p h o m is the pressure drop and t the time of passage of the liquid through the valve slit. The value of t is inversely proportional to the liquid velocity in the valve, v , which is, in turn, given by p h o m = \pv2 . Consequently, e = (p h o r a ) 3 / 2 and d32=Pho4V/5
[8.2.14]
Especially the dependence of droplet size on p h o m , is obeyed very well. Very small droplets can be obtained, down to d 32 = 0.1 ^m . A condition for [8.2.12] to hold is that the inertial forces acting on a droplet exceed the viscous forces. Considering that the relevant distance x equals d, [8.2.11] gives for the inertial stress (u)pS7vd , and the viscous stress is given by T]cVv . Hence, the condition becomes {u}dp/r]c =Redr >1
[8.2.15a]
Combination with [8.2.12] yields d>rfclyp
[8.2.15b]
If d = 7]2/yp, [8.2.10] and [8.2.12] give equal results. The viscous and the inertial stresses are equal, and we have a regime intermediate between TV and TI. Another condition for both regimes is that the turbulence is isotropic. However, even for Re as low as 10 000, the relations seem to hold well. Another aspect is that turbulence can be depressed by the presence of long polymer molecules in solution, with contour lengths » C 0. The result is that the smallest eddies are removed from the eddy spectrum. Consequently, the resulting value of d 32 is increased and that of c2 is decreased, at least in regime TI11. Turbulence depression can also be caused by the emulsion droplets present if cp > 0.05. A high value of
11
P. Walstra, Chem. Eng. Set 29 (1974) 882. This aspect is discussed in more detail by P. Walstra, P.E.A. Smulders (1998), General References. 21
8.44
EMULSIONS
into smaller ones. We have seen that, In the regime LV, break-up does not occur for a viscosity ratio X > 4 . For elongational flow, break-up appears possible for any value of X. However, a general condition for break-up is that r def (see table 8.2) is not longer than the time over which the stress acts,
r a c t . In practice, elongational flow is
generally induced by sharp constrictions in the vessel through which the emulsion is forced, and it turns out that r act is often too short for break-up of highly viscous drops to occur. The result is that the drops obtained are (much) larger than would correspond with the value of We c r . In turbulent flows, the value of r T
eddy '
w n e r e
c
is generally given by the lifetime of an eddy,
the value of the eddy size should equal the drop diameter d. In the
regime TV, d « I , and by use of [8.2.9] we obtain r e d d y = ?;J,/2e~1/2 . Taking r from table 8.2, it is found that T d e f /r e d d
= X . In other words, for A > 1, break-up
would not occur. This is an oversimplification, because drops can also be broken up by larger, and hence longer living, eddies. The author is unaware of any systematic theoretical study, but some experimental results are available. In a study with a high speed stirrer in the regime TV, rjo was varied, and for lvalues of 0.05, 1.5 and 31, the resulting d 32 values were about 3, 4, and 12 |jxn , respectively11. Another study concerned a high pressure homogenizer (regime TI, but low Re ). When j]D was varied from 6 to 300 mPa s (at r}c = 1), the resulting d 4 3 value ranged from 0.6 to 2.2 am ; further increase in rj had little effect21. Also in the regime TI, less efficient droplet break-up is to be expected for high X. The r
eddy
eddy =
time
should
e~3/5pl/5y2/5
be
calculated
for
d = £e .
From
[8.2.7]
we
obtain
. Putting this equal to the deformation time as given in table 8.2,
the following equation for the expected value of d at high droplet viscosity results: d = f
^4/£i/V/2
r def = r eddy
[8.2.16]
Various authors have shown the average diameter to increase with 7]D, but the exponent varied, and was generally < 0.75. Also, the exponent of e varied, generally being larger than 0.25. Also in this regime further study would be useful. For X values < 10, [8.2.12] is generally well obeyed, implying that the value of rjD has very little effect on the resulting d value. Methods to obtain small drops of high viscosity include the use of surfactants that give a very small interfacial tension, whereby relatively small d values are obtained at a given external stress (cf. sec. 8.2c), or by making a w-o emulsion (assuming the oil phase to be the most viscous one) of
11
H. Karbstein, Ph.D. thesis, University of Karlsruhe (1994). W.D. Pandolfe, J. Disp. Sci. Technol. 2 (1981) 459. 31 Sec e.g. B.W. Brooks, H.N. Richmond and M. Zerfa, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 6, p. 175, Royal Soc. Chem. (1998). 21
EMULSIONS
8.45
(iu) Bounded flow. Droplet break-up in the regime LB will now be considered. In unbounded flow theories, it is (implicitly) assumed that the velocity gradient is constant over a distance comparable to the droplet diameter. In bounded flow, the distance between a droplet and the vessel wall, or the next droplet, is O(d), which implies that Vu , and thereby the viscous stress, changes significantly over a distance d . Two situations will be discussed briefly. The first one is the flow in the valve of a very small high pressure homogenizer. Here, the valve slit L is very narrow, say < 10 |im . The situation has been analysed by Kiefer11. In the absence of particles, Poiseuille flow develops in the slit, the average Vu being, for example, 10 7 s~ 1 . Since d = O(L), the value of Vu over a drop will vary between 0 and the average Vu. The dynamic thrust of the continuous phase deforms the drop into a kind of hollow cone, which can be broken up into smaller particles. The critical drop size would be given by d5
=TTW-
[8 2 171
- -
Some conditions must be fulfilled for the equation to hold, the most important ones being, Re < 2000 and X > 5 . Further analysis of the flow in a homogenizer valve gives the result that d is approximately proportional to (PhOm'~°'9' a r e l a u o n i s experimentally well obeyed21. Another difference with the results obtained in a large homogenizer may be noticed. In the latter, a drop is broken up in several stages, possibly up to 50 times during one passage through the valve. In a very small valve, there may be only one break-up event per drop per passage. Hence, quite a wide drop size distribution results. To obtain a narrower distribution, with smaller droplets, the treatment should be repeated, say, five times. Another situation arises when the volume fraction of droplets is high. Already at
8.46
EMULSIONS
the continuous phase, was markedly visco elastic. The explanation of these effects is as yet unclear. 8.2c Roles of the surfactant Generally, emulsions cannot be made without surfactant . The effect of the surfactant on the obtained droplet size greatly varies with type and concentration, as illustrated in fig. 8.13. The plateau values of d 32 are roughly proportional to the plateau values of y3/5, as predicted by [8.2.12], but the curves vary greatly at lower surfactant concentrations. For example, the plateau value of d is reached at a lower concentration for amphiphiles than for polymers, despite the latter generally being more surface active; see fig. 8.2. Surfactants play various roles. They facilitate droplet break-up by lowering y, and thereby the Laplace pressure. They can affect the mode of, and increase the, resistance to deformation, especially via the surface dilational modulus, Kg. In some cases, they can facilitate droplet break-up by means of surface forces, a phenomenon designated, 'interfacial instability'. Most importantly, they counteract the (re)coalescence of just formed droplets during emulsifying. All of these phenomena depend on the local, and often transient, composition of the droplet surface layers.
Figure 8.13. Specific droplet surface area Av and average droplet size d 32 , as a function of total surfactant concentration, c , obtained under roughly the same conditions (o/w emulsions, cpai 0.2, regime TI, comparable power density) for various surfactants; PVA = polylvinyl alcohol). For the soy protein a plateau value is also reached, at ca;20kgm~ 3 . Approximate plateau values of yow of 3, 10 and 20 mNm~ 3 for the non-ionic, casein, and PVA, respectively. Assembled from various sources.
(i) Surfactant transport. This concerns primarily transport of surfactant from solution (generally the continuous phase) to the drops' surface. Transport is generally due to convection rather than to diffusion. Table 8.2 gives expressions for the An example of an exception is making a dilute emulsion of pure oil in pure water by ultrasonication; the droplets then acquire a substantial negative charge.
EMULSIONS
8.47
characteristic times needed, r a d s , derived from the flow equations . It is interesting to compare r a d s with the characteristic deformation time, r def , since the ratio of the two gives an indication of the value of F, and hence of y, during break-up of the drop. Formulae for r def are also given in table 8.2, derived from 18.2.4). This equation is based, however, on the assumption that 7/D » JJC , which is often not obeyed in the regimes LV and TV. This is because the deformation time cannot be shorter than the local value of 1 / Vu A few examples will be given for emulsions being formed of oil (77 = 60 mPa s) and water (1 mPa s); resulting d = 1 um ; plateau value of F= 1 mgm~ 2 ; mc = 2 k g m " 3 ; and at low values of
w/o
w/o
o/w
Regime
LV
TV
TI
rads= rdef =
100 10
100 10
1 10
us us
It is seen that for viscous forces r a d s > r d e f , whereas it is the other way round for inertial forces, at least in the present cases. The relations are approximate, primarily because the theory is approximate. Second, it may well be that transport of surfactant through a very thin layer close to the droplet is by diffusion, further increasing the adsorption time. The author is unaware of a definitive theory which takes this phenomenon into account. Probably, any retarding effect will be minor in the regime TI. Anyway, adsorption tends to be a fast process. Most polymeric surfactants will need a time far longer than r a d s or r d to attain an equilibrium conformation in the interface. During emulsification, the interfacial tension will temporarily vary from place to place at an interface. Such y- gradients can form because of local variation in adsorption of surfactant, or because a viscous stress acts at the interface. The latter occurs frequently during emulsification. If no external force is acting (any more), the distribution of surfactant at a drop surface will even out. This can occur by three mechanisms. The first is adsorption of surfactant from the bulk: characteristic times are given above, and these times depend strongly on the surfactant concentration (table 8.2). The second one is lateral diffusion in the interface, and the characteristic time for such a process over a distance z is given by z 2 / 4D° , where the value of the surface diffusion coefficient, Da can be approximated by the bulk diffusion coefficient in the most viscous phase. For a small molecule surfactant, it would be of order 5 x 10~ 12 m 2 s" 1 . Assuming that d = 1 um , and that z = d / 2 , we obtain a characteristic diffusion time of somewhat over 10 ms; for a polymeric surfactant, the time will be substantially longer. The third mechanism is motion of the interface owing to a
11 See, e.g., Lcvich (1962), General References, who gives full derivations, or P. Walstra, P.E.A. Smulders (1998), General References.
8.48
EMULSIONS
/-gradient, which can be described as a longitudinal wave11. (This is generally called, 'spreading' of surfactant over the interface, but actually the surface entrains the surfactant.) The characteristic time to even out a gradient over a distance z is given by Tlz) = r
z^3(T1pf/3(Ay)-2/3 2
2i1/2
[8.2.18]
77P = [(r? w p w ) +{%PO) J Assuming that Ay = 1 mN m" 1 , which is a value that will often be exceeded, and z = 0.5 um, a r(z) value of about 0.1 ms results. It may be concluded that for small droplets, evening out of y- gradients is a fast process and that either bulk diffusion or 'spreading' of surfactant over the interface will be the dominant mechanism. (it) Transient interfacial tension (y^). An important variable is the value of y during drop deformation and break-up, since that determines the final drop size. Characteristic times for transport towards the drop have been given, but even if the equations hold, the time needed to reach an equilibrium F value would roughly equal about 10 x Tads (for amphiphiles). Moreover, the concentration of surfactant in the continuous phase, c c , decreases considerably during the process, the decrease in d causes an increase in A v , the more so for a higher
J. Lucassen, Trans. Faraday Soc. 64 (1968) 2221.
EMULSIONS
8.49
surface excess (in mass units) is needed to obtain a substantial reduction of y (see fig. 8.2b). Since most polymers also give higher equilibrium values of y than do most amphiphiles, the droplets obtained will be substantially larger at the same power density. Consequently, a higher mass concentration is needed to obtain a plateau value of d 32 , as illustrated in fig. 8.13. However, this is probably not the only factor determining the difference in effectivity of both types of surfactant; see subs. (vi). (iii) The p.i.t. method. The idea of the phase-inversion temperature and of the p.i.t. method for making emulsions stems from Shinoda and co-workers11. It has been mentioned briefly in sec. 8. lb. Consider a system of oil, water (in roughly equal proportions), and 2-7 % of a surfactant. The latter must be a non-ionic, whose hydrophilic part consists of one or more poly(oxyethylene) chains, e.g., polyoxyethylene nonylphenyl ether. An idealized phase diagram, with temperature as the variable, is given in fig. 8.14. At low temperature, there is an oil phase and an aqueous phase which contains most of the surfactant. The surfactant is, for the most part, present in micelles, which are swollen and hence contain oil. Upon increasing the temperature, the micelles become elongated and contain more oil (assuming equilibrium to be reached); moreover, more
Figure 8.14. Idealized phase diagram of an oil water (non-ionic) surfactant system that shows a phase-inversion temperature. For explanation, see text. (Figure redrawn from Shinoda and Kunieda (1983).)
11 See e.g. K. Shinoda, J. Colloid Interface Set 24 (1967) 4, and K. Shinoda, and H. Sato, J. Colloid Interface Set 30 (1969) 258. A review on the p.i.t. and other surfactant properties is by K. Shinoda, H. Kunieda, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1, chapter 5, p. 337, Marcel Dekker (1983). A thermodynamic interpretation of the p.i.t. phenomenon is given by E. Ruckenstein, Langmuir 4 (1988) 1318.
8.50
EMULSIONS
surfactant will go to the oil phase. At still higher temperature an apparently homogeneous system (designated D) is formed. The interfacial tension between oil and water becomes very small, and the surfactant still tends to associate. However, the natural curvature of a monolayer is close to zero (it is concave towards the oil side at low temperature). The distribution quotient of surfactant between oil and water is of the order of unity. This is the region near the phase-inversion temperature, and it has been shown that here the system is a bicontinuous microemulsion". Above the p.i.t. region, the natural monolayer curvature is inverted, and most of the surfactant is in the oil, in the form of swollen inverted micelles. The shape of the phase diagram naturally depends on the oil to water ratio. The phase diagram, including the value of the p.i.t., also depends on the type of oil, the solvent quality of the water phase for the surfactant (e.g., salt content), and especially on the concentration and type of surfactant (e.g., length and number of POE chains, and length and unsaturation of the aliphatic chain). When the system is agitated in the p.i.t. region, an emulsion forms, of the w/o type if T > p.i.t., and o/w at T < p.i.t. Shaking by hand suffices to produce droplets down to d = 1 |im . Close to the p.i.t., the value of yow is very small, of the order of 10|iPas. According to the theories and experiments discussed in sec. 8.2b, an external stress of order 10 Pa would still be needed to obtain such small droplets, but such a high value is not nearly reached when shaking by hand. Presumably, droplet formation is substantially facilitated by interfacial instability: see subsection (v) below. In practice, o/w emulsions are often made by the p.i.t. method, because it saves energy and does not need expensive equipment. On the other hand, the high concentration of amphiphiles present is not always acceptable and may be costly. The system is brought to a temperature slightly below the p.i.t., and is agitated with sufficient intensity. Very small droplets can thus be obtained, even if rjD is high. However, they are very unstable towards coalescence, because of the very low interfacial tension (see section 8.3e). To avoid coalescence, the emulsion has to be cooled immediately to, say, 20 K below the p.i.t., e.g., by adding cold water if dilution is acceptable. Naturally, the emulsion should be kept at low temperature. If it is brought close to the p.i.t., rapid coalescence will occur; and above the p.i.t., true phase-inversion takes place21. Water in oil emulsions can also be made by a p.i.t. method, if the p.i.t. is low enough to make the process feasible31.
11
L. Taisnc, B. Cabane, Langmuir 14 (1998) 4744. Phase-inversion is extensively discussed by B.W. Brooks, H.N. Richmond, and M. Zerfa, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 6, p. 175, Royal Soc. Chem. (1998). 31 H. Sato, K. Shinoda, J. Colloid Interface Sci. 32 (1970) 647. 21
EMULSIONS
8.51
(ivj Effect on drop deformation. The main effect of the surfactant on droplet deformation and break-up, viz., the lowering of y, has been discussed above. There are some other effects related to the y gradients (see section 8.1c) that are formed by flow of the continuous phase along the drop. In principle, the presence of such a gradient can stop internal circulation in the drop, which would lower the energy needed for break-up. Moreover, the variation of y over the surface of a drop affects the deformation of the drop. In a largely theoretical study in the regime LV, the overall effect on break-up appeared to be small to moderate11. However, when tip streaming occurs (see section 8.2b sub (1)) in the regime LV, and the stresses are such that relatively large drops will result, the formation of small satellite droplets appears to be possible, since bimodal size distributions were obtained experimentally. However, in practice it took a very long emulsification time to arrive at a substantial volume fraction of small droplets21. Another effect is that the Gibbs energy needed for deformation of a drop equals the increase in Ay; generally, both A and y increase during deformation. Since d{Ay) = ydA + Ady, and the surface dilational modulus Kg = dy/dlnA , the following relation results d(Ay) = (y + Kg)dA
[8.2.19]
It would thus be as if y is enlarged by the value that Kg attains under the prevailing conditions. The situation has been studied in the regime LV on single drops, and
Figure 8.15. Effect of surfactant concentration, c c , on break-up of a single drop in simple-shear flow. The value x is the critical droplet size for break-up relative to that in the absence of surfactant, as observed and as calculated for the equilibrium value of y. The quantity y is the critical Weber number obtained over that predicted for y eq . (After results by Janssen et al. (1994) loc. clt.)
11
H.A. Stone, L.G. Leal, J. Fluid Mech. 64 (1990) 161, and W.J. Millikcn, L.G. Leal, J. Colloid Interface Sci. 166 (1994) 275. 21 F. Groeneweg, F. van Dieren, and W.G.M. Agterof, Colt. Surf. A91 (1994) 207.
8.52
EMULSIONS
under conditions where yow was close to its equilibrium value at the prevailing surfactant concentration11. The main result is given in fig. 8.15. It can be interpreted as if the interfacial tension during break-up equalled y+bK^, the value of the factor b being 0.23 in the case studied; it is uncertain why b < 1 (because flow in the drop is inhibited?). It is seen that the effect can nevertheless be substantial. To what extent this is also true in other regimes, or for polymeric, surfactants during practical emulsification is still unclear. (v) Interfacial Instability2*. This phenomenon is also called 'interfacial turbulence' and incorrectly 'spontaneous emulsification'. Consider two liquid phases (1 and 2) that are separated by an interface having a very small interfacial tension. Surfactant will diffuse from phase 1 to phase 2 if its chemical potential is higher in 1 than in 2. This means, in practice, that a surfactant is chosen that is soluble in both phases, but (far) more so in phase 2. It is dissolved in phase 1, and then the phases are brought into contact. Now an unstable situation arises. A small deformation of the interface will be enlarged if the bulge is toward phase 2. Because of depletion of surfactant in the bulge, y gradients are formed, and these cause Marangoni flow, as illustrated in fig. 8.16. Consequently, fairly long fingers of phase 1 can protrude into phase 2, and are then subject to Rayleigh instability, breaking into small drops. The extent to which interfacial instability occurs depends on some factors; a significant effect appears to be more likely if ^j > rj2- Moreover, the electrostatic charge of the interface, and the thickness of the electric double layer have a substantial effect . It is presumably a combination of gentle agitation (implying fairly long deformation times) and the generation of interfacial turbulence that can result in the formation of quite small drops. Prerequisites concerning the surfactant are (a), that it causes a very small yow (< 0.1 mPas), implying a suitable (amphiphilic) surfactant at high concentration; (b), that it is well soluble in the disperse phase, as such, or in micelles; and, (c), that it is nevertheless more soluble in the continuous phase. Especially for the
Figure 8.16. Illustration of an interfacial instability. Phase 1 originally contains the surfactant which is, however, more soluble in phase 2.
11
J.J.M. Janssen, A. Boon, and W.G.M. Agterof, Coll. Surf. A91 (1994) 141, and A.7.Ch.E.J. 40 (1994) 1929. 21 This phenomenon has been studied by C.V. Sternling, L.E. Scriven, A.I.Ch.E.J. 5 (1959) 514, and by J.H. Gouda, P. Joos, Chem. Eng. Set. 46 (1964) 521. Theory has been developed and checked by A. Sanfeld, M. Lin, A. Bois, I. Panaiotov, and J.F. Baret, Adv. Colloid Interface Set 20 (1984) 101.
EMULSIONS
8.53
making of emulsions with a very viscous disperse phase, e.g., of a bituminous oil in water, application of these principles may be useful. As mentioned, interfacial instability will also be of importance when applying the p.i.t. method of emulsification. (ui) Prevention of re-coalescence. If emulsification occurs at a given intensity, i.e., at a given value of e, and if the intensity is then changed to a lower value while emulsification proceeds, the average drop size increases in time, until a plateau value is reached. This has been observed by several workers, at least in the regimes TV and TI. It must be caused by coalescence of newly formed drops, and the phenomenon is generally called re-coalescence. For example, in systematic studies using a small high pressure homogenizer , the coalescence rate was higher for a higher value of
Figure 8.17. Average droplet size, d, and size distribution width, c, relative to their plateau values, as a function of agitation time or number of passes. In (a), the surfactant was a commercial amphiphile; in (b), a protein mixture. (Adapted from Walstra, (1983), loc. cit.) 11 S. Mohan, G. Narsimham, J. Colloid Interface Sci. 192 (1997) 1; and G. Narsimham, P. Goel, J. Colloid Interface Sci. 238 (2001) 420. 21 See P. Walstra (1983), General References.
8.54
EMULSIONS
Other methods for estimating the rate of re-coalescence depend on establishing the extent of mixing of oil from different droplets during emulsification, for example when droplets having two values of the refractive index, n , are present. To carry this out, two emulsions are made under identical conditions, but the (oil) drops differ in n value (rtj and n 2 ). The emulsions are mixed and the mixture is emulsified again at the same intensity, during various times or with an increasing number of passes. The emulsions obtained are diluted with an (aqueous) solution of high refractive index so that the continuous phase attains a value of n = (rij + n 2 ). The turbidity of the diluted mixtures is determined. If no mixing of oil, i.e., no re-coalescence, has occurred, the turbidity will equal that of the (dilute) original mixture; if re-coalescence would be complete, the turbidity will be virtually zero owing to refractive index matching. This allows estimation of the extent of re-coalescence1'. An alternative is to make two disperse phases, one of which contains a fluorescent probe plus an excimer, at such concentrations that dilution with pure oil virtually eliminates the fluorescent signal2'. The advantage of these methods is that the measurement of re-coalescence occurs under nearly the same conditions as the emulsification. Moreover, comparison of the variation in the average drop size obtained by, for example, varying surfactant type or concentration can be correlated with variation in the re-coalescence rate. Roughly the same trends are observed as with the other method; examples will be given below. Coalescence can occur if two droplets collide (a phenomenon that will occur frequently during emulsification): if and when the film between them becomes very thin, say a few nm, it will generally rupture, inducing coalescence. It is often assumed that prevention of close approach, and hence of re-coalescence, is due to colloidal forces, i.e., electrostatic or steric repulsion, caused by the presence of surfactant at the droplet surface. This can be questioned. The stress which presses droplets together during emulsification can be quite large, up to the value needed to result in droplets of that size (say, half the Laplace pressure of the drops). Considering a drop of d = 1 urn and with y = 0.01Nm~ 1 , the maximum stress will be 20 kPa. From DLVO theory, it can be calculated that the disjoining pressure between the droplets will, under most conditions, be < 200 Pa, even for drops with a plateau value of F. For steric repulsion, the values may be higher, but generally not much. Hence, the repulsive forces are generally far too weak to prevent re-coalescence. This is borne out by the following experimental result3'. An o/w emulsion was made with SDS under such conditions that insufficient SDS was present to give full monolayers ( F~ 0.06 -Tpl). Nevertheless, coalescence was not observed within a day. Adding 10 mmolar NaCl, which greatly reduces electrostatic repulsion, led to visible coalescence. When the same emulsion was made in the presence of 10 mmolar salt, it
11
L. Taisne, P. Walstra, and B. Cabane, J. Colloid Interface Sci. 184 (1996) 378. L. Lobo, in Proc. 2nd World Congress on Emulsion, Vol. 4 (1977) 75. 31 L. Taisne et al., (1996) loc. cit. 21
EMULSIONS
8.55
nevertheless gave exactly the same value of d32 as without the salt, i.e., 0.24 |im (to be sure, this emulsion also showed coalescence upon standing). To find out what mechanisms are responsible for diminishing re-coalescence, kinetic aspects should be taken into account: what happens with the droplets as a function of their surface excess during agitation? A full treatment is elaborate'' and we will mention only some trends, since accurate prediction is not possible, anyway, because of the complex and constantly changing situation during emulsification21. It is assumed that the surfactant is soluble in the continuous phase only. An important question concerns the transient value of F during a collision of drops. This will depend on the ratio of the characteristic adsorption time divided by the average time between encounters ('collisions') of a drop with another drop, renc . From hydrodynamic theory it follows that in all regimes in unbounded flow, the (average) ratio is given by,
WTenc=50V/ccd
I8-2-20!
Assuming / " j / c c = l | i m , ^ = 0.03, and d = 1 um , the ratio is 1.5. In most cases, e.g., with larger (p or smaller c c , the result will be that r ads is longer, or even much longer, than r enc . It then means that F is often substantially smaller than Fpl during a droplet encounter. On the other hand, when a drop starts to deform, it may already have acquired a substantial F value, and if the total surfactant concentration is high enough, many drops may even have F ~ F ^. As soon as a drop is deformed or even broken up, however, the value of F can become substantially decreased; nevertheless, adsorption does not start with a bare interface. Considering further that considerable statistical variation occurs, particularly in the value of r enc , there will be frequent occasions where F is low during an encounter. Owing to the large stress acting on them, the drops will often be deformed upon approach, forming a flat film that can subsequently drain. This is illustrated in fig. 8.18. The rate of drainage is an important variable. We will first consider a situation where the droplet surface is rigid in the lateral direction owing to the y gradient formed. The Reynolds equation can be applied, and for the present situation the linear drainage rate is given by _ dh _ 8ny2h3
(_ 8yh3 "j
2
dt ~3;7 c a F [~ 3;7ca3J
The following discussion is for the most part based on A.K. Chesters, Trans. Inst. Chem. Eng. E69.B (1991) 259; and on I.B. Ivanov, P.A. Kraichevsky, Coll. Surf. A128 (1997) 155. Also see Chesters (1991) and Ivanov and Kraichevsky (1997) in the General References. 21 Discussed by Walstra and Smulders (1998), General References.
8.56
EMULSIONS
Figure 8.18. Subsequent stages in the formation and drainage of a film between two drops that approach each other owing to a high stress. In frame (1), the motion of drops and continuous phase are indicated; in (2) is shown the distribution of surfactant molecules (indicated by dashes); and (3) defines the geometry of the system. Highly schematic and not to scale.
It is seen that a larger force (F) gives slower drainage, and hence a longer, not a shorter, drainage time. This is due to the higher stress causing a flat film to form of larger radius, R, and hence a greater resistance to flow. It is indeed frequently observed, though not always, that more intense agitation, and hence a higher local stress, goes along with less coalescence. The quantity between parentheses in [8.2.21] is valid for F = nya , which derives from the assumption that the stress acting on the drops equals half their Laplace pressure. Integration yields for the time needed to drain the film to a thickness h, the relation
'*-lfef
l8 2 221
-'
Again, let d = 1 ^m , ?/c = 1 mPas, and y = 10 mN m" 1 , the result is for drainage to 5 nm, r dr = 90 us ; for d = 3 }im , the value will be 2.5 ms. These times appear quite short, but the time during which a high stress will act on a droplet pair is even shorter. For the regime TI, [8.2.8] yields, assuming that (e = lOd , a lifetime of 1.5 jus for d = 1, a value of 8 us results for d = 3 /an. For the regime TV, roughly the same values are obtained. This means that for the present case, i.e., a rigid drop surface, the time during which the drops are strongly pressed together is generally too small to obtain an h value small enough for film rupture to occur. In other words, the drops move away from each other before they can coalesce, at least in turbulent flow. (This is probably also true in elongational flow, but not always in simple-shear flow.) For non-rigid surfaces the situation is different. The extreme is a surface that is fully mobile in the lateral direction. The drainage is then given by
EMULSIONS
8.57
h =h o exp(-t/t c h ) t
ch==7?cd/5' 3
regime TV U2
rch = pd (e/ric)
/10 y
[8.2.23]
regime TI
where hQ is the original distance between the drop surfaces, and tch is the characteristic drainage time. Assuming h/h Q =O.O5 (drainage from h = 100to 5 nm), d = 1 nm , 7=30 mN m"1 (since a fully mobile surface implies F close to zero) and 77 c =lmPas, we obtain for r dr in the regimes TV and TI, 0.1 and 0.05 (is , respectively; for d = 3 urn the results are about 0.3 |j.s in both regimes. This implies that the drainage times can be substantially shorter than the duration of the local stress, and hence that coalescence would occur readily, which does indeed happen in the absence of surfactant. In nearly all practical situations, the surface will be 'partially mobile', which implies that it does move under the prevalent tangential stress, but more slowly than in the absence of surfactant. The important variable is the Marangoni number; for a rigid surface Ma > 1 is assumed, for a fully mobile surface Ma ~ 0 . From [8.1.4] Ma can be calculated. In the present case, the stress acting in the film is about y/a , which leads to Ma ~ K^/y. Equation [8.1.5] gives an expression for K% and, in the present case, the value of the parameter t, will generally be negligible, owing to the very short time scale. This implies that the surfactant forms virtually a Langmuir monolayer. The result is dr
drc ydlnT
=_djnjl
dlnT
Calculation of the drainage times as a function of Ma is complicated, and there is some disagreement among authors. However, the drainage time is probably not much shorter than for a rigid surface (see [8.2.22]) if Madr > 0.5. Equations of state, as given, for example in fig. 8.2.b, allow calculation. A value of Madr > 0.5 will be reached for SDS at F~ 0.4 , and for (5-casein at about 1.1 mg m"2 . The latter value is probably too low: at the very short timescales involved during emulsification, the n values at a given low F must be even smaller than the equilibrium values given in the figure for this flexible protein. As mentioned earlier, a considerable difference is observed between the d 32 values obtained with amphiphiles and with polymers at the same intensity of agitation, and the same mass concentration of surfactant; the difference is especially large at a relatively low surfactant concentration. The above considerations point to the need for a sufficiently high Ma value during drainage of a film between droplets to prevent their re-coalescence. For polymers this is more difficult to achieve, they require a much higher surface concentration (in units of mass per unit of surface area) to attain a substantial Ma value. The latter will be due primarily to the large molar mass of polymers.
8.58
EMULSIONS
The importance of molar mass is further illustrated in the following case study. Two surfactants were used. The one was p-casein, a flexible protein of 209 amino acid residues, with part of the peptide chain being highly charged, and another, longer part being less charged and having a high concentration of hydrophobic side groups. The other was a fragment of the same (3-casein, obtained by selective hydrolysis, consisting of residues 29-106, and hence a much shorter chain; nevertheless, it has a similar amphiphilic character. Both of these surfactants were used to make emulsions, with the results shown in fig. 8.19. It is seen that the peptide gave smaller d 32 values at the same surfactant concentration. The rate of re-coalescence was estimated by the refractive index matching method, from the relative decrease in turbidity during homogenization (because a very small homogenizer was used, a great number of homogenization steps is needed to obtain a steady state). Figure 8.19b shows that the rate of re-coalescence is well correlated with the difference in d 32 . Figure 8.19c shows
Figure 8.19. Emulsions made with fi- casein (BC) and an amphiphilic peptide (AP); see text, (a) The resulting values of d^ as a function of total surfactant concentration c. (b) Relative turbidity as a function of the number of passes through the homogenizer valve, (c) Values of (A32 as a function of storage time t of emulsions made at c = 2 mg/ml. Emulsions of
11
After P.E.A. Smulders, P.W.J.R. Caessens, and P. Walstra, in E. Dickinson, J.M. Rodriguez Patino, Eds., Food Emulsions and Foams, Royal Soc. Chem. (1999) p. 61.
EMULSIONS
8.59
that the difference is inversely correlated with a difference in coalescence stability of the finished emulsions. It is also unlikely that the peptide gave a lower value of yow than the protein. Further studies on re-coalescence1' have shown that it occurred less if the emulsion was re-homogenized at a lower value of e. For SDS as the surfactant, the re-coalescence rate was very small if the surfactant was in excess. Good correlations were obtained between re-coalescence rate and the droplet size obtained when varying the surfactant concentration, using various peptides and proteins, and upon varying the pH. However, when comparing the proteins P-lactoglobulin and (3-casein, the former gave droplets that were smaller by some 25%, whereas the re-coalescence rates appeared to be the same. In summary, much of the differences given between various surfactants and surfactant concentrations can be ascribed to differences in re-coalescence, and the mechanisms involved are understood, at least in a qualitative sense. The issue is somewhat confused by the other effects a surfactant can have. To arrive at hard conclusions, more systematic studies would be needed, taking into account the considerations given above. (vii) Bancroft's rule. The reasoning in the above subsection is based on the assumption that there is one surfactant that is soluble in the continuous phase only . If the surfactant is in the drops, and is not transferred to the continuous phase, ygradients hardly form, since surfactant can readily diffuse to the thinnest spot in the film between approaching droplets. It has been shown2 that drainage will be practically along a mobile surface; in other words, Madr will be close to zero in nearly all cases. This provides directly an explanation for Bancroft's rule, when making an emulsion of two immiscible liquids and a surfactant, the liquid in which the surfactant is better soluble will become the continuous phase, since droplets containing surfactant will recoalesce much faster than those without dissolved surfactant. It should be added, however, that the reasoning given in subsection (vi) does not apply to the situation described in subsection (iii), where the p.i.t. method is described. Here, the time scales involved are much longer, the forces involved are very much smaller, and the interfacial stability described in subsection (v) plays an important role. A different treatment is needed, also, to explain Bancroft's rule31. 8.2d Formation of surface layers When making an emulsion with an amphiphilic surfactant (mixture), the surface 11 See L. Taisne, P. Walstra and B. Cabane, J. Colloid Interface Set 184 (1996) 378, for emulsification with SDS, and P.E.A. Smulders, Ph.D. Thesis, Wageningen University, (2000), for proteins and peptides. See I.B. Ivanov and P.A. Kralchevsky, mentioned in the General References. 31 See e.g. A. Kabalnov, H. Wennerstrom, Langmuir 12 (1996) 276.
8.60
EMULSIONS
excess and the composition of the adsorption layer will be governed by the Gibbs equation [8.1.2], or equations elaborated for more complicated systems. In principle, F can be calculated from the values of cp, the total surfactant concentration c , and d 32 or Ay . This is generally not so for polymeric surfactants. The main point is that equilibrium is generally not reached. Figure 8.20a shows an adsorption isotherm obtained for p1- casein (a flexible, almost random coil protein in solution) on a macroscopic surface, allowing sufficient time to obtain equilibrium. It also gives an apparent adsorption isotherm, calculated from F and c c values determined after emulsification, for a range of initial surfactant concentrations. The differences are striking. The jump in the 'macroscopic' curve around c c = 1 0 2 is unexplained (the increase in F is real, but it is not known precisely what the shape of the curve is). It appears that this increase in F does not correspond to a significant decrease in y • Neither is it known whether the / value on the emulsion droplets with F ~ 4 mg m~2 is lower than that observed in the macroscopic experiment (n ~ 20 mN m ] ) . If an emulsion has been made under conditions such that a relatively small F value results, e.g. ,2 mg m~2, addition of p- casein to the emulsion causes an Increase in F. The, 'emulsion' curve also depends on other factors, such as the value of cp during emulsification. If the data are plotted as in fig. 8.20b, the same curve is observed for different cp values. This means that the magnitude of F depends not only on c c , but also on the amount of continuous phase per unit droplet surface area. Figure 8.20b also shows that different proteins give different relations. By and large, for proteins that do not differ greatly in other properties, there is a correlation
Figure 8.20. Values of the surface excess, F, at the triglyceride oil water interface for some proteins, (a) (3- casein; F is plotted against protein concentration in the sub natant (left-hand curve) and in the continuous phase of an emulsion (right-hand curve); see text, (b) F plotted against total surfactant concentration over the surface area produced for various proteins, viz., (3- casein (BC), ovalbumin (OV), p- lactoglobulin (BL) and lysozyme (LY). The broken line indicates what the result would be if all of the protein were adsorbed . 11
For the most part, after P.E.A. Smulders, Ph.D. Thesis, Wageningen University (2000).
EMULSIONS
8.61
between molar mass and the plateau surface excess. The increase in r at high c/ Av values appears to be due to aggregation of the protein in the interface. (This cannot explain the differences in fig. 8.20a: (3-casein does not show aggregation.) 8.2e Effect of volume fraction Figure 8.21a shows d 32 as a function of homogenization pressure for various values of tp. The initial concentration of surfactant in the liquid making the continuous phase was constant. The regime was TI. For small (p values, [8.2.14] was obeyed, the slope of logd32 versus logp h being exactly 0.6. The higher is
Figure 8.21. Effects of droplet volume fraction, q>, on the resulting droplet size in emulsification. (a) Triglyceride oil in water emulsions made in a high pressure homogenizer at various pressures p h . The surfactant is milk protein. (After results by P. Walstra and G. Hof, unpublished.) (b) Toluene in water emulsions made by stirring in a small vessel at various values of (p. The dmax^ is the largest drop size observed in the emulsion by microscopy. The surfactant was SDS, present at high concentration. (Redrawn from results by S. Kumar, R. Kumar, and K.S. Gandhi, Chem. Eng. Set 46 (1991) 2483.)
8.62
EMULSIONS
d. Particles depress turbulence, the more so at a higher concentration (see section 8.2b sub (ii)); in particular, the smallest eddies will be removed from the spectrum. In the regime TI this will lead to an increase in drop size. e. The effective viscosity of the continuous phase, rjc, will increase. To be sure, its microscopic viscosity will remain the same, but the drops will sense a higher viscous stress. For the most part, the drops will have a tangentially rigid surface during agitation. This means that the velocity gradient between two drops will be enhanced at the same agitation intensity, as if T]c is increased. Moreover, the flow will have a stronger elongational component, which also means a higher viscous stress. If the drops are broken up by viscous forces, this may cause a greater reduction in average drop size. The results of fig. 8.21a can be explained by factors (a), (b), and possibly (d). The regime was probably TI in nearly all cases. The results of fig. 8.21b are more complex. Here, the power density was quite low; at low values of
We will largely follow an earlier review (P. Walstra, 1993) and chapters 1, 6, 7 and 9 in B.P. Binks (1998); sec General references.
EMULSIONS
8.63
8.3a Overview Figure 8.22 depicts the most important physical instabilities that can occur in emulsions. Ostwald ripening. Material in small drops can diffuse through the continuous phase toward larger drops. This leads to an increase in the average droplet size. The driving force is the difference in chemical potential of the material, caused by the difference in curvature, as expressed in the Kelvin equation. A prerequisite is a finite solubility of the disperse in the continuous phase. Aggregation. When droplets stay together for an appreciable time after having encountered each other as a result, for example, of Brownian motion, they are said to be aggregated. The aggregates can grow to considerable size, and possibly form a space filling network. The driving force is often van der Waals attraction between drops, or depletion interaction caused by polymer molecules or surfactant micelles present in the continuous phase; moreover, some kinds of bridging of drops can cause aggregation. Aggregation can be counteracted by colloidal repulsion. Hence, deaggregation can often be achieved by changing the disperse phase composition. Sedimentation of drops is caused by gravitational forces acting on the drops; it results in a decreased potential energy of the system. Sedimentation can be upwards, i.e., creaming, or downwards, i.e., settling. The final result is separation of the system into a cream layer (or sediment) of high
Figure 8.22. Illustration of changes in dispersity of an emulsion. Broken arrows denote the reverse change. Ostwald ripening and coalescence (actually, the second arrow in the lowest row) are irreversible.
8.64
EMULSIONS
The instabilities mentioned should be clearly distinguished, since they are governed by different variables and have different consequences. It is not always easy to distinguish them experimentally, especially in an early stage. Moreover, the processes can affect each other. All changes that cause an increase in particle size will enhance the sedimentation rate. This holds especially for aggregation, since its rate is, in turn, enhanced by sedimentation. Coalescence can only occur if two drops are quite close to each other. Consequently, it is enhanced in aggregates and in a cream layer or sediment. The decrease in Gibbs energy involved in the changes mentioned tends to be small. Consider an o/w emulsion of
EMULSIONS
8.65
crystalline phases with water, may, if present in a sufficiently high concentration, form such phases around oil droplets, and thereby enhance the stability against coalescence11. Small solid particles present in the continuous phase that adsorb onto the drops, can counteract coalescence. This 'Pickering stabilization' will be discussed in sec. 8.4. Particles present in the drops, especially triacylglycerol crystals, can in principle cause gross instability owing to 'partial coalescence'; this is discussed briefly in sec. 8.3c. 8.3b Ostwald ripening2) The solubility of the material in a small particle is generally increased over the bulk solubility c sat x . The increase in solubility is given by the so called Kelvin equation (although the form in which it is used is, in fact, due to Ostwald); see sees. 1.2.23c and IV.2.2e. For a spherical particle of radius a the equation reads; c
sat(a) = c sat~ e x P( x l / a '
x' = 2yM D /p D KT
[8.3.1]
where MD and p D are the molar mass and the density of the material in the disperse phase, respectively. The parameter x' is a characteristic scale, in units of length; the larger it is the greater the value of c s a t( a )/ c s a t o o • Consider a dodecane droplet in water. The interfacial tension is about 0.05 N m" 1 , MD = 0.17 kg mol" 1 , and pD = 745 kg m 3 ; at room temperature this results in x' = 10 nm . For a droplet of a = 0.5 (xm this results in a solubility ratio, c sat (a)/c satoo = 1.02 , i.e., significantly greater than unity. If the droplets in an emulsion are polydisperse, and the disperse phase has a finite solubility in the continuous phase (as is the case for dodecane in water), this will result in Ostwald ripening: the smaller drops decrease in size and eventually disappear, whereas the larger drops grow. The driving force is the difference in chemical potential, and hence in solubility, of the material between small and large drops. The material is transported by diffusion; the mass transfer process can be called isothermal distillation; and the result is disproportionation (more specifically coarsening) of the drop size distribution. (i) LSW theory. Theory for the change in droplet size distribution owing to Ostwald ripening in emulsions has been developed by Lifshits and Slezov31 and independently
11
See, e.g., S.E. Friberg, P. O. Jansson, and E. Cederberg, J. Colloid Interf. Sci. 55 (1976) 614. For the most part, based on reviews by A. Kabalnov, E.D. Shchukin, Adv. Colloid Interface Sci. 38 (1992) 69; J.G. Weers, chapter 9, p. 292, in B.P. Binks, Ed. (1998), General References; and A. Kabalnov, J. Disp. Sci. Technol. 22 (2001) 1 31 After I.M. Lifshits, V.V. Slezov, Zhur.Eksp. Teor. Fiz. 35 (1958) 474; transl. in Soviet Physics JETP 35 (1959) 331. The names arc also transcribed as Lifshitz, Slyezov and Slyozov.
8.66
EMULSIONS
by Wagner". This LSW theory proceeds on the following assumptions: - the drops are fixed in space; - the distance between drops is very much larger than d , which implies that
^
^
where u = a/aCT; a cr is the radius of a drop that, at the moment considered, neither shrinks nor grows. The size distribution is of a fixed shape, shown in fig. 8.23b. The change in aCI is now given by, da 3 —CL
dt
4x'Dc sat =
?^L
[8.3.3]
9pD
where x' is given by [8.3.1] and D Is the diffusion coefficient of the drop material in the continuous phase; c sat is in kg m~3 . Equation [8.3.3] holds for any type of average radius, though with a different numerical constant. The time needed to double the average volume of the drops would be given by a 3 r 0 divided by the r.h.s. of [8.3.3]. Figure 8.23a illustrates changes in droplet size. Small droplets disappear, and hence
Figure 8.23. Ostwald ripening according to the LSW theory, (a) Examples of the change in radius (a) with time (t) of some drops in an emulsion; the heavy line gives the change in a c r . (b) The normalized distribution J(u) .
11
C. Wagner, Z. Elektrochem.
3 5 (1961) 5 8 1 .
8.67
EMULSIONS
the number of drops per unit volume, Nv , decreases. After the LSW size distribution has been formed, the decrease is by a second order reaction, and is given at any time by dJVv x'Dc t o o v^ ^k^N2 dt pD
[8.3 A]
with a numerical constant close to unity. Table 8.3 gives some calculated examples of the rate of Ostwald ripening. The rate is proportional to a 3 and to D, and inversely proportional to y and to c sat ^ . The latter especially is an important variable. For example, emulsions of hexadecane or triglyceride oil in water (i.e., oils that are virtually insoluble in water), show negligible Ostwald ripening, despite the driving force being substantial. Table 8.3. Calculated examples of the time t * needed to double a 3 r due to Ostwald ripening according to LSW theory (uncorrected). Pe is a Peclet number; see [8.3.5]. Disperse phase Continuous phase
Benzene
a cr0 /|im
Water
Decane Water
Tetradecane Water
Water Triglyceride oil
5
0.5
0.5
1
35
10 a)
10a)
10a)
78
142
198
18
880
730
763
998
x'/nm
2.5
1.6
1.9
0.15
D/m 2 s-'
10-9
8X10"10
8x10-'°
2X10" 1 1
c
0.7
5xl0- 5
2.8 xlO" 7
1.4
t*
39 hours
37 days
16 years
6 days
Pe
1.4
14
12
150
MD/Da p D /kgm-
3
sat,~/ k g m ~ 3
Presence of surfactant The LSW theory is confirmed experimentally, insofar as it concerns the effects of the variables in the equations; the linear increase of a 3 with time is also well obeyed. However, the absolute rate is generally substantially higher than that predicted, e.g. by a factor of 5. It is difficult to establish quantitative relations experimentally. In order to obtain results in a reasonable time, very small drops (a down to 0.1 um) are often used, and it is then quite difficult to obtain reliable size distributions. Moreover, y may vary during the process. The main factors identified as being responsible for deviations from LSW theory are:
8.68
EMULSIONS
- The drops are not fixed in space, but show Brownian motion. Whether this affects the ripening rate can be derived from the value of a translational Peclet number for particle diffusion over molecular diffusion. It is given by11 Pe = —(3kT/m) 1/2
[8.3.5]
where m is the droplet mass, and D the molecular diffusion coefficient. If Pe > 1, Brownian motion will enhance Ostwald ripening. Table 8.3 also gives value of Pe and, except for the first example, its value is considerable. Quantitative relationships for the ripening rate have not been given, but enhancement by a factor of at least 2 has been observed experimentally. Convection in the emulsion, e.g. due to temperature fluctuation, can further enhance the ripening rate. - The assumption that the distance between drops is much larger than their diameter only holds for very low volume fractions. A theoretical treatment21 resulted in an increase of ripening rate by factors of 1.4, 1.75, and 2.2 for
EMULSIONS
8.69
effect may be that several non polymeric surfactants can form micelles, which possibly enhances mass-transport between drops; see above. Important effects can be caused by the action of a surface dilational modulus, KJ . Using the Laplace pressure as a variable, Gibbs derived that Ostwald ripening will stop if dpL da
—— =
d(2y/a) ; da
=
2a(dy/da)~2y , az
=
4Kg-2y n a1
u
lo.o.ba]
where use has been made of the relationship din A = (2 /a) da . The condition can be written as, K°>y/2
[8.3.6b]
This conclusion should hold if the following conditions are fulfilled: - The larger drop is only slightly larger than the smaller one; in other words, the drop size distribution is very narrow. In practice, the condition will be more like
Kg>y. - The value of KJ remains constant in time, implying that the surface is covered by a true Langmuir monolayer (i.e., K% is fully elastic). This is rarely the case, and the consequences will be discussed below. - The drop will remain spherical. However, the surfactant film of a shrinking droplet is being compressed and it will probably buckle; the drop will then collapse into a dented particle. The condition for buckling is roughly given by, a>SK°/y
18.3.7]
where 5 is the thickness of the adsorption layer. Assuming this to be 5 nm, and the value of K^/y to be 2-5, as Is often the case, a droplet of radius over 10-25 nm will tend to collapse. Droplet collapse has frequently been experimentally observed. For some surfactants, the value of K^ / y may become much higher on compression of the monolayer, so that larger droplets can escape collapse. The result may then be formation of a bimodal size distribution11. - The drops do not coalesce. The growth of drops covered by a Langmuir monolayer will result in a decrease of their surface excess F; if F remains low, it may promote droplet coalescence, as has indeed been observed21. Altogether, stopping Ostwald ripening by application of a surfactant that gives a high surface dilational modulus will rarely be successful. However, Ostwald ripening can, in principle, be stopped by making droplets that are covered by suitable solid particles, as applied in Pickering stabilization against coalescence. The particles will
11 21
See, e.g., M.B.J. Mcindcrs, W. Kloek, and T. van Vlict, Langmuir 17 (2001) 3923. E. Dickinson, C. Ritzoulis, Y. Yamamoto, and H. Logan, Coll. Surf. B12 (1999) 139.
3.70
EMULSIONS
not desorb upon shrinking of the droplet, and the fairly thick and stiff layer will not readily buckle11. In nearly all cases the surface dilational modulus is a complex number (see [8.1.5]), implying that the surface is viscoelastic. This may result in a substantial decrease in Ostwald ripening rate. One may use either a time dependent modulus 31
dependent surface dilational viscosity
or a strain rate
to calculate the ripening rate, but analytical
expressions are not available. Numerical methods have yielded results that agree rather well with experimental results. For emulsions stabilized with proteins, the rate may be decreased by a factor up to five . This applies especially to those globular proteins that exhibit intermolecular cross linking after adsorption. Bulk rheological properties may also affect the ripening rate 5 ' (apart from the effect of continuous phase viscosity on the diffusion coefficient). A bulk yield stress larger than the Laplace pressure of the smallest droplets will prevent droplets shrinking or growing in size. However, the yield stress may have to be quite high, e.g., 1 MPa, which is difficult to realize. (ili) Drops of mixed composition. If the disperse phase consists of two components of different solubility In the continuous phase, this may retard Ostwald ripening, owing to the development of a counteracting change in chemical potential. A simple example is given when one of the components is not soluble in the continuous phase, such as aqueous drops containing some salt in a triglyceride oil. Upon shrinking of a drop, its molar salt concentration c (in o s m o l m 3 )6) increases over the initial value c 0 , whereby the osmotic pressure in the drop increases according to /7
osm =c R T = c o K / a ) 3 R T
I8-3-8!
assuming ideal behaviour. No change in droplet size will occur if the increase in Laplace pressure owing to shrinkage equals the increase in /7 o s m . This point is reached if y= 1.5 acRT
[8.3.9]
Assuming p I O m N m " 1 and a = 1 (im , a value of c = 2.67 osmolm~3 would be
For foam bubbles this has been shown clearly by Z. Du, M.P. Bilbao Montoya, B.P. Binks, E. Dickinson, R. Ettelaie, and B.S. Murray, Langmuir 19 (2003) 3106. 21 M.B.J. Meinders, M.A. Bos, W.J. Lichtendonk, and T. van Vliet, in E. Dickinson, T. van Vliet, Eds., Food Colloids, Biopolymers and Materials. Royal Soc. Chem. (2003). 31 A.D. Rontcltap, B.R. Damste, M. dc Gee, and A. Prins, Coll. Surf. 47 (1990) 269. 41 See. E. Dickinson, et al., loc. cit. 51 W. Kloek, T. van Vliet, and M. Meinders, J. Colloid Interface Sci. 237 (2001) 158. The osmolarity is the osmotically active concentration; for instance, for ideally dilute NaCl and MgC^ solutions it equals 2c Na Q and 3cMgC[ , respectively.
EMULSIONS
8.71
sufficient to stop ripening; this corresponds to less than 0.01% NaCl. The presence of a 'trapped component', i.e., a component soluble in the disperse phase, but not in the continuous phase, can thus be an effective means of stopping Ostwald ripening. Equations for the rate of Ostwald ripening can be derived if all drops contain two components, 1 and 2, which both are soluble in the continuous phase. Assume that the volume fractions of the components in the original drops are / t and J2 (where _ / j + / 2 = l ) . When the condition / 2 < c 2 o o / c ] o o is fulfilled, the rate would be, according to LSW theory, given by da-?,.
4x\ D9c9
cr.^
t
/,
1 2 2,sat.°o-'2
[8.3.10]
dt 9p 2 The rate is thus given by the scale x' for component 1, and the solubility and diffusion coefficient of component 2. If the mentioned condition does not hold, the rate is given roughly by [8.3.3], calculated for component 1. If an emulsion consists of two kinds of drops containing different materials, which are both soluble in the continuous phase, then compositional ripening will occur1 (besides Ostwald ripening). The materials are exchanged between drops and, eventually, all drops will attain the same mixed composition. 8.3c Aggregation Colloidal interaction forces between particles, and the kinetics of aggregation, have been discussed extensively in chapters IV.3 and 4, and additional information is in chapters V.I, 2 and 3. Hence, we need not discuss aggregation in detail. In a sense, emulsions are ideal systems to study aggregation, although the polydispersity and the difficulty of estimating the colloidal interaction forces (see sec. 8.1g) may pose some problems. We will only mention points that are specific for (some types of) emulsions. Kinetics. Large drops, say > 3 (im , do not generally show perikinetic aggregation. Small velocity gradients resulting from temperature fluctuations (see [8.3.17]) can strongly promote orthokinetic aggregation. Assuming that the capture efficiency is the same in both cases, the ratio of the rates of orthokinetic over perikinetic aggregation is given by, ^ d h o J
peri
=
^ ^ 2kT
l8.3.ll]
where Vt> is the velocity gradient. In water at room temperature, drops of d = 3 |xm will aggregate faster by a factor of three, owing to a Vu of only 1 s" 1 . Moreover, the drops will generally show sedimentation and, owing to their polydispersity, the larger 11
See, e.g., L. Taisnc, P. Walstra, B. Cabane, J. Colloid Interface Set 184 (1996) 378.
8.72
EMULSIONS
drops overtake the smaller ones; this will strongly enhance the encounter frequency. Film formation. When two emulsion drops form an aggregate and the attractive forces are relatively strong, a flat film may form between the drops. This means that the net force keeping the drops together increases. This phenomenon will not, or will hardly affect the aggregation rate, but it may well decrease the possibility for deaggregation, for example, owing to a velocity gradient in the liquid. Film formation upon the encounter of drops will be discussed in sec. 8.3e. Dielectric permittivity of the continuous phase. The dielectric constant in the oil of a water-in-oil emulsion can be quite small, say 3. The water drops may have a surface charge, but the decay of the electrostatic potential with the distance from the drop surface is very weak; in other words, IIK is very large. Consequently, the repulsion is almost purely Coulombic. Since the variation with distance of the repulsive force is quite low, it provides very little stability to the drops. These aspects were discussed in sec. IV.3.11. It is often difficult to find small molecule surfactants that stabilize w/o emulsions; suitable copolymers will provide steric repulsion. Depletion interaction (see sees. 1.8 and 9). Substances considered essential for making stable emulsions may well cause depletion interaction, and hence droplet aggregation. To slow down sedimentation, high polymers are often added to increase the continuous phase's viscosity, but if they are non-adsorbing, they may also cause depletion interaction. For example, the addition of 0.02% xanthan gum may already cause aggregation, hence rapid sedimentation. Higher polymer concentrations may lead to formation of a particle gel; aggregation is then so fast that a space filling network of drops is formed before appreciable sedimentation occurs. Depletion interaction can also be caused at a high amphiphile concentration, when the amphiphiles form micelles1'. The depletion Gibbs energy per pair of drops is given roughly by AG
depl = - 2 ™ d 7 7 m ( 2 a m - h f ,
0
ad»am [8.3.12]
^m =
RT c
M
( m / m ) ( l + ^m)
where a = radius, h = distance between drops, 77 = osmotic pressure, c = concentration (mass/volume), M = molar mass, and
EMULSIONS
8.73
Mixtures of non-ionic amphiphiles and caseinate have also been incriminated as causing depletion flocculation in an emulsion11. Bridging. Small particles, or long polymer molecules, that adsorb onto the drops, but are preferentially wetted by the continuous phase, can be adsorbed simultaneously onto two drops, thereby forming a bridge between them. An example is given by bridging by protein aggregates in o-w emulsions. This can occur especially if the mentioned species form the sole surfactant and are present in insufficient quantity to fully cover the droplet surfaces. Bridge formation can then generally occur during emulsification, but rarely afterwards. The addition of some suitable amphiphile to the emulsion can cause desorption of the particles, and stirring then leads to deaggregation. Other bridges can form by reactions between specific groups of the adsorbed surfactants, especially with polymers. After all, steric repulsion caused by polymers does not preclude temporary partial overlap of the adsorbed layers. An example is found in the bridging by divalent cations associating with monovalent acid groups, or covalent - S - S - cross links between adsorbed proteins. These bridges generally form after emulsification. Partial coalescence. Crystals can form in some emulsion drops, for example, if the drops consist of triglyceride or paraffin oil. These oils are generally multicomponent mixtures and the proportion crystalline will greatly depend on the temperature and its history. Often, some crystals protrude a little from the droplet surface. Such a crystal can pierce the film between approaching droplets, and thereby form a bridge. If the contact angle is suitable, say between 110 and 170° as measured in the aqueous phase, the crystal will be wetted by oil. If so, there is oil oil contact between drops, and a driving force for complete coalescence. However, the latter will generally be prevented, since the crystals in a drop tend to form a space-filling network. Hence the term, 'partial coalescence' (also called 'clumping'). Heating the emulsion to cause the crystals to melt leads to the coalescence of each clump into a large drop. In a quiescent emulsion, aggregation due to partial coalescence is generally very slow, but on application of a velocity gradient the process can be very fast, millions of times faster than true coalescence of a similar emulsion without crystals. In simpleshear flow, the drops will roll over each other upon encounter, greatly enhancing the chance of a crystal piercing the film; piercing may also be promoted by the shear forces pressing the drops closer to each other. The rate and the consequences of partial coalescence depend in an intricate manner on several variables . 8.3d Sedimentation^ By the buoyancy principle of Archimedes, the net gravitational force acting on a single sphere of diameter d submerged in a liquid is given by 11
E. Dickinson, C. Ritzoulis, and M.J.W. Povey, J. Colloid Interface Set 212 (1999) 466. For a review see P. Walstra (1996). 31 For the most part based on P. Walstra (1996), General References. 21
8.74
EMULSIONS
F
b=^d39(pd-Pc)
18.3.13]
where g is the acceleration due to gravity. Hence, the sphere will move downward through the continuous liquid if the density difference is positive, or upward if it is negative. The moving sphere will be subject to a frictional drag force, according to Stokes, given by Ff = fv = 3ndi]cv where j
[8.3.14]
is the friction coefficient and v is the linear velocity of the sphere with
respect to the surrounding liquid. Soon after the drop starts moving, inertial forces will become negligible and the drop will sediment at a constant 'Stokes velocity'. This is obtained by putting Fb = F{:
vs=i^P 18 77c
18.3.15]
For example, an oil droplet of 2 (im in water, Ap = -70 kg m~3 , will attain a sedimentation rate i>s = -0.15 \im s" 1 , at room temperature, i.e., cream by 13 mm per day. Equation [8.3.15] can be modified for centrifugal sedimentation by replacing g by the centrifugal acceleration, Rco2 , where R is the effective centrifuge radius and co the rate of revolution in rad s" 1 . The Stokes equation is frequently used for calculating sedimentation rates, but is only valid under a very restricted range of conditions. (i) Prerequisites Jor the Stokes equation. The following conditions should be met for the equation to hold: - It concerns one sphere in an infinite amount of a homogeneous liquid. In practice this means that the volume fraction of the emulsion drops must be very small, say (p < 0.003 , and the vessel should have a diameter at least 50 times d . In most emulsions, hindered sedimentation occurs, which is discussed below. - The drops must be homogeneous perfect spheres. Many emulsion drops nearly meet this condition, but if the drops are quite small and covered with a polymeric surfactant, complications may arise. It is often possible to calculate approximate corrections. - The drop surface must be immobile in the tangential direction. As discussed in sec. 8.1c, this is nearly always the case. -The drop Reynolds number, given by dvpc/ric,
must be much smaller than
unity. This is virtually always the case, also for centrifugal sedimentation. - Brownian motion should not disturb the sedimentation. For very small drops, the
EMULSIONS
8.75
Stokes equation is no longer valid and a sedimentation equilibrium is (slowly) established. The concentration, c , of the droplets as a function of the vertical distance, z, is then given by
c,Z, = c o e X p [ z ^ N ]
,8.3.16,
Assuming the disturbance owing to Brownian motion to be negligible if c[z)/cQ < 1.02 , and taking Ap = -70 kg m~3 and the maximum z value at 0.2 m, the condition becomes d > 50 nm , which will nearly always be fulfilled. - Convection currents should not disturb sedimentation. Such currents can form, due to slight temperature fluctuations. To illustrate this, the free convection flow between two vertical plates with temperature difference, AT, at a mutual distance, x, will be considered. The average velocity gradient is given by" VVJWXAT 32 7]c
,8.3.17]
where ji is the volume expansion coefficient. For water at room temperature and x = 0.05 m , Vu would roughly equal 3AT; a temperature difference of 0.1 K would then produce a gradient of 0.3 s" 1 . The stress caused by a gradient will equal Vu x j]c , and it should be much smaller than the net gravitational stress, which equals 4F b Ind2 . This leads to the condition, d»Vt>—^-r ng\Ap\
[8.3.18]
implying that for oil drops in water, the value of d, in um , should be much larger than 2 Vu in s" 1 . For a AT value of 0.1 K, this implies d » 0.6 um . This is a more restrictive criterion than the one for negligible Brownian motion. It is indeed observed that sedimentation of 1 um droplets in water hardly occurs, unless the temperature is kept precisely constant. - The continuous phase should be a Newton liquid, implying that the viscosity is independent of the viscous stress (in a stress controlled rheometer) or of Vu (in a shear rate controlled instrument). In many emulsions, the liquid is not Newtonian, the apparent viscosity rja generally decreasing with increasing stress. The stress acting on a sedimenting drop will, to a first approximation, be equal to F b divided by the droplet surface area, i.e., dg|Ap|/6 . For an oil drop of 5 um size in water, this stress would then be about 6xlO" 4 Pa. The viscosity used to calculate the Stokes velocity should therefore be measured at that value of the shear stress. However, most rheometers cannot give results at a stress below 0.1 or even 1 Pa, and the apparent viscosity measured may then be a considerable underestimation of the value prevailing 11
R.B. Bird, W.E. Stewart, and E.W. Lightfoot, Transport Phenomena, Wiley (1981) p. 217.
8.76
EMULSIONS
during sedimentation, by up to some orders of magnitude. It should be noted that the reasoning given here is an oversimplification, because the liquid flow around a sedimenting droplet has an elongational component; especially for visco elastic liquids this can make a substantial difference. It may even be that the liquid has a yield stress that is greater than the net gravitational stress acting on the drop, given approximately by dg|Ap|, but far smaller than the smallest stress to be applied in the rheometer. Such a yield stress is not then detectable, although it can completely prevent sedimentation.
(ii) Hindered sedimentation.
The Stokes equation gives the velocity of the drop
relative to the surrounding liquid. Since we are generally interested in the velocity relative to the vessel containing the emulsion, a correction for the displacement of continuous phase must be made: the Stokes velocity us has to be multiplied by (1 - (p). However, much larger corrections are needed. Batchelor attempted to derive a rigorous equation for homodisperse spheres of
v
. The result is
— =1-6.55?)
[8.3.19]
The factor -6.55 is composed of the following three terms, (i) An amount -5.50 for counterflow of continuous phase, owing to the drops' displacing liquid and, portantly,
more im-
dragging liquid along with them; (ii) an amount -1.55 owing to the overlap
of the flow disturbances caused by drops sedimenting near to each other; (iii) an amount +0.50 caused by group sedimentation (see below). The relationship only fits (approximately) for very small
[8.3.20]
V
S
where the exponent n should presumably equal 6.55. This equation fits for far larger (p values, but the exponent tends to be significantly larger: see, e.g., fig. 8.24, solid line, where n = 8.6 . Published values for n vary mostly between 7 and 9. A major factor explaining the high values of the exponent is polydispersity.
In a
theoretical study on mixtures of spheres of two sizes, again at low
increases for the smaller spheres, and
decreases for the larger ones; however, the total sedimentation rate was calculated to decrease. Experimental studies have also shown a decrease in sedimentation rate with increasing polydispersity.
11 21 31
G.K. Batchelor, J. Fluid Mech. 52 (1972) 245. J.F. Richardson, W.N. Zaki, Trans. Inst. Chem. Eng. 32 (1954) 35. G.K. Bachelor, C.S. Wen, J. Fluid Mech. 124 (1982) 495.
3.77
EMULSIONS
Figure 8.24. Creaming rate relative to the Stokes velocity v/vs of mixtures of skim milk and cream, of various oil volume fractions ip. Average droplet size, dg 3 =1.5 ^m . ( • ) Creaming under gravity; (o) ibid, in a centrifuge at 200 xg " .
The average drop size that should be taken for polydisperse systems is d 53 , as is explained in sec. 8.1e. (ili) Group sedimentation. Figure 8.24 gives creaming results on a series of o/w emulsions of various (p, both under gravity and in a centrifuge. The conditions (geometry, and time of creaming) were chosen so that both sets of experiments would give the same total creaming according to the Stokes equation. All of the prerequisites for the Stokes equation mentioned in subsection (i), except the first one, were fulfilled, so that it would be expected that values extrapolated to tp = 0 would be equal for the two sets. This was indeed observed. It is also seen that [8.3.20] is well obeyed for gravity creaming, but that the results for centrifugal creaming were very different. This must be related to differences in group sedimentation. As was mentioned in subsection (ii), Batchelor has considered this phenomenon. If two drops are quite close to each other, the total friction factor of the doublet was calculated to be smaller than twice that of a single drop, especially if the one drop is above the other. If so, the sedimentation rate will be enhanced. Batchelor assumed that the mutual positions of the drops would not change during the process, but that is unlikely. Drops diffuse, i.e., they move by Brownian motion, which may bring them into a mutual position in which sedimentation is enhanced. In principle, this would increase the overall sedimentation rate. Whether the increase is significant depends on the time scales involved, i.e., whether the time during which two drops are close to each other is long enough for a substantial increase in sedimentation to occur. It should be taken into account that the sedimentation distance is proportional to time, whereas the distance of diffusion is proportional to the square root of time. The tendency for group sedimentation to be significant may be expressed in a translational Peclet number, i.e. the ratio of the time needed for a droplet to diffuse over a distance equal to d, over that for sedimentation over the same distance. From the Einstein and
11
After results by P. Walstra, H. Oortwijn, Neth. Milk Dairy J. 29 (1975) 263.
8.78
EMULSIONS
Stokes equations we derive, pe =
^-!-± 6kT
[8.3.21]
for gravity creaming; in a centrifuge, g should be replaced by Rco2 . If Pe » 1, group sedimentation will be of importance. In the example of fig. 8.24, Pe ~ 0.5 for gravity creaming and about 100 in the centrifuge. It is seen in the figure that near
aggr=udropl-8JV(1-1/d)
[8.3.22]
where N is the number of (equal sized) drops in the aggregate, and d is the fractal dimensionality (which will generally be about 2); the proportionality factor 1.8 is approximate, and depends somewhat on the value of d. To give an example, for N = 100 , v would be increased by a factor of 18. 11
R. Johnc, Chem. Ing. Techn. 38 (1966) 428. E. Dickinson, J. Colloid Interface Sci. 73 (1980) 578. 31 L.G.B. Brcmer, Ph.D. Thesis, Wagcningen University, 1992. 21
EMULSIONS
8.79
In practice, the situation is more complicated. The drops are polydisperse. Larger drops will overtake smaller ones, leading to enhanced aggregation. Larger aggregates will overtake smaller ones and single drops, which will enhance aggregation even more. Altogether, the rate of aggregation and sedimentation will increase ever faster, and a sediment or cream layer is rapidly formed. See also fig. 8.25.
Figure 8.25. Examples of the concentration profile developing during creaming of emulsions. The drop concentration is given as volume fraction, y . a s a function of height, z. The numbers near the curves denote the time after creaming started (say, in hours), (a) Calculated for a strictly homodisperse emulsion, (b) Polydisperse emulsion, no aggregation, (c) Polydisperse emulsion with aggregating drops, (d) Percentage of the total drop volume creamed as a function of time t, for cases a c; t 5 0 is the time needed for half of the total drop volume to reach the cream layer. Approximate examples, meant to illustrate trends .
If the drops are not too large, v is not very small, and aggregation is relatively fast, fractal aggregation will lead to the formation of a space filling particle network, i.e., a gel. This will stop sedimentation. However, gravity may cause consolidation of the gel, and then a layer of continuous phase will be formed at the top or the bottom of the vessel. (v) Sedimentation profiles. In practice, one is often interested in the development of a sedimentation profile, asking what is the volume fraction of drops as a function of height in the vessel as a function of time? This will be illustrated for creaming, with reference to fig. 8.25. If all drops are of the same size, a sharp demarcation plane forms between a liquid 11
From P. Walstra, 2003, Physical Chemistry of Foods, Marcel Dekkcr (2003).
8.80
EMULSIONS
with
z
= cr
53 18/
' ' J(
[8.3.23]
where z cr is the creaming distance and f[
[8.3.241
Sec, e.g., M.J.W. Povey, in E. Dickinson, Ed., New Physico chemical Techniques Jor the Characterization of Complex Food Systems, Blackie Academic (1995), chapter 8, p. 196. 21 Largely based on Walstra, 1996; on chapters 1 (by B.P. Binks), 7 (by A.S. Kabalnov), 8 (by B. Deminiere, A. Colin, F.L. Caldera, and J. Bibcttc) and 10 (by D.N. Petsev) in Binks (1998); and on I.B. Ivanov, P.A. Kralchevsky (1997); see General References. 31 For the thinning and rupture of isolated planparallcl liquid films, see sec. 6.4c.
EMULSIONS
8.81
where Af is the area of the film and AGact is an activation Gibbs energy for rupture. The natural frequency va (units m~2 s"1) is the presumed rupture rate if AG = 0. It is often supposed that it equals the frequency at which a single surfactant molecule leaves the adsorption layer. Applying the absolute rate theory for molecular reactions, the frequency for an area occvipied by a single molecule would be given by , va=kT/hz2
[8.3.24a]
where z is the molecular dimension. Another theory11 is based on the fact that transverse thermal waves develop spontaneously on a liquid interface (as follows from surface light scattering studies; sec. III. 1.10), and assumes that a wave of length equal to z may cause the removal of molecules from the surfactant layer. The frequency 0) then would be in the ultrasound region, and co=vs/z
where vs is the sound velocity.
2
The occurrence per unit area of one molecule, z , is then given by va=o)/z2 = vs/z3
[8.3.24b]
Assuming that vs = 1 km s" 1 and z = 1 nm, [8.3.24a] yields va ~ 10 3 1 , and [8.3.24b] about 10 30 m~2 s" 1 . It is not clear to what extent the equations are correct, since it is difficult to obtain reliable data for AGact. Nevertheless, [8.3.24] gives the main factors determining the rupture frequency, and especially that it will be proportional to the area of the film. In practice, one is often interested in an overall coalescence rate, expressed, for example, as the decrease in the number of drops per unit volume, JV, with time. Plots of 1/JV against time are often reasonably linear. Another method involves the determination of the average size of the drops, e.g. d 32 , as a function of time. Numerous factors affect the coalescence rate, and it is not possible to characterize the overall situation with a set of simple equations. This is mainly because several regimes for coalescence can be distinguished, and the main variables determining coalescence rates are not all the same in the various regimes. (i) Regimes. Rather than defining a great number of regimes, five variables are given whose values are used to distinguish regimes. a. Time available for film
rupture. When the drops are aggregated, or closely
packed as in a sediment layer, the time available is as long as the emulsion is kept. Short times prevail when the drops remain separate until they encounter each other by Brownian motion (perikinetic), or as a result of agitation (orthokinetic). The time span during which two drops are then close to each other depends on several factors, such as droplet diameter, colloidal interaction forces, and hydrodynamic conditions; but it is generally below 1 s. For small droplets (order of a |am ) and in the absence of
11
Due to P.G. dc Gennes, J. Prost, The Physics of Liquid Crystals, Clarendon Press (1993) p. 597.
8.82
EMULSIONS
colloidal interactions, Brownian motion would, on average, cause the droplets to be within a distance of roughly 10 nm for a time of the order of (in SI units) 107 times T]cd . For droplets of 1 um in water, this would amount to 10 ms. In practice, most emulsions with drops below 10 um in diameter, which have been made with a suitable surfactant and do not show aggregation, will not exhibit coalescence, provided that sedimentation is prevented; the latter can be effected by slowly rotating the vessel during storage. Table 8.4. Role of the Weber number ([8.3.25]) in the deformation of emulsion drops; calculated examples. Laplace pressure y = 12 mN m"l 7 = 3 mN m~l
a =0.25 um a =6 um
PL PL
Local stress
a = external stress x{a/h)
1. Van der Waals attraction3' A = 5x10-21 J, h = 10 n m
cr= A/12reh 3
A = 5 x l 0 - 2 J,
^Vu = 10Pa,
2 a ~ 10 Pa a~ 5xlO 3 Pa
h= 3 n m
2. Hydrodynamic shear stress
a•= rflvalh 3 a- 10 Pa
a/h = 100
3. Stress in a cream layerbl
= 105 Pa =10 3 Pa
cr=cpcreamApgHa/ h
Ap = 7 0 k g m ~ 3 , H = 1 0 m m , a/h = 100 Same in centrifuge at 1000 x g
5xlO 2 Pa <J~
5xlO 5 Pa
It is assumed that strong repulsion occurs below distance h. A= Hamaker constant. cp = volume fraction; Ap = density difference; H = height in cream layer. b. Droplet deformation. Drops which encounter each other closely as a result of an external force may become deformed, as illustrated in fig. 8.18. External forces may be colloidal (net attraction), hydrodynamic, or gravitational. They are counteracted by the Laplace pressure of the drops. The balance can be expressed in a dimensionless Weber number, given by ffe =
^ext=(^Hxt=^exL pL pL 2yh
[8
325]
where A is a stress concentration factor that applies to undeformed drops, and which equals the drop radius over the smallest separation distance, h ; crext is the external stress. For We « 1 , the droplets will not be deformed, meaning that the film area remains quite small, of the order of a h . With increasing We, deformation gradually becomes stronger and at We = 1 , a flat (i.e., parallel sided) film starts to form. For
EMULSIONS
8.83
W e » 1 , a relatively large flat film is formed. Generally, coalescence occurs much faster at We > 1 than at We « 1, other things being equal. Some calculated examples of We values are given in table 8.4. It is seen that for small drops and not very low interfacial tension, a « p L , hence We « 1, unless quite a high external stress is applied. c. Intensity of agitation. With an increasing agitation intensity of an emulsion, the droplet encounter rate is enhanced (see [8.3.11]); the external stress acting on a droplet pair increases (table 8.4, point 2); and the time needed for a film between drops to thin to a given thickness increases for We » 1, and decreases for We < 1. Coalescence during intense agitation was discussed in sec. 8.2c, sub (vi); the conclusions obtained there also hold if the agitation is not so intense as to cause droplet break-up. Altogether, the coalescence rate can be higher during agitation than in a quiescent emulsion, and the relation between coalescence rate and some variables are different (see below). d. Interface rigidity. This is governed by the (transient) value of K^ , as discussed in sec. 8.1c. The rigidity affects the rate of drainage of the film between drops, which is determined by the Marangoni number (Ma, see [8.1.4]). From the discussion in sees. 8.1c and 8.3c, sub (vi), it can be concluded that in nearly all real situations Ma > 1, implying that the interface is rigid. In extreme cases, e.g., an o/w emulsion with a very low concentration of SDS and a low ionic strength (electrostatic repulsion will then nevertheless prevent droplet aggregation), Ma can be small. Another, rather hypothetical, case of a low K^| is the use of a very pure small molecule surfactant at a very high concentration. e. Surfactant category. There is a vast literature on coalescence theory, partly backed up by experimental results, but In nearly all of these cases it is (often implicitly) assumed that the surfactant is a (small-molecule) amphiphile. Macromolecular surfactants do not fit the theories; for example, [8.3.24a-b] for the natural frequency of film rupture cannot apply, and the value of the film thickness, which appears in all film rupture theories, is unclear. For an emulsion stabilized with solid particles the factors governing stability are different again. All combinations of the two or three differences mentioned under each heading would give rise to a very large number of regimes. Although several of these combinations cannot, or hardly, occur, there still remains a substantial number of regimes to be distinguished in practice. It is often tried in practice to predict the coalescence stability of an emulsion from the result of a test in which the coalescence rate is greatly enhanced. The prime example is centrifuging the emulsion, or pressing the droplets together by osmotic deswelling of the emulsion. It may now be clear that this will generally alter the coalescence regime by the two criteria (a) and (b). The change in Weber number, especially, will cause the prediction to be unreliable.
8.84
EMULSIONS
(ii) Film formation and drainage. All theories for film rupture state that, for a given system, the probability of rupture will be greater for a thinner film. Moreover, the probability would be proportional to the film's area ([8.3.241). Hence, film formation and thinning are important aspects. For the thinning of isolated plane-parallel films, see sec. 6.4a. Film geometry. Whether a flat film is formed or not, is determined by the value of the Weber number [8.3.251, but the interaction forces change when deformation occurs. The complete equilibrium geometry of the film can in principle be calculated if all interaction forces are known11. Factors to be considered include; (i) the colloidal interactions, e.g., van der Waals attraction, depletion interaction, electrostatic and steric repulsion; and (ii) the deformation Gibbs energy, which is due to an increase of interfacial area and to the bending energy of the surfactant monolayers (at the contour lines of the flat film). For the latter, only the mean bending modulus need be taken into account (see sec. III. 1.15), the Gauss curvature being negligible. The Gibbs energy change owing to compression/dilation of the monolayers tends to be negligible. The effect of a constant external stress can also be taken into account. The deformation Gibbs energy is an important factor in determining the radius of the film formed, but it does not affect the equilibrium thickness. The latter follows from the balance of the colloidal and external interaction forces. If no repulsive interaction forces would act (except hard core repulsion) a film of (nearly) zero thickness would be formed. Such a 'black' film can, in principle, be quite stable, but in most practical situations the film would rupture before a black film has formed. It may be noted further that the formation of a flat film implies the droplets' being irreversibly aggregated (as long as conditions remain constant). Drainage rate2). The rate of drainage is discussed in sec. 8.2c, sub (vi). Equation [8.2.21] would hold for We > 1, Ma > 1 , and h« R«a (see fig. 8.18), a regime that often prevails when coalescence occurs in an emulsion. In deriving the equation, one should take into account that the force, F , and a are related, because the excess pressure in the flat film between drops will equal the Laplace pressure of the drop; s i n c e R « a , p L will be close to that of the undeformed drop. Note, also, that the drainage rate will be smaller for a larger F acting on the droplet pair. This is because a larger value of F will cause a larger film radius, R , and hence more resistance to drainage, especially for small values of h where drainage is hindered most. If the force acting on the droplet pair is due to colloidal interactions, its magnitude will change (considerably) with decreasing h, depending on the shape of the G{h) curve, and the calculation of the drainage time is intricate. For droplets in a simple-shear field, F will be, to a first approximation, proportional to r]cVv, and hence constant. Integration of [8.2.21] for this case leads to an 11 A full account is given by D.N. Petsev, chapter 10 in Binks (1998), General References. See also I.B. Ivanov, K.D. Danov, and P.A. Kralchcvsky, Colloids Surfaces A152 (1999) 161. 21 For the most part, after A.K. Chcsters, Trans. Inst. Chem. Eng. 69-A (1991) 259.
EMULSIONS
8.85
expression for the time needed for the film to thin to a given value of h : t[h) =
3n/c2a4Vv \6y2h2
[8.3.26]
One should note the enormous effect of drop radius. (If the two drops are of unequal radii, the relation 2/a = l/a1 + l/a2 can be used.) Assuming that 77C=1CT3 P a s , Vu = 10 3 s" 1 , and / = 3 mN m" 1 , t(h = 5 nm) for spheres of a = \ and 10 urn, would be 0.8 (is and 8 ms, respectively. The first conclusion may be that the times are quite short. However, by comparing them to the time during which the drops are close together (which is of order 1/Vu , hence 1 ms in the present case), it can be seen that small drops readily come close to each other within that time, but large ones do not. It is indeed generally observed that, for dilute emulsions that are agitated, the coalescence rate is much slower for larger drops and a higher velocity gradient. As discussed in sec. 6.4, large films (R > 50 urn ) often show uneven drainage: some thin 'channels' and thicker 'islands' are formed. In such a case, [8.3.26] does not applyFinally, the drainage rate is not always a relevant variable. For example, for small droplets that form aggregates owing to perikinetic flocculation, or for drops that are in a sediment layer, where the time available for drainage is quite long.
Figure 8.26. Cross-section through part of a film of (average) thickness h between two emulsion droplets, (a) Illustration of hole formation, (b) Properties of a peristaltic wave developing in the film.
(Hi) Film rupture. This subject is extensively discussed in sees. 6.4-6. Here, a few aspects will be mentioned that are of special importance for coalescence in emulsions. One characteristic of the film, then, is that it is quite small, the radius rarely being larger than a few um , and often as small as 0.1 |J.m . De Vrles theory. The first attempt to calculate the Gibbs energy for the formation of a small hole or pore in a film was by de Vries11. Suppose that a hole has formed, as 11
A.J. dc Vries, Rec. Trav. Chim. 77 (1958) 383.
8.86
EMULSIONS
illustrated in fig. 8.26a. At point (1) in the figure, the Laplace pressure equals y(2/h-l/R) ; the two main curvatures at this point have different signs. At point (2), p L = 0 , and if the pressure at (1) is > 0, liquid will flow from (1) to (2) and the hole will grow spontaneously. This can happen if 2/h>\/R . To achieve this, the area of the o/w interface has to be increased by an amount of the order h 2 ; hence, the Gibbs energy of the system would be increased by an amount yh2 . The latter value can be interpreted as the activation energy for hole formation, and hence for coalescence to occur. The magnitude of h is determined by the net colloidal repulsion. Assume, for example, that y = 5 mN m"1 and h = 10 nm. This results in AGact = 5xlO~ 19 J or 125 kT, which would mean that the film cannot rupture. If h = 3 nm, the result would be 11 kT. Assuming the film area to be 0.01 (im , this would lead, according to [8.3.24], to a rupture probability of 1011 s" 1 , i.e., immediate rupture. However, the theory needs considerable modification. It has been pointed out, for example, that the interfacial tension need not be constant during hole formation. Moreover, the compression of the monolayers is counteracted by the surface dilational modulus; a correction as given in [8.2.19] has to be applied. Kabalnov-Wennerstrom theory^. This theory applies for the regime where the surfactant is an amphiphile, We » 1, Ma > 1, and agitation is absent or weak. When a hole is formed in a film at a small h value (fig. 8.26a), the local curvature of the monolayer is very strong. This can go along with a considerable increase in bending Gibbs energy, which should be added to the de Vries value of AGact, especially if the interfacial tension is low. Both the mean and the Gauss bending modulus are important. The values of these moduli depend closely on the natural curvature of the monolayer. The latter depends on the shape of the amphiphile molecules21. A (truncated) cone shape can lead to a high, and a cylindrical shape to zero natural curvature. Moreover, a difference between film curvature and natural curvature leads to an increase in interfacial tension. All of these effects, as well as the Gibbs energy of compression of the monolayer, are taken into account in the theory, which is therefore intricate. The theory is especially successful for systems that exhibit a phase-inversion temperature (sees. 8.1b, sub (i) and 8.2c, sub (iv)), and nicely explains the mechanism causing phaseinversion, as a function of surfactant properties. Vrij-Scheludko theory3). As discussed in relation to [8.3.24], transverse thermal waves can develop spontaneously on an interface, but waves of a much longer wavelength are considered here. Such waves tend to be strongly damped, since they cause 11
The original article is, A. Kabalnov, H. Wennerstrom, Langmuir 12 (1996) 276. An updated review is by A. Kabalnov, chapter 7 in B.P. Binks (1998), see General References. 21 See, e.g., J.N. Israelachvilli, Intermolecular and Surface Forces, 2nd ed.. Academic Press (1992). 31 Developed independently by A. Vrij, Discuss. Faraday Soc. 42 (1966) 1966, and A.D. Schcludko, Adv. Colloid Interface Sci. 1 (1967) 391.
EMULSIONS
8.87
local differences in the Laplace pressure. If, however, a net attractive force acts between the two faces of a film, and if this force is stronger for a lower value of h, the waves on the two surfaces can become coupled, as illustrated in fig. 8.26b. Van der Waals forces acting across the film fulfil these criteria, provided that the film is thin enough. The condition for coupling depends on a number of conditions; in most cases of interest it is given approximately by (AR2\V4
h<0.15 ^ -
[8.3.27]
Assuming A = 5xlO~ 2 1 J, R = 1 |im, and 7 = 3mNm~ 1 , the critical value of h is about 5 nm. The condition for hcr implies, inter alia, that the Vrij-Scheludko theory only applies if We » 1. Coupled waves can cause local thinning of the film; hence, the film can rupture at an average h value that would provide stability, according to the theory discussed above. Whether this occurs depends on a balance offerees. First, the Laplace pressure gradients induced in the film cause flow of continuous phase from a thick to a thin spot. In other words, the wave tends to be damped. The larger is the film radius R , the longer can be the wavelength of the film, and hence the smaller the curvature of the interfaces for the same wave amplitude, the smaller the Laplace pressure gradient, and the weaker the damping. Consequently, waves with a wavelength X ~ 1R will dominate. Second, repulsive forces generally act between the faces of the film, and the total interaction Gibbs energy as a function of film thickness, AG[h), must be taken into account. The condition for the amplitude to grow, in which case the film will eventually rupture (or form a metastable black film), is given by d2G(h) 5—<
2n2y 7T-
[8.3.28]
dh2 R2 Considering a DLVO type of interaction, [8.3.28] predicts that the film will always rupture for a value of h corresponding to the primary minimum in the interaction curve, whereas the film will generally be stable in the secondary minimum. The theory given above has its limitations. First, it does not apply to very large films, say R > 50 |im , as mentioned1'. This can be neglected for nearly all emulsions: the drops are too small. Second, it is implicitly assumed that the interfaces are rigid. In an extension of the theory21 the effect of the surface dilational modulus is introduced. It turns out that the Kg values encountered in practice are nearly always sufficient to have a rigid interface. For the simple case, where the only colloidal interaction working is van der Waals attraction, the condition is Kg > A12nh2 . 11 See also, B.P. Radocv, A.D. Scheludko, and E.D. Manev, J. Colloid Interface Set 95 (1983) 254. A. Vrij, F.T. Hcsselink, J. Lucassen, and M. van den Tempel, Proc. Kon. Akad. Wetensch. B73 (1970) 124.
8.88
EMULSIONS
Hence, the critical magnitude for the modulus is virtually always below 1 mN m"1 . Only when the surfactant concentration is very low, may this value not be reached; if so, coalescence can occur already at a large value of h , provided condition [8.3.27] is fulfilled. Third, the theory would need modification if the main attractive force does not increase with decreasing h , as is the case for depletion interaction. Other aspects. The theories for film rupture do not generally provide good quantitative prediction of coalescence rates, but most trends are predicted. However, in some cases, variables other than those considered seem to control the coalescence process. Examples follow. A black film may form, rather than a hole in the film. Black films can have a considerable life time. Whether this occurs in films between small emulsion droplets and, if so, under what conditions, is poorly known. This is in contrast to isolated, plane-parallel films, for which black film formation is an important stabilization mechanism, see sec. 6.4d. Some amphiphiles can give rise to the formation of a laminar liquid crystalline phase at a relatively low concentration; this also depends on the further composition of the solvent. Such a surfactant may then form multilayers around emulsion droplets, if its concentration is high enough, and such multilayers provide a very high stability against coalescence1'. The explanation is still a matter of debate. Polymers. The rupture theories discussed do not apply if the surfactant used is a polymer. The author is not aware of a sound theory. Nevertheless, it is well known that polymeric surfactants can provide long term stability (several years) against coalescence, even for We » 1 and in a sediment layer. This is presumably a result of steric repulsion, acting over a relatively long range, keeping the droplets far away from each other and thereby precluding film rupture. Other things being equal, polymers of a higher degree of polymerization (in the range of, say, ten to several hundreds) tend to provide greater stability. Surfactant layer coherence. Several workers have observed that the coalescence rate of emulsions is correlated with the value of the surface shear viscosity, rfg (often called by the misleading term, 'film strength'), as measured at a macroscopic interfacial layer of the surfactant used in making the emulsion. An example is given by water in crude oil emulsions in which the surfactant consists of high molar mass asphaltenes in the mineral oil21. Another example is o/w emulsions stabilized by various globular proteins31; some of these proteins are known to form intermolecular cross links in an adsorption layer, which also appears to correlate with the value of T]G . The correlations are not perfect, and some clear exceptions to the rule have been
11
See, e.g., S. Friberg, L. Mandell, and M. Larsson, J. Colloid Interface Set 29 (1969) 155. P. Bechcr, p. 257 in H.F. Eicke, G.D. Parfitt, Eds., Interfacial Phenomena in Apolar Media, Marcel Dekker (1987). 31 E. Dickinson, B.S. Murray, and G. Stainsby, J. Chem. Soc. Faraday Trans. I 84 (1988) 871. 21
EMULSIONS
8.89
observed. It has recently been reported11 that failure, i.e., yielding or fracture, of the adsorption layer can occur during the measurement of r]% , which implies that the viscosity is measured in a 'destroyed' layer. In other words, the result is deceptive. It may be better to directly measure two-dimensional stress and strain at failure and correlate the results with proneness to coalescence. It may well be that strong coherence of the adsorption layer prevents rupture of the film surfaces, and thereby rupture of the film itself. (iv) Concluding remarks. Several factors determine whether coalescence can occur, and those dominating during emulsion formation and during storage are quite different. Duringjbrmation, the Marangoni number (at very short time scale) is the essential parameter. If Ma > 1, rapid drainage of the film between approaching droplets Is prevented, and coalescence will rarely occur. On the other hand, during storage the Weber number is the essential parameter. If We < 1, the film formed between droplets is generally too small in area or not thin enough to allow film rupture. There are a few exceptions to this dichotomy. If a 'finished' emulsion is agitated during storage (at an intensity that is insufficient to cause droplet break-up) an intermediate situation arises. This is, to some extent, discussed in subsec. (ii), but only for simple-shear flow. For elongatlonal and turbulent flows, the relations are partly different. Another exception relates to emulsion formation and stability at conditions close to the phase-inversion temperature. Here, We » 1 , and any agitation is quite weak; it follows that stability during formation and storage are, for the most part, governed by the same variables. Finally, Pickering emulsions appear to provide an exception; see sec. 8.4. Some practical variables will now be considered. Unless mentioned otherwise, it is assumed that the droplets are not too large, say < 10 |j.m, and are covered by a monolayer of surfactant; agitation is assumed to be negligible. Droplet size. For a larger value of d, the value of We is larger, and the area of the film forming between two droplets will be larger, also at constant We . This will mean faster coalescence, because the probability of film rupture is enhanced. This phenomenon has been observed by several workers. One study concerned very dilute protein stabilized emulsions, where the surface excess was much below its plateau value, so that coalescence did occur. The oil drops were allowed to cream to an oil layer on top of the emulsion. Drops reaching that position were observed by a microscope, and the average time needed for coalescence of a drop with the oil layer was noted. The time ranged from about 35 s for 2 jim drops, to 2.5 s for 10 [im drops 2 '. In another investigation, amphiphile-stabilized drops in a highly concentrated emulsion (e.g., cp= 0.85, hence We»1) were studied. The coalescence rate was 11 T. van Vliet, G.A. van Aken, M.A. Bos, and A.H. Martin, p. 176 in E. Dickinson, T. van Vliet, Eds., Food Colloids, Biopolymers and Materials, Royal Soc. Chem. (2003). 21 See E. Dickinson, et al. (1988), loc dt.
8.90
EMULSIONS
emulsion (e.g.,
11
B. Deminiere et al., chapter 8 in B.P. Binks (1998), General References. A good example is given by, J.A.M.H. Hofman, H.N. Stein, J. Colloid Interface Sci. 147 (1991) 508.
2)
EMULSIONS
8.91
8.4 Case Study: Pickering Emulsions1 Emulsions that are stabilized by adsorbed small solid particles have been mentioned in sec. 8.3b, sub (ii), as they can provide considerable stability against Ostwald ripening. Such adsorbed particles can also stabilize an emulsion against coalescence; this mechanism is called Pickering stabilization21. The particles should be located in the o/w interface, protruding further into the continuous phase than in the disperse phase. The latter condition implies that 0 < a < 90°, where a is the contact angle as measured in the continuous phase; see sec. III.5.lie. Moreover, the Gibbs energy for desorption of a particle needs to be large. For a sphere of radius a , the energy of its transfer to the continuous phase is given by AG a ^ 0 = 7ia2 r o w (l-cos«) 2
[8.4.1]
Assuming yow = 0.03 N in" 1 , a = 60°, and a = 3 nm, the value of AG equals about 50 kT, which would mean that desorption due to Brownian motion is virtually impossible. Hence, even quite small particles can be used. (For transfer of the particle to the disperse phase, the minus sign in the equation must be replaced by a plus sign). There is another factor that affects coalescence. If two drops are pushed close together, adsorbed particles can move laterally, away from the point of closest approach, and this can cause formation of a bare patch on the drop(s), which may, in turn, induce coalescence. The lateral motion can be counteracted by the formation of surface tension gradients, which needs a significant surface dilational modulus. The value of K° depends on, (i), the particle surface coverage, which must be large, say, 9>0.7 (depending on particle shape and on lateral interaction forces between the adsorbed particles); and, (ii), on the presence, type, and surface excess of surfactant(s) in the interface. However, the stability and the formation of the emulsion cannot be seen as independent processes. To make a Pickering emulsion, agitation is needed, but the intensity of agitation mostly does not determine drop size. The limiting factors are the concentration and size of the solid particles. The droplet diameter always exceeds the particle diameter, by a factor ranging mostly between 30 and 300, even if the particle concentration is not limiting. It has frequently been observed that the contact angle should not be much below 90° to obtain relatively small droplets (d = 50 a ). The explanation of these phenomena is yet unclear, to the author's knowledge. In most of the older studies on Pickering stabilization, relatively large solid parFor the most part, based on S. Levinc, B.D. Bower, and S.J. Partridge, Colloids Surfaces 38 (1989) 325 and 345; D.E. Tambe, M.M. Sharma, Adv. Colloid Interface Sci. 52 (1994) 1; B.P. Binks, S.O. Lumsdon, Langmuir 16 (2000) 8622; and N.X. Yan, M.R. Gray, and J.H. Masliyah, Colloids Surfaces A193 (2001) 97. 21 After S.U. Pickering, J. Chem. Soc. 91 (1907) 2001, although W. Ramsdcn, Proc. Roy. Soc. (London) 72 (1903) 156, was earlier in describing the phenomena.
8.92
EMULSIONS
tides were used, diameter O(l nm ), resulting in large drops. Generally, a surfactant is added, ostensibly to change and fine tune the contact angle. It cannot be ruled out, however, that the surfactant also helps In forming y gradients during emulsification, thereby counteracting drop re-coalescence; see sec. 8.2c, sub (vl). More recently, several far smaller particles (about 10 nm or even less) have become available, and some of these nanopartlcles, especially those made of silica, can be tailor made so that they produce the desired contact angle. Pickering emulsions can now be made in the absence of an added surfactant. Probably, the nanoparticles In the interface can produce a significant surface pressure, owing to their relatively large molar surface excess. For particles of 5 nm diameter and 0= 0.75, [III.3.4.39] yields n= 2.0 mN m" 1 . Moreover, the particles are large enough to form a true Langmuir monolayer. Hence, a small but significant surface dilational modulus can result, allowing the formation of interfacial tension gradients, provided that the droplets are small. Another difference is that the coarse emulsions tend to be unstable, unless weak attractive forces act between the particles in the Interface. The latter has been shown to result in larger values of the surface dilational modulus. The attraction should not be strong, since the particles would then aggregate in solution, thereby precluding the formation of an emulsion, since the aggregates are generally too large to allow full coverage of the drops. Hence, careful optimization of product properties and process conditions are needed to obtain a stable emulsion. Such problems do not (or need not) occur when nanoparticles are used. This may, again, be due to the nanoparticles producing a sufficiently large surface dilational modulus. Moreover, the drops obtained are far smaller (provided that the agitation is sufficiently intense) and hence the Weber number for coalescence will be smaller. For nanoparticles, the contact angle should also be close to 90° to allow formation of small drops. Theories have been proposed to explain some of the phenomena observed, and these have certainly enhanced understanding. However, a definitive and all embracing theory of the formation and stability of Pickering emulsions is not yet available. This means that the 'state of the art' Is , as for nearly all types of emulsions, that the application of the fundamentals of surface and colloid science has produced considerable and useful understanding of the processes Involved, and of the properties resulting, but several significant topics still need further research. 8.5 General References R. Aveyard, B.P. Binks, and J.H. Clint, Emulsions, Stabilized Solely by Colloidal Particles. Adv. Colloid Interface Sci. 100 (2003) 503-546. (Review, emphasizing the wetting and thermodynamic properties of well-defined particles at the o/w interface; the paper also includes a section on triple emulsions (w-o-w and o-w-o).)
EMULSIONS
8.93
P. Becher, Emulsions: Theory and Practice, Reinhold Publ. (1953). (Presumably the first book giving substantial understanding and a range of properties.) B.P. Binks, Ed., Modern Aspects of Emulsion Science, Royal Soc. Chem. (1998). (Most aspects discussed in this chapter are treated, largely based on fundamental theory and in great detail. Especially recommended are chs. 1, by B.P. Binks, on recent advances; 2, by P. Walstra, P.E.A. Smulders, on emulsion formation; 5, by E. Dickinson, on emulsion rheology; 6, by B.W. Brooks et al., on phase-inversion; 7, by A.S. Kabalnov on coalescence; 8, by B. Deminiere et al., on coalescence in concentrated emulsions; 9, by J.G. Weers, on Ostwald ripening and related phenomena; and 10, by D.M. Petsev, on the various interaction forces between droplets.) A.K. Chesters, The modelling of coalescence processes in fluid liquid dispersions: A review of current understanding, Trans. Inst. Chem. Eng. 69-A (1991) 259. (A very clear and complete review.) J.T. Davies, Turbulence Phenomena, Academic Press (1972). (Much information on droplet break-up, particularly in the turbulent regime.) Encyclopedia of Emulsion Technology, Vol. 1, Basic Theory (1983); Vol. 2, Applications (1985); Vol. 3, Basic Theory Measurement Applications (1988); Vol. 4 (1996); P. Becher, Ed., Marcel Dekker. (Contains several chapters on themes discussed in this chapter and also treats a range of applications.) I.B. Ivanov, P.A. Kralchevsky, Stability of emulsions under equilibrium and dynamic conditions, Colloids Surfaces A, 128 (1997) 155. (A critical analysis of theories and results on hydrodynamic and thermodynamic aspects.) I.B. Ivanov, K.D. Danov, and P.A. Kralchevsky, Flocculation and coalescence of micron sized emulsion droplets, Colloids Surfaces A152 (1999) 161. (An extension of part of the previous article.) V.G. Levich, Physicochemical Hydrodynamics, Prentice-Hall (1962). (Includes much information on droplet comminution under various conditions.) E.H. Lucassen-Reynders, Ed., Anionic Surfactants: Physical Chemistry of Surfactant Action, Surfactant Series 11, Dekker Marcel (1981). (Contains treatments of fundamental aspects important for emulsion formation and stability.) P. Sherman, Ed., Emulsion Science, Academic Press, London (1968). (Gives fundamentals of emulsion formation, stability, and various properties. Especially chapter 1, by E.S.R. Gopal, on emulsion formation is still worth reading.)
8.94 H.N. Stein, Ed., The Preparation
EMULSIONS of Dispersions,
Chem. Eng. Sci. 46 (2), (1993),
pp. 201-460. (A symposium report containing seven contributions on emulsion formation.) P. Walstra, Formation of Emulsions, ch. 2, pp. 57-127, in; P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1, Marcel Dekker (1983). (A fairly comprehensive review.) P. Walstra, Principles of Emulsion Formation. Chem. Eng. Sci. 46 (1993) 333-350. (A review with some emphasis on the time scales of the various processes.) P. Walstra, Emulsion Stability, ch. 1, pp. 1 62, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 4, Marcel Dekker, 1996. (A review)
Appendix 1 Self-consistent field modelling a) Basics b) Lattices and transition probabilities c) The self-consistent potentials d) Chain Hamiltonian e) Boundary conditions f) The standard state grand potential g) The compressible W5 model for the aqueous phase
A. 1 A. 2 A.5 A.7 A. 8 A. 10 A. 11
This Page is Intentionally Left Blank
APPENDIX 1 SELF-CONSISTENT FIELD MODELLING A.1 Basics At the base of self-consistent field analysis is the mean-field partition function Q[N,V,T) of a system with specified geometry and molecular composition. The characteristic function is the Helmholtz energy F(N,V,T) = -kT\nQ(N,V,T). Quite generally this Helmholtz energy can be cast in the form: F[u,f\ = kT^
ntIn—±
Y Y uA(r)(pAlr) +F^Hp]
[A. 1.1]
It is composed of entropic contributions (the first two terms on the r.h.s.) and one enthalpic term. Very characteristic for a SCF theory is the second contribution on the r.h.s. of [A. 1.1], which features the so-called self-consistent potentials u, conjugated to the dimensionless segment concentration (volume fraction) profiles (p. The self-consistent potentials, also called the segment potentials, operate as external potentials felt by the molecular entities, i.e. they appear in the Boltzmann weights that determine the distributions of the segments. The term "self consistent" is linked to this potential because the external potentials are made functions of the segment densities. The quantities u and
A1.2
densities and all interaction energies are known. Again, we use the notation [cp\ to remind us about this. It is quite important to specify for a given model all contributions that are accounted for. Therefore, one should mention this for each type of application. Usually there are contributions from packing effects (excluded volume), contributions from the short-range (nearest-neighbour! interactions, a term due to the electrostatic contribution and one due to the polarization of segments in an electric field. Optimizing the Helmholtz energy with the constraint that at each coordinate the volume fractions add up to unity (incompressibility constraint) gives the characteristics to which equilibrated systems must obey for all segment types A and coordinates r
-—— = -kTy ' -
3F
=O
[A. 1.2]
aF"11
= -uAr) + =0 A 3
[A. 1.3]
The coupled segment potential energy and volume fraction profiles that optimize the Helmholtz energy are known as the self-consistent field solutions. The density profiles are of course directly relevant because they may be used to find structural information on association colloids. The SCF profiles may also be used to compute the Helmholtz energy of the system ([A. 1.1]) and it is possible to also extract other relevant thermodynamic variables. It is essential to mention that the standard state grand potential em can also be accurately computed. In principle, the parameters that are needed in th^ SCF modeling are measurable quantities. However, as in coarse-grained MD or MC. the exact values of these parameters will depend on the choices made in the modeling, e.g. the level of coarsegraining. As a result, it is not practical to tabulate them, although one can make educated guesses. As a rule, pair interactions should be consistent in that the values needed to interpret one type of experiment should be identical for another type when the pairs are encountered under exactly the same conditions ( p, T, density, ...). For example, in self-assembly problems one uses the c.m.c. data to find good estimates for the FH interaction parameter for the hydrocarbon-water contacts because the c.m.c. depends very sensitively on the length of the hydrocarbon chain.
A. 2 Lattices and transition probabilities
In the following, we will elaborate a discrete (lattice) version of the SCF approach in the spirit of Scheutjens and Fleer1'. One important issue is that the geometry of the system
" G.J. Fleer, M.A. Cohen Stuart. J.M.H.M. Scheutjens, T. Cosgrovc. and B. Vincent. Polymers at Interfaces. Chapman & Hall (1993).
A1.3
Figure A l . l . Schematic representation of various lattice geometries used in the SCF calculations. The radial directions are indicated by r = 1,2,3,-•• . The variable z = 1,2,3,••• is used to refer to layer numbers in non-curved (flat) directions. In the case where there are two non-curved directions, the variable x = 1,2,3,-•• is also used, a) Flat lattice with flat lattice layers with one gradient; b) Cylindrical coordinate system with one gradient; c) Spherical geometry with one gradient; d) Flat geometry with two gradients; e,f) Cylinder geometry with two gradients. is an input parameter. In other words, to allow the evaluation of the partition function it is necessary to fix the coordinate system and the most appropriate way of doing that is by using a lattice. All lattice sites have a characteristic length given by I and each lattice site has a volume v0 = I3 . Referring to fig. A l . l , one uses lattices with flat, cylindrical or spherical geometry In which for the first two geometries one can choose between one and two gradients in density. For the radial directions (in the cylindrical and spherical lattices), we use the layer numbers r = 1,--,M , where r = 1 is the central lattice layer. An important quantity in each geometry is the number of lattice sites per coordinate. In the flat and ID cylindrical cases, the number of sites per layer is infinite. In these cases, we consider the number of sites per unit area (flat) or the number of sites per unit length (cylindrical). In the one-gradient lattices (flat, cylindrical and spherical) we have just one coordinate and in two-gradient systems there are two. L{r) = — l[rl}3 - [(r -1)!]3) 3vo\ I
(spherical)
[A.2.1 ]
A1.4
L{r) = — l[rl}2 -l(r-l)l]2)
(cylindrical)
[A.2.2]
L(z) = l
(flat)
[A. 2.3]
L(z,r)=—([r!] 2 -[(r-l)i] 2 )
(2-gradient cylindrical)
[A.2.4]
L(z, x) = 1
(2-gradient
[ A. 2.5 ]
flat)
Next, it is useful to compute local, geometry-dependent averages; these are also known as site fractions. Site fractions can be taken for quantities that are given as a function of the coordinate r , e.g. the volume fraction
(spherical and cylindrical)
[A.2.6]
r'
(0z)) = ]T Kz.z'J^z')
(flat)
[A.2.7]
(2-gradient cylindrical)
[A.2.8]
(2-gradient flat)
[A.2.9]
z'
(0z,r))= Y A{z,r,z',r')
(qAz,x)) = ^ l(z,x,z',x')
The transition probabilities X are subject to constraints, e.g., for spherical symmetry, ]£ - A(r,r') = 1 , where the sum typically extends only over the nearest neighbours. Otherwise stated, the transition probabilities A(r,r') have only non-zero values when r and r' are neighbouring coordinate s. It is important that the transition probabilities should obey the internal balance equation, which says that the probability of having contacts between sites r and r' counted from the r -side equals that counted from the r' -side L(r)A(r,r') = L(r')/l(r',r)
(spherical and cylindrical)
[A.2.10]
L(zU(z,z') = L(z'U(z',z)
(flat)
[A.2.11]
L(z,r)A(z,r,z',r') = L(z',r'U(z',r',z,r)
(2-gradient cylindrical)
[A.2.12]
L{z,x)Mz,x,z',x') = L{z',x')Mz',x',z,x)
(2-gradient
[A.2.13]
flat)
This internal balance is obeyed when the transition probabilities are scaled with the contact area of a site at coordinate r with coordinate r ' . Here we choose to present the contact areas in units I2
A1.5
a(r + l,r) = 4;r((r + l)2 - r 2 )
(spherical)
[A.2.14]
a(r + l,r) = 2/rr
(cylindrical)
[A.2.15]
a(z + l,z) = l
(flat)
[A.2.16]
a(z,r + l,z',r) = 27rr
(2-gradient cylindrical)
[A.2.17]
a(z,x,z',x') = l
(2-gradient
[A.2.18]
flat)
and the transition probabilities for spherical and cylindrical geometry are
A(r,r+1) = ! ^ l ± ! l 3 L(T)
1A.2.19]
A(r>r_l) =
[A.2.20]
l£!LIZ» 3 L{r)
Mr,r) = 1 - Mr,r + 1) - Mr,r -1)
[A.2.21 ]
For flat geometry, Mz,z + 1) = Mz,z-l) = Mz,z)-
[A.2.22]
For cylindrical geometry with two gradients, we use A(z,r,z/,r + l) = - a ( Z ' r ' Z ' r + 1) 9 L(z,r)
z ' = (z-1), z, (z + 1)
[A.2.23]
A(z,r,z < ,r-l) = - a ( z ' r ' Z ' r - 1 ) 9 L(z,r)
z ' = (z-1), z, (z +1)
[A.2.24]
Mz,r,z,r) = l-3Mz,r,z,r + l)-3Mz,r,z,r-l)
[A.2.25]
and finally we have for a flat geometry with two gradients Mz,x,z',x') = -
[A.2.26]
for z' = (z -1), z, (z +1) and x' = {x-1), x, (x +1). Note that in the 2-gradient systems we allow for 9 contacts (steps), which in the limit of no curvature are all weighted equally. Experience has shown that this Ansatz gives the smallest lattice artifacts. A.3 The self-consistent potentials The next step is to specify the terms that are accounted for in the self-consistent segment potential. One should account for all the volume work necessary to take a segment A from the bulk to coordinate r . Applying [A. 1.3], we may find
A1.6
^
= u'(r) + X Z A B ( ( ^ B ( r ) ) - ^ ) + ^ A ^ l
+
_l_£o(£A_i)E(r)2
iA.3.!,
where the sum over B runs over all segment types. The first term on the r.h.s. expresses the work necessary to create a vacant site at coordinate r and is coupled to the compressibility relation ]Tp A (r) = l
[A.3.2]
A
which is imposed at every coordinate r . Below we automatically implement the incompressibility constraint on the volume fractions in the bulk and therefore the work to annihilate a site in the bulk is zero, i.e. u'b = 0 by definition. The second term in the r.h.s. of [A.3.1] accounts for the short-range nearestneighbour contacts that are lost in the bulk and created at coordinate (r). In the volume fraction gradients, we use the site averages as elaborated on in sec. A.2. Shortrange interactions are parameterized by the well-known Flory-Huggins exchange interaction parameters. The third term in the r.h.s. of [A.3.1 ] gives the usual electrostatic contribution to the segment potential. In a classical Poisson-Boltzmann theory this is the only term in the segment potential. This theory neglects the other terms as expressed in [A.3.3] because the size of the ions is ignored. We explicitly account here for the volume of the molecular components and therefore the additional terms are essential. The vA is the valence (including the sign) of the segment A, e is the elementary charge and y/[r) is the electrostatic potential at coordinate r . It will be clear that the electrostatic potential follows from the Poisson equation that will be used in the form coVf(r)xV^(r) = -p(r)
[A.3.3]
The discretization of the Poisson equation, necessary to apply it to a discrete charge and variable dielectric permittivity profile, has its intricacies, which we will not elaborate on here11. The space charge density profile p(r) is found simply from the volume fraction profiles p(r) = £ p A ( r ) v A e
[A.3.4]
A
For the dielectric permittivity profile, we use the volume fraction-weighted average £ £
o^
= « o X
[A.3.5]
A
where EA is the relative dielectric permittivity of a pure phase of segments of type A. The fourth term in the r.h.s. of [A.3.1] accounts for the work to polarize a unit of 1
J. van Male, Self-Consistent-Field Theory for Chain Molecules: Extensions. Computational Aspects and Applications. PhD thesis Wageningen University (2003).
A1.7 type A when it is moved from the bulk where the electric field E(r) = -d y/(r)/dr is zero to a position r where the electric field is finite. For more details see ref.1). A.4 Chain Hamiltonian The final ingredient is information regarding the chain Hamiltonian. The chain Hamiltonian exactly specifies how the volume fraction profiles follow from the segment potentials. Let us introduce a chain molecule of type i with segment ranking numbers s = l,---,JV1, where Ni is the total number of segments of the molecule. Here we will present the procedure in the simplest case when the chains are linear and the segments freely connected. In the literature, extensions to branched molecules and to semi-flexible chains are available21. The type of each segment needs to be specified. To this end, we introduce the chain architecture operator 5^ , which assumes the value unity when segment s of molecule i is of segment type A and zero otherwise. The full set of 8 values are an input for the model. For most surfactant systems, and especially for polymeric surfactants, it is sufficient to treat the molecules in a freely connected segment level (zeroth order Markov scheme). To this end, we introduce segment type-dependent distribution functions Gh(r) = exp[-(uA(r)/)cT)], which are generalized to molecule type and ranking numberdependent distribution functions Gj(r,s) = £ A GA(r)<5A. . These distribution functions are used to generate two complementary end-point distribution functions, i.e. Gj(r,s|l) and Gjfr.sliV). These end-point distribution functions are generated recursively by a propagator formalism Gj(r,s 11) = Gi(r,s)(Gi(r,s-l
| 1))
Gjfr.s | N) = Gi(r,s)(Gi(r}s + l | IV))
[A.4.1] [A.4.2]
which are started by Gj(r,l 11) = Gj(r,l) and Gt{r,N | JV) = Gt{r,N), respectively. By this construction, the end-point distribution Gj(r,s|l) contains the combined statistical weights to find chain fragments from s' = 1 •••,s in the system with the constraint that the segment s is at coordinate r . The other end can assume all possible positions such that the fragment with length s can indeed reach the coordinate r . A corresponding definition can be given for the complementary end-point distribution functions. The combination of two complementary end-point distribution functions gives access to the measurable volume fraction profiles
11
J. van Male, loc. clt. F.A.M. Leermakers, J.M.H.M. Scheutjens, J. Chem. Phys. 89 (1988) 3264; L.A. Meijer, F.A.M. Leermakers, and J. Lyklema, J. Chem. Phys. 110 (1999) 6560. 21
A1.8
where the denominator corrects for the fact that the segment potential of segment s at coordinate r is already accounted for in both end-point distribution functions specified in the nominator. The normalization constant C{ is related to the single chain partition function Qt = J]rL(r)Gi(r,l | N). When the total number of segments of molecule i is given by 8i = 2 s X r ^ ' r ' ^ i ' r ' s ' •w e have q=-^-
[A.4.4]
As [A.4.3] should also be correct when it is applied to a homogeneous bulk with volume fraction cpP , it follows that Ct = cpP I JVt. This means that for given 0i one can compute
This completes the set of equations. The segment potentials may be used to calculate, via the propagators [A.4.2] and the composition law [A.4.3], the segment type-dependent volume fraction profiles. It is rather trivial that the segment typedependent volume fractions follow from
A1.9 higher numbers, whereas at the upper boundaries the first coordinate inside the system is at a lower coordinate . When studying adsorption issues, an impenetrable wall exists at one or more system boundaries. This implies that all molecular entities that are in the system cannot have a finite concentration at the coordinate of the wall. We often refer to this boundary condition as absorbing with the rationale that all conformations of molecules that intersect with the system boundary are removed (have been absorbed) from the statistical sum. We can implement this by freezing a surface component in the boundary with a volume fraction of unity, i.e.
A1.10
Figure A1.2. Illustration of the boundary conditions (b.c.) used in lattice-based selfconsistent field models. Several allowed conformations are shown of oligomers consisting of five segments. The first layer "outside" the system is labeled z b . The layers in the system are numbered z = 1, 2, 3, 4,... In the absorbing b.c. there is an impenetrable wall (black bar) at z = z b , and the molecules in the system cannot enter this region. Conformations that cross the boundary (not shown) are omitted from the partition function. In the reflecting b.c, we distinguish two cases. The mirror plane is indicated by the dashed line: (i) the mirror 1 b.c. layer z = z b is made identical to z = 1, (ii) the mirror 2 b.c. layer z = z b is made identical to z = 2. The arrows point to the mirror-imaged layers next to the boundary. In the reflecting b.c. we show a pair of conformations and their mirror images. No conformations are omitted from the partition function, however because of the imposed symmetry the statistical weight of various conformations are affected by the b.c.
boundary conditions will be electroneutral overall. The boundary conditions issue is illustrated in fig. Al .2. For self-assembly systems the reflecting boundary condition is frequently used. One should be aware of the fact that the system size sets the average distance between micelles, membranes etc. By systematically varying the system size one can therefore study the effect of micelle-micelle interactions. Then not only the structural properties of the association colloids will depend on the proximity of neighbouring entitles, but also the thermodynamlc quantities will depend on the interactions. For open systems the most important thermodynamic parameter is the grand potential. A. 6 The standard state grand potential Once a self-consistent field solution of the SCF equations is obtained, which obeys the compressibility constraint, including the requirement of overall electroneutrality and the appropriate boundary conditions, it is possible to evaluate the grand potential of the translationally restricted micellar system. It follows from a sum over the grand potential density co{r) £m = X UrMr)
[A.6.1 ]
r
where the local grand potential density may be interpreted as the local tangential
Al.ll pressure or [pN(r)~pT(r)]{a0/v0) (where ao/vo 3 volume (u 0 = ( ) of a lattice site) and is given by
i
*
is the area (a 0 °= I2) and v0 the
A
4 l I ^K^B^-^l-^^W^Blj + i^1^1 1
A B
Z
[A 6 21
-'
]
where it should be stressed that in inhomogeneous systems the grand potential density is not uniquely defined. This is due to that last term of [A.6.2], which contains the site fractions. In principle the site fractions give a non-local contribution to the grand potential density, and the double counting of all pair interactions and the correction by the factor of 2 as specified in [A.6.2] is just one of the ways to do the correct bookkeeping of the pair interactions. Alternatives are possible. The implication is that the exact local values of co{r) depends on this choice. However, when the appropriate sum is taken over all sites as given in [A.6.1], the standard state grand potential is independent of such choice. A. 7 The compressible W5 model for the aqueous phase In most of the SCF modelling of association colloids, we have treated the water phase as a mixture between water and free volume where the water molecules occupy five sites. This solvent model is called the "W5 compressible model". The strategy to formulate this model was to confront predictions based on the W5 model with the results of the more elaborate model discussed in sec. 4.3b(ii). Such an internal comparison between similar theories leads to reasonable values for the interaction parameters of the new model. As this model was specially developed for chapter 4, it is appropriate to discuss it in more detail. The solvent phase was taken to be composed of W5 clusters, where a central unit is surrounded by four similar ones with free volume V (which occupies just one site). We fix the FH interaction parameter at X^y = 2.5. This value is high enough such that the free volume- W5 system has a strong solubility gap. In line with the SCF results of sec. 4.3b(ii), the free volume-rich phase has the composition ( (p^ = 0.00016472, (fJy] = 0.99984 ) and the corresponding water-rich phase has the composition ( ^ * ' = 0.95140 , cpl™] = 0.048605 ). This means that there is about 5% of free volume in the water phase. The interfacial tension may also be evaluated from this "water"-vapour interface. In dimensionless units yxfi IkT = 0.66535 , corresponds to y ~ 69 mN/m, when for the characteristic length of the lattice site 1 = 0.2 nm is used. This value is sufficiently close to experimental values. In passing, we note that such a compressible water model poses the options either to present results in the lattice-gas mode, i.e. by converting the chemical potential of
A1.12 the free volume to the pressure and convert the chemical potentials of the various molecular components correspondingly, or to remain working in the incompressible model where the free volume is interpreted as just one extra component with a chemical potential associated to it. Within the lattice approach, the latter is by far the most convenient. In all calculations of chapter 4, the amount of free volume in the water phase was fixed at p ^ ' = 0.048605 (resembling a constant pressure ensemble), and the free volume was allowed to distribute itself throughout the system according to the interactions. When the W5 system is equilibrated against an alkane phase ( CN ), we should again have a strong solubility gap. It appeared below that Xcw = 11 leads to a correct c.m.c. as a function of the tail length. Comparison with the detailed SCF model, in which the free-volume uptake in the SOPC bilayer is approximately 50% higher than in the water phase, points to %cv = 2 . The interfacial tension between the alkyl phase ( N = 36) and the W5 system is (with these parameters) given by y ~ 52 mN/m. Again, this is in semiquantitative agreement with experimental data for the macroscopic surface tension.
CUMULATIVE SUBJECT INDEX OF VOLUMES I (FUNDAMENTALS), II and III (INTERFACES) AND IV and V (COLLOIDS) In this index bold face print refers to chapters or sections; app., and fig. mean appendix, and figure, respectively. The roman numerals I, II, III, IV and V refer to Volumes I, II, III, IV and V, respectively. When a subject is referred to a chapter or section, specific pages of that chapter or section are usually not repeated. Sometimes a reference is made even though the entry is not explicitly mentioned on the page indicated. Entries in square brackets [..] refer to equations. The following abbreviations are used: (intr.) = introduced; (def.) = definition of the entry, ff = and following page(s). Combinations are mostly listed under the main term (example: for negative adsorption, see adsorption, negative), except where only the combination as such makes sense or is commonly used (example: capillary rise). Entries with 'surface' are often also found under 'interface' except where one of the two is uncommon. Entries to incidentally mentioned subjects are avoided. For the spelling of non-English names, see the preface to each volume. To avoid undue expansion of this index, chemical substances are mostly grouped together; for instance, for butanol, palmitic acid, sodium dodecylsulphate, hexane and dimyristoylphosphatidylethanol amine (DMPE) look under alcohols, fatty acids, surfactant, (anionic), alkanes and (phospho-)lipids. absorption bands; 1.7.14 absorption coefficient; 1.7.13 absorption index; 1.7.13 absorption (of radiation); see electromagnetic radiation acceptor (in semiconductor); II.3.172ff 'acid rain1; 11.3.166,11.3.221 acoustic waves; 1.7.44, II.4.5d acoustophoresis; II.4.5d activity coefficient; 1.2.18a, 1.2.18b, 1.3.50 Debye-Huckel theory; 1.5.2a, I.5.2b of (single) ions; 1.5.1a, 1.5.1b, I.flg. 5.2, [1.5.2.28] (Davies) activator (in flotation); III.5.97 additivity (in coagulation); IV.3.9k adhesion; II.2.5, II.5.97, 01.5.4, III.5.2 (also see: wetting, adhesional, work of adhesion) adhesive joints; III.5.17 adiabatic (process); 1.2.3 (def.) admittance spectrum; II.3.93, II.fig. 3.30, II.3.97 adsorbate; I.1.17(def.), 1.3.17 ideal; 1.1.17 thickness; II.2.63, II.2.76, II.2.79 ellipsometric; 1.7.10b
2
SUBJECT INDEX
(see further: adsorption of polymers) adsorbent; I.1.18(def.) adsorption; I.1.4ff(intr.), Il.chapters 1-3 and 5, III.chapter 4, V.chapter 1 and diffusion; I.6.5d, I.6.5e, II. 1.6, see adsorption, kinetics in emulsions; see V.chapter 8, energy; I.1.19(intr.), I.3.23ff, [1.4.6.1], II.1.22, II.1.3c, II.1.3f, II.1.44ff, II.1.50ff, II.3.6d-e, II.chapter 5, V.chapters 1, 3 for heterogeneous surface; II. 1.104ff enthalpy; 11.1.3c, II.1.3d, II.1.3f, II.2.5b, Il.figs. 2.26-28, Il.figs. 3.60-61 isosteric, II.1.3c, II.1.3d, II.1.28, Il.figs. 1.8-1.10, II.2.26, II.2.49 entropy; 1.3.30, II.1.21, II.1.22, II.1.3c, II. 1.29, II.1.3f, II.fig. 1.11, II.1.43, II.1.52, Il.flg. 1.16, II. 1.65, V.chapters 1, 3 from solution; 1.2.73, 1.2.85, Il.chapters 2 and 5, III.chapter 4, V.chapters 1, 3 basic features; II.2.2, Il.flg. 2.1 composite nature; II.2.2, II.2.3-2.6 dilute solutions; II.2.4, II.2.7, II.5.7, II.5.8 electrosorptlon; II.3.12 exchange nature; II.2.1 experiments; II.2.5, II.5.6, III.chapter 4, V.chapters 1, 3 in emulsions; V.8.2d functional; II. 1.18, II. 1.33 Gibbs energy; 1.2.74, II.1.3e, II.1.3f, II.1.45, III.chapter 4, IV.chapter 3 V.chapters 1, 3 heat; II.1.3c, II.1.28, Il.flg. 1.7, Ill.flg. 4.16 (also see: enthalpy) Helmholtz energy; II. 1.21, II. 1.23 heterogeneous surfaces; II. 1.7 hysteresis; Il.flg. 1.13, II. 1.42, II. 1.82, Il.figs. 1.31-33, Il.flg. 1.35, Il.flg. 1.39, II.1.6e, II.5.26, II.5.7d, V.flgs. 3.16-17, V.3.38ff, V.fig. 3.20 kinetics; II.1.45ff, II.2.8, II.5.3c, III.4.5 localized; I.3.5b, I.3.6d, II.1.7, II.1.5a, 11.1.5b, II.l.Bd, II.1.5e, II.1.5f, Il.flg. 1.21 mobile; 1.3.5d, II.1.7, II.1.5c partially mobile; II.1.5d, II.1.5e, Il.flg. 1.18, Il.flg. 1.21 negative; I.1.3(intr.), 1.1.4, 1.1.5, 1.1.21, 1.2.85, 1.5.93, II.3.5b, II.3.7e, Il.flg. 3.40, II.5.20ff, II.5.3e, Il.flg. 5.6, Il.flg. 5.9, IV.fig. 5.3 physical; II. 1.18, Il.flg. 1.13 presentation of data; II.1.4, Il.flg. 1.12
SUBJECT INDEX adsorption (continued), residence time (adsorbed molecules); II.1.46ff specific; see specific adsorption standard deviations; II. 1.45 superequivalent; II.3.62(def.), IV,3.9j, see further: specific adsorption t-plot; II.figs. 1.27-28, II. 1.88, Il.fig. 1.34 (statistical) thermodynamics; I.2.20e, 1.2.22, II. 1.3, II.2.3, II.3.6d-e, II.3.12, II.5.5 «-plot; II. 1.90 (also see: adsorption isotherm (equation), calorimetry; for adsorption at liquidfluid interfaces, see (Gibbs) monolayers. For specific examples see under the chemical name of the adsorbate) adsorption of: atoms; 1.3.5b, 1.3.5c, II.chapter 1 biopolymers; 1.1.2, Il.fig. 5.26b, Il.fig. 5.29, V.chapter 3 gases and vapours; II.chapter 1 functional; II. 1.18, (for relation to wetting, see III.5.3b, III.fig. 5.16 ions; 1.1.20, II.3.6d-e also see: double layers, electric polyelectrolytes; II.5.8, V.2.3c, V.fig. 2.19 charge compensation; II.5.85 chemical and electric contributions; II.5.84ff, Il.fig. 5.32;, Il.figs. 5.38-39 grafting; V.2.3c isotherms; Il.figs. 5.32-33, Il.figs. 5.35-40 multilayer; V.2.6e, V.fig. 2.37 multi-Stern layer; II.5.87 profiles; Il.fig. 5.34 theory; II.5.5g polymers; 1.1.2, 1.1.19, 1.1.23, I.fig. 1.18, 1.1.27, 1.2.72, Il.fig. 4.42, H.chapter 5, Il.fig. 5.1, Il.fig. 5.6, V.chapter 1 applications; II.5.9 bound fraction; II.5.18, II.5.71, Il.fig. 5.22, V.chapter 1 dispersity effects and fractionation; II.5.3d, Il.fig. 5.8, II.5.7c, Il.figs. 5.29-31, V.I.91 energy parameter; II.5.28, V.chapter 1 equilibrium aspects; V.fig. 1.51 experimental techniques; II.5.6 film rupture (effect on); V.8.88 hysteresis: II.5.26. II.5.7d, V.1.12d isotherms: Il.figs. 5.7-8. Il.fig. 5.22. Il.figs. 5.25-28
3
4
SUBJECT INDEX
adsorption of polymers (continued), kinetics; II.5.3c layer thickness; II.5.6b, II.fig. 5.19, II.figs. 5.23-25, II.5.72ff, V.fig. 1.51 electrokinetic; II.4.128, II.fig. 4.42, II.5.63ff ellipsometric; II.5.64ff hydrodynamic; II.5.61ff, II.figs. 5.24-25 steric; II.5.65ff (also see, profiles) loops; II.5.18, II.5.32, figs. II 5.19-23, II.5.70ff, V.fig. 1.6, V.1.6c negative (= depletion); II.5.20, II.5.3e, Il.flg. 5.9, III.2.58 (self adsorption), V.1.8, V.1.9 profiles; II.5.18, Il.flg. 5.6, Il.flg. 5.10, II.5.40, II.figs. 5.15-16, II.5.6c, Il.figs. 5.19-20, III.3.8e, V.fig. 1.1, V.I.6, V.I.5, V.1.8, V.1.9, V.I.11, V.fig. 1.30. scaling; V.I.68, V.2.3 tails; II.5.18, II.5.32 , Il.figs. 5.19-23, II.5.70ff, V.fig. 1.6 theory; II.5.4, V.chapter 1 diffusion equation; II.5.32, (V.I.4.1] (Edwards) excluded volume effect; II.5.5b GSA (ground state approximation); V. 1.2 lattice theories; II.5.30ff Monte Carlo; II.5.30 scaling; II.5.4c Scheutjens-Fleer theory; II.5.31, II.5.5, see self-consistent field theory square gradient method; II.5.33 statistics; II.5.29 trains; Il.flg. 2.27, II.5.18, II.5.32, Il.flg. 5.19, Il.figs. 5.21-23, II.5.70ff, V.chapter 1 proteins; V.chapter 3 competition; V.3.8 dispersion forces; V.3.20 driving forces; V.3.3c electrostatics; V.3.19-20 equations of state (2D); V.fig. 3.24 fluid interfaces; V.3.7 hydration/dehydration; V.3.21-22 hydrophobic interactions; V.3.4b hysteresis; V.3.3 and 3.5 kinetics; V.3.3a, V.fig. 3.6 principles; V.3.1-3 reconformation; see structural alterations
SUBJECT INDEX adsorption of proteins (continued), relaxation; V.3.3b, V.3.32ff, V.fig. 3.18 reversibility; see hysteresis and relaxation structural alterations; V.3.4, V.3.22ff, V.figs. 3.15-19, V.3.6c, V.fig. 3.28 surfactants; 1.1.25, III. 1.14b ionic; II.3.12, III.4.6d non-ionic; II.2.7d, III.4.6c adsorption isosters; II.1.3d, Il.fig. 1.9, II.1.37, Il.fig. 1.12f, II.2.26 adsorption isotherm (equation); I.1.17ff(intr.), I.fig. 1.12, II.1.3ff, Il.fig. 1.12, II.1.5, Il.app. 1 Brunauer-Emmett-Teller (BET); I.3.5f, II.1.5f, [II. 1.5.47], [II. 1.5.50], Il.fig. 1.24a classification, adsorption from dilute solution; II.2.7b, Il.fig. 2.24 gas adsorption; II. 1.4b surface excess; II.2.3c, Il.fig. 2.8, II.2.4 composite; 1.2.85 (see surface excess) Dubinin-Radushkevich; [II. 1.5.56] electrosorption; II.3.12b, II.5.5g Frenkel-Halsey-Hill; [II. 1.5.55] Freundlich; 1.1.19, II.1.2, [II.1.7.7] (generalized), Il.fig. 2.24c Frumkin-Fowler-Guggenheim (FFG); I.3.8d, II.1.5e, Il.fig. 1.19, II. 1.64, Il.fig. 1.43, [II.A1.5a], II.2.65, H.3.195 for surface excess isotherm; II.2.4d for specific adsorption of ions; II.3.6d Harkins-Jura; [II. 1.5.57] Henry; 1.1.19, 1.2.73, 1.6.65, II.l.2, [II.Al.la], Il.fig. 2.24a heterogeneous surface; Il.fig. 1.43 high affinity; 1.1.19, Il.fig. 2.24d, Il.figs. 5.7-8, Il.fig. 5.26, Il.fig. 5.29, Il.fig. 5.31, V.chapter 1, 3 Hill-De Boer = Van der Waals; see there individual; see partial Langmuir; 1.1.20, 1.2.74, I.3.6d, I.fig. 3.2, 1.3.46, II.1.2, II.1.28, II. 1.4a, II.1.5a, II.1.5b, Il.figs. 1.14-17, II.1.5d, II.1.5e, [II.1.7.7] (generalized), [II.A1.2a], Il.fig. 2.24b, II.2.86, II.3.196 binary mixture; II.2.4b, II.2.4c, Il.fig. 2.11 local; II.1.104, Il.fig. 1.43, II.1.108 one-dimensional; 1.3.8a Ostwald-Kipling; II.2.3b, [II.2.3.6], [II.2.6.1] partial (= individual); Il.fig. 2.9, Il.figs. 2.11-14
5
SUBJECT INDEX
6
adsorption isotherm (continued), partially mobile; II.1.5d, Il.fig. 1.18 potential theories; II. 1.73 quasi-chemical; I.3.8e, II.1.57, [II.A1.6a], II.3.196, V.2.21 standard deviation; 1.3.36 statistical thermodynamics; II.1.3ff, II.1.5, II.1.6, II.1.105, II.2.4, 11.3.12,11.5.5, V.chapter 1 surface excess; II.2.3, II.figs. 2.11-14, Il.figs. 2.18-23 molecules of different sizes; I1.2.4e, Il.fig. 2.14 relation to interfacial tension; II.2.4f multilayer; II.2.44ff Szyzskowski; [III.4.3.14] Temkin; [II. 1.7.6] thermodynamics; I.2.20e, II. 1.3, II.2.3, II.3.6d-e, II.3.12, II.5.5 Van der Waals (= Hill-De Boer); II.1.59, [II.1.5.27], Il.fig. 1.20a, Il.fig. 1.23, [II.A1.7a] virial; I.3.8f, II. 1.58, [II.A1.4a] Volmer; II.1.5c, Il.figs. 1.15-17, [II.A1.3a] (also see: Gibbs1 adsorption law) adsorptive; 1.1.17-18 (def.) Aerosil; see silica aerosol; I.1.5(def.), 1.1.6, V.8.33 AES = Auger electron spectroscopy AFM = atomic force microscopy = SFM, scanning force microscopy ageing; 1.2.99 aggregation; 1.1.2, 1.1.6, IV.fig. 4.23 colloids; IV.2.2C, IV.4.5 emulsions; V.8.3c surfactants; see V.chapter 4 (see the pertaining systems) agitation (foam formation); V.fig. 7.9 air bubbles, electrophoresis; II.4.130ff, IV.3.115 floating; I.fig. 1.4 foam stability; V.7.10, V.7.19, V.fig. 7 submersion of; 1.1.1, 1.1.2, 1.1.11, III.fig. 5.19 albumin, adsorption; V.fig. 3.20, V.fig. 3.22, V.figs. 3.24-27, V.fig. 3.28, V.fig. 3.30 negative adsorption; IV.fig. 5.3 second virial coefficient; IV.fig. 5.29
SUBJECT INDEX alcohols, surface dynamics and rheology; III.figs. 3.46, 47, III.4.3c, III.figs. 4.12-16, III.table 4.1 surface entropy; III.fig. 2.15 surface tension; III.4.3c Volta potential; Ill.fig. 4.14, III.fig. 4.24 Alexander-de Gennes model (brushes); V.I.56, V.1.60 alkanes, surface entropy; Ill.fig. 2.15 surface dynamics and rheology; III.figs. 46, 47 surface tension, data; Ill.fig. 2.17, Ill.table 2.3, Ill.fig. 4.10 simulations; III.2.41-43, Ill.figs. 2.12-13 lattice theory; III.2.60 ff, Ill.fig. 2.17-19 alkylpolyglucoside microemulstions; V.5.6b aluminum oxide; IV.fig. 2.2d adsorption of HC1; Il.fig. 1.10 adsorption of water vapour; Il.fig. 1.28 boehmite preparation; PV.2.4c double layer; Il.table 3.8 gibbsite preparation; IV.2.4c point of zero charge; Il.app. 3b alveoles; V.6.8 amphipathic; I.1.23(def.) amphiphilic; I.1.23(def.) amphipolar; I.1.23(def.) analytical ultracentrifugation; IV.2.54ff analyzer; 1.7.27, I.fig. 7.7, 1.7.98 anionic surfactants, see surfactants anisotropic media; 1.7.14 birefringence; 1.7.97, 1.7.100 scattering; 1.7.8c anisotropy; 1.7.8c, 1.7.14 of colloidal particles; see particles (colloidal), shape annealed (polyelectrolytes); V.2.2 antagonism (in coagulation); IV.3.9k antifoam; V.7.6 antithixotropy; IV.6.14 Antonow's rule; [III.2.11.121, III.5.76 apolar media, double lavers; II.3.11
7
8
SUBJECT INDEX
apolar media (continued), electrokinetics; II.4.50 solvation; I.5.3f arabic gum; see gum arable Archimedes principle (for interaction in a medium); 1.1.30, 1.4.42, 1.4.47, 1.4.50, I.4.69ff, I.fig. 4.15, IV.3.91ff d'Arcy's law (flow in porous media); I.6.4f, IV.2.32ff, IV.4.50-51 area; see surface area association colloids; I.1.6(intr.), I.1.23ff, IV. 1.5 association colloids, general, esp. modelling; V.chapter 4 bending (of worm-like micelles); V.4.6d bending and vesicles; V.4.7d, V.fig. 4.38, V.fig. 4.47 bilayers (lipid); V.4.37ff, V.fig. 4.7, V.4.7c, V.4.99ff critical micellization concentration; V.4.1c, V.table 4.1. Many examples in V.chapter 4 cylindrical micelles; V.4.6, V.figs. 4.25-30, V.4.115ff disc-like micelles; V.4.7b, V.figs. 4.33-34 end-cap energy; V.4.6c, V.fig. 4.27, V.fig. 4.45 extensive introduction; V.4.1a ionic; V.4.5, V.figs. 4.18-24, V.fig. 4.37 kinetics of micellization; V.4.10 lamellar phases; V.4.7 lamellar phases, interactions; V.4.8 mass action model; V.4.2b micelles, size fluctuations; V.4.2d mixed micelles; V.4.9a, V.figs. 4.42-43 molecular simulations (MD, Monte Carlo); V.4.30, V.4.33ff, V.figs. 4.5-6 non-ionic; V.4.4, V.figs. 4.9-14, V.figs. 4.25-29, V.4.7, V.figs. 4.31-36 pluronics; V.4.4e, V.figs. 4.15-16 profiles; V.fig. 4.1, V.figs. 4.6-7, V.figs. 4.10-12, V.fig. 4.15, V.figs. 4.20-21, V.fig. 4.26, V.fig. 4.32, V.fig. 4.35, V.fig. 4.46 quasi-macroscopic models; V.4.3c, V.fig. 4.8, V.4.5e second c.m.c; V.4.6e self-consistent field theory; V.4.3b, V.fig. 4.6, V.fig. 4.9, V.4.4, V.4.5, V.4.6b, appendix V. 1 solubillzation; V.4.9b, V.fig. 4.46 surfactant packing parameter; V.4.1d, [V.4.1.4] thermodynamics (classical); V.4.2 thermodynamics of small systems; V.4.19ff undulation forces; V.4.8a, V.figs. 4.39-40 vesicles; V.figs. 4.35-38
SUBJECT INDEX association constant; I.5.2d association of ions; see ion association association of water; 1.5.3c atomic force microscopy; 1.7.90, II.1.12ff, ILfigs. 1.3-4, II. 1.91, Il.fig. 2.17, Ill.table 3.5, III.3.7d, Ill.figs. 3.66-68, III.fig. 5.28. V.fig. 1.43, V.fig. 3.14 ATR = attenuated total reflection, attenuated total reflection; 1.7.81, II.1.18, II.2.54, Il.fig. 2.16, V.fig. 3.18 Auger (electron) spectroscopy; 1.7.11a, I.table 7.4 autocorrelation function; Lapp. 11.1 autophobicity; II. 1.80, II. 1.101 averaging; I.3.1a(intr.) azeotrope (in adsorption from solution); II.2.23 or-method (porous surfaces); II. 1.89-90 bacteria, Corynebacterium; Il.fig. 4.39 halophilic; 1.1.27 Nitrosobacter, Nitrosomonas; II.3.122 BAM = Brewster angle microscopy Bancroft rule; III. 1.84, III.4.97, V.8.5, V.8.59ff barium sulphate; II.table 1.3 barometric distribution; 1.1.20 barrier crossing; IV.fig. 4.4 barycentric derivative; 1.6.5 Bashforth-Adams tables (for capillarity); III.l.lSff Batchelor eq. (viscosity); [IV.6.9.9] Baxter model (adhesive hard sphere); IV.5.41, IV.figs. 5.22-26, IV.fig. 5.30 BBGKY = Bogolubov-Born-Green-Kirkwood-Yvon (recurrency expression); II.3.53ff BDDT = Brunauer-Deming-Deming-Teller, II.(isotherm classification); II.1.4b beam splitters; 1.7.98 beating (optical) = optical mixing beating (mechanical, foam preparation); V.7.13 bending; Ill.fig. 1.34 bending moment of interfaces; 1.2.91, V.8.86 bending moduli of interfaces; [III.1.10.2], III.1.55, III.1.15, III.1.79, III.tables 1.6 and 1.7 (data); III.4.7, V.4.6d, V.4.7c, V.5.5a, V.fig. 5.34 Bernouilli's law; [V.8.2.2] Berthelot principle (for interaction between different particles); 1.4.3Iff, [III.2.11.18] BET = Brunauer-Emmett-Teller; see adsorption isotherm BET-transformed; II. 1.69 Bethe-Guggenheim (approx.) = quasi-chemical biaxiality (anisotropic systems); 1.7.97
9
10
SUBJECT INDEX
bicontinuity in microemulsions; see there bilayers; V.4.7 Bingham fluid; IV.fig. 6.5, IV.fig. 6.17, IV.fig. 6.21 Bingham viscosity; IV.6.12, 6.40 binodal; 1.2.68, II.5.12, II.fig. 5.4, III.2.26, IV..5.64ff, IV.fig. 5.62 binominal; [IV.A. 1.4] biological activity; V.fig. 3.29 biomineralization; IV.2.38 biopolymers; 1.1.2, 1.1.27, II.fig. 5.5, V.chapter 3 birefringence; 1.7.14 Bjerrum length; [1.5.2.30a] for polyelectrolytes; [II.5.2.23], V.2.9 Bjerrum theory (ion association); I.5.2d black body (radiation); 1.7.22 black film; see film, liquid blob; II.5.11, V.fig. 1.31 blood clotting; V.3.52ff body (or volume) forces; I.1.8ff(intr.), 1.4.2 Bohr magneton; 1.7.95 boiling point elevation; 1.2.74 Boltzmann equation, II.Boltzmann factor; I.3.10ff, [11.3.5.4], II.3.216 dynamic; II.3.217 Boltzmann's law (for entropy); [1.2.8.2], [1.3.3.7] (also see: Poisson-Boltzmann equation, theory) Booth eq. (prim, electroviscous effect); [IV.6.9.16] Born-Bjerrum equation (solvation); [1.5.3.4] Born equations (solvation); 1.5.3b, II.3.123 Born repulsion; 1.4.5 Bose-Einstein statistics; 1.3.12 Boyle point; 1.2.51, 1.2.64, II.5.6 Boyle-Gay Lussac law; 1.2.51 Bragg-Williams approximation; 1.2.62, I.3.8d, I.3.49ff, II.1.56ff, V.2.17 Bredigsols; IV.2.37 Brewster's angle; 1.7.74, II.2.51, V.fig. 6.4 Brewster angle microscopy; III.table 3.5, III.fig. 3.56 bridging; V.chapter 1, especially V.8.1.11 V.8.73 bridging (foam destruction); V.7.32 Brillouin lines: 1.7.44 Bronsted (acids, bases); 1.5.65, II.2.7. IV.4.2b Brownian motion: 1.3.34, 1.4.2, 1.6.3a. 1.6.3d, Lapp, l i e , IV.2.7, IV.4.2b in a force field; 1.6.3b, IV.4.3b
SUBJECT INDEX
Brownian motion (continued), rotational; 1.6.73 BSA = bovine serum albumine; see albumine brushes (at interfaces); III.3.4J, IV.4.5, V.I.11, V.fig. 1.28, 29, V.l.llg, V.2.3c bubbles; see air bubbles Burgers element; III.3.129 Cabannes factor; 1.7.53 Cabosil; see silica Cahn electrobalance; III. 1.44 Cahn-Hilliard theory (for interfacial tension); HI.2.6, V. 1.7 calcium carbonate (structure factor); IV.fig. 5.33 calomel reference electrode; 1.5.85 calorimetry (adsorption); 11.1.3c, Il.fig. 1.7, II. 1.29, II.2.5b, II.5.60 canal surface viscometer; III.3.183-184 capacitance, electric; 1.4.51, 1.5.13, II.3.7c, II.3.94, II.3.106 differential; I.5.13(def.), 1.5.15, 1.5.100, II.3.10, II.3.21, Il.fig. 3.5, II.3.29, Il.fig. 3.9, II.3.33, II.3.36, II.3.6c, Il.fig. 3.22, Il.figs. 3.42-43, Il.fig. 3.49, Il.figs. 3.50-51, Il.fig. 3.53, II.3.149, II.5.60-61 integral; I.5.13(def.), 1.5.15, 1.5.59, II.3.10, II.3.6c capillaries (electrokinetics in), electrokinetic velocity profile; Il.figs. 4.15-16 electro-osmosis; II.4.21ff electrophoresis; II.4.132 streaming current; II.4.3d streaming potential; II.4.3d capillary bridges; III. 1.49, III. 1.84, III.5.11d capillary condensation; II.1.42, II.1.6, Il.figs. 1.32-33, Il.fig. 1.35, Il.fig. 1.39 capillary depression; I.1.8ff, I.fig. 1.1, III.fig. 1.4b, III.5.4e capillary electrometer; II.3.139ff, Il.fig. 3.47, III. 1.20 capillary length; [III. 1.3.3], Ill.table 1.1 capillary number; [III.5.8.1], also see: [V.8.2.3] capillary osmosis; II.4.9 capillary phenomena (general); 1.1.3, 1.2.23, II.1.6d, II.1.6e, III.1.1, III.1.1, III.1.2 capillary pressure; I.1.8ff, I.fig. 1.9, I.fig. 1.10, 1.2.23, II. 1.6, V.6.25, V.7.5, V.8.2 Young and Laplace's law; I.1.9(intr.), 1.1.12, 1.1.15, 1.2.23b, especially [1.2.23.19], II.1.85ff, II.1.99, III.1.1, [III.1.1.2] capillary rise; 1.1.2, I.1.8ff, I.fig. 1.1, II.1.6e, III.1.3, Ill.flg. 1.4a, III.1.83, III.5.4e capillary waves; III.2.9c, III.3.6g, III.3.10 capture efficiency (orthokinetic); IV.fig. 4.18
11
12
SUBJECT INDEX
carrier wave; 1.7.38 carbon, graphite, AFM image; II.fig. 1.3 adsorption of: benzene, n-hexane; II.fig. 1.8 carbon tetrachloride; II.figs. 1.22-23 hexane + hexadecane; II.fig. 2.20 krypton; II.fig. 1.29 long alkanes from n-heptane; II.figs. 2.28-29 n-heptane-cyclohexane mixture; II.fig. 2.18 octadecanol; Il.flg. 2.17 pentane + decane; II.fig. 2.20 rubber; Il.fig. 5.31 water vapour; Il.fig. 1.11, Il.fig. 1.28 soot; IV. 1.3 immersion (= wetting) enthalpy; II.table 1.3 carboxymethyl cellulose solutions, viscosity; IV.fig. 6.35, V.fig. 2.33 Carnahan-Starling equation [1.3.9.31], [IV.5.4.14] casein (fj, K ); IV.fig. 5.18, V.figs. 3.24-27, V.fig. 8.19, V.fig. 8.20 Casimir-Polder equation (for retarded Van der Waals forces); [1.4.6.35], [1.4.7.9] Cassie equation; [III.5.5.2] caterpillar trough; III.fig. 3.74 cation exchange capacity; 1.5.99, II.3.165ff cationic surfactants, see surfactants CBF = common black film; see films, liquid c.c.c. = coagulation, critical concentration, see colloid stability CD = circular dichroism; see dichroism c.e.c. = cation exchange capacity cell (galvanic); see galvanic cells centrifugation potential (gradient); II.table 4.4, II.4.6-7, IV.2.54 ceramics; IV. 1.6, IV.3.185 Hamaker constants, IV.app. 3 chain crystallization; V.8.6 chain statistics; II.5; V.I, V.2.3 Chandrasekhar equation; [1.6.3.20] Chapman-Kolmogorov equation; [1.6.3.13], [IV.4.2.5] characteristic curve (in potential theory for gas adsorption); II. 1.74 characteristic functions (in statistical thermodynamics); 1.3.3, [1.3.3.8], III.table 3.2 (in Langmuir monolayers), III.fig. 3.14 charge (electric); 1.5.3, 1.5.9, I.flg. 5.1 (also see: double layer, surface charge density, space charge (density))
SUBJECT INDEX
charge-determining ions; 1.5.5b, II.3.7, 0.3.8, II.3.84ff, II.3.89, II.3.147ff charge reversal; II.3.62, see further overcharging charged (colloidal) particles; II.chapter 3, II.chapter 4 concentration polarization; II.3.206, II.3.13c, II.4.6c, [II.4.6.53], II.4.8, III.4.8.22] contribution to conductivity; II.figs. 4.37-39 contribution to dielectric permittivity; II.figs. 4.37-39 far fields; II.3.207, II.3.13b, II.3.211, II.3.217, II.4.18-19, II.4.6, II.4.8 fluxes; II.3.215ff, II.4.6, II.4.8 (in) alternating fields; II.4.8 interaction; see interaction between colloids induced dipole moment; II.3.206, II.3.210, II.3.212ff, II.4.8 local equilibrium; II.3.213, II.4.79 near field; 11.3.211,11.4.6 polarization field; II.3.207, II.3.209, II.3.211ff, II.4.18-19, II.4.70, II.4.87, II.4.8 polarization in external field; II.3.13, Il.fig. 3.86, Il.fig. 3.88, II.4.3a, Il.fig. 4.2, II.4.18ff, II.4.6, II.4.8 relaxation; II.3.13d, Il.fig. 3.89, II.4.6c, II.4.8 charging of double layers; I.5.17ff, 1.5.7, II.3.5, IV.3.1, IV.3.2, V.2.3, V.3.8ff (also see: under double layer, electric: Gibbs energy; for specific examples see under the chemical name) charging parameter; 1.5.17, 1.5.106 cheese rheology; IV.6.15, IV.6.19 chemical potential (intr.); 1.2.1 Iff, 1.2.35 dependence on curvature; 1.2.23c dependence on pressure; I.2.41ff dependence on temperature; 1.2.40 (of) polymers; II.5.9 chemisorption; II. 1.6, II. 1.32, II. 1.18, Il.fig. 1.9, II.2.85 chirality; 1.7.100, III.3.216 cholesterol monolayers; III.fig. 3.13, III.3.8d, Hl.figs. 3.92-93 chromatography; II.2.47ff, II.2.88 eluate; II.2.48 field flow fractionation; IV.2.61ff high performance liquid (HPLC); II.2.47 retention volume; II.2.48 chymotrypsin; V.fig. 3.29 c.i.p. = common intersection point circular dichroism (CD); see dichroism circular polarization; see electromagnetic radiation, polarization
13
14
SUBJECT INDEX
Clapeyron equation chemical equilibrium; [1.2.21.11 and 12] gas adsorption; [1.3.39] solubility; [1.2.20.6] in pores; II. 1.99 two-dimensional; III.3.38 Clausius-Mosotti equation (for polarization of a gas); [1.4.4.10] clay minerals (general); 1.5.99 cation exchange capacity; 1.5.99, II.3.165ff double layer; H.fig. 3.1c, II.3.8, II.3.10d electrokinetics; II.3.168 isomorphic substitution; II.3.2, II.3.165 structures; II.3.163ff, D.figs. 3.66-67 swelling; II.3.163ff cleaving of solid surfaces; 1.2.99 closure relations; 1.3.69, IV.5.3d cloud point; V.4.11 cloud seeding; II.3.130 CLSM = confocal laser scanning microscopy; see CSLM cluster integral; 1.3.65 c.m.c. = critical micellization concentration; see micellization coacervation; IV.5.95ff, IV.fig. 5.64 coagulation; I.1.6(intr.), 1.1.7, 1.1.28, 1.4.7, 1.7.61, IV.3.9 also see; colloid stability coagulation (flocculation) kinetics; IV.2.2d, IV.flgs. 3.65-66, IV.chapter 4, V.I.85 critical concentration; II.3.129ff, IV. 1.11 fractal formation; IV.4.5 irregular series; II.3.62; see further, overcharging orthokinetic; IV.5b particle size effects; IV.4.32 perikinetic (def.); IV.4.37 rapid; IV.4.3a slow; IV. 4.3b surface roughness effects; IV.4.32 coalescence, (emulsions); V.8.3b, V.8.53ff, V.8.64 partial; V.8.73 coalescence (foams); V.7.10 coherence (of radiation, electromagnetic waves); 1.7.15, I.7.22ff, 1.7.69 coherence time; 1.7.22 coherent neutron scattering; 1.7.70 cohesion (in liquids); 1.4.5c
SUBJECT INDEX
cohesion (work of); 1.4.47, III.5.2, III.fig. 5.9 cohesion pressure; 1.3.69 cohesion (or cohesive) energy; 1.4.46 coil (of polymer molecules); 1.1.26, I.fig. 1.17, II.5.2 co-injection (foaming); V.7.13 co-ions; I.1.21(def.) Cole-Cole diagram; 1.4.35, Il.fig. 3.30b collapse (monolayers) (def.); III.3.23, III.fig. 3.46, III.3.226 collector (in flotation); 1.1.25, III.5.97 colligative properties (intr.); I.2.20f collision broadening (spectral lines); 1.7.22 colloids (general); I.1.5ff(def.), 1.1.2, IV.chapter 1, Volume IV (particulate colloids), Volume V (hydrophilic colloids) characterization; IV.2.3 colour; IV.2.39 light scattering; IV.2.3b microscopies; IV.2.3a, IV.2.41ff sedimentation; IV.2.3d; see further separate entry surface area; III.3.131ff, IV.2.3c concentrated; IV.chapter 5 electron micrographs; Il.fig. 1.1, IV.figs. 2.1, 2.2 and 2.4 emulsions; V.chapter 8 in external fields; II.3.13, Il.chapter 4 fractionation; IV.2.2h history; IV. 1.4 hydrophilic; I.1.7(def.); IV. 1.11 (def.), Volume V (general) hydrophobic; I.1.7(def.); IV.1.11 (def.), Volume IV (general) stability; 1.1.6,1.1.22ff (also see: colloid stability, general) interaction between colloids and macrobodies, interaction curves; I.fig. 6.2, IV.chapter 3, IV.chapter 4, IV.chapter 5 atomic force microscopy; II.1.13, IV.3.12 constant charge vs. constant potential; 1.5.108, IV.3.8ff, IV.fig. 3.1 density correlation functions: Lapp, l i e , IV.chapter 5 depletion; IV.3.10 Deryagin approximation; I.4.60ff, IV.3.2, IV.3.7c Deryagin-Landau-Verwey-Overbeek (DLVO) theory; IV.chapter 3 Deryagin-Landau-Verwey-Overbeek extended (DLVOE) theory; IV.3.9 in films; V.6.5 disjoining pressure: [V.I.1.2 and 3], IV.chapter 3 dynamics; IV.chapter 4, IV.4.3 (def.)
15
16
SUBJECT INDEX
colloids, interaction between colloids and macrobodies, interaction curves (continued), electric; IV.chapter 3 comparison of models; IV.3.7f diffuse, constant charge; IV.3.4 diffuse, constant potential; IV.3.3 Gouy-Stern layers, regulation; IV.3.5, IV.figs. 3.24-28, IV.3.9d spherical double layers; IV.3.7, IV.fig. 3.30 energy barrier; IV.3.9, IV.fig. 4.4 external field; IV.3.10, IV.4.5, IV.5.30 forced interaction; IV.3.6, IV.3.10, IV.4.3, IV.4.5b, IV.5.3c hetero-interaction; IV.3.4, IV.3.6 hydrodynamics (influence); IV.4.5b kinetics; IV.chapter 4 induction; IV.fig. 3.21 linear superposition approximation (LSA); IV.3.12 Maxwell stress; IV.3.22 measurement; 1.4.8, Il.figs. 2.2-3, II.3.56ff, IV.3.12 nanoparticles; IV.3.81 orthokinetic; IV.4.5b primary minimum; I.fig. 4.2, IV.3.9 regulation; IV.3.4, IV.3.5 rheological consequences; IV.6.13 secondary minimum; I.fig. 4.2, IV.3.9 simplified models; IV.5.2c, IV.5.4, IV.5.5, IV.5.6a, IV.5.81 solvent structure-mediated; 1.5.3, 1.5.4, IV.3.5, IV.3.8c surface force measurements; IV.3.12 surface roughness; IV.3.82ff tabulation of electrical interactions; IV.app. 2 thermodynamics; IV.3.2, V.I.I timescales; IV.4. Iff two-dimensional; III.3.241 virial approach; 1.7.8b, IV.chapter 5 irreversible = hydrophic, lyophobic mills; IV.2.2g mixtures; IV.5.7c, IV.5.8C preparation; 1.1.6, 1.2.100, IV.chapter 2 (general) by comminution; IV.2.29 by condensation = by precipitation
SUBJECT INDEX
colloids, preparation (continued), by precipitation; V.2.2a (homogeneous), IV.2.2b (kinetics), IV.2.2c, IV.2.2f (heterogeneous) dlspersity; IV.2.2d examples of sol preparations; IV.2.4 fractionation; IV.2.2h, IV.2.54ff also see separate entry nucleatlon and growth; IV.2.2b, IV.2.2f particle growth; IV.2.2d size control; IV.2.2a size distributions; see separate entry sol-gel processing; IV.2.2J reversible = hydrophilic, lyophilic solvent structure contribution; II.figs. 2.2-3, II.2.10, III.3.8c solubility; IV.2.2e wetting; III.5.4h (also see: particles, charged (colloidal) particles) colloid stability, stabilization; IV.chapter 3, IV.chapter 4 adsorption versus depletion; IV.3.2, V.I.10 bridging; V.1.6b, V.fig. 1.6 brushes; V.I. 11, V.fig. 1.29, V . I . l l g by (bio)polymers (including steric stabilization); I.1.2(intr.), 1.1.7, I.flg. 1.18, I.1.27ff, I.fig. 2.11, II.5.96ff, IV. 1.3, IV. 1.4, IV.fig. 1.2, V.chapter 1 (general), V.6.5c by polyelectrolytes; V.2.7 by surfactants; 1.1.25 case studies; IV.3.13 concentration profiles; V. 1.5 critical coagulation concentration (c.c.c); IV.3.98, IV.3.9e, IV.table 3.2 depletion (flocculation); IV. 1.5, V.1.8, V.1.9, V . l . l l h disjoining pressure; IV.chapter 3, V.1.6, V.1.7 DLVO theory; I.1.21ff (intr.), 1.3.59, IV.1.14, IV.chapter 3 DLVOE theory; IV.chapter 3, esp. 3.9 emulsions; V.8.1g equilibrium aspects; V.I, V. 1.12 flocculation kinetics; V.1.12e general; I.1.6ff(intr.), I.1.21ff, I.flg. 1.14, 1.2.71, I.fig. 2.11, 1.4.8, IV.chapter 3 gravity influence; IV.3.10a Gibbs energy; IV.chapter 3, V.l.S, V.figs. 1.7-13, V.1.7, V.1.8b-d, V.1.9c, V . l . l l f grand potential; V.I.I, V.l.3-4, V.1.6 Helmholtz energy; V.I.I, V.1.4, V.1.10, V.1.57ff, V.1.65ff
17
18
SUBJECT INDEX
colloid stability, stabilization (continued), influence electric field; IV.3.10b irregular series; see overcharging lyotropic series; see separate entry magnetic forces; IV.3.10c measurement; IV.3.102ff, IV.3.I2 mushroom interaction; V . l . l l f nonaqueous media; IV.3.11 orthokinetic coagulation, flocculation; IV.4.5b, V. 1.84 Ostwald ripening; IV.2.2e, V.8.3b rheological consequences; IV.6.13 and point of zero charge; II.3.106 Schulze-Hardy rule; IV.3.9e tethered; V.I. 11 (also see: colloids interaction, Van der Waals interaction) colloid titration, calorimetric; II.3.98 conductometric; II.3.88, Il.fig. 3.20 polyelectrolytes; V.2.2d potentiometric; I.5.100ff, I.fig. 5.17, II.3.7, II.3.85, Il.fig. 3.29, II.3.151, Il.figs. 3.57-59, II.5.60, IV.fig. 3.75 proteins; V.3.5ff, V.figs. 3.15-17 colloid vibration potential; Il.table 4.4, II.4.7, II.4.3e, II.4.5d colloidal dispersion; IV. 1.9 (def.) comminution of big particles; IV.2.2g common intersection point (in colloid titration curves); II.3.8a, Il.fig. 3.34, Il.figs. 3.57-59, Il.figs. 3.63-64, Il.fig. 3.77, Il.fig. 3.80, II.3.206 complex coacervate micelles; V.2.6f, V.fig. 2.39 complex coacervation; IV.5.95, V.2.6c, V.figs. 2.34-36 complex quantities; Lapp. 8 compliance; see (interfacial) rheology composition law (polymer adsorption); II.5.40 compositional ripening (emulsions); V.8.71 compressibility; 1.7.46, [III.2.11.4] Ornstein-Zernike equation; [1.3.9.32], [IV.5.2.7] two-dimensional; see interfacial rheology compression; IV.6.2 compression modulus; IV.6.6 concentrated polymer regime; II.5.9, Il.fig. 5.3, IV.6.11, IV.6.12 concentration profiles (ions); Il.fig. 3.8, Il.fig. 3.20, III.3.4h, IV.fig. 3.1 concentration profiles (polym. ads.); V. 1.5
SUBJECT INDEX
condensation, counterions; V.2.2a condensation, homogeneous, 1.2.23d condensation (method for preparing colloids); IV. 1.2, IV.2.2 condensation, (two-dimensional); 1.3.43, Il.figs 1.3.5-8, I.3.47ff, I.3.53ff, II.1.59ff, Il.fig. 1.20, Il.figs. 1.31-33, II.fig. 1.35, Il.fig. 1.39, Il.fig. 1.42, II.2.66 conduction (electrolytes); 1.6.6 (also see: surface conductance, surface conductivity) conduction bond (solids); Il.fig. 3.68, II.3.173 conductivity (electrolytes), limiting; 1.6.6a, I.table 6.5 molar; 1.6.6a (also see: surface conductivity) conductivity of, capillaries and plugs; II.4.55ff, II.4.7, Il.fig. 4.34 colloids: a.c. measurements; 11.4.5e, Il.figs. 4.21-22, 11.4.8, il.figs. 4.37-39, IV.fig. 4.16 colloids (non-aqueous); IV.3.134 Ions and ionic solutions; 1.5.51, 1.6.6a, I.table 6.5, I.6.6b microemulsions; V.5.3f polyelectrolytes; V.2.5b, V.2.5c thin films; V.6.2g, V.fig. 6.37 water; 1.5.43 conductometric titration of colloids and polyelectrolytes; see colloid titration configurations; 1.3.11, I.3.29ff, II.5.2, V.2.3 also see; polymer adsorption configuration integrals; I.3.9a, [1.3.9.6], IV.chapter 5 configurational energy; 1.3.46 configurational entropy; 1.2.52, 1.3.30, V.3.2a confocal laser scanning microscopy; 1.7.91, IV.5.92 conformation; II.5.1, see under polymers, polyelectrolytes, proteins congruence (adsorption from binary mixtures); II.2.3e charge; II.3.198 electrosorption; II.3.198, Il.fig. 3.81 pH; II.3.155, II.3.198 temperature; II.2.27, II.3.156 conjugate acid (def.); 1.5.65 conjugate base (def); 1.5.65 conjugate force; see force conservation (of energy); see energy conservation (of momentum); 1.6.1b, IV.6.4
19
20
SUBJECT INDEX
conservation laws; see hydrodynamics conservative force (def.); 1.4.1 consistency test (adsorption from binary mixtures); II.2.3e, II.fig. 2.19 contact angle; 1.1.3, 1.1.8, I.fig. 1.1, Il.figs. 1.40-41, Ill.fig. 1.1, Ill.chapter 5 and enthalpy of wetting; II. 1.29, III.5.2 data; Ill.app. 4 heterogeneous precipitation; IV.2.2f hysteresis; Ill.fig. 1.20, III.1.41ff, III.5.5, III.5.4, Ill.fig. 5.4, III.5.9-10, III.5.40 measurement (general); III.5.4, (in films) V.6.2e, V.6.3e captive bubbles; III.5.4b capillary rise/depression; III.5.4e fibers; III.5.4g films, liquid; V.6.2e, V.6.3e, V.figs. 6.18-19, V.fig. 6.38 individual particles; III.5.4h objects in interface; III.5.4c powders, porous materials; III.5.4i pressure compensation; III.5.50 sessile drops; III.5.4b spinning drop; III. 1.53 tilted plates; III.5.4d contact angle, advancing; see hysteresis contact angle, dynamics; III.5.8 contact angle, interpretation; III.5.7 contact angle, receding; see hysteresis continuity equations; 1.6.la, IV.6.1, IV.6.2 contrast matching (in neutron scattering); 1.7.70 convection; 1.6.37, V.8.75 convective diffusion; 1.6.7c convolution; I.A10.3 co-operativity; II. 1.48 coordination number; 1.3.45, 1.4.46, 1.5.3c copolymers in microemulsions; V.5.6e copper phtalocyanine pigment; IV.fig. 1.5 cordierite; IV.fig. 2.2a core (of micelles) = interior part corona (of micelles) = exterior part correlation coefficient; I.3.9e correlation function; IV.5.2a direct; IV.5.16 pair; 1.3.66, II.3.5Iff
SUBJECT INDEX
correlation function (continued), time (-dependent), 1.6.31, I.7.6c, I.7.6d, 1.7.7, Lapp. 11, II.2.14, IV.2.46ff total; [1.3.9.23], II.3.51ff, IV.5.6ff, IV.5.16 (various examples in chapter IV. 5) correlation length; 1.7.46, II.1.94, II.5.11, III.2.27, IV.4.4, V.5.39, V.5.3h correlation time, rotational; 1.5.44 correlator; 1.7.6c, I.7.6d corresponding states; III.2.51, III.2.53, V.5.19ff, V.5.55 corrosion inhibition; II.3.224 Cotton-Mouton effect; 1.7.100 Cottrell equations; [1.6.5.20, 11.21], I.fig. 6.15a Couette viscometers/rheometers; IV.6.7b Coulomb's law, Coulomb interactions; [1.4.3.1], 1.4.38, 1.5.11, 1.5.16, 1.5.17, 1.5.21, II.3.36, II.3.48ff countercharge; I.1.20(def.), II.3.2, II.3.7 see, electric double layers counterions; I.1.21(def.), see, double layer (ionic components of charge), lyotropic sequences, specific adsorption coupling parameter (Kirkwood); 1.3.68 copper phtalocyanate; II.table 1.3, IV.fig. 1.5 creep flow; 1.6.45, IV.6.6b for interfacial creep, see interfacial rheology critical coagulation concentration (c.c.c); see colloid stability critical micellization concentration; see micellization critical opalescence; 1.3.37, 1.3.69, I.7.7c, IV.2.41 interfacial; 1.7.83 critical point or critical temperature; IV.fig. 2.3, IV.fig. 5.41, IV.fig. 5.43, IV.5.7a, IV.fig. 5.62 for polymer demixing; II.5.12 in pores; Il.fig. 1.39 two-dimensional; I.3.49ff, 1.3.53, II. 1.109 critical radius (nucleation); 1.2.101, IV.2.2b cross coefficients (in irreversible thermodynamics); 1.6.12, 1.6.2 cross differentiation (principles); 1.2.14c cross-section (molecular); see surface area cryogenic transmission electron microscopy; IV.2.42-43 crystal defects (in semiconductor); II.3.172 crystal growth; II.5.97, IV.2.2 crystallization (of concentrated colloids); IV.5.8a
21
22
SUBJECT INDEX
CSLM = CLSM = confocal laser scanning microscopy; 1.7.91 Curie temperature; IV.3.124 curvature (of interfaces); I.2.23a.III.1.4ff, III.1.17, III.1.15, V.4.7c, V.5.24ff, V.fig. 5.26, V.fig. 5.27b influence on chemical potential; 1.2.23c mathematical description; III. 1.78 radius of; 1.2.23a, figs. 1.2.14-15, III.1.4ff, III.1.2 spontaneous; III.1.78, V.5.25, V.fig. 5.14 (also see: bending moment, bending modulus) cut-off length (gel); IV.4.49 CVP = colloid vibration potential Dalton's law; [1.2.17.2] damping (of oscillations); I.4.37ff see further interfacial rheology, wave damping dashpot (and spring); III.fig. 3.50, IV.6.6, see further (interfacial) rheology; Maxwell element and Kelvin (or Voigt) element; Darcy's law; see d'Arcy's law Davies equation (for ionic activity coefficient); [1.5.2.28] Deborah number; 1.2.6, 1.2.86,1.5.77,1.6.2, II. 1.8, III. 1.32, III. 1.35, III.3.5, III.3.12, HI.3.90, III.4.62, IV.4.33ff, IV.6.16-17, [IV.6.4.3] De Broglie wavelength; 1.3.23, 1.7.24 Debye equation (for polarization of gases); [1.4.4.8] Debye-Falkenhagen effect; 1.5.60, 1.6.6c, II.4.111 Debye-Hiickel approximation (intr.); 1.5.19 Debye-Hiickel limiting law; [1.5.2.22] Debye-Huckel theory, for pair interaction; IV.3.3d, IV.3.68 for polyelectrolytes; V.2.2d for strong electrolytes; 1.5.2, 1.6.6b Debye length; [1.5.2.10], I.5.19(def.), I.table 5.2, II.3.19, [II.3.5.7]ff, [II.3.10.22] Debye-Van der Waals forces; see Van der Waals forces Debye relaxation; 1.6.73 decomposition; see demixing deep channel surface shear viscometer; III.fig. 3.70 defoaming; IV.7.6 deformation (in rheology); IV.6.1, IV.6.2 degeneracy; 1.3.4 degree of dissociation; 1.5.30, II.3.76, II.5.56, V.2.2d degrees of freedom (intr.); 1.2.36 de-inking; III.5.102 delayed (elastic) recovery; IV.fig. 6.12
SUBJECT INDEX
23
delta formation (relation to colloid stability); 1.1.1, 1.1.2, 1.1.7, IV. 1.2, IV.3.184 demixing; 1.2.19, IH.6e, IV.fig. 2.3, IV.5.7a binodal; 1.2.68, Il.fig. 5.4 critical; 1.2.19, II.1.6e spinodal; 1.2.68, Il.fig. 5.4 two-dimensional; III.3.4e density correlation functions; Lapp, l i b , Lapp, l i e density functional; [III.2.5.18), III.2.34, Ill.app. 3 density profiles, liquid-fluid interfaces; III.2.4, III.2.5, [III.2.5.31], III.fig. 2.6, Ill.fig. 3.29 see also concentration profiles, distribution functions (of liquids near solids and of fluid interfaces), adsorption of polymers depletion (adsorption) = adsorption, negative layer thickness; V.1.8, V.I.9 depletion interaction; IV.3.10, IV.5.79ff, V.1.8, V.1.9, V.8.72ff depolarization (of polarized interface); II.3.137 deposition; IV.4.3, IV.4.43-44 depolarization ratio; 1.7.54 Deryagin approximation (to compute interactions between non-flat colloids); 1.4.61, I.fig. 4.13, 1.4.64, IV.3.2, IV.3.7c Deryagin-Landau-Verwey-Overbeek (DLVO) theory; see colloids, interaction and colloid stability desalination; 1.1.3 (also see: salt-sieving) Descartes' law = Snell's law desorption; I.1.5(def.) detectors (for radiation); 1.7.1c quadratic; I.7.36ff detergency; III.5.101 dewetting; III.5.4, III.5.10, V.fig. 5.30, V.fig. 7.14 dextrane, adsorption on silver iodide; Il.fig. 5.26b, II.5.80ff, Il.fig. 5.29 DFG = difference frequency generation; III.3.7c.v dialysate; I.5.86ff dialysis; IV.2.32 diamagnetism; IV.3.124 dichroism; 1.7.98, II.2.56 circular (CD); 1.7.99, V.fig. 3.28, see chapter V.3 (general) dielectric displacement; I.4.5f, 1.7.9, IV.3.124 dielectric dispersion of sols, low and high frequency; II.3.219, Il.fig. 3.89. II.4.1 10
24
SUBJECT INDEX
dielectric dispersion of sols (continued), measurements; II.4.5e, II.fig. 4.21 theory; II.4.8, IV.4.5a, IV.figs. 4.16-17 dielectric drag; 1.5.51 dielectric increment, of colloids; II.4.8, Il.fig. 4.37-39, IV.figs. 4.16-17 of ions; I.table 5.10 dielectric permittivity (dielectric constant); 1.4.10, I.4.4e, 1.4.5a, I.4.5f, 1.5.11, I.table 5.1, I.5.3e, 1.7.2,1.7.6 complex formalism; I.4.4e, 1.7.2c, Lapp. 8 emulsions; V.8.19ff measurement; 1.4.24, I.4.5f, II.4.5e, ILfigs. 4.21-22 relation to polarization; I.4.23ff relation to refractive index; 1.7.12 dielectric polarization; see polarization dielectric relaxation; I.4.4e, I.4.5f, II.4.8 dielectric saturation; 1.5.11 dielectrophoresis; II.4.51 differential scanning spectroscopy; V.3.26, V.fig. 3.28 diffraction, principles; 1.7.13, 1.7.24 diffraction colours; IV.2.40 diffuse charge, diffuse double layer; see double layer, diffuse (also see: surface charge) diffuse transmission spectroscopy; V.7.23 diffusing wave spectroscopy, IV.4.9, V.7.23 diffusion (coefficient); 1.6.3, 1.6.5, 1.7.15, Lapp. H e along surface; I.6.5g, II.2.14, II.2.29, HI.3.74 and correlation functions; Lapp. 11 and irreversible thermodynamics; 1.6.5a collective; 1.6.55, 1.7.15, Lapp, l i e colloids; IV.4.1ff, IV.4.2 concentrated sols; 1.7.66, 1.7.15, Lapp, l i e convective; 1.6.7c data; I.table 6.4 forced; 1.6.53, 1.6.7 hydrodynamic correction; 1.6.56, Lapp, l i e in condensed media; 1.6.56 in films; V.fig. 6.45 in gases; 1.6.55 in water; I.5.44ff model interpretation; 1.6.5b
SUBJECT INDEX
diffusion (coefficient) (continued). non-linear geometry. I.6.5f non-spherical particles; I.6.69ff, I.fig. 6.19 of colloids from dynamic light scattering; 1.7.8b, 1.7.8c, I.7.8d, 1.7.15 rotational; 1.5.44, 1.6.20, 1.6.53, I.6.70ff, 1.7.8c, 1.7.59 self; 1.5.44, 1.6.53, 1.7.15, Lapp, l i e semi-infinite; 1.6.59 thermal; 1.7.44, 1.7.48 to/from (almost) flat surface; I.6.5d, II.2.8, II.4.6c, II.4.8b to growing particles; IV.2.2c diffusion-controlled particle growth; IV.2.2c diffusion equation (theory for polymer adsorption); II.5.32ff diffusion impedance; II.3.96 diffusion layer; 1.6.63, 1.6.68 diffusion-limited aggregation (DLA); IV.2.20ff, IV.4.5 diffusion-limited cluster aggregation (DLCA); IV.4.45, IV.4.48, IV.fig. 4.25 diffusion potentials; I.5.5d, II.4.125 also see: potential difference diffusion relaxation; II.3.13, II.3.219, II.4.6, II.4.8 diffusiophoresis; 1.6.91, II.3.214, II.4.9 dilatant, dilatancy (rheology); IV.6.11 dilatometry (and surface excesses); II.2.7 dilational modulus, interfacial; see interfacial rheology dimple formation (in draining films); V.6.39ff, V.figs. 6.21-23 dipole field; 1.4.4b, II.3.13, II.4.6, II.4.8 dipole moment; I.4.4b(def.), I.7.3b data for molecules; I.table 4.1 of colloids; II.3.13 dipoles; 1.4.4b ideal (point dipole); 1.4.20 induced; 1.4.22, 1.4.27. 1.7.18. I.7.93ff, II.3.13, II.4.6 of colloids; II.3.13, Il.fig. 4.1, II.4.8 oscillation; 1.7.3b, II.4.8 permanent; 1.4.22, 1.4.27, 1.7.17 Dirac delta function; I.7.40(def.) disc centrifuge; IV.2.60ff discotic fluid (2D); III.3.62 discs (surfactants); V.4.7b
25
26
SUBJECT INDEX
disjoining pressure; I.4.6(def.), II.1.22, II.fig. 1.37, II.1.95ff, II.1.101. [II.2.2.1], II.5.65, III.3.176, III.5.7, III.5.14-15, III.5.23, III.fig. 5.12, III.fig. 5.15, IV. 1.3, V.I.I, V.6.5 (also see: films, interactions; for interacting colloids, see IV. chapter 3) dispersion, dielectric; see there for preparation of colloids; IV.2.2g of colloids; 1.1.5, IV. 1.2 of refractive index; 1.4.37,1.7.13 of transverse waves; III.3.116 dispersion forces; see Van der Waals forces dispersity (of colloids); see size distributions displacement, of particles; I.6.18ff, I.6.30ff dielectric; see there dissipation; 1.2.7, 1.2.22, 1.4.3, 1.4.34, 1.6.9, 1.6.13, I.6.35ff, 1.7.2c, 1.7.14, IV.6.3 dissociation constant; I.5.2d (in) double layers; II.3.65ff, II.3.72ff, II.3.76, II.3.82ff (in) polyelectrolytes; V.2.2d (in) proteins; V.3.5ff, V.table 3.1 relation to points of zero charge; II.3.8c, II.table 3.5 dissolution, heat of; 1.2.71 dissymmetry ratio (in scattering); 1.7.58, I.fig. 7.13 distal length (polym. ads.); V.I. 18 distributions; 1.3.la, 1.3.7, IV.app. 1 most probable; 1.3.7 Poisson; [IV.2.3.46] also see Gauss distribution distribution (partition) coefficient; 1.2.69 distribution (partition) equilibrium; 1.2.20a distribution function; I.3.9d, II.3.6b, II.3.9 direcVindirect; IV.5.21 higher order; I.3.9e in electrolytes; I.5.28ff, I.fig. 5.9, I.5.57ff in fluid interfaces; III.fig. 2.1, III.2.3, III.2.4, III.2.5, III,2.24, [III.2.5.30], III.fig. 2.6, [III.2.5.40] in liquids near surfaces; II. 1.94, H.flg. 1.38, [II.2.1.2], II.2.6.8, II.figs. 2.2-3, II.2.2b, II.figs. 2.4-8, II.3.6b, II.3.9 in water; 1.5.3c, I.fig. 5.6 pair; 1.3.71, II.3.5Iff, III.2.30 relation to pair interaction; IV.3.142ff
SUBJECT INDEX
distribution function (continued), radial (or pair correlation); I.3.9d, 1.7.66, Lapp, l i e , II.3.6b, IV.5.5, IV.5.14, IV.fig. 5.4, IV.fig. 5.8, IV.fig. 5.31 singlet; 1.3.71 dividing plane; see Gibbs dividing plane DLA = diffusion-limited aggregation DLCA = diffusion-limited cluster aggregation DLVO = Deryagin-Landau-Verwey-Overbeek (theory), see colloid stability, colloids interaction DLVOE = Deryagin-Landau-Verwey-Overbeek extended (theory), see colloid stability, colloids interaction DNA, (persistence length); Il.fig. 5.5 Donnan effect; 1.1.21, 1.5.90, 1.5.93, II.3.10, II.3.26, II.3.99, IV. 1.6, IV.5.2d, V.2.4 relation to suspension effect; I.5.5f Donnan e.m.f.; 1.5.88 Donnan potential; V.2.38ff donor (in semiconductor); II.3.172ff donor number; 1.5.65 Doppler broadening (spectral lines); 1.7.22 Doppler effect, Doppler shift; 1.7.16, I.fig. 7.6, 1.7.19, 1.7.45, 1.7.94, II.4.46 Dorn effect = sedimentation potential double layer, electric; I.1.20(def.), I.fig. 1.13,1.5.3ff, I.fig. 5.1, Il.chapter 3 in apolar media; II.3.11, IV.3.11 diffuse; 1.1.21, I.fig. 1.13, 1.5.3, I.fig. 5.1, II.3.5 (in) asymmetrical electrolytes; II.3.5c capacitance; II.3.21, Il.fig. 3.5, II.3.29, Il.fig. 3.10, II.3.33, II.3.36 (in) cavity; Il.fig. 3.16 charge; II.3.21, Il.fig. 3.4, II.3.29, Il.fig. 3.9, II.3.32ff, II.3.36, II.3.37, Il.fig. 3.12, Il.table 3.1, Il.table 3.2, II.3.40, Il.fig. 3.14, III.4.4 cylindrical; II.3.5f, II.5.14ff, V.2.2b, V.2.2c electrolyte mixtures; II.3.5d field strength; II.3.20, II.3.21, II.3.29, II.3.32 Gibbs energy; II.3.23, Il.fig. 3.6, IV.3.2 Gouy-Chapman theory; see there introduction; II.3.17ff ionic components; II.3.9ff. II.3.5b, Il.fig. 3.8, II.3.33, Il.fig. 3.11, Il.fig. 3.15, II.3.168 negative adsorption of co-ions; II.3.5b, II.3.7e, Il.fig. 3.33 potential distribution; II.3.24ff, Il.fig. 3.7. II.3.35. II.3.36, Il.fig. 3.12 III.4.4. IV.3.2-3.7, IV.3.11 spherical; II.3.5e. Il.table 3.1, Il.table 3.2
27
28
SUBJECT INDEX
double layer, electric; diffuse (continued), statistical thermodynamics; II.3.6b enthalpy of formation; 1.5.108, II.3.98, II.3.155ff, II.figs. 3.60-61, II.table 3.6 entropy; 1.5.109, Il.fig. 3.44, II.table 3.6 equivalent circuit; I.fig. 5.11, 11.3.7c, Il.fig. 3.31 examples; II.3.1, Il.fig. 3.1 Gibbs energy; 1.5.7, II.3.5, II.3.9, II.3.23, Il.table 3.6, II.3.142, II.3.146 Gouy-Stern model; II.3.6c, Il.figs. 3.20-26, II.3.6f, II.3.133ff, II.3.154, II.3.158 (also see: Gouy-Chapman theory, Stern layer) heterogeneity; II.3.83ff measurements; II.3.7 moment; [II.4.6.50] in monolayers; III.3.4h, III.fig. 3.17 in non-aqueous solvents; 1.5.66, II.3.36, IV.3.11 ionic components of charge; 1.5.2, I.5.90ff, I.5.6b, 1.6.88, II.3.5b, Il.fig. 3.8, Il.fig. 3.46, Il.figs. 3.53-55, Il.fig. 3.62 (also see: double layer, diffuse) Oosawa model; V.2.2b-c origin; II.3.2, II.3.110, II.3.117, II.3.155ff, II.3.158, V.2.2a overlap; 1.2.72, II.3.24 see colloid stability polarization in external field; II.3.13 (see further: charged (colloidal) particles) polarized vs. relaxed; 1.5.5b, II.3.1, II.3.4 polyelectrolytes; V.2.2 polyelectrolytic adsorbates; II.5.5g relaxation; 1.5.5b, I.6.6b, I.6.6c, II.3.94, II.3.13d, II.4.6c, II.4.8, IV.4.4 site binding; see Stern layer statistical thermodynamics; II.3.6b Stern layer; 1.5.9, 1.5.59, II.3.17, II.3.6c, II.3.6g, II.5.5g, IV.3.9d capacitance; II.3.59ff, Il.fig. 3.43, Il.fig. 3.50 charge and potential distribution; II.3.6c, II.3.6d, Il.figs. 3.20-21, II.4.71, Il.figs. 5.17-18, II.5.5g, IV.3.9d condensation; V.2.2a, V.2.2b-c Gibbs energy; II.3.6f, Il.fig. 3.26 site binding; II.3.6e, II.3.6g specific adsorption; II.3.6d zeroth-order; II.3.59, Il.fig. 3.20a, Il.fig. 3.21 surface conduction; see there
SUBJECT INDEX
double layer, electric (continued), thermodynamics; 1.5.6, II.3.4, II.3.1 lOff, II.table 3.6, II.3.138ff, II.3.155ff, Il.figs. 3.60-61 triple layer model; II.3.61(def.), II.3.6c two-state models; V 2.2a-b, also see Oosawa model (also see: capacitance, ionic components, surface charge. For specific examples; see under the chemical name of the material.) drainage (of films and foams); I.fig. 1.7, 1.1.15, V.6.4, V.7.10, V.7.3a, V.8.55ff, V.fig. 8.18 drilling fluids; V.7.35 drilling muds; II. 1.80 drop, break-up (emulsification); V.8.34ff, fig. V.8.10 in electric field; III.1.5, Ill.fig. 1.15 pendant; 1.1.11, I.fig. 1.3 pressure relaxation; III.3.188 rheology; III.3.187, Ill.fig. 3.72 sessile; I.fig. 1.1, III.figs. 5.1-2, III.5.4b, Ill.fig. 5.19 (see interfacial tension, measurement, III.chapter 1) Drude equations; I.7.78ff, II.2.52 dry foam; V.7.1, V.fig. 7.12 DSC = differential scanning calorimetry Dukhin number; [II.3.13.1], II.3.208ff, II.4.12, II.4.30, II.4.3f, II.4.35ff, II.4.59 Dupre equation; [III.5.2.2b and 2c] DWS = diffusing wave spectroscopy dynamic light scattering; see electromagnetic radiation, scattering dynamics and rheology; IV.6.4 Edwards equation; [V. 1.4.1] Egyptian ink and paints; I.I.Iff, IV. 1.3, IV.2.1 Einstein crystal; I.3.21ff, I.3.6a Einstein equation (for diffusion); 1.6.20, 1.6.30, [1.7.8.12a], 1.7.64, 1.7.66, 1.7.15 Einstein equation (for viscosity); [1V.3.10.2], IV.6.9a efficiency (thermodynamic); 1.2.9, 1.2.22 elastic (material); IV.fig. 6.10, IV.6.14 elastic aftereffect; IV.fig. 6.12 elastic force, between brushes; V . I . l i e in gels; V.2.3d, V.figs. 2.20-21 elastic recovery; IV.fig. 6.12 elasticity modulus; IV.6.7, IV.6.14
29
30
SUBJECT INDEX
electrical birefringence; 1.7.100 electric double layer; see double layer, electric electric capacitance; see capacitance electric charge; see charge electric current (density); 1.6.6 two-dimensional; I.6.6d electric field; 1.4.3, I.4.5f, 1.5.10, 1.7.6b caused by surface charge; 1.5.11; also see Gauss equation and electromagnetic waves; I.chapter 7 electric potential; see potential electroacoustics; II.4.3e, II.4.5d, Il.fig. 4.20 electrocapillary curves; I.5.96ff, 1.5.99, I.fig. 5.16, II.3.139, Il.fig. 3.48, III.1.45 electrocapillary maximum; I.5.99ff, II.3.102, II.3.139ff, Il.fig. 3.48 electrochemical potential; I.5.1c, [1.5.1.18], 1.5.74, II.3.5, II.3.90 electrochemistry (general introduction); I.chapter 5 electrodialysis; II.4.132 electrokinetic charge; II.3.90, II.4.1, Il.fig. 4.13 electrokinetic consistency; II.4.58, II.4.6e, Il.table 4.2, II.4.6f, ll.table 4.3 electrokinetic phenomena; II.chapter 4, V.2.5 a.c. phenomena; II.4.8 advanced theory; II.4.6 applications; II.4.6i, II.4.10, II.5.63ff double layer relaxation; II.4.6c elementary theory; II.4.3, II.4.7a irreversible thermodynamics; see Onsager relations polyelectrolytes; V.2.5a survey; Il.table 4.1 techniques; II.4.5 xylene in water; III.fig. 4.21 (see further the specific electrokinetic phenomena) electrokinetic potential; I.5.75ff, 1.6.87, II.chapter 4, V.2.5 examples; Il.fig. 4.13, II.figs. 4.29-30, IV.fig. 3.64, IV.fig. 3.72 interpretation; II.4.1b, II.4.4 relation to t/ 1 ; II.4.41ff, Il.fig. 4.12 electrolytes; 1.5.1b, 1.5.2 electrocratic (colloid stability); IV. 1.2 electromagnetic radiation and waves; I.chapter 7, IV.2.3b absorption; 1.7.2c. 1.7.3, I.7.60ff secondary; 1.7.15 coherence; 1.7.15, 1.7.23 and oscillating dipoles; 1.7.3d
SUBJECT INDEX
electromagnetic radiation and waves (continued), detection; 1.7. l c in a medium; 1.7.2b, 1.7.2c intensity; 1.7.8, 1.7.33 interaction with matter; 1.7.3 in a vacuum; 1.7.la, 1.7.2a irradiance; 1.7.5,1.7.23 Maxwell equations; 1.7.2 phase shift; 1.7.3 polarization; 1.7.1a, I.fig. 7.2, I.fig. 7.4, 1.7.23, 1.7.26, I.fig. 7.8, 1.7.14 elliptical; 1.7.6, I.fig. 7.4, 1.7.98, III.3.7 circular; I.fig. 7.4 planar; I.fig. 7.4 scattering; 1.7.3 and absorption; I.7.60ff and fluctuations; 1.7.6b dynamic; 1.7.6c, I.fig. 7.10,1.7.6d, 1.7.7, 1.7.8, IV.2.46ff forced Rayleigh; 1.7.103 Guinier; IV.2.44ff inelastic; I.7.16(def.) Mie; I.7.60ff, V.8.22 of colloids; 1.7.8, II.4.46, IV.2.3b, IV.3.142ff, IV.5.21 of emulsions; V.8.17ff, fig V.8.6, V.table.8.1 of foams; V.fig. 7.2 of interfaces; 1.7.10c, III.1.10 of liquids; I.5.44ff, 1.7.7, 1.7.8a of microemulsions; IV.fig. 5.14, V.5.3d, V.5.3e plane; I.7.27(def.) quasi-elastic, QELS; 1.7.16, 1.7.6, 1.7.7, II.4.46, II.5.62 Raman = inelastic Rayleigh-Brillouin = QELS Rayleigh-Debye; I.7.8d, IV.2.44ff, V.8.17 secondary; 1.7.16
static; 1.7.33, I.7.6d, 1.7.7, 1.7.8 survey; I.table 7.3 wave vector; I.7.27(def.) (also see: neutron scattering, X-ray scattering) sources; 1.7.4 types of; I.fig. 7.1 electromotive force; 1.5.82 electronegative; 1.4.19, 1.4.48
31
32
SUBJECT INDEX
electroneutrality (electrolyte solutions); 1.5.1a; 1.5.lb electroneutrality of double layers; 1.5.4, 1.5.6a, II.3.6ff electron microscopies for sols; IV.2.42, also see the photographs in that chapter electron microscopies for microemulsions; V.5.3b electron pair acceptor; 1.5.65 electron pair donor; 1.5.65 electron spin resonance (ESR, principles); 1.7.16, 1.7.13 of aqueous electrolytes; I.5.54ff of interfaces; II.2.8, II.2.55, II.5.59 electro-osmosis; 1.6.12,1.6.16, II.4.1, Il.table 4.4, II.4.6, II.4.3b, Il.fig. 4.6, II.4.46, Il.fig. 4.15 in plug of arbitrary geometry; II.4.21-22, II.4.5b electro-osmotic dewatering; II.4.132 electro-osmotic flux; II.4.23 electro-osmotic pressure (gradient); Il.table 4.4, II.4.6, II.4.23ff electro-osmotic slip; II.4.19, II.4.21ff electro-osmotic volume flow; Il.table 4.4, II.4.6, II.4.22ff electrophoresis; II.chapter 4 advanced theory; II.4.6 anticonvectant; II.4.131 applications; II.4.10 elementary theory; II.4.3a experiments; II.4.5a polarization retardation; II.4.3a electrophoretic light scattering; II.4.46 electrophoretic mobility, velocity; II.4.4-5, Il.table 4.1, II.4.3a, Il.fig. 4.41, IV.figs. 3.623.64, IV.fig. 3.68, IV.fig. 3.74 cylindrical particles; Il.fig. 4.4, II.4.16 Dukhin-Semenikhin equation; [II.4.6.45], Il.fig. 4.29 Helmholtz-Smoluchowski equation; [II.4.3.4], II.4.12-14, II.4.17-19, Il.fig. 4.29 Henry; Il.fig. 4.4, 11.4.16 Hiickel-Onsager equation; [II.4.3.5] hydrodynamics; II.4.14ff, II.4.6 influence of surface conduction; Il.fig. 4.4, Il.fig. 4.31 irregular particles; II.4.6h. Il.fig. 4.33 measurement; II.4.5a electroacoustics; II.4.5d microelectrophoresis: II.4.45ff, II.figs. 4.14-16 moving boundary; II.4.5Iff, Il.fig. 4.17 Tiselius method; II.4.53
SUBJECT INDEX
electrophoretic mobility, velocity (continued), non-aqueous media; IV.3.135ff O'Brien-Hunter equation; [II.4.6.44] O'Brien-White; II.figs. 4.26-29 polyelectrolytes; V.2.5a ribonuclease; V.fig. 3.5 sol concentration effect; II.4.6g, II.fig. 4.32 stagnant layer thickness determination; II.4.128ff, II.fig. 4.42 verification of theories; II.4.6e electrophoretic deposition; II.4.132 electrophoretic retardation (in ionic conduction); 1.6.6b, II.4.3a electropositive; 1.4.19, 1.4.48 electrosonic amplitude; II.4.7, II.4.5d electrosorption; II.3.4(def.), II.3.12 electrostriction; 1.5.103 electroviscous effects; II.4.122ff, IV.6.9b, V.2.48ff electrowetting; IH.5.103 ellipsometric coefficients; 1.7.75, II.2.51, [II.2.5.7] ellipsometric thickness; [1.7.10.17], II.5.64 ellipsometry; I.7.10b, II.2.5c, II.5.64ff, III.2.47, Ill.table 3.5, III.3.141ff, V.fig. 6.5 elliptical polarization; see electromagnetic radiation, polarization eluate; see chromatography emission spectrum; 1.7.14 emulsification; 1.1.3, 1.6.45, III.3.237, III.4.97, V.8.2 emulsification failure boundary; V.5.21, V.5.49 emulsifier; V.8.2 emulsifier (biological); 1.1.3 emulsion films; V.6.1 (def), see films, liquid emulsions (general); V.chapter 8 emulsions; I.1.3(intr.), I.1.5(def.), 1.2.98, IV. 1.9 (def. + classif) aggregation; V.8.63, V.8.78ff characterization; V.8.1 coalescence; V.8.3e, V.8.64 creaming; V.fig. 8.25 dielectric properties; V.8.19ff drop size distribution; V.8.1e, V.8.66ff, V.fig. 8.25 formation; V.8.2, V.fig. 8.9, V.fig. 8.10, V.table 8.2 interfacial layers; V.8.2d multiple: IV. 1.9 optical properties; IV.fig. 5.63, V.8.17ff Ostwald ripening; V.8.3b, 8.63
33
34
SUBJECT INDEX
emulsions (continued), phase inversion; V.8.64 Pickering stabilization; III.5.99, V.8.4 preparation; see formation sedimentation; V.8.3d, V.8.63 stability; V.8.3, V.fig. 8.22 type; V.8.3 viscosity; V.8.15, V.8.29 (also see; microemulsions) endothermic; see process energy (principles); 1.2.4, Lapp. 3, Lapp. 4 absorption; 1.4.4e, 1.7.3 configurational; 1.3.46 conservation; 1.2.8 interfacial; 1.2.5, 1.2.11, Lapp. 5, Il.table 1.2 levels (semiconductors); II.fig. 3.68-69 mixing (polymers); [H.5.2.12] of radiation; 1.7.5 (also see: adsorption, interaction) engulfment; 111.5.11c enhanced oil recovery; 1.1.1,1.1.3,1.1.11, III. 1.84, V.7.35 ensemble (intr.); 1.3.lc canonical; 1.3. l c grand (canonical); 1.3.lc microcanonical; 1.3. l c entanglements; IV.6.67ff enthalpy (principles); 1.2.6, Lapp. 3, Lapp. 4 interfacial; 1.2.6, 1.2.11, Lapp. 5, Il.table 1.2 of chemical reactions; 1.2.21 of dissolution; 1.2.20c, 1.5.3a of electric double layer; 1.5.108, II.3.98, II.3.155ff, Il.figs. 3.60-61 of hydration; I.table 5.4 of transfer; 1.2.69 of wetting; see wetting (also see: adsorption, enthalpy) entropy (principles); I.2.8(intr.), 1.2.9, Lapp. 3, Lapp. 4 absolute; 1.2.24, 1.3.16 configurational; 1.2.52, 1.3.30 interfacial; 1.2.9. 1.2.42, 1.2.83, Lapp. 5. II. 1.2. Il.table 1.2, Il.fig. 3.44 intrinsic: 1.2.52 of electric double layers: 1.5.109, Il.fig. 3.44
SUBJECT INDEX
entropy (continued), of mixing; 1.2.53, 1.3.28, [II.5.2.11 ] of solvation (hydration); 1.5.3, I.table 5.4 production of; 1.6.2a, I.6.2b statistical interpretation; 1.3.16ff (also see: adsorption entropy) environment, double layer effects; II.3.220ff environmental scanning electron microscopy (ESEM); IV.2.42 Eotvos equation (for surface tension); [III.2.11.1 ] Eotvos number; [III. 1.3c] EPR = electron spin resonance equation of motion; 1.6. l b , IV.6.1, IV.6.2 equation of state, BET; [II. 1.5.49], Il.fig. 1.24b Boyle-Gay Lussac (ideal); [1.1.3.4], 1.2.17a Carnahan-Starling; [1.3.9.31] hard sphere fluid; [1.3.9.26] one-dimensional; [1.3.8.5] Percus-Yevick; [1.3.9.29 and 30] Van der Waals; [1.2.18.26], [1.3.9.28] two-dimensional; 1.1.17, 1.3.42, I.3.8d, II.1.3, II.1.3b, II.1.39, II.1.45, Il.app. 1, III.3.4, Ill.table 3.3, III.4.2, III.4.3 double layer; II.3.14 electrosorption; II.3.197 Frumkin-Fowler-Guggenheim (FFG); I.3.46ff, [II.A1.5b] Henry; [II.Al.lb] Hill-De Boer = Van der Waals; see there Langmuir; I.3.6d, [1.3.6.23], II. 1.45, [II. 1.5.10], Il.fig. 1.15b, [II.A1.2b] polymer monolayers; III.3.4i, [III.3.4.56] protein monolayers; V.fig. 3.24 quasi-chemical; I.3.8e, [II.A1.6b] sols; [IV.3.12.8] Van der Waals; II. 1.51, II. 1.59, [II. 1.5.28], Il.fig. 1.20b, [II.A1.7b and c], III.3.4e (two dimensional) virial; [II.A1.4b] Volmer [1.5.23], Il.fig. 1.15b, [II.A1.3b] equilibrium (general); 1.2.3, 1.2.8, 1.2.12, 1.3.7, IV.6.3a chemical; 1.2.21 frozen; 1.2.8 local; 1.6.2. 1.6.2a mechanical: 1.2.22, V.6.3b
35
36
SUBJECT INDEX
equilibrium (continued), membrane; I.2.33(def.), I.5.5f, III.3.29 metastable; 1.2.7, 1.2.68 partial; 1.2.7 restricted; V.I.12c stable; 1.2.7, 1.2.19 equilibrium constant; 1.2.77, II.3.6e equilibrium criteria; 1.2.12 equipartltion (of energy); 1.6.18 equipotential planes (in gas adsorption theory); II.fig. 1.25 equivalent circuits (electrochemistry); I.fig. 5.11, II.3.7c, Il.fig. 3.31 equivalent circuits (rheology); III.3.6i, IV.6.6 also see: interfacial rheology error function; [1.6.5.25], [II.4.6.37] Esin-Markov coefficients; I.5.6d, 1.5.102, II.3.15ff, II.3.26, II.3.103ff, II.3.136, Il.fig. 3.45, II.3.144ff ESA = electrosonic amplitude ESCA (electron spectroscopy for chemical analysis) = XP(E)S ESEM = environmental scanning electron microscopy ESR = electron spin resonance Euler-Lagrange equation; [III.2.5.25], [III.A3.9] Euler's theorem; 1.2.28, 1.2.14a evanescent waves; 1.7.75, II.1.18, II.2.54, IV.3.157 evaporation; Hl.fig. 2.16 (heat), III.2.55 (entropy) prevention (water conservation); III.3.239 EXAFS = extended X-ray absorption fine structure exchange (adsorption from binary mixtures); II.2.1, II.2.3, II.2.4 constant; [II.2.3.16] excess functions and quantities; I.2.18b(intr.) in regular solutions; 1.2.18c, I.3.46ff (also see: interfacial excesses) excimers; III.3.165 excluded volume; see polymers exothermic; see process expansion coefficient (2D); [III.3.3.2], [III.3.4.5] extended X-ray absorption fine structure (EXAFS); 1.7.88 extraction (of films); V.6.4e extrapolation length (polymer ads.); V. 1.8, V.fig. 1.2 Fabry-Perot interferometer; 1.7.36, I.7.6d falling film (interfacial rheology); III.3.189ff falling meniscus (for measuring interfacial tensions); III.1.11
SUBJECT INDEX
Faraday's gold sols; IV. 1.13, IV.2.27 fats (metabolism); 1.1.1, 1.1.3, 1.1.7 fatty acids; III.tables 3.7a, 3.7b (overview) fatty acid monolayers; III.fig. 1.32 (interfacial tension, dynamics), III.fig. 3.58, III.fig. 3.60, Ill.fig. 3.68, Ill.fig. 3.76, III.figs. 3.78-87, Ill.fig. 4.26 fatty amine monolayers; III.3.212, fig. III.3.88 FCS = fluorescence correlation specroscopy; III.3.7c.iv FEM = field emission (electron) microscopy; 1.7.lib Fermi-Dirac statistics; 1.3.12, II.3.172 Fermi (energy) level; II.3.172, II.fig. 3.68, II.3.174 ferrofluids; IV.fig. 2.2c, IV.3.10c preparation; IV.2.4d ferromagnetism; IV.3.124 FFEM = freeze fracture electron microscopy; see transmission electron microscopy FFF (sediment) = (sedimentation) field flow fractionation FFG = Frumkin-Fowler-Guggenheim; see adsorption isotherm fibres, wetting; Ill.fig. 5.2, III.5.4g fibronectin (adsorption); V.fig. 3.14 Fick's first law; I.table 6.1, 1.6.54, I.6.5d, I.6.5e, 1.6.67 two-dimensional; 1.6.69 Fick's second law; 1.6.55, I.6.5d, I.6.5e, II.4.78, II.4.115 field emission techniques; 1.7.lib field strength, electric; I.4.12ff (also see: double layer) film balance; 1.1.16, III.3.3, III.figs. 3.4, 3.5 film drainage; V.6.4, V.fig. 8.18, V.8.84 film formation; V. chapter 6, V.8.72 film tension; I.1.12ff, 1.2.5ff, II.1.95ff, V.6.3b, V.fig. 6.17 films, liquid (free); 1.1.6, I.1.12ff, I.fig. 1.4 up to and including fig. 1.11, I.fig. 2.1, 1.2.5, IV. 1.3, V.chapter 6 (general) black or common black; IV. 1.3, V.6.6, V.fig. 6.8, V.fig. 6.10, V.fig. 6.25, V.fig. 6.33, V.fig. 6.37, V.fig. 6.43, V.8.88 Brewster's angle; V.fig. 6.4 colours; 1.7.80 conductivity; V.6.2g, V.fig. 6.37 contact angle; V.6.2e, V.6.3e, V.fig. 6.38 diffusion; V.6.7 disjoining pressure; V.6.2, V.6.2d, V.6.25, V.fig. 6.27, V.6.5, V.figs. 6.31-32, V.6.6a, V.fig. 6.34 elasticity; V.6.2f ellipsometry; V.fig. 6.4
37
38
SUBJECT INDEX
films, liquid (free) (continued), emulsion; V.6.2c, V.fig. 6.9 experiments; V.6.2 flow in them, drainage; 1.1.15, I.6.4d, IV. 1.3, V.6.4 FTIR; V.fig. 6.6 interferometry; V.6.2a macroscopic; V.6.2b, V.6.4e microscopic; V.6.2c multiple reflections; 1.7.80 Newton (black); III.5.39, IV. 1.3, V.6.2, V.fig. 6.33, V.6.6b, V.6.6c, V.figs. 6.43-44 oscillatory forces; see stratification reflectometry; V.fig. 6.4-6.5 permeability; V.6.2b, V.6.7a-b, V.figs. 6.43-44 phospholipid; V.6.7c pinch-off (in foam breaking); V.7.32 rupture; V.6.4c, V.6.4d, V.6.3b, V.6.6c, V.7.3b, V.7.6, also see; stability critical thickness; V.6.4c, V.fig. 6.23, V.fig. 6.25 stability; 1.6.45, V.7.3, V.fig. 7.14, V.8.81ff, V.8.85ff, V.fig. 8.26 Kabalnov-Wennerstrom theory; V.8.86 Vrij-Scheludko theory; V.6.4c, V.8.87 de Vries theory; V.8.85 stratification; V.6.2, V.6.5d thermodynamics; V.6.3 thickness; V.6.2a (def.), V.6.3a,b, V.6.4, V.fig. 6.28 thinning; III.5.3d Van der Waals forces; 1.4.72, III.5.3 X-ray; V.6.9 also see; foams films, liquid (on solid supports), disjoining pressure; II. 1.22, [II.1.3.15], II.1.6d, [II.1.6.18], III.fig. 1.14, III.5.3 (also see; wetting films) filtration; IV.2.32ff filtration law; [IV.2.2.66] FIM = field-ion microscopy; 1.7.lib fire fighting; V.7.35-36 first curvature (of interfaces), see mean curvature First Law of thermodynamics; see thermodynamics First Postulate of statistical thermodynamics; see statistical thermodynamics fish diagrams; see microemulsions flatbands (semiconductors); H.3.174, Il.fig. 3.69
SUBJECT INDEX
flickering clusters; 1.5.42 FLIM = fluorescence lifetime imaging microscopy floating (objects on liquids); I.l.lOff, I.fig. 1.2, I.fig. 1.4 flocculation, flocculants; 1.1.28, II.2.88, II.5.97, IV. 1.5 see colloid stability, especially by polymers flocculation kinetics; V.1.12e orthokinetic; V.1.84 Flory-Huggins interaction parameter (x); 1.3.45, 0.1.56, II.2.34, II.5.5ff, V.1.4, V.1.8 flotation; 1.1.25, II.2.88, III.4.4d, III.5.lib flow; see viscous flow flow birefringence = streaming birefringence fluctuations; 1.2.68, 1.2.102, 1.3.1, 1.3.7, 1.4.26, 1.7.26 in micelles; V.4.2d; more illustrations in V.chapter 4 in surfaces; I.7.81ff, II.1.45 of dielectric permittivity; 1.7.6 fluctuation potential; II.3.52 fluidity; I.6.52(def.) fluorescence; 1.7.14 in surfaces and adsorbates; 1.7.1 la, II.2.54, II.2.80, Ill.table 3.5 fluorescence correlation spectroscopy; III.3.7c.iv fluorescence lifetime imaging microscopy; III.3.7c.iv fluorescence recovery (after photobleaching) (FRAP); 1.7.104, III.3.7c.iv fluorescence resonance energy transfer; III.3.7c.iv fluorophores; Ill.fig. 3.63 flux; I.6.5ff, 1.6.11, 1.6.2b, 1.6.32 around polarized double layers; II.3.13c, IV.4.2-4.4 radial; 1.6.68 foam; I.fig. 1.7, 1.2.98, IV.1.11 (classif.), breaking; III.5.99, V.7.6 characterization; V.table 7.1, V.7.4 coalescence; V.7.3b, V.7.6 definition; V.7.1 drainage; V.7.3a dry; V.7.1 foam film; V.6.1 (def.), see films, liquid foam fractionation; V.7.35 flotation; V.7.35 (also see: films, liquid) formation; V.7.2, V.figs. 7.6-7 general; V.chapter 7 old; V.7.2
39
40
SUBJECT INDEX
foam (continued), optical properties; V.7.5b Ostwald ripening; V.7.3c polyhedral; V.figs. 7.1-3, V.7.1b, V.fig. 7.13 rheology; V.7.5a stability; V.7.2, V.7.3 structure; V.7.1a, V.7.1b, V.fig. 7.1 wet; V.7.2 young; V.7.2 foaming agents; V.7.1c fog; 1.1.6 Fokker-Planck equation; 1.6.3c, 1.6.73, IV.4.7, IV.4.9 forces (principles); conjugate; 1.6.12, I.6.2b, conservative; 1.4.2 directional; 1.6.1, I.6.3b, electrostatic; 1.4.2 fields; 1.4.3 generalized interpretation; 1.6.1 Iff, 1.6.54 hydrophobic; see hydrophobic interaction and bonding internal vs. external; 1.6.13 mechanical vs. thermodynamic; 1.2.27, 1.4.2, 1.4.49 steric; 1.4.2 stochastic; 1.4.2, 1.6.1, 1.6.3 Van der Waals; see Van der Waals forces vectorial interpretation; 1.4.3a, Lapp. 7 (also see: interaction and interaction force) force constant; 1.4.44 force (effect) microscope; 1.7.90 forced interaction; IV.3.10, IV.4.5, IV.5.3a forced Rayleigh scattering; 1.7.103 forced wetting; III.5.4, III.fig. 5.5 form drag; 1.6.47 form factor (in scattering); IV.2.44, IV.3.143, IV.5.21 Fourier transform infrared (spectroscopy) (FTIR); II.2.53, V.fig. 6.6 Fourier transforms; Lapp. 10 fractal dimensionality; IV.2.21ff, IV.4.6, IV.6.75 fractal growth; IV.2.21ff, IV.4.6 fractal structures; IV.4.6, IV.6.75ff, IV.fig. 6.40 fractionation; IV.2.7, IV.2.2h adsorptives of different sizes; II.2.45
SUBJECT INDEX
FRAP = fluorescence recovery after photobleaching free energy; see Helmholtz energy free enthalpy; see Gibbs energy free path (mean); 1.6.55 freezing point depression; 1.2.75 Fredholm integrals; II. 1.108 Frenkel defects; II.3.173 Fresnel equations; 1.7.10a Fresnel interfaces; 1.7.73, II.2.5c FRET = fluorescence resonance energy transfer Freundlich adsorption isotherm; see adsorption isotherm friction; 1.4.3, 1.6.10 friction coefficient (intr.); I.6.21ff, I.6.30ff, 1.6.56, 1.7.101, Lapp. 11 frictional drag (intr.); 1.6.47 froth = foam FRS = forced Rayleigh scattering FTIR = Fourier transform infrared (spectroscopy) function of state (def.); 1.2.9, 1.2.14c functional adsorption; see adsorption, functional; III.2.22, III.2.34, Ill.app. 3 fundamental constants (table); Lapp. 1 funnel technique (interfacial rheology); III.3.185 Fuoss-Onsager equation for limiting conductivity; [1.6.6.29] Fuoss theory for ion association; I.5.2d Galvani potential; see potential galvanic cells; I.5.5e gas pockets (and nucleation); V.7.9, V.fig. 7.5 gases, adsorption on solids; chapter 1 ideal; 1.1.17, 1.2.40, 1.2.17a, 1.3.If, I.3.6b, I.3.6c, 1.3.58 non-ideal; 1.3.9 Gauss distribution; 1.3.39, 1.6.21, 1.6.3c, 1.6.63, [IV.2.3.41] for surface heterogeneity; II. 1.106, II.fig. 1.43 Gauss' law (surface charge and field strength); I.4.53ff, Lapp. 7e, II.3.59, IV.3.11, IV.3.54 Gauss(ian) beams; 1.7.23 Gauss(ian) curvature; I.2.90(def.), ffl.l.4ff, III.1.15, V.4.94ff, V.5.24 (also see: bending moduus) gel, gelation; IV.4.48ff, IV.fig. 4.26, IV.5.86ff. IV.6.13, IV.6.14. V.2.3d, V.5.86 gel electrophoresis; II.4.131 gel. thermoreversible; IV. 1.6
41
42
SUBJECT INDEX
generic phenomena, properties; I.5.67(def.), II.3.6 Gibbs adsorption law; I.1.5(intr.), 1.1.16, 1.1.25, 1.2.13, 1.2.22, II.1.2, II.2.27 Gibbs adsorption law for; charged species; 1.5.3, II.3.12a curved interfaces; 1.2.94 dissociated monolayer; 1.5.94 electrosorption; II.3.12d liquid films; V.6.3d, [V.6.3.34] narrow pores; II. 1.96 polarized charged interfaces; 1.5.6c, II.3.4, II.3.138 relaxed (reversible) charged interfaces; 1.5.6b, II.3.4, II.3.138 water-air interfaces; II.3.178 see III.chapter 4 for many examples Gibbs convention (for locating dividing plane); 1.2.5, I.fig. 2.3, V.6.3 Gibbs dividing plane; 1.2.5 (intr.), 1.2.22, II. 1.2, II.2.4, II.2.64, III.4.4, V.6.2a, V.6.3 for curved interfaces; 1.2.23b Gibbs-Duhem relation; 1.2.10, 1.2.13, 1.2.78, 1.2.84, 1.5.92, [II. 1.3.35-36], V.6.3c, [V.6.3.20], [V.6.3.36, 37] Gibbs elasticity; III.3.30, V.6.2f Gibbs energy; 1.2.10, Lapp. 3, Lapp. 4 interfacial; 1.1.4,1.2.10, 1.2.11, Lapp. 5, Il.table 1.2, V.1.3, V.8.2a, V.8.64 self; 1.4.5c, 1.5.17, 1.5.3a, 1.5.3b statistical interpretation; 1.3.18, [1.3.3.14] Gibbs energy of; adsorption; see adsorption, II.Gibbs energy chemical substances; 1.2.77 colloid interaction; IV.chapter 3, V.chapter 1 double layer; 1.5.7, II.3.5, II.3.9, II.3.23, Il.fig. 3.6, Il.table 3.6 II.fig. 3.26, II.3.142, II.3.146, II.3.155ff, III.3.4h, IV.chapter 3 ions; I.5.171T, (also see: solvation, hydration) liquid films; V.6.3 nucleation; I.2.23d; IV.2.2b, IV.2.2f polarization; 1.4.55 proteins; V.3.2 solvation (hydration); 1.4.45, 1.5.3a, I.table 5.4, 1.5.3f transfer; I.5.3f, I.table 5.11,1.5.5a see further the pertaining system Gibbs free energy = Gibbs energy Gibbs triangle = phase diagram, ternary Gibbs-Helmholtz relations; 1.2.41. 1.2.15, 1.2.61. 1.2.78. II.3.156, III.3.34
SUBJECT INDEX
Gibbs-Kelvin equation; see Kelvin equation Gibbs-Thomson equation; see Kelvin equation Girifalco-Good-Fowkes theory (for interfacial tensions); III.2.lib, III.table 2.3 glass, double layer; II.fig. 3.64 glass electrodes; II.3.224 glass transition point; IV.5.90 goethite; see iron oxide gold sols; IV. 1.2, IV. 1.3, IV.1.13; IV.2.27, IV.3.185 Gouy-Chapman length; [V.2.2.13] Gouy-Chapman theory (diffuse double layers); 1.5.16, 1.5.18, II.3.5 (in) cavities; II.3.5g cylindrical surfaces; II.3.5f, V.2.2 defects; II.3.6a flat surfaces; II.3.5a-d, II.fig. 5.18 improvements; II.3.6b, Il.figs. 3.18-19 spherical surfaces; II.3.5e (also see: double layer, Gouy-Stern model) grafting (macromolecules); V.I.10, V.I.11, V.3.9, V.fig. 3.30 grand potential; III.2.7, V.I.I, V.l.3-4, V.1.8c, association colloids; V.chapter 4 translationally restricted; V.4.25ff also see interfacial grand potential graphite; see carbon Graphon, graphite; see carbon gravity (influence on colloid stability); IV.3.10a, see further sedimentation grazing incidence; II.2.56, III.3.147, III.3.152, Ill.fig. 59 ground state approximation (Edwards eq.); V. 1.11 ground water table; 1.1.1 growth (particles, drops); V.8.31 GSA = ground state approximation guar solution; IV.fig. 6.34 Guggenheim convention (for treating interfacial excesses); 1.2.15, I.fig. 2.4 Guinier radius; IV.2.45 gum arabic; 1.1.2, 1.1.7, IV. 1.3 Gurvitsch's rule; II. 1.94 gyration radius; see radius of gyration haematite (a-Fe2O3); see iron oxides Hageman factor; V.3.52 Hagen-Poiseuille law; 1.6.42, II.4.47, II.5.62, IV.6.7a
43
44
SUBJECT INDEX
Hamaker constant, Hamaker function; 1.3.45, 1.4.59, [1.4.7.7], 1.4.79, II.2.5, II.3.129, (tables) Lapp. 9, IV.app. 3 and interfacial tensions; III.2.5c Hamiltonian; I.3.57ff, II.3.47ff hard core or hard sphere interaction (molecules or colloids); 1.3.65, 1.4.5, 1.4.42, IV.5.4 hard sphere liquid; IV.5.3 heat, statistical interpretation; 1.3.16 (also see; enthalpy) heat capacity; 1.2.7, 1.3.36, 1.5.42 interfacial; 1.2.7, III.2.74 of ions; I.table 5.6 heavy metal pollution; II.3.221 Helfrich equation (bending); [III.1.78], [III.1.15.1 and 2], [V.5.5.1] helix (a); V.fig. 3.2 Helmholtz energy; 1.2.10, Lapp. 3, Lapp. 4 interfacial; 1.2.10, 1.2.11, Lapp. 5, Il.table 1.2, V.1.3b, V.1.4 liquid films, V.6.3 statistical interpretation; 1.3.17, II.especially [1.3.3.10] see further the pertaining system Helmholtz free energy = Helmholtz energy Helmholtz planes; see inner Helmholtz plane and outer Helmholtz plane Helmholtz-Smoluchowski equation; see electrophoretic mobility hematite = haematite, see iron oxides Henderson equation, for liquid junction potential; [1.6.7.11 ] for solvent structure contribution to disjoining pressure; [1.6.13] for-^-potential; [II.3.9.9] Henderson-Hasselbalch equation, plot; [1.5.2.34], [II.3.6.52, 11.53], II.3.88ff, V.2.19ff, V.fig. 2.12 Henry adsorption isotherm; see adsorption isotherm Henry constant, for adsorption; 1.1.19,1.2.71 for solubility of a gas; 1.2.20b Henry's law for gas solubility; 1.2.20b Herschel-Bulkley (rheological model); [IV.6.3.4] Hess' law; 1.2.16 heterodispersity (of colloids); I.7.8e, IV.5.3 Iff, IV.fig. 6.27 heterodyne beating; see optical mixing heterogeneity of surfaces; see surface, heterogeneity
SUBJECT INDEX
hetero-interaction; 1.4.72, I.fig. 4.17, IV.3.4, IV.3.12, IV.flg. 3.58, IV.flg. 3.61. IV.5.3Iff, IV.flg. 5.64 hexadecylpyridinium chloride (adsorption); II.fig. 2.22 higher-order Tyndall spectra (HOTS); 1.7.61 Hill plot; II. 1.48 HLB = hydrophile-lipophile balance HNC = hypernetted chain Hofmeister series = lyotropic series holes (in semiconductors); II.3.171 homodisperse colloids; 1.1.14, 1.1.28, 1.7.53, IV.1.6, IV.flg. 1.3 homodyne beating; see optical mixing homogeneous condensation, see condensation homogenizer; V.fig. 8.17 homointeraction; 1.4.72 homopolyelectrolytes; II.5.1 homopolymers; II.5.1 honeycomb symmetry; 1.1.14 Hooke (material, law); IV.6.9, IV.6.13 HOTS = higher order Tyndall spectra HPLC = high performance liquid chromatography; see chromatography HSA = human serum albumin; see albumin Hiickel-Onsager equation; see electrophoretic mobility Huggins constant; IV.6.61, V.2.48 Huggins equation (rheol.); [IV.6.11.9], [V.2.4.3] Huygens oscillator; V.fig. 1.26 hyaluronic acid (charge); V.fig. 2.7 hydration; 1.2.58,1.5.3, II.3.121, II.table3.7 (also see: solvation) hydration number; 1.5.50 hydraulic radius; 1.6.50, II. 1.84 hydrodynamic radius, layer thickness; 1.7.50, II.5.61 hydrodynamics; 1.6.1, conservation laws; 1.6.la, 1.6.1b, II.4.6, II.4.8, IV.6.1 in colloid interaction; IV.4.5b in electrophoresis; II.4.3, II.4.6 in emulsification; V.8.2b in polarized double layers; II.3.215ff, II.4.6 hydrogen bonding, hydrogen bridges; I.4.5d, 1.5.3c hydrophile-lipophile balance (HLB); V.4.1b, V.8.5 hydrophilic: 1.1.7, 1.1.23, II.table 1.3 (also see: colloids)
45
46
SUBJECT INDEX
hydrophilicity/phobicity; Il.table 1.3, III.5.5, III.5.11, III.5.1 la hydrophobic; 1.1.2, 1.1.7, 1.1.23, Il.table 1.3 (also see: colloids) hydrophobic interaction and -bonding, hydrophobic effect; 1.1.30, I.4.5e, 1.5.3c, 1.5.4, I.table 5.12, II.2.7d, II.3.12d, V.3.2b, V.3.6b (also see: lyotropic sequences) hyperbolic functions; Lapp. 1.2 hypernetted chain (HNC); 1.3.69, [IV.5.3.41] hysteresis; 1.2.7 in rheology; IV.6.3b, IV.fig. 6.7 also see, (hysteresis of) adsorption, contact angles, monolayers ideal dilute (polymer solution); II.5.9 i.e.p. = isoelectric point IgG = immunoglobulin iHp = inner Helmholtz plane illites; II.3.165 image charges; II.3.48ff, Il.fig. 3.17, Il.table 3.3 imaginary quantities; Lapp. 8 imaging techniques; 1.7.lib (also see: transmission electron microscopy, atomic force microscopy, surface force microscopy, etc.) immersion method (to determine points of zero charge); II.3.105 immersion heats or enthalpies; II. 1.29, Il.table 1.3, II.2.5, II.2.6, II.2.7, II.2.3d, Il.fig. 2.10, Il.fig. 2.20, II.3.98, II.3.114, III.5.2 (also see: wetting, immersional) immunoglobulins, adsorption; V.fig. 3.13, V.fig. 3.18 impedance (spectrum); II.3.92, Il.fig. 3.30, II.3.149, II.4.8 incident angle (for radiation); 1.7.10a incident plane (for radiation); 1.7.10a index of refraction; see refractive index indicator electrode; 1.5.82 indifferent (ions, electrolytes); II.3.6, II.3.103, IV. 1.11 inertia; II.4.2 infrared spectroscopy; 1.7.12, II.2.8, II.2.71ff, II.5.57; III.3.7c.i, III.fig. 3.62 infrared reflection-absorption spectroscopy; III.3.7c.i injection (emulsification); V.8.31 injection (foaming); V.7.8ff ink (Egyptian); 1.1.1, 1.1.2, 1.1.7, 1.1.27, IV. 1.3, IV.2.1 ink jet printing; III. 1.84 inner Helmholtz plane (intr.); II.3.61ff insoluble monolayers, see monolayers, Langmuir
SUBJECT INDEX
interaction (principles), energy, force; 1.4.2, 1.4.8, Il.figs. 2.2-3 multiparticle, Born-Green-Yvon; 1.3.69 Carnahan-Starling; I.3.69ff hypernetted chain: 1.3.69 Percus-Yevick; 1.3.69 relation to distribution functions; I.3.9d, I.3.9e relation to virial coefficients; 1.3.9c pairwise; 1.3.8, 1.3.9, 1.4.1, 1.4.2, 1.4.3, 1.4.4 tabulation for electric repulsion; IV.app. 2 potential; see interaction energy sign; 1.4.4 solvent structure-originated; 1.5.15, 1.5.3, 1.5.4, II.table 1.5.12, II. 1.95-96, H.fig. 2.2, II.3.184ff, IV.3.8C see further, colloids, interaction interaction between colloids and macrobodies, see colloid stability, Van der Waals forces and colloids, interaction interaction between ions; 1.5.2 interaction between molecules and surfaces; [1.4.6.1], II.1.5, II.chapters 1, 2 and 5 interaction curves; I.fig. 3.4, I.fig. 4.1, figs. 1.4.2-3 interaction energy parameter; I.3.40ff, 1.3.43-45, [1.3.8.9], II.1.56, II.2.34, II.5.5ff excess; 1.3.45, especially [1.3.8.9] interaction forces, general introduction; I.chapter 4 interactions inside proteins; V.3.3 interaction in solution (excess nature of); 1.1.29, 1.4.5, 1.4.6b, 1.4.7 interface; I.1.3(intr.) curved; see capillary phenomena of tension; 1.2.94
optical study; 1.7.10, 1.7.11, II.2.5c reflection of light; 1.7.10a, II.2.5c refraction of light; 1.7.10a scattering; 1.7.10c (also see: purity criteria) interfacial area; see surface area interfacial charge; see surface charge interfacial concentration (intr.); 1.1.5, see further, interfacial excess interfacial energy; III.2.9a, Ill.fig. 2.14, Ill.fig. 2.16 (relation to heat of evaporation)
47
48
SUBJECT INDEX
interfacial entropy; III.2.9a, Ill.fig. 2.14-1§5, HI.2.55 (relation to entropy of vaporization), III.4.2d, Ill.fig. 4.19 interfacial excess; 1.2.5, 1.2.42, 1.2.22, III. 1.2, II.2.2, II.fig. 2.1, [II.2.1.2],
II.2.3, III.2.2, V.6.3, [V.6.3.24] isotherm; II.2.3, III.4.2 interfacial Gibbs energy; III.2.2, V.I.3 interfacial grand potential; III.2.2, III.5.18ff, V.I.I, V.l.3-4, V.1.6, V.1.8, V.1.9c, V.I.53 (also see: adsorption, Gibbs adsorption law, surface excess) interfacial polarization; see potential difference, x interfacial potential jump (x); see potential difference, x interfacial potentials; 1.5.5, II.3.9, III.4.4 also see: electrokinetic potential; Gouy-Chapman theory; monolayers, ionized; Poisson's law; potential difference between adjacent phases interfacial pressure; see surface pressure interfacial rheology; III.3.6, Ill.table 3.5 Burgers element; 111,3.129 compliance; Ill.table 3.4, III.3.105 compressibility; [III.3.3.1], [III.3.4.3], III.3.93 compression; III.3.83, III.3.91 creep; Ill.fig. 3.40, III.3.61 deformation types; Ill.fig. 3.34 dilation; III.3.81, III.3.83, III.3.91, Ill.fig. 3.38 dilational elasticity (modulus); [III.3.4.4], III.3.40, III.3.81, Ill.table 3.4ff, [III.6.18-19], [III.3.6.34-39], III.3.6g, Ill.fig. 3.48, Ill.fig. 3.87, Ill.fig. 3.91, III.4.5, Ill.figs. 4.26-27, Ill.table 4.3, Ill.fig. 4.38, V.fig. 3.27, V.7.6, V.7.16, V.S.lc, V.fig. 8.4, V.8.69, V.8.83, V.8.88 distance coefficient; III.3.113, Ill.fig. 3.44, III.3.117ff distance damping; III.3.112 elasticity; III. 1.55, III.3.82, Ill.table 3.4 emulsification; V.8.2b, V.fig. 8.11, V.8.47ff equivalent mechanical circuits; III.3.6i experimental methods; III.3.6f, IH.3.7e Fourier transform method; III.4.59ff Gibbs monolayers; Ill.table 4.2 Kelvin element = Voigt element Kelvin equation (damping); [III.3.6.63] kinetics; III.4.5 loss angle; [III.3.6.12], [HI.3.6.41a], [III.3.6.66], [III.3.6.74]
SUBJECT INDEX
Interfacial rheology (continued), Marangoni effect; 1.1.2 (intr.), 1.1.17, I.6.4.43ff, III. 1.35, III. 1.72, III.3.81 III.3.6e, Ill.flgs. 3.35-37, III.3.239, V.6.2, V.6.37ff, V.7.6, V.8.1c, V.flg. 8.3, V.8.48, V.8.52ff Maxwell element; Ill.flgs. 3.51-52 proteins; V.3.7, V.figs. 3.25-27 relation to Interfacial tension; III.3.92, III.3.6d recovery; Ill.flg. 3.40, Ill.fig. 3.52 relaxation (Gibbs monolayers); Ill.fig. 4.20 relaxation (Langmulr monolayer); III.3.6h, III.3.61 respiratory stress syndrome; III.3.238-239 shear modulus; III.3.84, IH.3.92 shear viscosity; III.3.84, III.3.92, [III.3.6.20], III.3.6g, Ill.fig. 3.91, V.8.89 stress relaxation; Ill.fig. 3.39, III.3.61 stress tensor; III.3.9Iff, Ill.table 3.4, V.figs. 3.25-27 time damping; III.3.113 viscosity; III. 1.55, III.3.82, Ill.table 3.4, V.figs. 3.25-27 Voigt element; III.3.94, Ill.flgs. 3.51-52 wave damping and propagation; III.1.1, III. 1.58, III.3.6g, Ill.fig. 3.42-44, 111.3.183,111.3.185, also see: loss angle interfacial science (first review); 1.1.2, 1.1.3, Volumes II and III interfacial tension, surface tension; I.1.4(intr.), 1.1.25, I.fig. 1.16 binary mixtures; III.4.2 data; Ill.app. 1, III.1.12 dynamic conditions; III. 1.14b, Ill.fig. 1.31, Ill.fig. 1.32, V.fig. 8.3, V.8.1c, V.8.48ff Interpretation; III.chapter 2, III.3.6d Cahn-Hilliard; III.2.6 capillary waves; III.2.9c distribution function; III.2.4, [III.2.4.6-8], III.2.24 empirical; III.2.11 and geometric means; III.2.lib and grand potential; III.2.7 Hamaker-de Boer approximation; III.2.5c lattice theory; III.2.10 mechanical; III. 1.3, V.6.3b pressure tensor (interfacial); III.2.3, [III.2.3.5], III.3.6d, V.6.3b scaling; [III.2.5.35], V.5.74ff simulations; III.2.7, Ill.flgs. 2.9-10, Ill.table 2.2 statistical thermodynamics; III.2.4, III.2.30-31, III.2.51ff
49
50
SUBJECT INDEX
Interfacial tension, surface tension, interpretation (continued), thermodynamic or quasithermodynamic; I.2.10ff, I.2.26ff, 1.2.11, 1.2.91, III.2.2, III.2.9 van der Waals; III.2.5 measurement; 1.1.11, 1.2.5, 1.2.96, II.3.139, III.chapter 1 'Bugler method'; III. 1.49 capillary rise; III.1.3, (differential) III.1.19 (inawedge), Ill.fig. 1.10 captive drops; Ill.fig. 1.11, further see sessile and hanging drops drop oscillations; III. 1.58 drop weight; III.1.6, III.1.72ff drops in a gradient; III.1.5 (also see: growing drops, sessile drops, spinning drops, etc.) du Nouy ring; III. 1.8b, III.1.72ff dynamic (conditions); III.1.14, III. 1.3-4, III.1.53ff (also see; interfacial tension, relaxation) falling drop; Ill.fig. 1.17 (see further, drop weight) falling meniscus; III.1.11, Ill.fig. 1.26 growing drops; III. 1.74 hanging drops; see sessile drops maximum bubble pressure; III.1.7, III. 1.72ff micropipette; III. 1.57 'Padday's pencil'; III. 1.48 pendant drop = hanging drop; see sessile and hanging drops rheology; III. 1.57 sessile and hanging drops; III. 1.4, III.1.72ff sphere tensiometry; III. 1.48 spinning bubbles; see spinning drops spinning drops and bubbles; III.1.9, Ill.fig. 1.24 surface light scattering; III. 1.10 tensiometers; III. 1.8 wave damping; Wilhelmy plate; 111.1.8a, III.1.72ff of curved interfaces; 1.2.23, II.1.6d, III.1.1, III.1.15, V.5.4 of electrolyte solutions; III.4.4 of films; II.1.95ff of microemulsions; V.5.4 of solid surfaces; 1.2.24, III. 1.5 measurement; III.1.13 pressure dependendence; III.2.9b relation to adsorption from binary mixtures; II.2.4f
SUBJECT INDEX
interfacial tension, surface tension, measurement (continued), relation to compressibility; III.2.11a, [III.2.11.6-7] relation to interfacial Helmholtz, Gibbs or internal energy; 1.2.11, Lapp. 5, [Ill.l.la.b] relation to molar volume; III.2.11a relation to surface light scattering; 1.7.10c, III.1.10 relation to work of cohesion; 1.4.47 relaxation; III. 1.14, III.fig. 1.29 temperature dependence; 1.2.42, [11.1.3,42], III.2.9a, V.fig. 8.1, V.5.4 (also see: capillary phenomena, monolayers, wetting, interfacial rheology) interfacial turbulence; V.8.52ff interfacial viscometers; III.3.180ff, Ill.figs. 3.69-71 interfacial work; 1.2.10, 1.2.3,1.3.17 interference (intr.); 1.7.8,1.7.15 interferometry (contact angle); III.5.43ff ion association (in solution); I.5.2d ion binding; I.5.3ff, II.3.6d-e ion condensation; II.4.43, V.2.5 ion correlations; II.3.6b ion exchange; II.3.35, II.3.168ff ion mass spectroscopy (SIMS); 1.7.lla, I.table 7.4 ion pairs; 1.5.3 ion scattering spectroscopy (ISS); 1.7.lla, I.table 7.4, II. 1.15 ion specificity; see lyotropic series ion transfer (resistance); II.3.95, IV.4.19-20 ion vibration potential; II.4.29 ionic atmosphere; 1.5.16, see double layer, diffuse ionic components of charge; see double layer, electric ionization, ionization (Gibbs) energy; 1.4.29, I.5.34ff, 1.7.86 ionomer; V.2.70 ions, activity coefficient; see there bound vs. free; 1.5.la, II.3.6, III.3.4h, V.2.2, V.2.5a hydration; 1.5.3 hydrophilic; 1.5.47 hydrophobic; 1.5.46 radii; I.table 5.4, I.fig. 5.7 solvation; 1.5.3 structure-breaking; 1.5.47 structure-forming; 1.5.47 transfer; I.5.3f, II.3.9
51
52
SUBJECT INDEX
ions (continued), volumes; I.tables 5.7 and 8 ion-solvent interaction; see hydration, solvation ionic surfactants; see surfactants IRAS = infrared reflection-absorption spectroscopy iron oxides goethite (a-FeOOH), electrokinetic charge; II.fig. 4.13 point of zero charge; II.table 3.5, Il.app. 3b haematite (a-Fe2O3); Il.fig. 1.1, IV. fig. 2.1b adsorption of fatty acids from heptane; Il.fig. 2.26 conductivity of sols; Il.fig. 4.38 dielectric relaxation of sols; Il.fig. 4.38 double layer; II.3.94, Il.table 3.6, II.figs. 3.59-62, Il.table 3.8 electrokinetic properties; Il.table 4.3 point of zero charge; Il.table 3.5, Il.fig. 3.37, Il.app. 3b immersion, wetting; Il.table 1.1 irradiance (intr.); 1.7.5 IRRAS = infrared reflection-absorption spectroscopy irregular coagulation series; see coagulation irreversible thermodynamics; see thermodynamics irreversible colloids = colloids, lyophobic irreversible process; see process, natural Ising problem; 1.3.40,1.3.43, V.2.20 isobaric (process, def.); 1.2.3 isochoric (process, def.); 1.2.3 isoconduction; II.3.215 isoelectric point; II.3.8, II.3.103, II.3.106, Il.fig. 3.78, Il.fig. 4.41, II.4.127ff relation to point of zero charge; II.3.8b, Il.fig. 3.35 isoelectric focusing; II.4.131ff isodisperse = homodisperse isomorphic substitution (in clay minerals); II.3.2, II.3.165 isosteric (process); 1.2.3 isosteric heat of adsorption; see adsorption, isosteric enthalpy isotachophoresis; II.4.131 isothermal (process); 1.2.3 destination; see Ostwald ripening reversible work; 1.2.27 ISS = ion scattering spectroscopy Jones-Dole equation, coefficients (viscosity of electrolytes); 1.5.52, I.table 5.9, I.6.78ff, IV.6.53-54
SUBJECT INDEX
kaolinite; II.3.164, II.fig. 3.66, IV.flg. 2.2b wetting; II.table 1.34, Keesom-Van der Waals forces; see Van der Waals forces Kelvin cells; V.7.5 Kelvin element (rheol.) = Voigt element; see interfacial rheology Kelvin equation; [1.2.23.24], [II. 1.64, [II.1.6.17], [III.1.13.3], [IV.2.2.50], IV.2.2e Kelvin equation (wave damping); [III.3.6.63] Kerr effect; 1.7.100 Kiessing fringes; III.3.150, Ill.fig. 3.58 kinetics (coagulation); IV.4.3 kinetics (micellization); V.4.10 Kirkwood equation (electric polarization); [1.4.5.22] Kirkwood-Buff equation (interfacial tension); [III.2.4.6-9] Kirkwood-Frohlich equation (electric polarization); [1.4.5.23] Kolmogorov theory (for emulsification); V.8.41ff Kozeny equation; [1.6.4.39], II.4.55 Kozeny-Carman equation; [1.6.4.41], II.4.55, [IV.2.2.67] Krafft temperature; V.4.13, V.8.5 Kramers-Kronig relations; [1.4.4.31, 11.32], 1.4.36, 1.4.77, 1.7.13, II.3.93 Krieger-Dougherty eq. (viscosity); [IV.6.9.10], V.8.15 Kugelschaum; V.7.2 Kuhn segment = statistical chain element lactalbumin (a), adsorption; V.figs. 3.15-18 lactoglobulin ((3 ); V.figs. 3.25-26, V.fig. 8.20 Lambert-Beer's law; 1.7.13, 1.7.41 Landau-Ginzburg analysis (microemulsions); V.5.38ff Langevin equation (for forced stochastic processes); [1.6.3.4], 1.6.3d, Lapp. 11.2, IV.4.2 Langmuir adsorption isotherm; see adsorption isotherm Langmuir-Blodgett layers; II.2.56, III. 1.42, III.3.7a Langmuir monolayers, see monolayers Langmuir trough; see film balance Laplace's law = Young and Laplace's law, see capillary pressure Laplace pressure = capillary pressure Laplace transformations; Lapp. 10 lasers; 1.7.4c laser-Doppler microscopy = QELS; see electromagnetic radiation latex, (pi. latices or latexes); 1.1.6, 1.5.99, II.3.87, II.fig. 3.29, IV. 1.9 compressibility and scattering; IV.fig. 5.32 conductivity; Il.table 4.2, II.fig. 4.34 crystallization; IV.5.8a, IV.flg. 5.58
53
54
SUBJECT INDEX
latex (continued), distribution function; IV.fig. 5.4 electro-osmosis; II.table 4.2 electrophoresls; Il.flg. 4.29, H.table 4.2 rheology; IV.fig.6.28-29 sedimentation; IV.fig. 5.60 stability; IV.3.13c, IV.fig. 4.8b, IV.figs. 4.19-21 streaming potential; Il.flg. 4.30, Il.flg. 4.35 structure factor; IV.fig. 5.19, IV.fig. 5.34, IV.fig. 5.36 surface charge; Il.flg. 3.29, IV.fig. 3.75 lattice statistics; I.3.6d, I.3.6e, I.3.8b, II.chapter 5, V.chapter 1 polymer adsorption; II.5.30ff random walk; 1.6.3d LC = liquid condensed (2D phase); III.3.3b LE = liquid expanded (2D phase); III.3.3b lead oxides; IV. 1.3 LEED = low-energy electron diffraction Lennard-Jones pair interaction energy; I.fig. 4.9, 1.4.5b, [1.4.5.1] in adsorbates; [II. 1.1.14], II. 1.74 in liquids near solids; II.figs. 2.4-5 structure factor; IV.fig. 5.5 leucocytes; III.5.100 leukemia; III.5.100 Levich equations (for convective diffusion); I.6.92ff levitation; IV.3.12d Lewis acids, bases; 1.5.65, II.3.185 Lifshits theory; see Van der Waals interaction between colloids Lifshits-Slezov-Wagner (LSW) theory (Ostwald ripening); IV.2.26, V.8.66ff light scattering; see electromagnetic radiation line tension; III. 1.6, III.5.5, III.5.6 lipase; 1.1.3 lipids (phospho-); III.table 3.8 lipids (phospho-) films; V.6.67ff, V.fig. 6,35, V.fig. 6.41, V.fig. 6.45 lipids (phospho-) monolayers; Ill.fig. 3.8, Ill.fig. 3.12, Ill.fig. 3.14, Ill.fig. 3.29, III.3.140, Ill.fig. 3.55, IH.figs. 3.62-62, Ill.fig. 3.67, III.3.8c, Ill.figs. 3.89-91, III.3.238 Lippmann capillary electrometer; see capillary electrometer Lippmann equation (for electrocapillary curves); [1.5.6.17], 1.5.100, 1.5.108, II.3.138 liquid bridges, see capillary bridges liquid junction potentials; see potential difference
SUBJECT INDEX
liquid-fluid interface, general; Ill.chapter 1.2, III.2.8 (thickness), III.2.9b density profile; Ill.fig. 2.1, III.2.3, [III.2.5.31], III.2.8, III.figs. 2.11-13, Ill.fig. 2.19 double layer; II.3.10g thermodynamics; III.2.2 liquids, apolar, double layers; II.3.11, II.4.50 solvation; I.5.3f in pores; I1.1.6d near surfaces; II.1.6d, II.3.123ff, II.4.38, H.fig. 4.11 London-Van der Waals forces; see Van der Waals forces longitudinal waves; III.3.110, see interfacial rheology loops; see adsorption of polymers Lorentzian peaks; 1.7.50 Lorenz-Lorentz equation (electric polarization); [1.4.5.21], 1.7.43 loss angle (rheology), III.3.90, [III.4.5.44] see interfacial rheology low-energy electron diffraction (LEED); 1.7.25, 1.7.86, II.fig. 1.2, II.l.l Iff LSA = linear superposition approximation, see colloids, interaction LSW (theory) = Lifshits-Slezov-Wagner (theory) lunar soil, spherules in; 1.1.1, 1.1.2 sorption of methanol; II.fig. 1.35 lung surfactant; 1.1.1, 1.1.2, III.3.219, III.3.238, V.6.8 lyotropic series; I.5.66ff(intr.) in coagulation; 1.5.67, IV.3.91, IV.table 3.3 in ionic binding, double layer charge or double layer capacitance; I.5.66ff, II.3.15, II.3.109, II.3.132, H.fig. 3.41, II.3.135, Il.fig. 3.53, Il.fig. 3.55, II.3.147, Il.fig. 3.75, II.3.10h, Il.table 3.8, II.3.203, III.3.207ff, Ill.figs. 3.85-86, III.4.89ff, IV.3.149, IV.table 3.6, IV.3.162ff, IV.3.170, IV.3.173, IV.figs. 3.72-73, V.2.22, V.2.53, V.2.68 lysozyme; IV.fig. 5.27, V.fig. 3.24-25, V.fig. 8.20 macromolecules; see polymers, polyelectrolytes, proteins macropores; see pores magnetic birefringe; 1.7.100 magnetic colloids; IV.3.10c magnetic fields; 1.7.1a, 1.7.2, 1.7.16, 1.7.13, IV.3.10c magnetic induction; 1.7.2, IV.3.10c magnetic permeability; 1.7.9, IV.S.lOc magnetite; IV.fig. 2.2, IV.2.4d, IV.fig. Al.l
55
56
SUBJECT INDEX
magnetization; 1.7.9, IV.3.10c Mandelstam equation (for surface scattering); [1.7.10.24], [III. 1.10.11, V.6.43 manganese dioxide, double layer; Il.table 3.8, Il.app. 3b Marangoni effect; see interfacial rheology Marangoni number; V.6.38, V.7.15ff, [V.8.1.4], V.8.57, V.8.83, V.8.89 marginal regime (polymer concentration); II.5.9, II.fig. 5.3 marginal regeneration (in films); V.6.4e Mark-Houwink eq. (rheology); [IV.6.11.8] Markov chain; 1.6.24, V.A1.7 Markov process, first order; 1.6.24, II.5.30 Martin eq. (rheology); [IV.6.11.14] masers; 1.7.4c mass action (micelle formation); V.4.2b mass conservation (in hydrodynamics); 1.6.la, IV.6.1 mass, reduced; 1.4.44 maximum term method (statistical thermodynamics); 1.3.37 Maxwell-Boltzmann statistics, distribution; 1.3.12,1.6.26ff, II.3.172, IV.4.6ff Maxwell element, see rheology and interfacial rheology Maxwell equations (for electromagnetic waves); 1.7.2 Maxwell (-Wagner) relaxation; 1.6.84, II.3.219, Il.fig. 3.89, IV.4.23 Mayer function; I.3.60ff, I.3.64ff mayonnaise; 1.1.6 mean curvature (of interfaces); 1.2.23a, III.1.1, III.1.15 mean field theories; II.5.7, II.5.29, III.2.S membrane emulsification; V.8.32, V.fig. 8.8 membrane equilibrium; see equilibrium mercury-solution interface, double layer; 1.5.6c, II.3.10b, II.figs. 3.48-55, Il.table 3.8 interfacial tension; II.3.138ff, Il.fig. 3,48 mercury sulphide; IV. 1.3 mesopores; see pores mesoscopic = colloidal metabolism; see fats metal(s), contact angles on ...; III.table A4.1 Hamaker constants; Lapp. 9.4, IV.app. 3 points of zero charge; Il.app. 3a sol preparation: IV.2.37 methylviologen; [II.3.14.1], II.3.224
SUBJECT INDEX
mica; II.3.165, IV.fig. 3.56 interaction; IV.3.12b, IV.fig. 3.57 micelle; I.1.6(def.), 1.1.24,1.fig. 1.15, III.4.6 complex coacervate; V.2.6f reverse (or inverted); 1.1.25 for general discussion, see V.chapter 4 micellization, critical concentration of (c.m.c); I.1.24ff(iritr.), III.4.6a, Ill.table 4.4, IV. 1.5 determination; I.1.25ff, V.4.1c critical temperature; V.4.11 for general discussion, see V.chapter 4 microelectrophoresis; II.4.45ff, Il.figs. 4.14-16, IV.fig. 5.14 microemulsions; 1.1.3,1.1.7,1.2.68, IV.1.6, V.chapter 5 (mostly non-ionic) applications; V.5.6 bending; V.5.5a, V.fig. 5.34, V.fig. 5.36, V.5.93 bicontinuity; V.5.1 Iff, V.5.30 conductometry; V.S.3f, V.fig. 5.21 correlation lengths; V.5.39, V.5.3h, V.fig. 5.25, V.fig. 5.27a, V.5.4e, V.5.5 curvatures; see microstructure diffusion; V.5.3e, V.fig. 5.20 efficiency boosting; V.5.6e emulsification failure boundary; V.5.21, V.5.49 experimental methods; V.5.3 fish diagrams; V.figs. 5.6, 5.13, 5.22, 5.38-39, 5.41, 5.43, 5.47 Gibbs triangle = ternary phase diagram interfacial tensions; V.5.4, V.figs. 5.28-33, V.5.5, V.fig. 5.37 Landau-Ginzburg approach; V.5.38ff microstructures; V.5.3 middle phase (= one of the Winsor states); V.5.2g optimalization; V.5.2d, V.5.2e phase behaviour (-diagrams); V.5.2, V.5.4b binary systems; V.5.2a quaternary systems; V.5.84, V.fig. 5.39, V.fig. 5.43 quinary systems; V.5.87, V.fig. 5.43 ternary systems; V.5.2b phase inversion (intr.); V.5.3 (many examples in chapter V.5) phase inversion temperature: V.5.6ff phase trajectories; V.5.2g and elsewhere in V.chapter 5 plumbers nightmare = bicontinuity scaling; V.5.2h, V.fig. 5.27, V.5.4e, V.5.73ff. V.fig. 5.37 surfactant solubility; V.5.2f, and elsewhere in V.chapter 5
57
58
SUBJECT INDEX
microemulsions (continued), theory; V.5.5 wetting; V.5.4c, V.fig. 5.30 Winsor states (def.); V.5.1, V.fig. 5.1 micropores; see pores microscopies of sols; IV.2.4Iff middle phase (microemulsions); V.5.2g Mie theory (light scattering); see electromagnetic radiation milk; V.8.1 mixtures, athermal; 1.2.55 colloidal; IV.5.7c, IV.5.8c homogeneous, thermodynamics; 1.2.16 ideal; 1.2.17 non-ideal; 1.2.18 mobile films; V.6.48 mobility (of ions); 1.6.6a molality; I.2.45(def.) molarity; I.2.45(def.) mole fraction; I.2.44(def.) molecular condensor; I.fig. 5.1, II.3.59 molecular dynamics; I.3.1e(intr.), association colloids; V.4.3a, V.fig. 4.7 electrolytes; I.5.57ff, I.fig. 5.9 liquids in pores; Il.fig. 1.38 liquids near surfaces; Il.figs. 2.5-7, II.3.55, Il.fig. 3.39 wetting; Ill.fig. 5.36 molecular mass (of colloids); see polymers, particles molecular sieve, adsorption of krypton; Il.fig. 1.19 adsorption of methane; Il.fig. 1.36 molecular state; see state molecular thermodynamics; see statistical thermodynamics moment, of a distribution; 1.3.7b of a double layer; [II.4.6.50] moment expansion; IV.2.45, IV.app. 1 momentum; I.3.57(def.) momentum conservation (in hydrodynamics); 1.6.lb, IV.6.4 (also see: transport of momentum) monatomic crystal: see Einstein crystal
SUBJECT INDEX
monochromatic (waves, radiation); 1.7.2 monochromator; 1.7.3 monodisperse = homodisperse monolayers (at liquid-fluid interfaces); I.fig. 1.15b adsorbed = Gibbs monolayers bending moduli; III.tables 1.6 and 7 binary mixtures; see Gibbs monolayers characterization; III.3.7 cholesterol; III.3.8d curved; see Gibbs monolayers; diffraction; III.3.7b dilute solution; see Gibbs monolayers electrolytes; see Gibbs monolayers in emulsions; V.8.1f film balances; III.3.3a fatty acids; alcohols; III.3.8b Gibbs (monolayers); III.chapter 4 binary mixtures; III.4.2, V.8.7 curved; III.4.7 dilute solutions; III.4.3 distinction from Langmuir monolayers; III.3.1 dynamics; III.1.14b, III.4.5 electrolytes; III.4.4 Gibbs equation; 1.5.94 ionized; II.3.2, Il.fig. 3.1b, V.6.5b proteins; V.3.7 rheology and kinetics; see interfacial rheology surfactants; III. 1.14b, III.4.6 temperature dependence; V.fig. 8.1 Langmuir (monolayers); III.chapter 3 Brewster angle microscopy; III.table 3.5 characteristic functions; III.table 3.2 cholesterol; III.3.8d collapse; Ill.fig. 3.46 diffraction; III.3.7b distinction from Gibbs monolayers; III.3.1 ellipsometry; III.table 3.5, III.3.7b energy-entropy compensation; III.3.37 fluorescence; III.table 3.5 hysteresis; III.3.13. III.3.8a, Ill.fig. 3.79 ionized; III.3.4h
59
60
SUBJECT INDEX
monolayers (at liquid-fluid interfaces), Langmuir (continued), Langmuir trough; HI.3.3a Langmuir-Blodgett; see there lattice theory; III.3.5e mixed; III.3.4f molecular dynamics; III.3.5d molecular thermodynamics; III.3.5 Monte Carlo; III.3.5c neutron reflection; Ill.table 3.5, III.3.7b optical techniques; III.3.7b-c permeation; III.3.238 phase behaviour; III.3.3 phospholipids; III.3.8c polymer brushes; III.3.4J, III.3.8f polymers; III.3.4i, III.3.8e, V.8.8, V.fig. 8.2, V.I.77 preparation; III.3.2 proteins; V.3.7 reflection; III.III.3.7b relaxation; III.3.6h, V.I. 12b reproducibility; III.3.8a, IILfig. 3.79 rheology; see interfacial rheology; scanning probe; III.3.7d, Ill.table 3.5 simulations; III.3.5c, III.3.5d spectroscopy; III.3.7c, Ill.table 3.5 thermodymamics; III.3.4 thermodynamics; III.3.4 transfer; III.3.7a, IILfig. 3.5,, III.figs. 3.53-54 Volta potential; see there, For the optical techniques see the entry in question X-ray diffraction; Ill.table 3.5, III.3.7b X-ray reflection; Ill.table 3.5, III.3.7b monolayer formers; III.3.200 monolayer spreading; III.3.2 montmorillonite; II.3.165 adsorption of alcohols + benzene; II.fig. 2.21 adsorption of methane + benzene; II.fig. 2.22 adsorption of poly(acryl amide); II.fig. 5.39b adsorption of water vapour; II.fig. 1.30 disjoining pressure; IV.fig. 3.55 Monte Carlo simulations; I.3.1e(intr.), I.fig. 5.4. 1.5.30, IV.fig. 5.16, IV.fig. 5.30 adsorbed liquids; II.fig. 2.4 adsorbed polymers; II.5.30
SUBJECT INDEX
Monte Carlo simulations (continued), association colloids; V.4.3a, V.figs. 4.5-4.6 electric double layer; II.fig. 3.18 Mountain lines; 1.7.45 moving boundary electrophoresis; 11:4.5 Iff, Il.fig. 4.17 mushrooms (polymer ads.); V.I.lid, V.flg. 1.27 muscovite; II.3.165 myoglobin (ads.); V.flg. 3.19 natural; see process nanoparticles = small colloids nanoscience = science of small colloids Navier-Stokes equation; [1.6.1.15], 1.6.51, II.4.18, [II.4.6.4]ff NBF = Newton black film; see films, liquid negative adsorption; see adsorption, negative Nernst-Einstein equation; [1.6.6.15], [II.3.13.14], [II.4.3.551 Nernst-Planck equation; I.6.7a, 1.6.89, II.2.85, [II.3.13.12], [II.4.6.2] Nernst's heat theorem; 1.2.24 Nernst's law for distribution equilibrium; 1.2.20a, 1.2.81 Nernst's law for electrode potential; 1.2.34, I.5.5c, I.5.5e, II.3.8, II.3.91, II.3.147ff, II.3.150 networks; IV.6.14 also see: gels, percolation Neumann triangle; [III.5.1.3], III.fig. 5.6 neutron reflection (by surfaces); II.2.7, II.5.66ff neutron scattering (by colloids); 1.7.9, 1.7.102, IV.fig. 5.25, IV.fig. 5.33, IV.fig. 5.36 Newton films; see films, liquid Newton(ian) fluids; 1.6.8,1.table 6.1,1.6.4a, III.3.6b, IV.6.1, IV.6.2, IV.fig. 6.5, IV.table 6.3 Newton's second law; [1.6.1.12], 1.6.4 NMR = nuclear magnetic resonance non-ionic surfactants; see surfactants non-linear optical techniques; Ill.table 3.5 non-Newton(ian) flow; 1.6.36, III.3.6b, IV.6.7a, IV.fig. 6.17 non-solvent; 1.1.27 normal stress; see stress nuclear magnetic resonance (NMR); 1.7.16, 1.7.13. 1.7.102 chemical shift; I.5.54ff, I.fig. 5.8 of emulsions; V.8.23 of interfaces; II.2.8, II.2.55. II.5.58ff. II.5.71 of microemulsions; V.5.3e, V.flg. 5.20 of pores; II. 1.90
61
62
SUBJECT INDEX
nuclear magnetic resonance (NMR) (continued), of water; I.5.54ff, I.fig. 5.8 (also see: spin, etc.) nucleation, in colloid preparation; IV.2.9ff, IV.2.2b, IV.2.2c in emulsification; V.8.31 heterogeneous; 1.2.100, II. 1.42 homogeneous; 1.2.23d. IV.2.9ff in pores; see capillary condensation number of realizations; 1.3.4 (def.) octupole moment; 1.4.19 odd-even parity; III.3.302, Ill.fig. 4.31, Ill.fig. 4.36 oHp = outer Helmholtz plane Odijk-Skolnick-Fixman theory (polyelectr.); V.2.27ff oil recovery, tertiary; see enhanced oil recovery ointments; 1.1.6, 1.1.28 Onsager formula (for limiting conductivity); [1.6.6.26] Onsager formula (for polarization); [1.4.5.20] Onsager relations (irreversible thermodynamics); 1.6.2b application to electrokinetics; I.6.2c, II.4.2, II.4.7, II.4.21, II.4.27, II.4.61, II.4.106 Onsager theorem (for approach to equilibrium); 1.7.44, 1.7.48, Lapp. 11.3, Lapp. 11.5, Lapp. 11.7-8 optical activity; I.7.99ff optical axes; 1.7.14 homodyne; 1.7.37, I.7.6d optical levitation; IV.3.157ff optical mixing (beating); 1.7.37 heterodyne; 1.7.37, I.7.6d optical trapping; IV.3.12d optical tweezers; IV.3.158 ordering parameter; 1.6.73, III.3.71ff, Ill.fig. 3.61, III.3.166, [III.3.7.13] open circuit potential; II.3.149 ore benification; 1.1.25, II.5.97 orientation of adsorbed molecules; II.2.55 Ornstein-Zernike equation (for compressibility); [1.3.9.32] Ornstein-Zernike equation (for correlation functions); [IV.5.3.19], [IV.5.3.33b] Ornstein-Zernike equation (for critical opalescence); [1.7.7.10] orthokinetic coagulation or flocculation; IV.2.41, IV.4.5b, IV.fig. 4.18, IV.4.47-48, V.1.84, V.fig. 1.50 oscillating drop; Ill.fig. 3.72
SUBJECT INDEX
oscillating liquid jet; III.figs. 1.28 and 29, III. 1.84 oscillation, harmonic; 1.4.38, 1.4.44, 1.7.3d, III.3.80ff, III.3.105ff, Ill.fig. 3.41, Ill.fig. 3.45 oscillator; I.7.3b oscillator, harmonic; 1.3.5a, 1.4.37 oscillator strength; 1.4.38 osmotic coefficient; 1.2.18a osmotic compressibility; IV.5.3d osmotic equilibrium; 1.2.34, IV.5.2 osmotic pressure; 1.2.34, 1.2.64, I.2.20d, IV.3.144, IV.5.46ff, V.I.5, V.I. 10, V.2.11ff, V.figs. 2.2-3, V.2.39 osmotic repulsion; 1.2.72, I.fig. 2.11 Ostwald equation; [1.2.23.25], II.1.19, [III.1.13.2] Ostwald ripening; 1.2.97, II.1.103, 11.3.110, IV.2.2e, V7.14ff, V.7.3c, V.8.3b Ostwald viscometer; IV.fig, 6.18 outer Helmholtz plane; II.3.17, II.3.59ff Overbeek equations (for retarded London-Van der Waals forces); [1.4.4.23a, b], 1.4.74 overcharging; H.fig. 3.20c, IV.3.9J, IV.3.164ff, IV.figs. 3.62-64, IV.fig. 3.67, IV.3.74 overflowing cylinder (in rheology); Ill.fig. 3.73, V.fig. 3.25 overpotential; 1.5.79 overrun (foams); V.7.7 oxides (ingeneral), contact angles on ...; III.table A4.3 double layer; 1.5.6a, 1.5.6b, II.3.71ff, Il.table 3.5, II.3.8, II.3.71ff, 11.3.10c point of zero charge; II.3.112, Il.app. 3b, Il.table 3.5 (for specific oxides, see under the chemical name) paints; 1.1.22, 1.1.28, IV.2.1ff, IV.3.185 pair correlation function; 1.3.66, II.3.5 Iff, IV.5.3 pair interaction; see interaction pair potential; see interaction Pallmann effect = suspension effect pancake (polymer ads.); V.I.11, V.fig. 1.27 paper electrophoresls; II.4.131 papermaking; IV.2.1 papyrus; IV. 1.3, fig. IV. 1.1 parachor; III.2.67 paramagnetism; IV.3.124
63
64
SUBJECT INDEX
parameter (def.), extensive; 1.2.10 intensive; 1.2.10 mechanical; 1.3.40 thermodynamic; 1.3.40 paraquat; [II.3.14.1] partial molar quantities; 1.2.46 particle-in-a-box problem; 1.3.23 particles (colloidal), form factor; 1.7.56,1.7.70ff interaction and rheology; IV.6.9, IV.fig. 6.26 networks; IV.6.14a shape; I.6.5g (2), 11.(3), 1.7.8c, I.7.8d, 1.7.69 size; 1.7.8, 1.7.26, 1.7.63, 1.7.67, IV.4.32 size distributions; see separate entry structure factor; I.7.64ff (see separate entry) also see: charged (colloidal) particles particle-wave duality; 1.7.5 partition; see distribution partition coefficients (micelles); V.4.9b partition function; I.3.2(intr.), 1.3.3, 1.3.4,1.3.5, Lapp. 6 canonical; I.3.2(intr.), 1.3.3,1.3.4, 1.3.5,1.3.51,1.3.54,1.3.59,1.3.63, Lapp. 6 for ideal gas; I.3.6b, I.3.6c for localized adsorbate; I.3.6d for subsystem; 1.3.5 grand (canonical); I.3.2(intr.), 1.3.3, 1.3.4, 1.3.31,1.3.33,1.3.54ff, 1.3.63, Lapp. 6, II. 1.95, II. 1.99 isobaric-isothermal; 1.3.13, 1.3.18,1.3.19 microcanonical; I.3.2(intr.), 1.3.3, 1.3.4 separable; 1.3.20 Pascal's law; 1.2.90, III.2.9 Paul! principle: 1.4.5, 1.4.42, II.3.172 PCS = photon correlation spectroscopy = QELS; see electromagnetic radiation Pearson's rule; II.3.185 Peclet number; 1.7.97, [IV.4.5.11], [V.8.1.18], [V.8.3.5], [V.8.3.21] pendant drop; see drop, pendant penetration depth (evanescent waves); [1.7.10.12] Percus-Yevick (PY) equations [1.3.9.29 and 30], [IV.5.3.40], IV.5.4b, IV.fig. 5.16 percolation (threshold); lV.5.86ff, IV.6.83 perikinetic coagulation (def.); IV.4.37
SUBJECT INDEX
65
period (of a wave); 1.7.4 permeability, of liquid films; V.6.2h of monolayers; III.3.239ff also see: porous plugs perpetual motion; 1.2.8 of second kind; 1.2.23 persistence length; see polymers/polyelectrolytes in solution for worm-like micelles; V.4.6d persistence parameter; see polymers/polyelectrolytes in solution persistence time; 1.5.45 PFM = polarized fluorescence microscopy phagocytosis; III.5.2, III.5.100 phase angle; Lapp. 8, see further loss angle phase diagrams of microemulsions; see V.chapter 5 phase diagrams of surfactants; V.4.1e, V.fig. 4.4 phase diagrams (2D); III.fig. 3.15, III.fig. 3.19, (see the K(A) curves in Ill.chapter 3) phase diagrams and nucleation; IV.fig. 2.3, IV.figs. 5.37-41, IV.fig. 5.43, IV.figs. 5.445.45, IV.figs. 5.49-53, IV.fig. 5.56, IV.fig. 5.62, IV.fig. 5.63c. phase integral; 1.3.57 phase inversion; V.5.2c, V.8.64 phase inversion temperature; V.5.6ff, V.8.5, V.8.49ff, V.fig. 8.14, V.8.89 phase rule (Gibbs); 1.2.13 phase separation, transitions and coexistence; 1.2.19, II.5.2e IV.5.7 in capillaries; II.1.6e in interfaces; III.2.18, III.3.3b, Ill.table 3.1, III.3.4d, III.3.217ff (also see: demixing, critical point, condensation: two-dimensional, polymers in solution and microemulsions) phase space; I.3.57(def.) phase stability (cone, colloids); IV.5.7 phase transitions (cone, colloids); IV.5.8 phenomenological approach; I.1.29(def.), 1.2.2 phosphate binding (soils); II.3.222 phospholipids, see lipids photobleaching; 1.7.103 photocatalysis; 0.3.222 photochromic probes; 1.7.103 photoconduction; II. 3.173 photocorrelation spectroscopy = QELS; see electromagnetic radiation photoelectric effect; 1.7.85 photographic emulsion: IV. 1.9
66
SUBJECT INDEX
photolysis of water; II.2.87, II.3.223 photons; 1.7.5, III.3.168 (counting) physisorption; II. 1.5, II. 1.18, II.1.3Off Pickering (emulsion stabilization); III.5.99, V.8.4 pipettes, emptying; III.fig. 1.8 p.i.t. = phase inversion temperature plant growth in arid regions; 1.1.1, 1.1.2 plastic behaviour (rheology); IV.6.3a Plateau border; 1.1.16, I.fig. 1.11, V.6.2, V.6.3, V.6.48, V.fig. 7.2 Plateau rules (foam structure); V.7.4 plumber's nightmare; V.4.18, see further microemulsions, bicontinuous pluronics (in micelles); V.4.4e point of zero charge; 1.5.90, I.5.6e, II.3.8, II.3.11, II.3.17, II.3.74, II.3.8, Il.app. 3, II.3.118, II.3.120, Il.fig. 3.61, II.3.162 experimental determination; II.3.8a influence of organic additives; II.3.12d, Il.figs. 3.77-80, Il.fig. 3.82, II.3.223 influence of specific adsorption; II.3.68ff, II.3.103ff, Il.fig. 3.34 interpretation; II.3.8c pristine; 1.5.102, II.3.8, II.3.103ff, Il.fig. 3.34, II.3.140, II.3.152 relation to isoelectric point; II.3.8b, Il.fig. 3.35 tabulation; app. II.3 temperature dependence; 3.75ff, II.3.115ff, Il.figs. 3.36-37 Poiseuille's law; see Hagen-Poiseuille's law Poisson-Boltzmann equation; 1.4.16, 1.5.18, [II.3.5.6], [II.3.5.44], [II.3.5.57ff] Poisson-Boltzmann theory; II.3.6a for flat interfaces = Gouy-Chapman theory for low potentials = Debye-Hiickel theory improvements; 1.5.2c, II.3.6b, Il.figs. 3.18-19 Poisson distribution; [IV.2.3.46] Poisson's law (electrostatics); 1.4.53, 1.5.10, [I.5.1.20-20a], II.3.19, II.3.35, [II.3.5.43], [II.3.6.14], II.3.211, II.4.18, II.4.70, [II.4.6.12]ff, II.4.115, II.5.55 Poisson ratio; IV.6.9 polarimeters; 1.7.99 polarizability (intr.); I.4.22ff, I.4.4d, I.4.4e, 1.7.18, 1.7.53, 1.7.94 data for molecules; I.table 4.2 molar; 1.4.24 polarization, dielectric (phenomenon); 1.4.4b, I.4.4e, I.4.5f of colloids; see dielectric dispersion of sols of interfaces; 1.5.5b, II.3.9, (see potential difference, %) polarization, dielectric (physical quantity); I.4.5f, 1.7.2, 1.7.6, IV.3.10 polarization (of radiation, etc.); see electromagnetic radiation
SUBJECT INDEX
polarized fluorescence microscopy; III.3.7c.iv polarizer; I.fig. 7.7, 1.7.99 polarizer-sample-analyzer; III.3.7b.i polar molecules; 1.4.4b poly(phenylene), V.fig.2.3 polyampholytes; II.5.13, V.2.1, V.2.6d polydispersity; 1.1.13, IV.2.2d, IV.figs. 5.15-17, IV.fig. 6.27 relative; IV.2.61, IV.app. 1 also see, size distributions polyelectrolytes in solution; 1.1.6, II.5.2f, V.chaper 2 (general) annealed; V.2.2 brushes; V.2.3c chain statistics; V.2.3 charge; II.5.14ff, V.2.2, V.fig. 2.4 chemical composition; V.2.1b, V.table 1 colloid flocculation; V.2.7a colloid stabilization; V.2.7b complex coacervation; V.2.6e, V.2.6f (compl. coac. micelles) complexation; V.2.6 concentration regimes; V.2.3, V.fig. 2.15 conductivity; V.2.5 configurations; V.2.3 dielectrics; V.2.5d dissociation; V.2.2d electrokinetics; V.2.5a gels; V.2.3d grafts, see adsorption of polyelectrolytes interaction; V.2.15, V.fig. 2.5 multilayers, see adsorption of polyelectrolytes persistence length; II.5.14, Il.fig. 5.5 phase separation; V.2.6 quenched; V.2.2 solubility; V.2.6a viscosity; V.2.4 polyelectrolyte adsorption; see adsorption of polyelectrolyte effect; II.3.71, II.3.76, V.2.2d polyelectrolyte gels; V.2.3d polyhedral (foams); see foams polymer brush, monolayers; III.3.8f, III.figs. 3.96-98, III.fig. 3.100, V.I.11 polymer colloids; see latex, latices
67
68
SUBJECT INDEX
polymer melt near a wall; II.5.45ff polymer surfactants; V. 1.51 polymers in solution; I.1.2(intr.), I.fig. 1.17, 1.3.34, IV.6.11-12; V.I.2 concentration regimes; II.5.9ff, Il.fig. 5.3, IV.6.11-12 conformation; Il.figs. 5.1-2, II.5. Iff, IV.fig. 6.30 end-to-end distance; II.5.4, Il.fig. 5.2, IV.6.61 entanglements; IV.6.67ff excluded volume (parameter); II.5.3, II.5.2b, II.5.5b, IV.6.62, [IV.6.11.5] expansion coefficient; II.5.6ff, IV.6.62 Flory-Fox constant; IV.6.63, V.2.51 ideal chain; II.5.3, II.5.2a intrinsic viscosity; IV.6.11 light scattering; I.7.56ff, I.7.62ff molecular mass; 1.7.26, I.7.62ff, I.7.68ff, IV.6.11, also see, size distributions networks; IV.6.14a overlap; II.5.2c, V.chapter 1 persistence (stiffness) parameter; II.5.4 phase separation; Il.fig. 5.3, II.5.2e, Il.fig. 5.4, V.2.3a reptation; IV.figs. 6.36-37 solvent quality; II.5.2b swollen chain; II.5.3, II.5.2b, II.5.9 thermodynamics; II.5.2c viscous flow; IV.6.12 (also see: radius of gyration, adsorption of polymers, colloid stability) polymers, contact angles on ...; III.table A4.2 poly(acrylic acid), adsorption; V.fig. 2.19 charge; V.fig. 2.12 transference number; V.fig. 2.30 viscosity; V.fig. 2.24 polyfacryl amide), adsorption on montmorillonite; Il.fig. 5.39b poly(diallyl dimethylamm. chloride), conductivity; V.fig. 2.31 poly(ethylene imine) (charge); V.fig. 2.8 poly(ethylene), AFM image; Il.fig. 1.3 poly(ethylene oxide or oxyethylene). adsorption on latex and SiO2; Il.fig. 5.25 adsorption of poly(styrene sulfonate); Il.fig. 5.35
SUBJECT INDEX
polyfmaleic acid) (charge): V.fig. 2.9 poly(methacrylic acid), adsorption on silver iodide; II.fig. 5.36-37 charge; V.figs. 2.11, 2.12 polyfmethacrylic ester), raonolayer; III.3.8e, III.figs. 3.94-95 poly(oxyethylene) (PEO), influence on interaction; V.figs. 1.43-45 surface pressue; V. 1.77 poly(phenylene); V.fig. 2.3 poly(styrene), adsorption on silica; II.fig. 5.28, II.fig. 5.30 mixtures with coated silica; IV.fig. 5.64 poly(styrene) latex, adsorption of CgifiP^E^ (non-ionic); Il.fig. 2.32 adsorption of Cg(|)P/13>E/27\ (non-ionic); Il.fig. 2.33b adsorption of lactalbumin; V.figs. 3.15-18 adsorption of poly(ethylene oxide or oxyethylene); Il.fig. 5.25a coagulation; IV.sec.3.13c, V.figs. 1.47-50 rheology; IV.figs. 6.28-29 poly(styrene sulfonate), adsorption on poly(oxymethylene); Il.fig. 5.35 brushes and stability; V.fig. 1.36 conductivity; V.fig. 2.32 diffusion; V.fig. 2.17 overlap cone; V.fig. 2.16 stabilization of mica; V.fig. 1.42 viscosity; V.fig. 2.23, V.fig. 2.25-28 polyfstyrene co 2-vinyl pyridine), AFM image; Il.fig. 1.4 polyfvinyl alcohol) (PVA); V.fig. 3.24 poly(vinyl chloride), flow behaviour; IV.fig. 6.31 poly(vinylpyrrolidone), adsorption on SiO2; Il.fig. 5.22 influence on flocculation; V.fig. 1.46 pores (in surfaces); II.1.6 ad- and desorption of gases; II.figs. 1.32-35 classification into macro-, Il.meso- and micropores; II.1.6a connectivity; II. 1.82 mesopore filling; II. 1.6b micropore filling; 11.1.82, II.1.6c molecular dynamics: Il.fig. 1.38
69
70
SUBJECT INDEX
pores (In surfaces) (continued), radius (effective); II. 1.84 size distribution; II. 1.85, II. 1.88 volume; II. 1.84ff (also see: porosity, hysteresis, capillary condensation) porosity, mesoporosity; II. 1.6b microporosity; II. 1.6c of plugs; I.6.50ff of surfaces; II.1.6, II.2.67, II.3.161 by mercury penetration; II. 1.90, II. 1.100 classification; II. 1.6a (also see: adsorption hysteresis, capillary condensation, pores (in surfaces)) porous plugs, permeability; I.6.4f, II. 1.90, II.4.55ff, II.fig. 4.18, IV.2.32ff, IV.fig. 4.26 electro-osmosis; II.4.3b, Il.fig. 4.6, II.4.5b, Il.fig. 4.18 other electrokinetic phenomena; II.4.5b, Il.fig. 4.18-19, II.4.7, Il.figs. 4.34-35 wetting; III.5.4i, III.5.9 porous surfaces; see pores (in surfaces) potential-determining ions; see surface ions or charge-determining ions potential difference (between adjacent phases); I.table 5.13, II.3.7b, II.3.138 X, 1.5.73-74, II.2.19, II.3.91, II.3.102, II.3.115, II.3.9, Il.figs. 3.38-39, Il.table 3.7, II.3.179, Il.fig. 3.75, Il.table 3.9, II.3.200ff, Il.fig. 3.79, III.2.47, [III.3.7.22], III.4.4a,b, III.fig. 4.20, Ill.fig. 4.24 electrokinetic ((); see electrokinetic potential Galvani; 1.5.5a, 1.5.5c, II.3.14, II.3.90, II.3.119ff, Il.fig. 3.38, II.3.138 liquid junction (diffusion); I.5.5d, I.fig. 5.12, 1.6.7b real; 1.5.75, II.3.121ff, Il.table 3.7 Volta; 1.5.5a, II.3.119ff, Il.fig. 3.38, Il.figs. 3.74-75, II.3.179, III.3.7f, Ill.fig. 3.75, Ill.fig. 3.76, Ill.fig. 3.85, Ill.fig. 3.88, Ill.fig. 4.14 (also see: suspension effect) potential of a force; 1.4.3b electric; 1.4.12, 1.5.3, 1.5.7ff, 1.5.10, I.fig. 5.1, II.3.3 interfacial; 1.5.5, II.3.6b in diffuse layer, Stern layer etc.; see there of mean force vs. mean potential; I.4.3c, 1.5.18, I.5.24ff, I.5.27ff, II.3.51ff, IV.5.2b potentiometric titration (of colloids); see colloid titration pouring (foaming); V.7.13 powder technology; II.5.97 powders (wetting); III.5.4b, III.5.9 Poynting vector; 1.7.5, 1.7.10, 1.7.97
SUBJECT INDEX
precursor film; III.5.8, III.fig. 3.35 preparation of colloids; IV.chapter 2 kinetics; IV.2.2c preparation of emulsions; V.8.2 pressure, two-dimensional; see surface pressure pressure tensor; III.2.3, IV.6.1, V.6.3b prefixes (table); Lapp. 2 primary minimum; see colloids, interaction primary structure (proteins); V.3.3 primitive (liquid model); I.5.1(def.) in conduction; 1.6.79 in diffusion; 1.6.56 in hydrodynamics; 1.6.Iff in solvation; 1.5.3b principal axes (in optics); 1.7.98 principal radii of curvature; see curvature, radius of probability; 1.3.1, I.3.2d, 1.3.3 probability distributions; I.3.2d, I.3.7b, IV.app. 1 process; I.2.3(def.) endothermic; 1.2.8 exothermic; 1.2.8 isobaric; 1.2.3 isochoric; 1.2.3 isosteric; 1.2.3 isothermal; 1.2.3 natural (or irreversible); 1.2.4, 1.2.8 reversible; 1.2.3,1.2.21 spontaneous; 1.2.4, 1.2.8 stochastic; 1.6.3 (also see: transport) protection, of colloids against aggregation; I.1.2(intr.), 1.1.27 of colloids against aggregation by (bio-) polymers; V.chapter 1 see further; colloid stability proteins; 1.1.23 conformation; V.3.2a, 2b relaxation at Interfaces; V.3.3b structure changes upon adsorption; V.3.4 structures in solution: V.3.2, V.table 3.3 proton acceptor; 1.5.65 proton donor; 1.5.65
71
72
SUBJECT INDEX
Prussian blue; IV.2.2 PSA = polarizer-sample-analyzer pseudoplastic behaviour (rheology); IV.6.3a pullulan (radius of gyration); V.fig. 2.14 pulmonary surfactant = lung surfactant purity criteria (of interfaces); III. 1.7, HI. 1.14c PY = Percus-Yevick p.z.c. = point of zero charge QELS = quasi-elastic (light) scattering; see electromagnetic radiation, scattering quadrupole moment; 1.4.19, 1.5.42 quartz; see silica, etc. quasi-chemical approximation (in statistical thermodynamics); I.3.8e quasi-elastic (light) scattering (QELS); see electromagnetic radiation, scattering quaternary structure (proteins); V.3.3ff quenched (polyelectrolytes); V.2.2 quenchers; III.3.165 radial distribution functions; see distribution function radiant intensity; 1.7.5 radiation; see electromagnetic radiation radius, Guinier; IV.2.45, IV.A1.4 hydraulic; 1.6.50 hydrodynamic (viscometric); 1.7.51; IV.6.9, [IV.6.9.2], IV.6.13, IV.A1.4 (of) gyration; 1.7.57, II.5.4, II.5.6ff, IV.6.11, V.2.3, V.fig. 2.14 ionic; I.table 5.4,1.fig. 5.7 Raman scattering; see electromagnetic radiation scattering Raman spectroscopy; 1.7.12, III.3.7c.ii Randies circuit; Il.fig. 3.31 random coil (intr.); II.5.3 random flight or random walk; 1.3.34, 1.6.3, II.5.3ff, Il.fig. 5.2, II.5.24, IV.4.2 random phase approximation (interaction); IV.5.50ff, IV.fig. 5.30 random sequential adsorption; V.3.16ff Raoult's law for vapour pressure lowering; 1.2.74, II. 1.70 rate of coagulation; IV.4.3 rate of strain; see strain tensor, [IV.6.1.5] Rayleigh-Brillouin scattering; see electromagnetic radiation scattering Rayleigh-Debye (-Gans) scattering; I.7.8d, 1.7.67 Rayleigh instability: III.5.1 Id, III.fig. 5.47 Rayleigh line; 1.7.44 Rayleigh ratio; [I.7.7.6](def.), 1.7.3 (table) recipes for sol preparation; IV.2.4
SUBJECT INDEX
red shift (of spectra); 1.7.19 reference electrode; 1.5.82 reference state; see state, standard reflection, multiple; 1.7.80, II. 1.18 total; 1.7.74 reflection angle; 1.7.72 reflection at interfaces; 1.7.10a, Ill.fig. 3.57 reflection coefficient; 1.7.73, II.2.50 reflection electron spectroscopy; see scanning electron spectroscopy reflectometry; 1.7.10b, II.2.5c, Il.figs. 2.15-16, II.5.64, III.2.47, V.figs. 6.4-6.5 refraction angle = transmission angle refraction by interfaces; 1.7.10a refractive index; 1.7.12,1.7.14 complex; 1.7.2c, 1.7.61, 1.7.98 regular solutions; 1.2.18c, II.2.30, II.5.8 regulation (of double layers), see colloids, interactions relaxation (time); I.4.4e, I.6.6c, II.3.13, II.4.10ff, II.4.6c, II.4.8, IV.4.4, IV.table 4.3 adsorbed proteins; V.3.3b colloid interaction; IV.4.4 Debye; 1.6.73 dielectric; see dielectric relaxation double layers; see there (in) external fields; IV.4.5 Maxwell (-Wagner); 1.6.84, II.3.219, Il.fig. 3.89, II.4.111, IV.4.23 mechanical; IV.6.4 of interfaces (electric); 1.5.5b in Langmuir monolayers; III.3.6h, III.figs. 3.46-47 retardation (in ionic conduction); 1.6.6b, 1.6.6c thermodynamic; 1.2.3 (also see: diffusion, rotational, double layer, relaxation) reptation; IV.6.69ff, IV.figs. 6.36-37 repulsion, electric; 1.1.2Iff osmotic; 1.2.72 (see further: colloids, interaction) residence time, and adsorption; II.1.46ff, [11.2.4.1] and hydration; 1.5.53 resonance band = absorption band resonance (electric): I.4.4e
73
74
SUBJECT INDEX
resonance frequency; 1.4.34 resonators; 1.7.4a respiratory distress syndrome (RDS); HI.3.238, V.6.87-88 retardation (of dispersion forces); see Van der Waals forces retention volume; see chromatography reversible, reversibility (in thermodynamic sense); 1.2.3, 1.2.9, 1.2.8 colloids = colloids, lyophilic interfaces (in electrical sense); 1.5.5b (also see: process; for adsorption reversibility, see (adsorption) hysteresis) Reynolds eq. (film thinning); see Stephan-Reynolds eq. Reynolds limit (wave damping); Ill.fig. 3.44, III.3.117ff Reynolds number; I.6.4b, I.table 6.2 rheology (general); IV.chapter 6 rheology; III.3.6b, IV. 1.3 (and) colloid interaction; IV.2.41, IV.3.144, IV.6.13 compliance; IV.6.23ff concentrated dispersions; IV.6.10 creep compliance; IV.6.24 creep; IV.6.6b, IV.fig. 6.12 definition; IV.6.1 descriptive; IV.6.3 dilute sols; IV.6.9 distribution functions; IV.3.144 electroviscous effects; IV.6.9b emulsions; V.8.15ff foams; V.7.5a (and) fractals; IV.6.13 fracture; IV.6.5, IV.figs. 6.8-6.9 (of) gels; IV.6.14 instrumentation; see measurements Kelvin element = Voigt element Maxwell element; IV.6.16, IV.fig. 6.11, IV.6.20-21 Maxwell modulus; IV.6.16 measurements; IV.6.6 constant strain rate; IV.6.6c creep; IV.6.6b dynamic; see oscillatory oscillatory; IV.6.6d stress relaxation; IV.6.6a also see, viscometers overshoot; IV.6.6c
SUBJECT INDEX
rheology (continued), particle networks; IV.6.14b polymer networks; IV.6.14a polymer solutions; IV.6.11, IV.6.12 principles; IV.6.1 quantities; IV.6.2 structure; IV.6.8 time scale effects; IV.6.4 Voigt element; IV.flg. 6.13 yield; IV.6.5 yield stress; IV.fig. 3.73, IV.6.3a yield value = yield stress also see: interfacial rheology, stress, strain rheopexy = antithixothropy ribonuclease; V.fig. 3.3, V.fig. 3.5 adsorption; V.fig. 3.23 rigid (films); V.6.48 ring trough (interfacial rheology); III.fig. 3.71 ringing gels; IV.2.41 RIS = rotational isomeric state RNase = ribonuclease Ross Miles test (foams); V.7.13ff rotating molecule; I.3.5e rotational correlation time; see correlation time rotational isomeric states (RIS); V.4.38 rotational diffusion (coefficient); see diffusion Rouse-Zimm theory (polym.); IV.6.64-65. Rowlinson-Widom equation (for surface tension); [III.2.5.40] RSA = random sequential adsorption RSD = respiratory distress syndrome rubber, adsorption on carbon black; II.fig. 5.31 rupture (of films); see films, stability ruthenium dioxide, double layer; II.fig. 3.56, Il.fig. 3.59 point of zero charge; Il.app. 3b XPS (= ESCA) spectrum; Il.fig. 1.5 rutile; see titanium dioxide saddle splay modulus = Gauss modulus Sackur-Tetrode equation; [1.3.1.9], 1.5.35, [III.2.9.12] (for surfaces), III.3.37 salt-sieving; 1.1.1, 1.1.3, 1.1.2Iff. II.3.28. 11.3.223, II.4.56. IV.1.6
75
76
SUBJECT INDEX
salting-out; 1.5.71 sal ting-in; 1.5.71 Sand equation; [1.6.5.23], I.flg. 6.15b SAMS = self assembled monolayers; III.3.240 SANS = small angle neutron scattering saponine (films); V.fig. 6.27 Saxen's rule (electrokinetics); 1.6.17, II.4.2 SAXS = small angle X-ray scattering SCAF = self-consistent anisotropic field scaling theory; II.5.11, II.5.4c, III.3.8e,f scanning electron microscopy (SEM); 1.7.lib, II.fig. 1.1 scanning, optical; III.3.7c.iv, III.3.7d scanning probe microscopy (SPM); III.table 3.5 (includes STM and AFM), III.3.7d scanning transmission electron microscopy (STEM); 1.7.lib scanning tunnelling microscope (STM); 1.7.90, II. 1.12, III.3.7d Scatchard plot; II. 1.48 scattering, length density; 1.7.70, II.5.66 length; 1.7.70, II.5.66 plane; I.7.27(def.) from surfaces; III. 1.10 wave vector; I.7.27(def.), III. 1.54 scattering of, neutrons; see neutron scattering, small angle neutron scattering (SANS) radiation; see electromagnetic radiation X-rays; see X-ray scattering Schiller layers; IV.2.40 Schottky defects; II.3.173 Schulze-Hardy rule (for coagulation of colloids); 1.5.67, 1.6.83, II.3.129ff, IV. 1.11, IV.3.9f SCF = self-consistent field Schrodinger equation; 1.3.1, 1.3.11, I.3.20ff, V.I.7 screening (of charges); 1.5.11, also see: double layer charge, capacitance, etc. Searle viscometers/rheometers; IV.6.7b second harmonics, generation; II.2.55, III.3.7c.v, III.figs. 3.64-65 second central moment; 1.3.35, IV.app. 1 Second Law of thermodynamics; see thermodynamics Second Postulate of statistical thermodynamics; see statistical thermodynamics second virial coefficient; see virial coefficient secondary ion mass spectroscopy (SIMS); 1.7.lla, I.table 7.4. II. 1.15, II.fig. 1.6
SUBJECT INDEX
secondary minimum; see colloids, interaction secondary structure (proteins); V.3.3 sediment, sedimentation; 1.1.2, 1.1.22, I.fig. 1.14, 1.6.48, IV. 1.2, IV.2.40ff, IV.2.3d, V.8.3d, V.8.23, V.fig. 8.24 equilibrium; IV.fig. 5.11, V.8.75 groupwise; V.8.77 hindered; V.8.76 sedimentation coefficient; IV.2.51 sedimentation current; II.4.24 sedimentation-diffusion equilibrium; IV.2.52ff sedimentation field flow fractionatin (FFF); IV.2.61ff sedimentation potential (gradient); II.table 4.1, II.4.6-7, II.4.3c sedimentation profiles; IV.2.3d, V.8.80 seeding, seeds (in nucleation); 1.2.100, IV.2.2f segment weighting factors; II.5.37ff selection rules, infrared; Liable 7.5 Raman; Liable 7.6 self-assembly; see IILchapter 3, V.chapter 4 self-avoiding walk; II.5.6 self-consistent anisotropic field (SCAF); V.4.39 self-consistent field (SCF) theory; II.5.29, II.5.5, V.1.4, V.appendix 1 association colloids; V.chapter 4 polymers; II.5.29, II.5.5, V.1.4 self-diffusion; see diffusion self Gibbs energy; see Gibbs energy self-similarity, in Brownian motion; 1.6.18 in fractal structures; IV.6.7Iff in scaling theory; II.5.34 SEM = scanning electron microscopy semiconductors; II.3.10e double layer; II.3.1Oe, Il.flgs. 3.60-72 intrinsic; II.3.170 n-type andp-type; H.3.173 semidilute (polymer) solution; II.fig. 5.3, II.5.2d settling ~ sedimentation SER = surface enhanced Raman spectroscopy SF = Scheutjens-Fleer (polymer adsorption theory) SFG = sum frequency generation
77
78
SUBJECT INDEX
SFM = scanning force microscopy, (including scanning, probe microscopy, scanning tunnelling microscopy); IV.3.12c, IV.table 3.5, IV.3.58 also see AFM = atomic force microscopy shaking (foam formation); V.7.13 shear rate; 1.6.32, IV.fig. 6.2, IV.fig. 6.3, IV.6.2, IV.table 6.1 shear stress; see stress shear thickening, thinning; III.3.87, IV.fig. 6.5, IV.6.3a SHG = second harmonic generation Shinoda cut (microemulsions); V.5.42, V.fig. 5.22b), V.5.46 silica, silicium dioxide, adsorption of poly(oxyethylene); Il.fig. 5.25b adsorption of poly(styrene); Il.fig. 5.28, Il.fig. 5.30 adsorption of poly(vinyl pyrrolidone); Il.fig. 5.22 crystals; fig. IV. 14 point of zero charge; Il.app. 3b Aerosil, adsorption of Cg
SUBJECT INDEX
79
silica sols; IV.fig. 2.1a, IV.2.4, IV.fig. 2.8, IV.3.13b, IV.fig. Al.l Ludox; IV.figs. 3.71-72 mobility; IV.fig. 3.68, IV.fig. 3.72, IV.fig. 3.74 preparation; IV.2.4a stability and structure; II.3.161-2, IV.3.13b, IV.fig. 5.13, IV.fig. 5.17, IV.figs. 5.24-25, IV.fig. 5.35, IV.fig. 5.62 Stober; IV.2.63 silicium dioxide-zirconium dioxide catalyst, SIMS spectrum; Il.fig. 1.6 silver bromide, point of zero charge; Il.app. 3c silver iodide, adsorption of alcohols; Il.figs. 3.77-79, Il.table 3.9 adsorption of dextrane; Il.fig. 5.26b, II.5.80ff, Il.fig. 5.29 adsorption of poly(methacrylic acid); Il.figs. 5.36-37 adsorption of tetraalkylammonium salts; Il.figs. 3.80-81 double layer; II.3.8, Il.fig. 3.28, Il.fig. 3.32, Il.table 3.6, II.3.10a, Il.figs. 3.40-46, II.3.112, Il.figs. 3.52-53, Il.fig. 3.56, Il.table 3.8, II.3.202ff, Il.figs. 3.77-81, Il.table 3.9 electrokinetic charge; Il.fig. 4.13 electrosorption; II.3.12d, Il.fig. 3.77-81, Il.table 3.9 negative adsorption of ions; Il.fig. 3.40 point of zero charge; II.3.1 lOff, Il.fig. 3.36, Il.figs. 3.41-43, Il.figs. 3.77-80, Il.app. 3c, Il.fig. 5.37 relaxation of double layers; IV.fig. 4.9 sols;IV.2.16 site binding (adsorption); II. 1.47-48, II.3.6e, II.3.159, Il.fig. 3.63 SIMS = secondary ion mass spectroscopy single ionic activities; 1.5.lb (also see: activity coefficient) size distributions and averages; IV. 1.9, IV.2.2d, IV.2.45ff, IV.2.3f, IV.2.61ff, IV.app. 1, V.8.4, V.8.1e, V.table 8.1, V.8.66 relative dispersity; IV.app. 1 self-sharpening; IV.2.17ff viscometric; IV.6.63 sky (blue colour); 1.3.34, 1.7.25 slip plane, slip process: 1.5.75, II.4.1b, Il.fig. 4.3 interpretation; II.4.4, V.2.5a sludges; 1.1.23 small-angle neutron scattering (SANS); 1.7.9b, IV.fig. 5.25, IV.fig. 5.31, IV.fig. 5.33, IV.fig. 5.36, V.5.3d, V.fig. 5.18, V.fig. 5.19 small-angle X-ray scattering (SAXS); 1.7.9a
80
SUBJECT INDEX
smectite, wetting; Il.table 1.3, II.3.165 smoke; 1.1.6 Smoluchowski eq. (electroviscous effect); [IV.6.9.15] Smoluchowski's theorem (electrokinetics); II.4.21-22 Smoluchowski's theory (coagulation); IV.4.3a Snell's law; 1.7.11,1.7.72 soap bubbles; I.fig. 1.4,1.figs. 1.9-10 soap films; see films, liquid sodium dodecylsulphate; see surfactants, anionic sodium laurylsulphate = sodium dodecylsulphate soft depletion; V . l . l l h soils (permeation in); 1.1.2,1.1.3,1.1.22ff, 1.1.28 soil structure; IV.3.184 sol; I.1.5(def.) ageing; 1.2.99, IV.fig. 2.8 colour; I.7.60ff, IV.2.39ff preparations; IV.2.4 sol-gel processing; IV.2.37 solar energy conversion; II.3.223 solid surfaces and interfaces, characterization; II.1.2 solid-liquid; II.2.2, Il.figs. 2.4-7 thermodynamics; 1.2.24 (for adsorption, double layers etc. see there; also see, interfacial tension of solid surfaces) solubility, of colloids; IV.2.2e, IV.fig. 2.7 of gas in liquid; I.2.2Ob of liquid in liquid; 1.2.20c of monomers in microemulsions; V.5.2f of small drops and particles; 1.2.23c of solid in liquid; I.2.20c of surfactants in microemlusions; V.chapter 5 solubilization; IV. 1.5, V.4.9b solubility parameter (Hildebrand); 1.4.47 solutions, principles, (ideally) dilute; 1.2.17c, 1.2.20 non-ideal; 1.2.18 also see regular solutions solvation; 1.2.58, 1.4.42, 1.4.5c. 1.5.3
SUBJECT INDEX
solvent, quality of; 1.1.7, I.1.25ff, I.fig. 1.17,1.1.30,11.5.2, Il.fig. 5.24 solvent structure; 1.5.3, 1.5.4 near surfaces; see distribution functions of liquids near surfaces (see; interactions, solvent structure-mediated) space charge (density); 1.5.9, Il.fig. 3.70 speciation; I.5.2(def.) specific binding, specifically bound charge; 1.5.3; see further, double layer, Stern specific vs. generic (properties, phenomena); 1.5.67, II.3.6 specific adsorption; I.5.6d, I.5.104ff, II.3.6, Il.fig. 3.20b, II.c, II.3.6d, II.3.6e, IV.3.9i criteria for absence or presence; 1.5.102, II.3.108ff, II.3.132 first and second kind; II.3.64 site binding models; II.3.6e speckle pattern; IV.2.46 spectral density; 1.7.34 (intr.) spectroscopy, of adsorbed proteins; V.3.4a of surfaces; 1.7.11,1.table 7.4, II.2.7ff, Ill.table 3.5, III.3.7c also see, microscopies spectrum analyzer; 1.7.37 (intr.) spin-echo techniques; 1.7.102 spin (electronic); 1.7.16, 1.7.13 spin labels; 1.7.13 spin-lattice relaxation; 1.7.96 spin (nuclear); 1.7.16, 1.7.13 spinning drop; see interfacial tension, measurement spinodal; 1.2.68, II.5.12, Il.fig. 5.4 decomposition; IV.2.8, IV.figs. 2.3-4, IV.5.65ff, IV.fig. 5.41, IV.fig. 5.62 SPM = scanning probe microscopy spin quantum number; 1.7.16 spin-spin relaxation; 1.7.96 spontaneous; see process spontaneous curvature; III.1.15, III.4.7 spreading; 1.1.8, III.3.2 rate of; III.3.12 spreading coefficient = spreading tension spreading parameter = spreading tension spreading tension; III.3.8, III.5.4, III.5.6. [III.5.1.1], III.5.15ff spring (and dashpot); III.fig.3.50. IV.6.20ff see further; (interfacial) rheology, Maxwell element and Kelvin (Voigt) element sputtering; II. 1.1 10
81
82
SUBJECT INDEX
square gradient (across interfaces); [III.2.5.28], [III.2.5.30], III.2.28ff, III.2.36, V.1.7, V.1.4e square gradient method (polymer adsorption); II.5.33ff, V.1.4e stability, stabilization, thermodynamics; 1.2.7,1.2.19 (see colloids, stability of, emulsions, stability; steric stabilization, state) stability ratio; IV.figs. 3.65-67, IV.fig. 3.71, IV.4.17, IV.fig. 4.14, IV.fig. 4.21 stagnant layer (electrokinetics); II.2.15, II.4.1b, II.4.4, II.4.128ff, V.2.5a standard deviations; I.3.7a, I.6.19ff, [II. 1.5.11], IV.5.33 starch; 1.1.2 Stark effect (for spectral lines); 1.7.16 state (def.), metastable; 1.2.7 molecular; 1.2.3, 1.3.2 stable; 1.2.7 standard; 1.2.4 thermodynamic; 1.2.3 (also see: equation of, function of) state variables; 1.2.3 stationary state; 1.6.8, 1.6.13, 1.6.15, II.4.2, II.4.6 statistical chain element; I.3.5f, II.5.5 statistical mechanics; see statistical thermodynamics statistical thermodynamics; I chapter 3 classical; 1.3.9 postulates; 1.3.2, I.3.1d (also see: adsorption isotherm, Fermi-Dirac, Maxwell-Boltzmann, interfacial tension: interpretation, self-consistent field theories and the various applications) Stefan-Ostwald rule (for surface tensions); [III.2.11.25] Stephan-Reynolds eq. (for film thinning); [V.6.4.2 and 3] STEM = scanning transmission electron microscopy step-weighted lattice walk; 1.6.28, II.5.5 steric stabilization; see colloid stability (by polymers) Stern layer; see double layer stiffness parameter = persistence parameter, see polymers in solution Stirling approximation; [1.3.6.5] STM = scanning tunnelling microscope Stober sols (silica); IV.2.63ff stochastic; see processes, forces Stockmayer-Fixman equation (vise); [V.2.4.11] Stokes' law; [1.6.4.30], 1.6.56, II.4.18, II.5.62, V.8.74ff Stokes limit (wave damping); III.fig. 3.44. III.3.117ff
SUBJECT INDEX
stopping mechanism (in micelle formation); V.4.8ff strain; III.3.6b, IV.6.2, IV.6.1 strain rate; III.3.85ff strain tensor; [IV.6.1.4] strain energy release; IV.fig. 6.9 strain hardening; IV.6.12 strain (rate) thinning; IV.6.12 stratification (in films); see films, liquid streaming, or flow birefringence; 1.7.97,1.7.100 streaming current; I.6.16ff, Il.table 4.4, II.4.7, II.4.3d, Il.fig. 4.8, II.4.55 streaming potential; I.6.16ff, Il.fig. 3.78, Il.table 4.4, II.4.7, II.4.3d, Il.fig. 4.8, II.4.5b, Il.fig. 4.30, Il.fig. 4.35, II.5.63 stress; IV.fig. 6.4 normal; 1.6.7, III.3.6b, IV.6.1, IV.fig. 6.1 shear; 1.6.7, 1.6.1 Iff. I.6.4a, III.3.6b, IV.fig. 6.1, V.8.2 stress overshoot; V.fig. 7.11 stress relaxation; IV.figs. 6.9-10, IV.6.6a modulus; IV.6.20-21 spectrum; IV.6.20-21 stress tensor; I.6.6ff, [III.3.6.1-2], [IV.6.1.1] also see: interfacial rheology stretching of solid surfaces; 1.2.103 structural forces; 1.4.2, II. 1.95 structure breaking; 1.5.38,1.5.3d structure factor; 1.3.67, 1.7.64, Lapp, l i e , IV.3.143, IV.5.3, IV.5.21, IV.figs. 5.5-5.6, IV.figs 5.16-18, IV.figs. 5.21-23, IV.5.49, [IV.5.6.8], IV.fig. 5.30, IV.figs. 5.33-36, V.2.26, structure of colloids; IV.chapter 5 rheology; IV.5.1, IV.6.8-10 structure of water; see water, structure structure promotion; 1.5.38, 1.5.3d substantial derivative; 1.6.5 subsystems (statistical thermodynamics); 1.3.5, 1.3.6 dependent; 1.3.5, 1.3.20,1.3.8 independent; 1.3.5, 1.3.20, 1.3.6 sulphur sol (preparation); IV.2.4b sum frequency generation; III.3.7c.v superadditivity (in coagulation); IV.3.9k supercooling; 1.2.23d supermolecular fluids; 1.7.63 supersaturation; 1.2.23d. IV.2.9ff, V.7.8
83
84
SUBJECT INDEX
supersaturation ratio; [IV.2.2.14], IV.2.15 surface; I.1.3(lntr.), acidity/basicity (dry surfaces); II. 1.18 characterization in general; II.1.2, II.table 1.1, II.2.2a external vs. internal; II. 1.6a heterogeneity; I.1.18(intr.), 1.5.106, II. 1.5, II.1.7, II.2.29, II.3.83 patchwise vs. random; II. 1.103ff 'high' vs. 'low' energy; II. 1.35 hydrophilicity/hydrophobicity; II. 1.19, II. 1.35, II.2.7, II.2.87, II.3.130 imaging techniques; 1.7.lib modulus; see interfacial rheology porosity; see porosity of surfaces reconstruction; II. 1.8 scattering by; 1.7.10c spectroscopic characterization; 1.7.11, II.1.9ff, II.figs. 1.1-6 (also see: interface, especially for 'wet' surfaces, equations of state) surface (or interfacial) area, (molecular, in monolayers), III.3.15, III.3.24, see further the TI(A) isotherms in Ill.chapter 3, Ill.fig. 3.16, Ill.fig. 3.82, III.3.84, V.5.3h(ii), V.fig. 5.24 specific; I.1.18(def.), 1.1.20, I.6.4f, II.1.5f, II.2.67, II.2.73, II.3.7e, II.3.127, II.3.131ff, IV.fig. 2.8, IV.fig. 2.10, IV.2.33ff, IV.2.3c surface charge (density); 1.1.20, 1.5.3, 1.5.9, 1.5.6, II.3.3, II.3.21, Il.fig. 3.18, Il.fig. 3.28, Il.fig. 3.41, Il.fig. 3.52, Il.figs. 3.56-59, Il.figs. 3.63-65, Il.figs. 3.69-70, Il.fig. 3.77, Il.fig. 3,.80, Il.fig. 3.82-83, IV.fig. 3.75 determination; I.5.6e, II.3.7a dipolar contribution; II.3.126 discrete nature; II.3.46, II.3.6e for clay mineral-type particles; II.3.10d for polarized interfaces; 1.5.6c, II.3.10b, II.3.163 for relaxed (reversible) interfaces; 1.5.6a, 11.3.10a, II.3,10c, II.3.163 formation thermodynamics; II.3.1 lOff Gouy-Stern layer; Il.figs. 3.23-25 relation to D and E; I.4.53ff site-binding models; II.3.6e (from) statistical theories; Il.fig. 3.18 (see double layer, diffuse, charge) surface concentration; see interfacial concentration surface conduction and conductivity; 1.5.4, I.6.6d, II.3.208, II.4.1, II.4.28, II.4.3f, Il.fig. 4.9, II.4.91, Il.table 4.3 behind slip plane; II.4.32ff, II.4.37ff, II.4.67, II.4.94, II.4.6f, Il.table 4.3
SUBJECT INDEX
surface conduction and conductivity (continued), in diffuse double layer; II.4.321T, II.fig. 4.10 Bikerman equations; [II.4.3.59]ff influence on ^-potential; II.4.6e, II.figs. 4.29-31, II.4.6f, II.table 4.3 measurements; II.4.5c surface correlation length; II.2.10 surface diffusion (coefficient); I.6.69ff, II.2.14, II.2.29 surface excess; see interfacial excess surface energy; see energy surface enhanced Raman spectroscopy; III.3.7c.ii surface equation of state; see : equation of state, two-dimensional surface force apparatus; 1.4.8, IV.3.12b surface forces versus body forces; 1.1,8ff, 1.4.2 surface ions; 1.5.89 surface modification; II.1.110, II.2.88, II.5.97, IV.2.2i surface porosity; see pores (in surfaces), porosity of surfaces surface potential; 1.5.5 ('surface potentials' of monolayers = Volta potentials; see under potentials) surface pressure; I.1.16(def.), 1.3.17,1.3.32ff, I.3.47ff, I.3.51ff, II. 1.3, II.1.3b, II.1.28, II. 1.51, Il.fig. 1.15, II.1.59ff, Il.app. 1, II.3.14, II.3.140, Ill.chapters 3, 4, V.fig. 8.2, V.8.8ff (also see: equation of state, two-dimensional; for measurement, see film balance) surface pressure isotherms; III.chapter 3 surface rheology; see interfacial rheology Marangoni effect; I.1.2(intr.), 1.1.17,1.6.43ff surface roughness, and Van der Waals forces; 1.4.68 in colloid interaction; IV.3.82ff in electrokinetics; II.4.39 in optics; 1.7.10 surface states (semiconductors); II.3.172ff, II.3.176 surface tension; see interfacial tension surface of tension; 1.2.93, V.6.5 surface undulations; 1.7.77, 1.7.10c surface wave; 1.7.75 surface work; see interfacial work surfactants; I.1.4(def.), I.1.6(intr.), I.1.23ff. II.figs. 1.1.15-16, II1.4.6a, III.table 4.4 anionic; I.1.23ff, I.fig. 1.15, III.4.6d adsorption of; Il.fig. 3.Id, also see monolayers
85
86
SUBJECT INDEX
surfactants, anionic (continued), in films; V. figs. 6.23-24, V.fig. 6.28, V.fig. 6.32, V.fig. 6.34, V.6.64ff, V.fig. 6.36, V.fig. 6.44 in microemulsions; V.5.60 monolayers; Ill.fig. 1.30, III.4.6d, Ill.fig. 4.36 association behaviour etc.; see V.chapter 4 bending moduli of monolayers; III.tables 1.6 and 7 cationic; 1.1.23, III.4.6d adsorption, see monolayers interfacial tension (dynamic and Theological); Ill.figs. 3.43-44, Ill.fig. 4.17 monolayers; III.4.6d, Ill.fig. 4.35, Ill.table 4.6, Ill.fig. 4.38 coalescence; V.8.83ff emulsifiers; V.8.1b, V.8.2c, V.fig. 8.13, V.8.15 interfacial tension (dynamic and Theological); Ill.fig. 1.31, III.4.6 monolayers; Ill.fig. 3.65, III.4.6 non-ionic; 1.1.23 adsorption of; see poly(styrene) latex, silica, and monolayers cloud point; Ill.fig. 4.29 in emulsions; V.fig. 8.13, V.fig. 8.14, V.8.49ff in films; V.fig. 6.26, V.6.66ff, V.fig. 6.40 in microemulsions; V.chapter 5 interfacial and surface tension (static, dynamic and Theological); Ill.fig. 1.31, Ill.table 4.5 monolayers; III.4.6c, Ill.figs. 4.30-34 packing parameter; V.4.1d, [V.4.1.4] surroundings (in thermodynamic sense); 1.2.2, 1.3.2 susceptibility (electric); 1.4.52 (intr.) suspension; 1.1.2, 1.1.22, I.fig. 1.14 ageing; 1.2.99 suspension effect; I.5.Sf, I.fig. 5.15, II.3.105 Svedbergequation (sedimentation); [IV.2.3.23] swelling; V.2.23, V.2.39, V.2.3d, V.4.115ff swollen dilute (polymer solution); II.5.9, II.fig. 5.3 synergism (in coagulation); IV.3.9k system, (in statistical sense); 1.3.la (in thermodynamic sense); 1.2.2 Szyzskowski isotherm; [III.4.3.14] tactoids; II. 1.80 tails; see adsorption of polymers t-plot; see adsorption
SUBJECT INDEX
tangential stress = shear stress; see stress Tate'slaw; [III. 1.6.1] Taylor number; 1.6.36 Taylor vortices; fig 1.6.8 Teflon, wetting; Il.table 1.3 tensiometers; III. 1.8 tensors; Lapp. 7f TEM = transmission electron microscopy ternary phase diagrams; V.chapter 5 tertiary oil recovery; see enhanced oil recovery tertiary structure (proteins); V.3.3ff tethered chains; V. 1.11 thermal diffusion; 1.6.12, 1.7.44, 1.7.48 thermal neutrons; 1.7.25 thermal wavelength; [1.3.5.14] thermodynamic state; 1.2.3 thermodynamics (general); I.chapter 2 First Law; 1.2.4, 1.4.3 irreversible; 1.6.2,1.6.5a, I.6.6a, 1.6.7 Second Law; 1.2.8 'Third Law1; 1.2.24 (also see: statistical thermodynamics) thermodynamics of small systems; V.4.2a theta (0) solvent; 1.6.28 thickness of adsorbed layers; see adsorbate thin liquid films; see films thlxotropy, thixotropic; 1.1.23, III.3.87, IV. 1.3, IV.6.14, IV.fig. 6.7 tilt angle; III .fig. 3.60 tilted plate; III.5.4d, Ill.fig. 5.24 time correlation functions; see correlation functions TIRF = total internal reflection fluorescence TIRM = total internal reflection microscopy titanium dioxide; Il.table 1.3, Il.table 3.6, IV.3.13a anatase, double layer; II.3.94 point of zero charge; Il.table 3.5, Il.app. 3b rutile, adsorption of anionic surfactants; II.figs. 3.82-83 adsorption of water vapour; II.fig. 1.9 double layer: II.figs. 3.58-60, II.fig. 3.63, Il.table 3.8, II.figs. 3.82-83 electrokinetic charge: II.fig. 4.13
87
88
SUBJECT INDEX
titanium dioxide, rutile (continued), mobility; IV.flgs. 3.62-64 point of zero charge; Il.table 3.5, II.fig. 3.37, II.fig. 3.82, Il.app. 3b stability ratios; IV.figs. 3.65-67 ^-potential; Il.fig. 3.63 topology (vesicle formation); V.4.7d torque; IV.6.7b total internal reflection fluorescence (TIRF); II.2.54, III.3.7c.iv total internal reflection microscopy (TIRM); IV.3.157 total reflection; 1.7.74, II.2.51, II.2.54 trains; see adsorption of polymers trajectories (of particles); IV.fig. 4.2, IV.flg. 4.20, V.fig. 8.7 transfer in galvanic cells; I.5.5e transfer, of molecules; 1.2.18a (ions to other phases); I.5.3f, 1.5.5 (molecules to other phases); 1.2.20, 1,4.47,1.6.44 (also see: transport, II.vlscous flow) transference numbers; see transport numbers transmission angle; 1.7.72 transmission coefficient; 1.7.73, II.2.50 transmission electron microscopy (TEM); 1.7.lib, Il.fig. 1.1, V.5.3b cryo-direct imaging; V.5.3b freeze fracture direct imaging; V.5.3b, V.figs. 5.15-17 transport processes; I.chapter 6 (also see: hydrodynamics, diffusion, conduction) transport, linear laws; I.6.1c, I.table 6.1 of charge; I.table 6.1 in double layers; IV.4.4a of heat; I.table 6.1 of mass; 1.6.1a, I.table 6.1 through interfaces; 1.6.44 of momentum; 1.6. lb, I.table 6.1,1.6.4a transport numbers; I.6.76ff, [1.6.6.14](def), V.2.5b, V.fig. 2.30 transverse waves; III.3.110, III.fig. 3.43, see further: interfacial rheology trapping (optical); IV.3.157 Traube's rule; 1.4.51 triboelectricity; II.3.187 Trouton number; V.8.37 Trouton ratio: IV.6.10
SUBJECT INDEX
Trouton's rule; III.2.54 (also for surfaces) tunnelling (of electrons); 1.7.90 turbidity; 1.7.41, 1.7.47, V.fig. 8.6 turbulence; I.6.4b, V.8.40ff Tyndall effect; 1.7.26, II.4.45, IV. 1.3 Ubbelohde vlscometer; IV.flg. 6.18 ultracentrlfuge; IV. 1.13 ultracentrlfugation, see sedimentation; IV.2.3d ultramlcroscope; 1.7.26, II.fig. 4.14 ultrasonic emulsificatlon; V.8.33 ultrasonic vibration potential; II.4.7, II.4.3e Ultra Turrax; V.flg. 8.17 uncertainty principle (Heisenberg); 1.3.4, 1.3.58 undulations (of fluid interfaces); III. 1.78 undulation forces (membranes, etc.); V.4.8 uniaxlality; 1.7.97 UPES = UPS = ultraviolet photo-electron spectroscopy; 1.7. l l a UVP = ultrasonic vibration potential vacancy (in semiconductor); II.3.171 valence band (solids); II.fig. 3.68, II.3.173 Van der Waals interactions; I.chapter 4, IV.3.8a between molecules (general); I.4.10ff, 1.4.4 additlvity; I.4.18ff London (or dispersion); 1.4.17, I.4.4d, I.table 4.3, I.4.4e, 1.4.5, 1.4.6, 1.4.7 Debye; 1.4.17, I.4.4c, I.table 4.3, 1.4.41 Keesom; 1.4.17, 1.4.4c, I.table 4.3, 1.4.41 retardation; 1.4.17, 1.4.31, I.4.78ff between colloids and macrobodles; 1.4.6, IV.3.9 Hamaker-De Boer; 1.4.6 Lifshlts; 1.4.7 measurement (direct); 1.4.8, I.fig. 4.19 macroscopic; see Lifshits microscopic; see Hamaker-De Boer repulsive; 1.4.72, 1.4.78 retardation; 1.4.6c in thin films; II. 1.101 Van der Waals equation of state; [1.2.18.26], [1.3.9.28], [1.4.4.1], III.2.17, [III.2.9.3] (reduced), IV.5.7a Van der Waals loops; 1.3.47, 1.4.17, Il.fig. 1.20, II. 1.101, Il.fig. 1.42, IV.5.7a, IV.fig. 5.37
89
90
SUBJECT INDEX
Van der Waals' theory (interfacial tensions); III.2.5, III.2.3.1 (generalized ) Van 't Hoffs law, for boiling point elevation; 1.2.74 for freezing point depression; 1.2.74 for osmotic pressure; I.2.20d, 1.7.50 vapour pressure, lowering; 1.2.74 of small drops; 1.2.23c variance; 1.3.35 vector, vector field etc.; Lapp. 7 velocity correlation function; Lapp, l l a , II.2.14 velocity distribution; 1.6.2Iff, I.6.3c, I.fig. 6.4 velocity persistence length; IV.4.5 vermiculite, wetting; II.table 1.3 vertical plate, wetting of; III. 1.3b end effect correction; III. 1.22 vibration; 1.3.5a, 1.4.44 vibrational spectroscopy; IILtable 3.5 virial coefficients; 1.2.18d, I.3.8f, I.3.9c second; I.2.18d, I.3.8f, I.3.9b, [1.3.9.12], [1.4.2.8-11], 1.7.51, 1.7.57, IV.table 5.1, IV.fig. 5.24, [IV.5.6.2], IV.fig. 5.29, IV.fig. 5.55 two-dimensional; [II. 1.5.24], II. 1.60 virial expansions; I.2.18d, I.3.9b, I.3.9c, I.5.27ff, 1.7.51, [II. 1.5.30], IV.5.2e viscoelasticiry; I.2.7(intr.), IH.3.88, IV.chapter 6, IV.6.1, IV.6.11, IV.6.14, IV.fig. 6.10, IV.fig. 6.12, IV.6.6, IV.fig. 6.15 viscoelectric coefficient; II.4.40 viscoelectric effect; II.4.40 viscometers, capillary; IV.6.7a Couette; IV.6.7b Ostwald; IV.fig. 6.18 rotation(al); I.6.36ff, IV.6.7b Searly; IV.6.7b Ubbelohde; IV.fig. 6.18 viscosity; 1.5.43, I.6.10ff, IV.6.1, IV.6.2 apparent; IV.6.11 definitions; IV.table 6.3 dispersity effect; IV.fig. 6.27 dynamic vs. kinematic; 1.6.11 Einstein; IV.6.9a extensions; IV.table 6.3. IV.tabie 6.4
SUBJECT INDEX
viscosity (continued), (in) electrokinetics; II.4.4 elongational; IV.6.9 emulsions; V.8.15ff, V.fig. 8.5 examples: I.table 6.3, I.6.4g, IV.figs. 6.28-29, IV.figs. 6.31-32, IV.figs. 6.34-35, IV.fig. 6.38, V.2.4 intrinsic; IV.6.47, IV.fig. 6.24, V.2.4c Newton, definition; IV.table 6.4 polyelectrolytes; V.2.4 polymer solutions; IV.6.11, 6.12 shear vs. elongational; 1.6.8 viscosity-averaged molecular mass; IV.6.63 viscous flow; 1.6.1, 1.6.4, IV.6.1, IV.6.2, IV.fig. 6.5, IV.6.8-13 around spheres; I.6.4e, II.4.6, II.4.8 between parallel plates; I.6.40ff, IV.fig. 6.3 dilational vs. rotational; I.fig. 6.7 due to Marangoni effects; 1.1.17,1.6.44 due to temperature gradients; 1.6.4c fluid-fluid interfaces; I.6.42ff in cylindrical tube; I.6.41ff in porous media; I.6.4f laminar linear; 1.6.4a, I.6.4d turbulent; I.6.4b Volta potential; see potential Volmer adsorption isotherm; see adsorption isotherm volumes, of ions (partial); I.table 5.7 Vonnegut equation (for spinning drops); [III. 1.9.6] vortices; 1.6.4b Vroman effect (protein ads.); V.3.52 Warburg coefficient; II.3.96 Ward-Tordai equation; [1.6.5.36], [II.1.1.15] Washburn equation; [III.5.4.4] III.5.57ff, III.5.84ff waste water treatment; IV.3.184 water, interactions in; I.4.5d, I.4.5e structure; 1.5.3c, 1.5.4, II.2.16 near surfaces; II.2.2c, II.figs. 2.6-7, II.3.122ff, Il.fig. 3.39. II.4.38ff water-air interface, double layer; II.3.10f, Il.fig. 3.78, III.4.4, III.fig. 4.20 reflectivity; III.fig. 3.57
91
92
SUBJECT INDEX
water-air interface (continued), surface relaxation; III.fig. 1.29 surface tension; III.1.12, HI.table 1.2 influence of electrolytes; II.3.180, II.fig. 3.73 influence of temperature; III.1.12b, Ill.table 1.3, III.table 1.4, Ill.fig. 1.27 simulation; Ill.table 2.2, Ill.figs. 2.12-13 waterglass; IV.2.2 wave damping; see interfacial rheology waves, electromagnetic; I.chapter 7 evanescent; 1.7.75 in a vacuum; 1.7.1 plane; 1.7.la films; V.8.87 polarization; 1.7. l a spherical; 1.7.1b surface; see surface wave (also see: electromagnetic radiation) wave vector; I.7.4(def.) wave vector transfer = wave vector Weber number; [V.8.2.3], V.fig. 8.12, V.fig. 8.15, [V.8.3.25], V.table 8.4, V.8.89 Wenzel equation; [III.5.5.1] wet foam; V.7.2, V.fig. 7.12 wetting (general); III.chapter 5 wetting; 1.1.2,1.1.8,1.1.3, I.fig. 1.13, Il.table 1.3, Ill.fig. 5.7 adhesional; II.2.5, III.5.2 and gas adsorption; II.1.19, III.5.3, Ill.fig. 5.16 and Van der Waals forces; 1.4.72 complete; III. 1.5, III.5.1, III.5.4 dynamics; III.5.8 enthalpy or heat; [II.1.3.43], Ill.table 1.3, Il.fig. 1.10b, II.2.6, II.2.3d, II.fig. 2.10, Il.fig. 2.20, III.5.2, III.5.20 entropy; II.2.7 immersional; II.2.5, Il.fig. 2.10, III.5.2 liquids by liquids; III.5.3 microemulsions; V.5.4c molecular dynamics; Ill.fig. 5.36 (and) nucleation; IV.fig. 2.9, IV.2.2f partial; III. 1.5, III.5.1, Ill.fig. 5.1, Ill.fig. 5.13, porosity; III.5.9 scales; III.5.5 selective; II.2.88
SUBJECT INDEX
93
wetting (continued), silica by water; III.5.3c surfactant influence; III.5.10 thermodynamics; III.5.2 wetting agents; III.5.86, III.5.10 wetting films; III.5.3 wetting transition; II. 1.101, II.fig. 1.41, III.5.14, III.5.30, Ill.fig. 5.14, V.5.60, V.fig. 5.30 Wiegner effect; see suspension effect Wien effect (in electrolytic conductance); 1.6.6c Wilson chamber; 1.2.100 wine tears; 1.1.1,1.1.2,1.1.17 wolfram surface, covered with palladium; II.fig. 1.2 work; 1.2.4 of adhesion; III.2.34, III.2.7 Iff isothermal reversible; 1.2.27 statistical interpretation; 1.3.15, [1.3.3.11] (also see: interfacial work, transfer, potentials) work function; 1.5.75, II.3.114, II.3.174 work hardening or softening; III.3.88 worm-like chain; V.2.27 Wulff relations; IV. 2.23 XAFS = X-ray absorption and fluorescence spectroscopy; 1.7. l l a XPS = XPES = X-ray photoelectron spectroscopy; 1.7.lla, I.table 7.4, II.1.15, H.flg. 1.5 X-ray reflection and diffraction; III.2.47, Ill.table 3.5, III.3.7b, IV.2.40 X-ray scattering; I.fig. 5.6,1.7.9a of films; V.6.9 yield stress; see rheology (of) emulsions; V.8.17 young foam; V.72 Young and Laplace's law; see capillary pressure Young's law for capillary pressure; see capillary pressure, Young and Laplace Young's law for contact angle; [III. 1.1.7], III.5.1b, [III.5.1.2), III.5.2 Yukawa pair interaction; IV.5.6a, IV.fig. 5.30 Z-average; 1.7.63 Zeeman effect (for spectral lines); 1.7.16 zeolithe; see molecular sieve zero point of charge; see point of zero charge zero point vibrations; 1.3.22, 1.4.29 Zimmplot; I.7.57ff, I.fig. 7.12. I.fig. 7.16 zinc oxide. SEM; II.fig. 1.1
94
SUBJECT INDEX
Zisman plot; III.fig. 5.42 zone electrophoresis; II.4.131 zwitterionic surfaces; II.3.74 a -helix; V.chapter 3 (3 -sheet; V.chapter 3 ^-parameter (polymers etc.); see Flory-Huggins interaction parameter 2"s-parameter (polymer ads.); [II.5.4.1], II.chapter 5 4, t ;II.5.40,II.fig.5.22 ^-potential; see potential ^-potential; see electrokinetic potential <9-point, ©-temperature; II.5.6ff, II.5.2b