You are on page 1of 44

COLLOIDS

AND
Colloids and Surfaces
A: Physicochemicaland Engineering Aspects 111 (1996) 75-118

ELSEVIER

SURFACES

Dilational and shear rheology of adsorption layers


at liquid interfaces
R. Miller "'*, R. Wfistneck b, j. Krfigel u, G. Kretzschmar a
"Max-Planck-Institutf~r Kolloid-und Grenzfl~chenforschung, Berlin, Germany
u Universiti~tPotsdam, Institutfar FestkOrperphysik, Potsdam, Germany
Received 26 July 1995; accepted 16 November 1995

Abstract

Dynamic properties of interfaces are of increasing interest in science and in practice as they give insight into
interactions and processes at interfaces rather than equilibrium properties. The general ideas on mechanical interfacial
properties as an important part of dynamic properties were established long ago by Gibbs and Boussinesq. Now on
the basis of new techniques, better experiments can be performed which allow a more and more quantitative
understanding. The mechanical behaviour of interfaces, modified by soluble adsorption layers or insoluble monolayers
of surfactants or polymers, is the subject of many actual studies. Computer-driven instruments using new sensors and
very sophisticated methodologies enable us to perform very complex and sensitive measurements which were impossible
until recently.
Numerous studies of interfacial shear and dilational rheology have been reported and use a large variety of
techniques. Shear experiments are most useful for polymer and mixed polymer-surfactant adsorption layers and
insoluble monolayers and give access to interaction forces in two-dimensional layers. Dilational interfacial properties
however are most frequently studied for soluble adsorption layers of surfactants and mixtures of polymers and
surfactants.
This overview gives an introduction to the interfacial rheology and discusses some specific theoretical aspects
necessary to interpret experiments. Experimental techniques to perform shear and dilational experiments at liquid
interfaces are summarised and only the most recent developments are described in more detail. Examples are given to
demonstrate how the experiments work and what output can be expected.

Keywords: Adsorption layers; Liquid interfaces; Proteins; Surface theology; Surfactants

1. Introduction
An understanding of the mechanical behaviour
of soluble adsorption layers or insoluble monolayers of surfactants and polymers and their
mixtures is fundamental to m a n y technologies.
Dilational and shear mechanical properties and
exchange of matter at the interface are crucial to
* Corresponding author. E-mail: miller@mpikg.ffa-berlin.de.
0927-7757/96/$15.00 1996 Elsevier Science B.V. All rights reserved
S S D I 0927-7757(95)03492-7

coating processes of photographic emulsions containing gelatine and surfactants [ 1,2]. The same
situation is true for the process of enhanced oil
recovery described by Dorshow and Swafford [33
and Kovscek et al. [4], in the food industry
[ 5 - 1 1 ] , and in flotation [12]. The relevance of
these properties to natural phenomena, such as
processes occurring at the atmosphere/ocean interface, is emphasised by Loglio and coworkers
[ 13-16 ] for example.

76

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

The effect of surface active compounds on formation and stabilisation of dispersed liquid systems
such as foams and emulsions, can be increasingly
more quantitatively described as a result of our
growing insight into such very complex processes,
A general discussion of the role of dynamic surface
properties in different areas of application has been
given by van den Tempel and Lucassen-Reynders
[17]. Foam and emulsion formation and stability
depend on various specific parameters and the
discussions are based on elementary processes,
such as stability of thin liquid films, coalescence of
two drops or bubbles, break-up process of drops
in flow fields, motion of bubbles and drops in a
solution, etc. The relevance of dilational properties
and exchange of matter mechanisms as well as
shear rheological parameters, to phenomena such
as foaming and emulsification, has already been
emphasised and is explained by the rapid expansion of surfaces during the generation of foams or
emulsions [ 18,19]. The rheology of wet foams has
been analysed by Wasan et al. [20] and a relation
was derived [ 21 ] by connecting the foam dilational
viscosity with the surface dilational elasticity. A
qualitative correlation between foam stability and
effective dilational elasticity or dynamic surface
tension, respectively, was found for some simple
surfactants by Matysa [22] and Garrett and
Moore [23].
Although it is usually accepted that knowledge
of surface rheology is essential for many technological processes there have not been many studies of
such properties. The reason is that there are
different surface rheological parameters which
reflect the different aspects of a complex behaviour,
It is not trivial to predict which of the parameters
are needed to understand and control a technological problem. As a mechanical load may affect an
interface in many different ways and may provoke
various reactions there are only few commercial
instruments available to determine such parameters. Each of these instruments has a restricted
range of application and can therefore be used
only for the study of a special rheological behaviour. This situation is also true for original set-ups
described in the literature. Development of several
new methods during the last few years will help to
overcome these experimental limitations,

The two-dimensional rheology is more or less


an analogue of the three-dimensional rheology.
The most striking difference between the two rheologies however is that adsorption and monolayers
are compressible. This particular property is of
importance in many practical cases.
This review gives the present state of our understanding of shear and dilational deformations in
interfacial rheology. Section 2 introduces some
rheological models and the terminology. In
Section 3 various classic and newly developed
techniques for interfacial shear studies are compared and some selected experimental examples
given.
In Section 4 the dilational behaviour of adsorption layers and monolayers at liquid interfaces is
discussed. Relaxation theories providing information on exchange of matter and dilational rheological parameters will be described for the damping
of harmonically generated disturbances as well as
relaxations to transient perturbations and methods
of experiment are described, based on the damping
of capillary and longitudinal waves, and oscillation
behaviour of bubbles. Transient relaxations with
pendent drop and bubble and drop pressure measurements are also shown and applications to
different interfaces, using surfactants, surfactant
mixtures, polymers and polymer-surfactant mixtures discussed.

2. Theoretical description of interfaciai rheology


A rheological process takes place whenever a
piece of matter is influenced by a field of force
leading to a straining displacement of the whole
particle or a part of it. Depending on the duration
and the strength of the influence, given by a
frequency and an amplitude, the reaction may vary
and may lead to quite different results. The application of a strong magnetic field within a time
interval of some 10-9 s will provoke a displacement
of electrons of a molecule (NMR spin resonance)
whereas energies in the range of 0.6-40 kcal M-1
will cause motions of different molecular parts. An
oscillation is usually characterised by the wavelength 2 or the wave number ~ and the frequency
v. These values are connected by ~ = 1/2 = v/c,

R. Miller et aL/Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

where c is the velocity of light and the energy is


given by hv. In the time scale of 10-7-10 -8 s the
N M R technique can be used to characterise the
motion of an entity in aggregates, for instance a
surface active molecule in a micelle. With another
time scale (influence of light) changes of the structure of molecules can be detected, for instance the
change of protein structure caused by an interaction with surfactant molecules, whereas a
mechanical motion in a time scale of some 100 Hz
can be used to characterise the exchange of matter
between the bulk and the interface. The influence
of a mechanical force of much lower frequency will
lead to a displacement of some parts of the structure of molecules, i.e. to a flow of a so-called
rheological unit. The extreme of such a mechanical
motion is the creep process where the frequency is
close to zero.
The behaviour of glass is an interesting example
to demonstrate the reaction of a material at
different frequency ranges. At small mechanical
force glass shows a nearly ideal elastic behaviour,
U p to a critical weight and velocity a falling small
steel ball will be rejected elastically by a plate of
glass. When the velocity or the weight becomes
too high the structure breaks. Gravity, however,
on the window of a church over several hundred
years produces a flow process which thins the glass
in the upper part of the window and thickens it
lower down.
The two different reactions, i.e. elasticity on the
one hand and flow on the other, can however take
place at the same time. This may result in a
damped elasticity in a material or in a flow accompanied by a rapid deflection at the first time of
stress; a j u m p which changes into a flow. Thus, the
results of rheological experiments always depend
on the time scale and the strength of stress applied
in the experiment. When taking into account the
reaction of interfaces some particularities should
be considered,
The interfacial tension of a pure liquid does not
change with the quantity of strain. This is a peculiarity of the interface and can be explained when
the location where the interfacial tension acts is
considered. The interfacial tension occurs as a
result of the interfacial interaction between the
molecules. At the same time the interface is an

77

open system and can be enlarged simply by a


displacement of molecules from the bulk into the
interface. This displacement however does not
change the value of surface tension. In the case of
surface active molecules adsorbed at the interface
a displacement needs some time. Although such a
displacement is comparatively rapid at sufficiently
high surfactant bulk concentration, the surface
tension can be influenced when the adsorption
layer is deformed to a certain extent, for example
during the measuring procedure. Lunkenheimer
and Wantke 1-24] have shown that the measurement of surface tension of a surfactant solution by
using the ring method can be strongly influenced
by the strain of the interface, especially when
vessels of small diameters are used. A dependence
on stress and strain already exists in such a case.
In contrast to usual solid structures or liquids,
the interface of a liquid may be easily expanded
and compressed. Both kinds of m o t i o n - - d i l a t i o n
and shear - - can therefore be easily achieved. The
reactions observed and the information we will get
depend on the kind of deformation or stress. The
oscillation of a bubble leads to dilation and compression of the interface. Assuming that soluble
surfactants are present in the bulk phase leads to
an exchange of matter between the bulk and the
bubble surface when the frequency becomes comparable to the time of the adsorption/desorption
process, whereas an oscillation with another
frequency may lead to a displacement within
the interface, i.e. a lateral displacement of some
molecules.
On the other hand a displacement within the
interface can be initiated by a displacement within
the bulk - - the motion of a canoe on a river's
current. That however is not always the case. The
motion within the hot m a g m a of the Earth - - a
rising of hot parts and sinking of cool parts in
ordered structures, the Marangoni instability - may take place without any dramatic motion in
the core of the Earth. An analogous phenomenon
can be observed when pouring milk into a cup of
coffee. When the milk is poured slowly close to
the side of the cup the milk can stream down and
up without displacing the coffee floating at the
surface. This shows how very different the flow
within the interface can be from the flow within

78

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

the bulk depending on the solidity of the structure


at the interface,
Confined to mechanical motions with a time
scale of some hundred hertz up to a very slow flow
of the interface - - a creep - - different reactions of
the structure at the interface can be investigated,
This may lead to a characterisation of some interactions between molecules forming these structures. From dilation and compression in a diffusion
time scale we can get information about an elasticity, Gibbs elasticity, which results in changes in
the excess of surface-active molecules. When the
motion is too fast no exchange of matter will occur,
whereas in cases of slow motion the interface is
always in equilibrium with the bulk phase. Any
elastic reaction in this case is the result of a
displacement within a network formed by the
molecules adsorbed or even of some domains of
molecules at the interface,
The structure of the adsorption layer is the result
of the interaction between all molecules and ions
in the region of the interface. These interactions
may in the simplest case be the interaction between
the solvent molecules. In the presence of other
molecules (surface active ones) the structure occurring depends on the interaction between these
adsorbed molecules as well as the interaction
between them and the solvent.
In more complicated cases different surface
active components can be adsorbed. The structure
of such mixtures at the interface is formed in a
process that may take a long time and be further
influenced by the adsorption process. So for
instance the formation of the interfacial structure
of a mixture of different surface active components
is determined by the adsorption, where in a first
step components may adsorb with a lower interfacial activity than the main component. These
molecules can be displaced by molecules that
adsorb later. At the same time an aggregation can
occur accompanied by different rearrangements of
the adsorbed molecules. The properties of such a
structure are time dependent and may not be in
equilibrium within the time of investigation.
Interfacial structures dissipate energy when
stressed. On a microscopic scale the dissipation is
connected with different processes of displacement.
Bonds within the adsorption layer may be stressed

or distorted. Some structures even become disconnected. Disconnected molecules may be laterally
displaced or may desorb and so be displaced by
other molecules. The laterally displaced molecules
and the newly adsorbed ones will enter into new
bindings with their neighbours in the adsorption
layer. In the case ofmacromolecules different loops
and trains may be reoriented. It is clear that the
structure of the adsorption layer is now different
even when the stress stops.
When the dissipated energy becomes too great
the structure cannot be restored. In this case an
avalanche-like destruction may occur often accompanied by drastic changes of the flow properties of
the adsorption layer.
Considering the picture of these processes it
becomes clear that experimental values determined
by rheological measurements may differ remarkably for the same system even when analogous
experimental conditions are applied. Therefore, for
comparing viscosities and elasticities of interfaces
not only the kind of stress, energy and amplitude
have to be taken into account but also the adsorption layer alteration time and its history.

2.1. Interfacial rheology of insoluble monolayers


The projection of three-dimensional rheology to
a surface was argued first by Boussinesq [25,26].
He intended to connect the hydrodynamic equations of two phases discontinuously and to formulate a two-dimensional theology for the interface.
The mathematics of two-dimensional rheology
is based on the interfacial stress tensor. An outstanding description of the theory was given by
Krotov [27].

2.1.1. Tensor algebraic description of the surface


rheology
In the two-dimensional case the contact forces
(stress) are given by four values Vi.j in a point,
where i is the component of force which acts
through a line unit perpendicular to j
~x~ ~xr
/7 =

(1)
7yx Try
Using the physical requirement of finite progressive

R. Miller et al./Colloids Surfaces A." Physicochem. Eng. Aspects 111 (1996) 75-118

velocity of an infinitesimally small surface element


it can be shown that this matrix is a tensor, i.e.
independent of the system coordinates. At the same
time it can be shown that in the case of rotation
7ik =3:kl, i.e. the tensor of interfacial tension is
symmetric.
Further, only small deviations from equilibrium
will be taken into account, because only in this
case are conditions of linearity observed and the
principles of superposition valid. The tensor may
be written as a sum of a scalar part in the absence
of any deviation from equilibrium:
?o

01 I

(2)

Using Eq. (4) to rotate the whole system the tensor


reads:

3:21

3:12
3:22--

~u\
---y|

1/Bux

[
2 \~yy

0
u~ =

1 (Bu=
- 2 \ By

8x J

Buy)

(6)

Bx J

This operation does not change the physical state


and so the anti-symmetric tensor ~ should be
excluded from the tensor of deflection and the
strain tensor results:

where $ is the two-dimensional unit tensor and a


deviation part:
A~--- 3:11- 3:0

79

Bu=

1 (8ux

Our

8x

2 \ ~yy

8x ) [

(3)

3:0

2 \ By

The aim of rheology is to work out which values


characterising an infinitesimally small surface eles e n t have to be altered to cause a reaction without
changing the parameters of the surface element, i.e.
to obtain coefficients of proportionality,
First we will focus on a purely elastic case,
consisting of an element of surface whose points
are deflected by a vector ff with components
u~(x, y, t) and Uy(X,y, t). Obviously AB:i,kcannot be
proportional to the components of if, because a
deflection of the whole system does not alter the
physical state. Such a deflection can be excluded
by differentiation of the components

Oux
8ux
du~ = -~x dx + ~y dy

B=,
By

BX]

(7)

which expresses the deformation # in the theory


of elasticity. The changes of t~ are directly connected to the stress tensor. There is however no
linear proportionality between these tensors. Only
in the one-dimensional case (Young coefficient)
can a deflection be directly proportional to a stress.
When the isotropic part of tensor #i, the trace of
the tensor, is extracted, we obtain

1 (Bu~

Buy~

2 \ 8x

By ]

l(dux
2 \ Bx

Bu,~
By J

and

B=,

B.,

du, = ~-x dx + ~-y dy

(4)

and constructing the corresponding tensor of


deflection:
OUx OUx

t~ -

8x
Bur

By
Bur

t'gX

By

- 2 \ Bx

(8)

The isotropic part of the tensor A2 can be written


as
ZI~7i = k i ~ i

(5)

By ]

(9)

On the other hand the deviator can be written as


Affd = ka~d

(10)

80

R. Miller et aL/Colloids Surfaces A. Physicochem. Eng. Aspects 111 (1996) 75-118

where

Here qs is the Newtonian shear viscosity and qa

1 (Ou~

Our~

1 (OUx Our~ I

~a- 1 (OUx ~?uv"]

1 (C3Ux Our'][

\ ~3y

x J

- 2 \ ~-x

the[25,26]dilationalasviscosity, first defined by Boussinesq

(11)

0y / [

ki and kd are the coefficients which characterise


the physical nature of the surface elasticity. For
the case of a spread monolayer at a liquid interface
(insoluble in the bulk phase) slowly stressed in all
directions the stress tensor reads
zJ~ = Z~i = A ~
(12)
and the strain tensor becomes
0u~
{~= {oi= ~-x ~

( 13 )

From Eq. (9) the scalar results

(18)

It should be mentioned that another formulation


of a two-dimensional rheology is possible and can
be found in the literature [28,29].

2.1.2. Visco-elastic effects


When the deformation of an insoluble monolayer does not result from infinitesimally slow
motion the elastic effects are usually accompanied
by viscous effects. Provided the motion is sufficiently slow to preserve linearity and the principle
of superimposition, the stress tensors are
A~i = 2E~i + 2 qd {0i

(19)

and

OUx
A~ = ki ~?x

rla-(dA/dt)/A

A~;d= 2Gs#d + 2rh6d


(14)

(20)

for a dilation 3Ux/C~X= AA/2A, with A the starting


area, and AA a small deviation of surface area
from equilibrium,
Under dilation or compression of a surface the
modulus of elasticity is defined by

In reality these parameters may often be superimposed, i.e. Ay = Avi + A~d. Based on this approach
different rheological models can be derived to
describe the large diversity of rheological behaviour. Using Eqs. (19) and (20) and applying them
to the deformations given by Eqs. (9), (15) and
(16) leads to

/1~2
E - AA/A

~i = ~~ i + ~ 1 fJ A~i dt+ C~

(15)

with ki = 2E. In this case the modulus E equals the


Marangoni elasticity modulus. Analogously to k~
for shearing, kd=2G~. Therefore an ideal elastic
two-dimensional area can be characterised by two
parameters. For ideal fluid interfaces we get G~ = 0.
For every process connected to energy dissipation viscosity effects have to be considered. When
replacing the vector of deflection ~ by the vector
of deflection velocity if, analogously to the former
case of interfacial properties of an ideal fluid, the
two relations for dilation and shear result

A ffi =

2~i

where C is an integration constant. Differentiation


of Eq. (2l)gives
1 A~i + --1 A~i = 2~i
~
qd
and
1
1
GssA~d + ~s A2)d= 2~d

(22)

(23)

(16)

For the two extreme cases, deformation in every


direction (~d = 0) and shear flow (Ovx/Oy# 0), the
following scalar equations result

(17)

1 dAy AV 1 dAA
E d ~ - + r/d - A dt , or

and

A,~a=2rl~(oa

(21)

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

dA 7
- -

dt

Ay
+

--

"Ci

dAA
=

81

and for shearing

- -

(24)

Adt

dyx, + ?xy=(G~+G,)(O2u~
dt

Td

1 c?u,~

\C~t 0y + ~ 0y /

(30)

with the relaxation time Ti = ~la/E, and


The values 0i and 0a are the retardation times.

OVx
+
Gs dt
qs
Oy
1 dvxr

7xy

(25)

or.
d~r
~r
~2ux
-- + -- = Gs-

dt

zd

(26)

Ot ~y

with Td = q,/G~. These equations are called the


Maxwell model. For fast deformation of time At,
the relations At <<zi and At <<r a are valid, so that
the second parts of the equations can be neglected,
i.e. ideal elastic behaviour results. On the other
hand, when At >>"/7i and At >>zd the first parts may
be neglected; i.e. ideal viscous flow is observed,
The parameters E, r/d, G~ and rh are non-equilibrium values, i.e. relaxation moduli.
In the same manner rheological equations can
be derived separately for the isotropic parts and
the deviators. One possible case is the addition of
forces according to Eqs. (9), (22) and (10), (23) for
a three-parameter model. Instead of Eq. (21) one
may use an additional parameter e'. Using the
tensor equation A~ = 2e'~i we get the two relationships (dTi/~ls)
~'
= (2e/t/~i'~ and (A~/E)= (2e'/E)~oi.
With Eq. (22) and maintaining the symbol A~i for
the sum tensor leads to

E
A~i +

--

2Ee'
A]3 i --

/~d

2(E + e')(~i +

- /~d

~i

(27)

where E + e' is the dynamic modulus with e' the


relaxation independent modulus. Analogously for
the case At>>z~ we get
A~d +

G~
--

A~a = 2(Q + G's)q~d

(28)

rl~

A7

So far only insoluble monolayers have been


described. The formalism of interfacial rheology
for soluble adsorption layers is similar. For shear
deformations exactly the same models can be used.
The dilational rheology of insoluble monolayers
can be applied to soluble layers only if the characteristic time of interfacial disturbances is much
shorter than the exchange of adsorbed molecules
with the adjacent bulk phase. Differences arise
when a soluble adsorption layer is compressed and
expanded in a time less than or comparable with
the characteristic adsorption time. The characteristic adsorption time of a surfactant depends on
surface activity and concentration and varies
over many orders of magnitude in a homologous series of surfactants. The consideration of
compressions/dilations of soluble adsorption layers
is reviewed in detail in Section 4.

2.3. Interpretation of interfacial flow processes


The interfacial shear viscosity was discussed in
terms of changing the position of a rheological
unit. First approaches were given by Frenkel [30],
Ewell and Eyring [31], and Moorew and Eyring
[32]. Following this approach, in the range of its
melting point, a liquid can be treated like a cornposite solid body, i.e. containing quasi-crystalline
domains. Due to the Brownian molecular movement molecules at the surface may change their
position. The possibility of a rheological unit staying in a definite position is

z = z o exp(U/k T)

For compression or dilation into every direction


we obtain

day

2.2. Interfacial theology of adsorption layers

(dAA

d---t--+ -Ti
- = (E + e') \ ~

1~)
+ Oii

(29)

(31)

with U the energy of activation and To the period


of oscillation (of the order of 10-~2 10-13 s according to Boyd and Harkins [33]). The possibility of
it changing its position is based on the assumption
of gaps around a molecule. As a first approach the

R. Milleret al./ColloidsSurfacesA."Physicochem,Eng. Aspects 111 (1996) 75-118

82

change of position is independent of other processes. U depends on temperature and the distance
between molecules and is assumed to be equal for
all position changes. 6 is the average amplitude
of thermal oscillation that may lead to position
changes. The self-diffusion coefficient can be
defined by
62
62
D - 6~ - 6% e x p ( - U/kT)
(32)
Applying the Stokes law on a molecular level we
obtain

kT
D-

(33)

6zu/r
with r the radius of a molecule. If r does not differ
much from 6 the bulk viscosity results

kT%
=~
exp(U/kT)

(34)

Similar equations were derived by Prandtl [34]


and Andrade [35]. The basic idea of flow is that
the relative movement of the molecules is directed
by the influence of an external force, i.e. a directed
self-diffusion process. The flow is connected to
overcoming a potential barrier which divides two
possible equilibrium states. The activation process
can be treated on the basis of rate processes [36].
Applied to the interface this leads to

rls = C exp(AF/k T)

(35)

where r/s is the interfacial shear viscosity, C =


constant, and AF is the free energy of activation.
Starting from the first surface rheological experiments it was assumed that the molecules of a
monolayer join a surface flow not as single molecules but as complexes consisting of monolayer
and solvent molecules [37-41]. When we assume
a Couette flow and the area of a kinetic unit being
dA = dx dy with dx the length in the direction of
the flow, the surface shear viscosity can be written
as

~ = Vxy/(O2u~/at Oy)

a well and a barrier becomes ~ 1/2 dx. The shear


stress influencing a kinetic unit is ~xr dy. The energy
of the kinetic unit by a motion over ~ 1/2 dx is
then

7xydxdy=yxyA
(37)
In hydrodynamic equilibrium (Y~r= 0) the number
of moving kinetic units in every direction
per second is
kT

K, = ~ - e x p ( - A F / k T )

(38)

where h is the Planck constant. The coefficient of


transmission is 1 and the tunnel effect is negligible.
The number of kinetic units overcoming the barrier
in the flow direction is
K1 = k exp(7~y dx dy/2kT)
(39)
and in the opposite direction
K2 = k e x p ( - ~ r dx dy/2kT)
The velocity increment is

(40)

du~ = (K1 - K2) dx

(41)

and therefore
, {~xr dx dy'~
dux = 2 dxK, slnn~ 2 ~ T )

(42)

The following relation for the interfacial shear


viscosity results
r/~=

~r
= ~yhexp(AF/kT)
2K, slnn / _ - 7 ~ /
2kTsinh

(43)

Together with Eq.(36) the Eyring equation (43)


can be rearranged to

(rl~O2ux/c3y
~t~
2kT
/

sinh\

__ (632Ux/C3y &) exp(AF/kT)


2kT
(44)

For x < 0.25 the approximation sinh(x) = x can be


used so that for rh(~32uJOy &)A < 2kT we obtain

(36)

Assuming furthermore a symmetric potential


between two neighbouring potential wells with a
distance of the order of dx, the distance between

th=~ex p ~

(45)

Eq. (45) is valid only for Newtonian flow. For non-

R. Milleret al./Colloids SurfacesA: Physicochem.Eng. Aspects111 (1996) 75-118


Newtonian flow the viscosity is reduced 1-36] by
increasing 7xr and
h
/ d E - ~ A T x y \|
t/8 = ~ exp ~
kT
]
(46)
where is a coefficient determined by yxr~ = 0 for
7xr ~ 0 and ~ --, 1 for high 7xr. This approach does
not take into account possible structural changes.
Rehbinder [42] introduced the term structure
viscosity for decreasing viscosity when the deformation velocity increases at constant film pressure,
The character of the interfacial structure of a
monolayer depends on the interaction between the
kinetic units. In some cases decreasing viscosity is
caused by reorientation of kinetic units in the flow
field [43]. A structural change of a polymeric
system for instance contains an additional time
dependence, i.e. the system is thixotropic. For a
thixotropic destruction we have
[
z = to exp

(
-b
)]
- a exp O2u~/ffy Ot

(47)

For the Eyring flow without destruction a = b = 0


and z becomes a maximum. For a Rehbinder flow
a, b # 0, the relaxation time z changes during the
flow process, and a splitting of the activation
energy into an entropic and enthalpic term
becomes useful. For two different structures characterised by S* and S* (linearity provided) through
the relation

OZux
S* = S * +av~y ~y&
the viscosity becomes
r/8 = ~ exp

exp

(48)

- ay~r

(49)

For a flow without destruction Eq. (49)simplifies


to
h
(~
)
t/8 = ~ exp
- aTxr
(50)
i.e. a decreasing potential
Y~r. When a structure is
increases, accompanied by
shear viscosity. If kinetic

barrier with increasing


destroyed the entropy
a decreasing interfacial
units in a shear field

83

orient, then the configuration entropy decreases.


Furthermore the formation of a new structure, for
instance for polyelectrolytes [44], as well as the
formation of fluctuating networks of hydrogen
bridges have been discussed [45].

3. Interfaeial shear rheology


Interfacial shear rheological studies of adsorption layers, monolayers or other organised thin
liquid layers along with other interfacial investigations are useful for the discussion of their structure
at liquid/gas and liquid/liquid interfaces. This section gives an overview of different experimental
techniques and presents some selected examples of
systems studied in the literature.
Numerous experimental techniques have been
employed for the measurement of interfacial shear
viscosity. To avoid the presence of surface tension
gradients during the generation of surface flow two
different types of surface viscometers are typically
used, involving: (i) the determination of surface
velocity profiles (indirect methods) and (ii) the
determination of torsional stress values (direct
methods).

3.1. Indirect measurements ofinterfacial shear


rheological properties
Indirect methods are based on the determination
of surface velocity profiles measured by different
experimental techniques, for example canal or
rotating wall viscometers.

3.1.1. Canal surface viscometer


The measurement of r/s can be made in a manner
entirely analogous to the Hagen-Poiseuille law for
the bulk viscosity of liquids. It is based on the
determination of the flow rate of a film through a
narrow canal or slit under a two-dimensional
pressure difference Arc. The classical canal surface
viscometers, first proposed by Myers and Harkins
[-46] and Dervichian and Joly [47] are based on
this idea. A schematic sketch is given in Fig. 1.
The surface viscosity r/~ is calculated from the
rate of film flow Q through the canal of width a

84

R. Miller et al. /Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

--------r

~
Fig. 1. Canal surface viscometer,
and length L according to Eq. (51)

Ariaa
qs - 12LQ

(51 )

viscometer, was proposed by Mannheimer and


Schechter [51-53] and is a modification of
Davies' device.
The deep-channel surface viscometer consists of
two concentric, stationary vertical cylinders and a
rotating flat-bottomed dish containing the liquid
(see Fig. 2). The cylinders are placed so that they
almost touch the bottom of the dish. The dish
rotates at a known angular velocity, shearing the
fluid between the channel walls, and enhancing the
effect on the interface. The centreline surface
motion within the channel is monitored by a very
small PTFE particle placed within the fluid interface. From the experimentally measured centreline
surface velocity, the interfacial shear viscosity ~/s
can be calculated.
For laminar, Newtonian, time-independent flow
in the channel, the centreline velocity in the channel
is given by Pintar et al. [54] as

A corrected form was given by Harkins and


Kirkwood [48]

Ana3
qs-12LQ

arl
n

8e =~o(1 + zt ~/~)

V~-

(53)

(52)

including a correction term for the drag of the


underlying viscous liquid having a bulk viscosity
q. Eq. (52) is valid only if the canal is narrow,
L >> a, and the walls are smooth and parallel. Other
assumptions are that there is no slip of the film
along the walls, and that there is Newtonian flow
and constant surface viscosity within the canal,
Hfihnerfuss [49] suggested a redefinition of the
parameter "surface viscosity", which includes the
coupled system surface film/film-induced water
layer. Based upon this definition it is possible to
measure time-dependent surface viscosity values of
both relaxing and relaxed surface films by means
of the canal method,

n (l+n

Eq. (54) relates the surface shear viscosity to the


experimentally available velocities V* and F~ for
the pure and loaded interfaces:
t/o__yo (V*
"~
qs =
\ V~ - 1]
(54)
n
The analysis of Pintar et al. [54] was based on
assumptions that do not very closely approach
inner fixed

cylinder

~m
r.

3.1.2. Deep-channelsurfacev i s c o m e t e r
It is not easy to achieve all the above mentioned
theoretical conditions of a canal surface viscometer
experimentally. This early technique is not free
from dilational motion. To avoid this complication,
Davies [50] proposed a modification called the
double knife-edge viscometer. This type of viscometer is relatively insensitive. The modern canal surface viscometer, called the deep-channel surface

q-~)

channel

outer fixed

cylinder

Yo m /
|

T
x
~Y

~
%

~
rotating sha~

Fig. 2. Deep-channel surface viscometer.

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

reality. First the assumption of a flat shape of the


fluid interface can not hold if the separation width
of the concentric cylinders is close to the value of
the outer cylinder radius. Another deviation from
a linear canal occurs when the contact angle
between the interface and the walls is greater than
90 . A further source of deviation of the analysis
from experiments is the neglect of the small gap
between the rotating dish and channel walls,
A theoretical analysis of interfacial shear viscosity measurements at liquid/liquid interfaces with
an improved deep-channel surface viscometer is
presented by Wasan et al. [55]. This analysis takes
into account the finite depth of the two liquids
and permits interpretation of the experimental results in terms of interfacial shear viscosities
at the liquid/gas interface in addition to the
liquid/liquid interface. The results of the analysis
clearly show a coupling of interfacial and bulk
fluid flows. Measurements of the interfacial velocities at xylene/water and hexadecane/water interfaces were performed. The interfacial flow data for
pure systems were in excellent agreement with
the theory.
Gupta and Wasan [56] presented the results of
extensive measurements of the surface shear viscosity of a number of different soluble surfactant
systems. None of the soluble single-component
surfactants investigated yielded significant surface
shear viscosities. However, dodecanoic acid and
dodecyl alcohol, practically insoluble in water but
solubilised by sodium dodecylsulphate (SDS),
yielded significant surface shear viscosities at certain concentrations and proportions. Relatively
insoluble substances, or suchsubstances solubilised
by other substances, usually cause significant surface shear viscosity at liquid/gas interfaces. The
surface rheological behaviour is often dependent
on the rapid aging of surfactant solutions. A paper
by Mohan et al. [57] deals with the interpretation
of the interfacial shear flow data in the deepchannel surface viscometer in terms of aging effects
on surface viscosity. It was found that the surface
shear viscosity increases considerably with aging
of dilute aqueous surfactant solutions. For the
same age of the interface, the surface shear viscosity
increases with surfactant concentration,
Deemer et al. [-58] have extended the analysis

85

for the liquid-gas deep-channel surface viscometer


to the liquid-liquid-gas case with two interfaces.
They discuss a class of interfacial studies with
surface stress as a function of the surface deformation rate. The analysis takes into account the
viscous effects in all three phases. By measuring
the gas-liquid surface velocity distribution, the
liquid-liquid interfacial behaviour can be determined, as long as the shear stress distribution at
the gas/liquid interface has already been found.
Experimental results for a surfactant solutiondecane-air system are also discussed.
The effects of salt concentration and nature on
the surface viscosity of monolayers of different
diethanolaminederivativesofn-alkylsuccinicanhydrides have been measured with a deep-channel
surface viscometer by Chattopadhyay et al. [59].
To study the molecular mechanism for destabilisation of foams by organic ions Blute et al. [60]
measured the surface viscosity of SDS solutions in
presence and absence of organic salts with this
technique.
3.1.3. Rotating wall knife-edge surface viscometer

The rotating wall knife-edge surface viscometer


introduced by Goodrich et al. [61] measures the
ratio of displacement between the surface velocity
and the angular velocity of the rotating outer ring.
From this ratio the interfacial shear viscosity can
be calculated. The method is capable of achieving a high sensitivity ( 10-5 mN s m- 1 < r/S < 0.1
mN s m-l), does not suffer from the difficulties of
placing a knife "at" the interface, and does not
break the plane of the interface. As with the deepchannel surface viscometer, a small particle is
placed within the surface to discern the rotational
speed of the fluid interface.
The basic experimental design of the rotating
wall knife-edge surface viscometer is depicted in
Fig. 3. A cylindrical dish is filled with a liquid to
a level at least as high as the radius of the dish.
The upper surface of the liquid contacts the wall
of the container along a knife-edge ring that has
been inserted into the wall. Rotation of the ring
creates a viscometric flow in the vicinity of the
interface. A small talc or Teflon particle is placed
in the centre part of the interface, and the period
of rotation is monitored. From the ratio of the

R. Milleret aL/ColloidsSurfacesA: Physicochem.Eng.Aspects111 (1996) 75-118

86

" ~

3.1.4. Transient rotating cylinder apparatus

where (t2p/f2r) is the ratio of particle to ring


angular rotation. For practical proposes, a leastsquare polynomial fit of the general integral expression [61] relating to the angular rotation of the
talc particle I2p and the surface shear viscosity qs
is used, applicable to arbitrarily large interfacial
shear viscosity for the range of dimensionless rotation speeds 0 ~<I2p/12r ~<0.7, and furnishes the relation

Krieg et al. [64] proposed a method to measure


surface shear viscosity of a Newtonian liquid/gas
interface which consists in monitoring the decay
of surface motion after a sudden cessation of a rigid
body motion of a rotating cylindrical container
with fluid. The analysis of Krieg et al. [64] was
extended to liquid/liquid interfaces by Hassager
and Westborg [65]. Both fluid phases are assumed
to be Newtonian and the interface too. The two
key parameters for the design and sensitivity of
this type of viscometer are the height and radius
of the cylinder. The angular displacements of interfacial elements were obtained from direct visual
observation and from video recordings of the
experiments. Small polystyrene spheres were used
as tracer particles. Analysis and experiments on
the influence of the radial position of a surface or
interfacial element on the angular displacement
were carried out. The results show that the point
of observation should be chosen close to the centre
of the container. By applying oscillatory shear
deformations of the cylinder apparatus, the interfacial viscoelastic response can be observed. The
theoretical basis of a frequency response method
to determine the material properties of a viscoelastic liquid/gas interface has been presented by
Prieditis et al. [66]. The method involves measurements, at the surface, of the amplitude change and
phase shift of oscillations imposed on a cup of
liquid. The method is most sensitive for geometries
where cup radius-to-liquid height ratios are
about 0.5.

t/s = qRr 0.5631 ~ -

3.2. Direct interfacial shear methods

]'
liquid

I
~

rota~nag
~

2~z~

Fig. 3. Rotatingwall knife-edgesurfaceviscometer.


displacement velocity to the applied angular velocity of the rotational ring, the interfacial shear
viscosity may be determined,
The simple geometry of Fig. 3 is used in the
theoretical analysis. The theoretical analysis of
Goodrich et al. [61] affords a simple analytical
expression. For very small interfacial shear viscosities and under the conditions of a small gap the
surface viscosity near the cylindrical axis is asymptotically given by
qs = 0"5631(g2p/Or)r/Rr
(55)

+ 1.1189

\Q,/

-0"6254/O\'|""
~-~P)//O\?-I
\ O r } + 3.4489

(56)

which is valid to within an accuracy of 0.3%.


Some experiments with this device on stearic
acid monolayers spread on water have been
reported by Poskanzer and Goodrich [62,63]. It
was found that the surface shear viscosity of these
monolayers is very low until high surface concentrations are reached. Over all ranges of surface
concentration studied the rheology of stearic acid
monolayers is Newtonian.

In contrast to the indirect methods the direct


methods determine the torsional stress values of
the interface, which is stressed by a body just
touching it, under investigation.

3.2.1. Torsional stressmeasurements


The determination of surface shear viscosity by
measuring the damping of an oscillating torque
pendulum touching the interface is one of the
oldest methods in surface rheology [67].
The classical knife-edge surface viscometer of
Brown et al. [68] consists of a knife-edge bob

R. Miller et aL/Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

suspended from a torsion wire such that the circular knife just touches the surface of a solution
contained in a cylindrical vessel. The measuring
vessel is forced to rotate, and the torsional stress
on the knife-edge is measured in order to determine
the surface shear viscosity. Mannheimer and
Burton [69] improved the theoretical analysis of
the typical "knife-edge" torsional viscometer. Based
on a similar principle, two more versions, the
double knife-edge surface viscometer [50,70] and
the blunt knife-edge surface viscometer [71,72],
have been developed. The sensitivity is of the order
of qs >~0.01 m N s m - 1.
Many surface shear viscometers described in the
literature utilise torsional stress measurements
applied to a rotating disk within, or near to, a
liquid interface [-72-76]. The disk interfacial shear
viscometer consists of a fiat, circular disk suspended from a torsion wire into the interface. The
measuring vessel is rotated with angular velocity u
and a torque is exerted on the disk by both the
surfactant film and the viscous liquid,
If a ring or disk is forced to oscillate in the
interface, the rheological parameters are better
defined, particularly when the interface is bounded
by a concentric outer ring [77]. The equation of
motion for either free [41,78] or forced [78]
oscillation has been solved. In the case of free
oscillation, with the assumption of the additivity
of resistance due to the interface and the adjoining
phase, the surface shear viscosity can be determined
by measuring the decrements of damping and
periods of oscillation for a free and surfactantloaded interface. In forced oscillation experiments
the amplitude ratio and phase angle are the experimentally determined variables. These methods of
measuring the surface viscosity are in principle
invalid since the viscous interactions between the
interface and the adjoining phases are not taken
into account. The improved theoretical analysis of
Mannheimer and Burton [69] estimated an additional torque due to the viscous interaction
between the flowing monolayer and the substrate.
Calculations based on this analysis indicate that
surface viscosities reported from current torsional
formulae - - all of which assume that the film slips
over the substrate - - may be in serious error.
Examples are also presented that illustrate how

87

comparative torque measurements can lead to


anomalous results when the effects of temperature
on surface viscosity are considered.

3.2.2. Knife-edge surface viscometers


The knife-edge surface viscometer of Brown et al.
[68] is a rotational viscometer (Fig. 4). It consists
of a knife-edge bob suspended from a torsional
wire such that the circular knife just touches the
interface of a surfactant solution contained in a
cylindrical cup. The cup is rotated and the torsional
stress on the bob measured to determine the interfacial shear viscosity. Only the steady state solution
is observed in this case. A constant rate of torque
produces a constant rate of shear in the film. The
surface torsion pendulum operated in this way is
the two-dimensional analogue of the well known
bulk rotational viscometer (Couette type apparatus). The film can be sheared at a constant rate
only if it is either purely viscous or after the elastic
forces have been ruptured. The limiting stress
necessary to overcome the elastic forces is called
the surface yield stress. In this case the shearing
element will show a constant angular displacement
in the steady state. Measurements at constant rate
of torque in the steady state cannot therefore yield
information on the elastic properties of the surface film.

//////
torsio.,~e
R
I
R

air

RI

liquid

,o .(~

>

Fig. 4. Single knife-edge surface viscometer, after Brown et al.


[68].

88

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

Mannheimer and Burton [69] offer the


following expression for the torque upon the torsional wire in the single knife-edge experimental
apparatus
knife torque

na2Rrqg2r
-

[~1
i

[qq~i(a~i/Rr--)]2
1

1}

(~i)2 + i coth(h~i/Rr)

[jo(~i)] 2

(57)
where J, (a~i/RO and Jo(~i) are the first- and
zeroth-order Bessel functions of the first kind,
respectively, and ~ are the roots of J,(~)= 0.
The double knife-edge (schematically shown in
Fig. 5) and blunt knife-edge surface viscometers
are similar in design and practically the same as
the single knife-edge viscometer discussed earlier,
The theoretical description of the respective viscometric flows has been provided by Lifschutz et al.
[70] and Briley et al. [72].
With each of these knife-edge viscometers it is
crucially important that the knife makes contact
with, but not break, the fluid interface. For this
reason the knife-edge has to be made non-wetting,
though the problem of knife placement is still
apparently cumbersome [69].

3.2.3. Blunt knife surface viscometer


Surface rheological measurements with flat rings
and blunt edges suspended on a torsion wire have
been carried out by several authors, e.g. Inokuchi
[79-83], Tachibana et al. [84], Motomura and
Matuura [85], Biswas and Haydon [86,87], and
Patil et al. [88,89].
An axisymmetric technique for measuring the
shear modulus, viscosity and surface tension of a
spread monolayer using a small blunt knife PTFE
edge was introduced by Abraham et al. [90]. This
apparatus enables one to measure simultaneously
the surface shear modulus, the surface shear viscosity and the surface pressure of a monomolecular
film on water under a hydrostatic compression. It
is particularly suitable for measurements of the
static shear modulus of insoluble monolayers. A
capillary wave generator/detector system is also
incorporated to determine the surface pressure.
The trough is a PTFE cup with sloping interior
sides. As the cup is raised the film on the water
surface is uniformly compressed. There are no
barrier seals and the circular symmetry is mainrained at all times. A small blunt knife edge suspended at the centre from a fine torsion wire is
used to measure the rheological properties of the
monolayer (see Fig. 6).
To measure the quasi-static shear properties, the
cup is rotated through a small angle and the
residual deflection of the blunt knife edge is deter-

torsion

torsionwire
R

R
air
-

Rl
Rz

Fig. 5. Double knife-edge surface viscometer.

fiquid

o)~/-

co

Fig. 6. Blunt knife-edgesurface viscometer, after Abraham et al.

89

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

mined. The dynamic shear response is probed


by studying transient torsional oscillations following a sudden angular displacement of the torsion wire. The apparatus has a resolution of
5 x 10-4 mN s m - 1 of shear viscosity and
1 10 -4 mN m -1 of static shear modulus. First
measurements of classic monolayer systems (stearyl
alcohols and nonadecanoic acid) spread on water
have been made as a function of density by
Abraham et al. [91], who have published a detailed
description of several improvements of this sensitive apparatus [92]. An exact solution to the
viscoelastic damping of a torsion pendulum by a
surface film developed for the device of Abraham
et al. [90] has been given by Feng et al. [93]. In
earlier works by Abraham and coworkers an
approximate analytical solution was used, assureing that the viscous penetration depth into the
substrate phase is much less than the wavelength
of the surface viscoelastic shear wave. A theoretical
basis for the investigation of viscoelastic behaviour
of surface films has been established by obtaining
the exact solution to the boundary value problem
of a film-covered fluid set in free oscillatory motion
by a torsion pendulum rotor in a shear trough,
The surface shear viscosity of valinomycin
spread on buffered aqueous substrates as a function
of surface pressure has been determined by
Abraham and Ketterson [94]. Studies of viscoelastic measurements of dimyristoyl phosphatidylserine and bovine brain phosphatidylserine
monolayers at the air/water interface have also
been published by Abraham and Ketterson [95].
Peng et al. [96] measured the shear moduli and
viscosities of monolayers of poly(methyl methacrylate), poly(vinyl stearate), octadecanol and
behenic acid at various stages of compression. The
viscoelastic properties of the polymer monolayers
were found to have much higher values than those
of the single amphiphiles. The authors found a
clear correlation between the shear viscoelasticity
of the monolayers and the development of surface
pressure gradients during compression.

[72-76,86,87]. This type of viscometer is moderately sensitive (disk viscometers are typically sensitive to interfacial shear viscosity values in the range
t/s ~>0.01 mN s m-l). In addition, such viscometers
generally require numerical evaluation of measured
parameters. Nevertheless, torsional measurements
avoid the difficulties of tracking the interfacial
velocity - - the primary drawback of the deepchannel surface viscometer. Torsional devices are,
therefore, useful practically for the measurement
of interracial shear viscosities of highly viscous
surfactant monolayers [97].
The classic disk surface viscometer apparatus is
depicted in Fig. 7. A thin, fiat, circular cylindrical
disk is rotated within the plane of an interface with
angular velocity f2r. A torque is exerted upon the
disk by both the surfactant film and the viscous
liquid. According to the theoretical model, the
torque exerted may be decomposed into a lineal
surface traction torque along the rim of the disk
(through which the interfacial shear viscosity t/s
imparts a direct influence), and a tractional torque
owing to familiarity of the bulk-phase liquid along
the base of the disk. Standard analyses do not
account for contact angle effects nor for the effects
of a finite disk thickness.
In the limit of zero Boussinesq number
(58)

Bo---~O

t/Rr

torsion

R
air
a
liquid

I I
tO~

3.2.4. Disk surface viscometer

Many surface viscometers exist in the literature


which utilise torsional stress measurements upon
a rotating disc within or near to a liquid interface

Fig. 7. Flat disc surfaceviscometer.

90

R. Miller et al. /Colloids Surfaces A: Physicochem. Eng. Aspects I l l (1996) 75-118

at
a
---,0
h

/,4,
and

a
--~0

(59)

Rr

the torque exerted upon the rotating disk is provided to leading order by the expression [73]

liquid2

disc torque = ( 16/3 )a 3 r/O r

liquid l

(60)

In the limit of large Boussinesq number


Bo--,oe,

at~0

and

R~0

(61)

the following relation holds:


disc torque ---a(8/3)aa~lf2r + 4rca2t/s~2r
a

co
(62)

General torque expressions are provided by Briley


et al. [72] and Shail [74], the latter provides
additionally the higher-order terms of the above
expressions. Theoretical results for liquid-liquid
systems are provided by Oh and Slattery [75].
A second type of disk surface viscometer [98,99]
employs a thin disk rotating below the fluid interface. This technique may exhibit an advantage
relative to the preceding designs because it avoids
the delicate placement of the disk "in" the interface,
It also eliminates possible contact angle complications. The sensitivity of this configuration to interfacial shear viscosity is reduced however because
the disc is placed in the bulk fluid at a finite
distance from the interface. A similar analysis has
been carried out by Davis and O'Neill [76] for a
rotating sphere beneath a surfactant-loaded fluid
interface,
A variation of the disk viscometer is the biconical
bob interfacial viscometer depicted in Fig. 8. The
theoretical analysis [75] for this device is quite
similar to that of the disc viscometer, and the
technique may be regarded as possessing similar
advantages and limitations,
A theoretical analysis for the disk torsional
interfacial viscometer with oscillatory motion is
given by Ray et al. [ 100]. A sinusoidal oscillation
is imposed upon the angular velocity of the dish.
The bob is induced to oscillate with the same
frequency as the dish. From measurements of the
amplitude ratio and phase angle between these two
oscillations, the surface shear viscosity can be

t ~
Fig. 8. Biconicaldisc interfacial viscometer.

determined. The theoretical analysis is confirmed


by experimental measurements and the typical
behaviour of an interface between the two liquid
phases is obtained.
3.2.5. An automated interfacial torsion shear
rheometer
Equipment to measure interfacial rheological
behaviour at very low shear rates is described by
Kr~igelet al. [ 101 ]. As we have already mentioned,
rheological investigation of adsorption layers at
fluid interfaces requires equipment which does not
disturb the structure of these layers during the
measuring procedure. For this reason the torsion
pendulum technique is preferred in surface shear
rheology. This newly designed method allows
experiments with very small mechanical deformations of the adsorption layer. The principle of the
rheometer is based on a ring with a sharp edge
hanging from a torsion wire. When applying an
impulsive torque by an instantaneous movement
of the torsion head, the pendulum performs
damped oscillations with damping factor fl and
circular frequency e). This kind of experiment
provides simultaneously information on the surface
shear coefficient of viscosity and the surface shear
modulus of rigidity from a single experiment.
The schematic set-up of the rheometer, available
as the commercial instrument ISR1 from LAUDA,

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

Germany, is shown in Fig. 9. The active part


(stepper motor, gearing, motor controller) comprises a drive for the deflection, at which the
torsion wire with a circular measuring body is
fixed. The body touches the liquid surface or is
situated at the liquid/liquid interface. The angular
position of the body is registered by means of a
mini laser and a position-sensitive photosensor
with an accuracy of +__0.01 at deflection angles of
2 . The whole instrument is computer-driven and
the software controls all calibrations and measurements and acquires and analyses the experimental data.
The application of this torsion pendulum apparatus and its sensitivity and accuracy have been
demonstrated by different model experiments
[ 101-104]. The instrument is expected to be applicable also to liquid/liquid interfaces. It is very
sensitive so it can be used to investigate insoluble
monomolecular layers as well [105].

3.2.6. Other techniques


Another knife-edge surface viscometer has been
proposed by Shahin [ 106], in which a thin rod is
placed within the planar surface of a solution
contained in a trough. A magnetic field forces the

rod to move within the interface. Small particles


are placed within the wake of the moving rod to
track the interfacial shear field. By establishing a
relation between the force exerted upon the rod
and the shear field created, interfacial shear viscosity (as well as non-Newtonian interfacial rheological properties) may potentially be detected. Shahin
[106] cites a sensitivity to the interfacial shear
viscosity of 10-5 mN s m-~, though an inconclusive amount of data has been reported.
Another surface shear viscometer, proposed by
Gaub and McConnell [107], uses a thin glass disk
equipped with a small magnet which is made to
rotate by an external magnetic field. The surface
viscosity is measured by determining the viscous
drag on the rotating disk that floats on top of a
spread monolayer. The surface viscosity is calculated from the torque at the steady-state condition.
An interesting and unique peculiarity of interfacial rheology is represented by the device for
characterising the interfacial behaviour of foam
lamellas designed by Zotova and Trapeznikov
[108]. A foam lamella was positioned in the gap
between a ring hanging on a torsion wire and a
second ring on the wall of a climate chamber. The
surface viscosity determined relates to the two

motor lii ~_ steppermotor


con~
withgearing

steppermotor
processorcard
minilaser
~
mirror , ~
knife-edge ."......................... dq
'~
~
water PSD
--jacket
I

lifting

91

I D/Ai

A/Dconverter

measuring
. vessel
Fig. 9. Automated interfacial shear rheometer ISR1.

92

R. Miller et al./Colloids Surfaces A." Physicochem. Eng. Aspects 111 (1996) 75-118

interface layers. The device has been used to characterise both surfactant 1-109,110] and protein
lamellas [ 1l 1,112].

3.3. Selected experimental results of surface shear


rheological studies
Couette type surface rheometers are the most
widely used instruments for investigating surface
shear rheological properties. The measuring
interval can easily be changed across a wide range,
which is very important because of large differences
in the surface rheological behaviour of different
surface active materials. Very different techniques
have been developed to characterise these properties and to determine both viscous and elastic
parameters over a broad time interval. A short but
very instructive summary of such studies has been
given by Izmailova et al. [ 113 ]. Protein and coadsorbing systems consisting of proteins and surfactants have been used most often to determine
viscoelastic and viscoplastic properties, thixotropic
effects [ 1t4], relaxation spectra [86,87], and creep
compliance [81,115-125]. The results of such
investigations are applied for example in the photographic industry [126] and in food research
[127-133].
In the following sections a few experimental
results are presented to demonstrate the application of different techniques to various chemical
systems. The authors realise how subjective and
incomplete this selection is.

3.3.1. Insoluble surfactant monolayers


Abraham and coworkers have performed shear
rheological measurements on insoluble monolayers
with high precision. The surface viscosity of valinomycin spread on aqueous substrates (buffered to
pH 6.7 with triphosphate) has been determined by
Abraham and Ketterson [94] over a surface pressure range from 7 to 3 0 m N m -1. The surface
pressure-molecular area diagrams H(A) in all
cases, with and without Na or K ions in the
water, are monotonic and almost identical. The
surface shear viscosity remained essentially constant at 1.8 mN s m-1 over the entire compression
range, regardless of the presence of Na + or K
ions. There was no effect which could be attributed

to a phase transition or a conformational change


so it was concluded that all monolayers had essentially the same conformation.
Abraham and Ketterson [95] also measured the
surface shear viscoelastic properties of dimyristoyl
phosphatidylserine (DMPS) and bovine brain
phosphatidylserine (BBPS) monolayers at the
air/water interface. The substances were spread on
buffered aqueous substrates and compressed slowly
in discrete steps. At a number of compression
stages, the response of the film to a shearing stress
was seen to be a function of time. From the data
collected both the surface elasticity and the surface
viscosity were calculated by using a Maxwell viscoelastic model. The ~-A diagram of D M P S shows
a slight pH dependence while that of BBPS, on
the other hand, is essentially pH independent. The
viscosity displayed by D M P S is attributed to weak
networks formed by hydrogen bonding between
water and the serine moiety of the phospholipid.
The large number of double bonds in BBPS
expands the monolayer, which reduces both head
and tail group interactions and lowers the viscosity.
Peng et al. [96] measured the interfacial shear
moduli and viscosities of monolayers at various
stages of compression of two polymers, poly(methyl methacrylate) (PMMA) and poly(vinyl
stearate) (PVS) and of two simple amphiphiles,
octadecanol and behenic acid. The monolayers of
the simple amphiphiles exhibited low values of
these viscoelastic properties. The surface shear
elastic moduli have values below 0 . 2 m N m 1
and the surface shear viscosities are below
0.25 mN s m - 1. Changes with increasing surface
pressure appear to correlate with changes in monolayer phase. Fig. 10 illustrates that the shear viscosities of the monolayers of small amphiphiles show
abrupt rises at about 15 mN s m -1 for octadecanol
and about 5 mN s m-1 for behenic acid. At higher compressions the values drop sharply down
to about 2 9 m N s m -1 for octadecanol and
22 mN s m -1 for behenic acid.
The interfacial viscoelastic properties of the polymer monolayers have much higher values than
those of the simple amphiphiles: for PMMA, the
values of interfacial shear modulus and viscosity
were up to 20 mN m - 1 and 700 mN s m - 1, respectively, for PVS up to 100 mN m-1 and 5 N s m-1

93

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

0,25

0,2

0,15

0,0
0

_
5

10

15

. . . .
20

25

30

35

40

surface pressure [mN/m]


Fig. 10. Shear viscosities of octadecanol, 0, and behenic acid, II, monolayers as a function of surface pressure, according to Peng
et al. [96].
respectively. The interracial viscoelastic properties
of P M M A show a dependence on the molar mass
when compared on an area per molecule basis but
little dependence when compared at the same
surface pressures. A clear correlation between the
inter facial shear viscoelasticity of the monolayers
and the development of surface pressure gradients
during compression could be demonstrated,
HOhnerfuss [49] and Ht~hnerfuss et al. [134]
measured the time-dependent surface viscosity
values of both relaxing and relaxed surface films
by means of the canal method. Monolayer film
viscosities for fatty acids have been investigated by
many experimentalists, who clarified the effect of
alkyl chain length and metal ions added to a water
phase. It has been found that the monolayer property is profoundly affected by their hydrophobic
and hydrophilic interactions. In recent studies
Hahnerfuss [135,136] outlined the optimum
hydrophobic interaction between the hydrophobic
alkyl chains that results in an increase in surface
viscosity. The surface viscosity values measured for
a series of homologous compounds by a canal

surface viscometer were compared and optimum


hydrophobic interactions were found.
Using a modified version of the above described
torsion shear rheometer the surface shear viscosity
for D P P C and D M P E were measured. The results
are shown in Fig. 11 as a plot of t/s vs. the film
pressure ~. The determined shear elasticity was
independent of the surface pressure and close to
zero. At small surface pressure the shear viscosity
is close to zero. With increasing rr values the
viscosity increases first almost~ linearly and then
abruptly to reach a plateau value at very high
surface pressures.
The shear viscosity is only insignificantly
changed up to the end of the coexistence region.
Beyond this point the shear viscosity increases
linearly. For D M P E a further abrupt increase to
very high viscosity values is observed after the
transition point from a liquid condensed (LC) to
a solid-like state at a pressure of about ~ =
21 mN m -1. For D P P C a similar behaviour is
observed. Although a transition from the LC-phase
to a solid-like film at about ~ = 32 mN m-1 was

94

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects Ill (1996) 75-118


2500

"~ 2ooo

"oo 1500
co

1o o o

500
U3

}r~[34zIt3--u
0

10

~m~-m~4mJ
20

30

40

50

Surface Pressure lmN/m]

Fig. 11. Surface shear viscosities of spread DPPC, U], and DMPE, m, monolayers as function of surface pressure, according to
Kr~igelet al. [105].
not reported, the shear viscosity increases sharply,
Maybe the monolayer undergoes a phase transition
which has not been observed yet. X-ray reflection
in this range of surface pressure should be best
suited to give evidence of such a process if it exists,
A comparison of these results for D P P C with
the dilational rheological behaviour of the same
monolayer will be given in Section 4.4. It can be
shown that both shear and dilational rheology
correlate with the morphology of the monolayer
as observed by optical methods, such as Brewster
angle microscopy [137,138].
3.3.2. Polymer adsorption layers

Proteins are the most frequently studied polymers at liquid interfaces. Therheologicalbehaviour
of protein adsorption layers has been extensively
studied by Dickinson and coworkers. Dickinson
et al. [ 131] measured the time dependence of the
interfacial viscosity of adsorbed films of casein and
gelatine at the oil/water interface using a Couettetype torsion wire surface viscometer operating at
low shear rate. The surface shear viscosity is calculated from the steady-state biconical disk deflection
of the torsional pendulum. The measured surface
viscosity of a mixed protein film adsorbed at the
interface between n-hexadecane and a dilute aqueous solution of casein and gelatine is found to

depend sensitively on the age of the film and the


protein composition in the bulk phase.
Dickinson et al. [132] also studied the influence
of the adsorbed protein film interfacial rheology
on coalescence stability of emulsion droplets at a
planar oil/water interface. Distributions of coalescence times as a function of droplet size are presented for systems with n-hexadecane as the oil
phase and a protein solution of B-casein, x-casein
or lysozyme as the aqueous phase. When the planar
interface is aged for just 20 min there is no measurable difference in the distribution of the coalescence
time for the three proteins. However, when the
planar interface is aged for 72 h there is a sizeable fraction of droplets, different for each protein, which do not coalesce at all, and the
relative efficiencies of the adsorbed proteins in
preventing coalescence lie in the order: lysozyme > x-casein > B-casein. Time-dependent surface shear viscosities at the oil/water interface are
reported for adsorbed films of the same three
proteins under exactly the same experimental conditions. After 20 min the surface viscosity values
for all three proteins are low (< 1 m N s m - 1), but
after 72 h their values are distinctly different:
200 m N s m -1 for lysozyme, 60 mN s m -1 for
x-casein and < 1 m N s m - ~ for B-casein. These
results are consistent with the view that there is a

95

R~ Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

positive correlation between coalescence stability


and the mechanical strength of protein films
adsorbed at the oil/water interface,
Time-dependent surface viscosities are reported
by Dickinson et al. [133] for films adsorbed from
binary mixtures of the proteins ~-lactalbumin,
/3-1actoglobulin and /%casein. The measurements
were made at a planer interface between
n-tetradecane and various aqueous protein solutions. Differences in the behaviour between simultaneous and sequential exposure of the pairs of
proteins to the interface were investigated. Some
experiments were performed with chemically modifled /3-1actoglobulin samples whose disulphide
bonds had been split and blocked. Displacement
of one protein by another (e.g. ~-lactalbumin by
/~-casein) is indicated by a sudden drop in surface
viscosity immediately after addition of the second
protein. In systems containing/3-1actoglobulin, the
long-term surface viscosity is very sensitive to the
adsorption time of/3-1actoglobulin prior to addition of the second protein. Blocking the disulphide
bonds in /3-1actoglobulin leads to a much faster
approach to a steady-state surface viscosity. Fig. 12

shows the effect of adding the second protein. It


leads to a "freezing" in of the surface shear viscosity
of the system to the value of that for the
/%lactoglobulin film at the time of the addition - whether it be 5 min, 30 min, 4 h or 24 h. This
behaviour is interpreted by the authors in terms
of a much more rapid unfolding of the disordered
molecules of modified /3-1actoglobulin at the
oil/water interface. It is concluded that surface
viscosity experiments give useful and sensitive
information about competitive adsorption and
cooperative interactions in mixed protein films.
The interfacial shear rheological properties of
a two-dimensional polymerisation process were
studied by Rehage and Veyessie [139-141].
Diacrylate diesters are surface active and form
monomolecular films at the oil/water interface.
These molecules polymerise and cross-link in a
monolayer after UV-irradiation. The static shear
modulus of the two-dimensional network was measured with a surface torsional viscometer by
Miyano and Veyessie [142]. A clear transition
from a viscous solution to a cross-linked gel state
has been observed. The two-dimensional structures

1400
1200

1000
800
600
400
200
0
0

I
10

I
20

I
30

i
40

I
50

60

time [h]
Fig. 12. Effectof added c~-lactalbumin(10 -3 wt%) to a ]Y-lactoglobulinfilm (10 -3 wt%) adsorbed at the n-tetradecane/waterinterface;
time of addition (a) 5 min, (b) 30 min, (c) 4 h, (d) 24 h, according to Dickinson et al. [133].

96

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

behave like rubber-elastic systems. The elastic


properties depend strongly on the concentration
of the cross-linking agent. Through polymerisation
of mixed films of monoacrylates and diacrylates
two-dimensional structures with defined rheological properties can be designed,
3.3.3. M i x e d protein-surfactant adsorption layers

Minimal surface shear viscosities of 1 ktN s m- 1


were detected with sufficient accuracy using an
equipment to study creep compliance [143] in
agreement with the theoretical analysis for an
interface-touching ring [118]. Using this instrument an experimental study of surface rheology of
adsorbed gelatine layers was carried out [116].
After determining the rheological parameters the
support solution (gelatine in phosphate buffer) was
substituted by a gelatine-free phosphate buffer
solution. The difference in the parameters calculated before and after the substitution of the bulk
phase are used to prove some model assumptions
of protein adsorption and the physical values of
the surface rheological parameters. The calculated
rheological parameters allow us to distinguish
between three different concentration ranges of
gelatine adsorption: the range of reversible adsorption, the saturation and the formation of irreversible adsorption layers, and a range of molecular
rearrangements. In agreement with other experiments a possible multilayer formation was
discussed,
In another study by Wiastneck and coworkers
[117,125] adsorption layers of gelatine solutions
with different average molecular weight and ash
contents were modified by the addition of different
surfactants. It was shown that the addition of
anionic and cationic surfactants leads to extreme
changes of the creep compliance behaviour in the
range of gelatin~surfactant complex adsorption.
The surface shear viscosity was increased by more
than three orders of magnitude. Strongly pronounced surface viscoplastic properties confirm the
existence of a saturated adsorption layer, containing gelatine-surfactant complexes of high surface
activity. At surfactant concentrations higher than
the corresponding CMC both the surface tension
and the surface viscosity become comparable to

the values obtained for surfactant solutions without


gelatine. This is explained by a decrease of the
surface activity of the complexes at higher surfactant concentrations where they are then replaced
by surfactant molecules. These different ranges of
gelatine-surfactant adsorption are confirmed by
measurements of the adsorption layer thickness
[144], the triple-helical structure of gelatine in
surfactant solutions [145-147] and the interfacial
tension at toluene/water and diethyl phthalate/
water interfaces [148]. Fig. 13 illustrates the
behaviour of different gelatine types as a function
of SDS concentration.
Insoluble monolayers of n-octadecanoic acid
show a typical Newtonian flow behaviour. When
spreading these monolayers on a gelatine adsorption layer the rheological behaviour is changed
dramatically by electrostatic interactions. This
interaction was prevented by addition of both
soluble anionic and cationic surfactants to the
gelatine solution [119-121].
The automated torsion pendulum instrument
described by Kr/igel et al. [101] was used to
determine the surface shear properties ofmacromolecular adsorption layers at fluid interfaces. The
experimental results for human serum albumin in
an aqueous buffer solution (pH 6.5, 0.99 g 1 1
Na2HPO 4 2H20 and 1.76 g 1-1 KH2PO4 ) are displayed in Fig. 14.
The surface shear viscosity of a 0.001 mg m1-1
HA solution increases with adsorption time and
levels off at about 150 min while the surface
shear elasticity still increases. The same picture
is obtained for the higher concentration of
0.1 mgm1-1 HA, only the absolute values are
about three orders of magnitude higher. Further
studies using the automated surface shear rheometer were performed with food proteins [103] and
polymeric humic substances [ 104].

4. Interfaeial dilation rheology


While only transient relaxation methods were
recently discussed by Miller et al. [149], in this
section the two types of relaxation techniques,
based on harmonic and transient area changes,
respectively, will be described from the theoretical

P~ Miller et aL/Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

97

10000

1000

100

-~

10

0,1

0,01

- =

0,001

1,00E-05

1,00E-04

1,00E-03

1,00E-02

1,00E-01

1,00E+00

SDS-concentration [mol/l]
Fig. 13. Surface shear viscosity of different gelatine types as a function of SDS concentration, according to WOstneck and MOller [144].

70

6O

50

o~ 40
30

2o

10

-"

I50

~
100

I
150

I
200

250

adsorption time [min]


Fig. 14. Surface shear viscosity of human serum albumin (HA) as a function of time, HA concentration 0.001 mg ml, II, and
0.1 mg ml -t, O; according to Kr~igel et al. [101].

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

98

as well as experimental point of view. It will be


shown that the exchange of matter functions are
generally applicable to both types of relaxations,
as pointed out by Loglio et al. [150]. Finally,
experimental details of several methods and examples of results from the literature, for surfactants
as well as biopolymers, will be discussed. A comprehensive discussion of dilational rheology is also
given by Dukhin et al. [ 12].

4.1. The difference in adsorption kinetics and


relaxation processes of surfactants and polymers at
interfaces
Before starting with the description of present
theories of interfacial relaxations and dilational
elasticity, we want to point out how different the
adsorption kinetics are and what an enormous
effect surface active impurities have on dilational
interfacial parameters. The general difference consists in the composition of the adsorption layer.
Adsorption kinetics processes usually start from
an uncovered interface. The species having the
highest concentration and correspondingly lowest
surface activity adsorb first. The best means of
estimating the rate of adsorption at the beginning
of the process is the ratio of surface concentration
Fo over bulk concentration Co. To compare the
adsorption rate of two surface active compounds
a rough estimate can be used to determine the
time needed by a surfactant to reach 95% of the
equilibrium adsorption,
--

tad

-~\Co/

(63)

This adsorption time does not, of course, represent


the time necessary to establish adsorption equilibrium but is only a characteristic adsorption time
for comparisons. As a good approximation for
many surfactants, the ratio Fo/co can be calculated
from the well-known Langmuir isotherm

Fo
Fo~/aL
Co 1 + (co/aL)

~<

Fo~
ae

(64)

Thus, the characteristic adsorption time is obtained

from
~ fFooX~ 2
/--/
tad~\aL/

(65)

Typical values of the diffusion coefficient and


the maximum surface concentration are D =
5 1 0 - 6 cm 2 s -1 and Fo~ = 5 x 10 -1 mol cm -2.
The characteristic adsorption time of sodium octylsulphate with a L = 3 . 5 10-6molcm -3 comes
to tad ~ 0.00004 S, for sodium dodecylsulphate
tad ~ 0.009 S, almost three orders of magnitude
higher.
From these approximations one can conclude
that in time intervals of the order O(100 tad) the
adsorption kinetics is controlled by the corresponding surfactant, for sodium dodecylsulphate
up to some seconds. At longer times, the effect of
surface active contaminations have to be taken
into consideration, as comprehensively discussed
for example by Lunkenheimer and Miller
[143,151,152]. At times t>> 100 tad, the properties
of the adsorption layer can already be dominated
by a contamination.
In contrast to adsorption, the process of desorption starts from a more or less pre-established
equilibrium adsorption layer. As the compressions
performed to initiate desorption processes are usually large, the situation is comparable to the process of adsorption, although it is much more
affected by surface active impurities.
If kinetics processes of adsorption or desorption
were observed in time scales t >> 100 tad, it is difficult to distinguish between impurity effects and
specific effects of the surfactant system, such as
electrostatic retardations, phase transitions in
adsorption layers, conformational changes, structure formations etc.
In relaxation experiments the situation is quite
different. These experiments are usually performed
when the equilibrium state of adsorption has been
established. Relaxations are defined as processes
setting in after the equilibrium state has been
disturbed a little. The disturbance of an adsorption
layer can be produced by changing the surface
concentration, the pressure or temperature or other
parameters. Relaxations then start to re-establish
an equilibrium state of the system. More general

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

99

discussions of relaxation processes are given in


Dukhin et al. [12], for example,
When surface active contaminations are looked
at again, the relaxation behaviour after an interfacial area disturbance is controlled simultaneously
by the main surfactant, the concentration of contamination and the ratio between them. The last
depends on the time allowed for the establishment
of equilibrium. In real systems only a pseudoequilibrium is ever reached,
In the ideal case where no contamination is
present in the system, relaxations enable study of
the dynamic behaviour of the adsorption layer,
Such investigations yield information about
adsorption mechanisms as well as interfacial interactions and transitions of coexisting phases. Due
to the small deviation from equilibrium, theories
of relaxation experiments are usually easier to
derive because linearisations are justified. Thus,
complex processes are better studied by relaxations
than by adsorption kinetics.

investigate the surface relaxation of soluble adsorption layers due to harmonic disturbances is the
oscillating bubble method. The technique involves
the generation of radial oscillations of a gas bubble
at the top of a capillary immersed into the solution
under study. The first set-up was described by
Lunkenheimer and Kretzschmar [172] and
Wantke et al. [173] followed by newly designed
instruments using a novel pressure transducer technique to register the pressure changes inside bubbles or drops [174-181].
Assuming isotropic area deformations the diffusional flux at the interface is given by

4.2. Interfacial relaxation techniques

Ot - D ~x 2

The two types of relaxation techniques, based


on harmonic and transient area changes, respectively, will be described in this paragraph from the
theoretical point of view. It will be shown that the
exchange of matter functions are generally applica-

If area changes are performed in an anisotropic


way, for example in trough experiments, the theoretical model has to take into account the lateral
transport of adsorbed molecules [ 182-184]. The
solution to the problem has the general form [ 185 ]

ble to both types of relaxations, as pointed out by


Loglio et al. [150].

c(x, t) = Co + ~ exp(/~x) exp(io)t)

1 d(FA)
Oc
A d ~ - D ~ x at x = 0

(66)

where A(t) is the time function of the interracial


area. The transport of molecules in the bulk by
diffusion is given by Fick's diffusion law
Oc
~2c
(67)

(68)

The boundary condition (66) can be rearranged to


4.2.1. Techniques based on harmonic interfacial
disturbances
The method of studying the damping of capillary
waves at interfaces is the classic version of all
relaxation methods at interfaces. Harmonic waves
are generated mechanically, thermally or electrically. The response of the system is usually mechanically measured in terms of a relative damping of
the propagated wave. Detailed descriptions of capillary wave techniques including specific theories
are given in numerous reviews [17,153-156].
Recent work has been focused on theoretical modifications to this technique [157-166] as well as on
experimental improvements and new equipment
[134,162,164,167-171].
One of the more recently developed methods to

d In F
[Oc/Ox
]- 1
~ n A - - L 1 + D (dF/d-d-~c/&ij/

(69)

Recalling the definition, Eq. (15), of the dilational


elasticity
d7 d In F
E - d In F d In A

(70)

the following relation is obtained

E-

dy [
dc Oc/~x ] - 1
d ~n F L 1 + D dF ~ J

(71)

The introduction of Eq. (68) leads to the expression

100

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

for E(ia)) as the final result


E(ie)) = E

x~
+ x/~0

(72)

with
(d~nn F )
d?

E=--

and

\~/
e)O=(dd~C~]2D
-~

(73)

A
For the special case of a Langmuir type equation
of state (64) to describe the equilibrium adsorption
behaviour of the system, the following exchange of
matter function is obtained
E(ia)) = RTFo~ Co
(1 + CO/aL) z
ae ~ 1 ~ +

(74)

obtained analogously with the method described


above, but the exchange function contains both
the diffusion coefficients of the surfactant m the
respective phase and the distribution coefficient of
the surfactant between the two liquids (see Refs.
[12,149].
The models of the complex elasticity, given by
Eqs. (77) and (78), are derived for surfactant
solutions well below the critical micelle concentration, CMC. Thus it is assumed that no aggregates exist in the solution bulk. Lucassen [186]
derived the corresponding function for the case of
micellar solutions and obtained
E(iog) = E [1 +(1 - i ) (
L

Fo~/at

To model the exchange of matter at the interface


of a mixed surfactant solution the same principle
can be used as for the system containing only one
surfactant. However, one term for each of the m
surface active compounds in the system is needed,
Garrett and Joos [18] derived the complex elasticity modulus given by

E = d ln-----~ i=~ l"T d ln A

X/(1 --ix)(1 + 3~/1 --ix) 2 + ik(52 - 1)7


1 - i~+ -6~1--- ix

(79)
with
{
x = k ~1+

6=

"]
n2

cn I
elf

kd
k o.)

(8o)

~=

with the total surface concentration defined by


m

Fx = ~', F~
(76)
i=1
The solution to Eq. (75) is found in the same way
as described for a single surfactant system and was
given in its general form also in Ref. [18] as
E(iog)--E ~ F/
~
i=I/-'T N/~---~-N//~i0

(77)

with
f dc;'x
2Di
(78)
e)~o=/---:\dFj
! 2
At liquid/liquid interfaces one must consider that
the surfactant is usually soluble in both adjacent
phases. The exchange of matter takes place in both
bulk phases and the diffusion laws must be therefore applied in both of these phases. The result is

This function has to be used to describe the


relaxation behaviour of surfactant solutions above
the CMC, or in solutions containing premicetles
or other aggregates,
The capillary wave method was developed more
than five decades ago. Many authors have contributed to the theory, such as Levich [187-189],
Hansen [190], Hansen and Mann [191], van den
Tempel and van de Riet [192], Lucassen and
Hansen [193], Stenvot & Langevin [159]. New
generalisations of experimental peculiarities were
made, for example to cylindrical waves [164] or
with respect to the analysis of the wave propagation [194]. As the damping of capillary waves
is a useful method only for small values of the
dilational elasticity, a second wave method was
developed, the longitudinal wave method. The
theory of the damping effects of longitudinal waves
was developed by van den Tempel and van de Riet

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

[ 192], Lucassen [ 195,196], van den Tempel [ 197],


Lucassen and van den Tempel [185,198]. Both
theories allow a quantitative interpretation of
experimental data and provide information about
the dilational elasticity modulus and the amount
of the exchange of matter, erroneously named
dilational viscosity. More details on the wave
damping methods and a discussion of this fact is
given elsewhere (Ref. [ 12], Chapter 3).
The theory of pulsating bubbles in surfactant
solutions is also well developed. Wantke and
coworkers [173,199] derived a hydrodynamic
equation which describes the pressure balance in
the present system. Combined with a model which
describes the exchange of matter at the oscillating
bubble, such as the model given by Eq. (72), the
dilational elasticity and the exchange of matter
mechanism can be obtained. The theory has been
recently generalised by Johnson and Stebe [200].
According to their results, the determination of the
frequency dependence of the phase lag between
oscillation generation and pressure response
should allow us to differentiate between the
exchange of surfactant and the surface dilational
viscosity. There are no experiments available to
verify this hypothesis so far.

101

vation of interfacial response functions has recently


been discussed by Miller et al. [149]. Joos and
van Uffelen [208,209] and Loglio et al. [150,210]
have shown that exchange of matter functions
derived for harmonic disturbances can be applied
to transient ones.
As the result for a diffusion-controlled exchange
of matter, from Lucassen and van den Tempel's
theory [ 185], the following functions appear when
assuming a trapezoidal area change [208]:
f2E exp(2e)ot)erfc(2x/2~mot)
ATl(t) = 2e)--~
+ 212E
_ _ w/t
x/2zw)

DE
2o9o'
0 < t < tl

(81)

Ay2(t) = A71(t) - AT~(t - tl)


ta < t < t~ + t2
Ay3(t) = Ayz(t) - A y t ( t -

(82)

tx -t2)

tl + t2 < t < 2tl + t2

(83)

Ay4(t) = A73(t) - Ay~(t - 2tl - t2)


4.2.2. Techniques based on transient interfacial
disturbances

Both the capillary wave and the oscillating


bubble methods use harmonic disturbances of the
equilibrium adsorption layer to generate relaxation
processes. Their frequency intervals are very
different and sometime only reach a few oscillations
per minute [177,178]. Methods applicable to arbitrary area changes to induce relaxation processes
are the Langmuir trough technique [183], the
elastic ring [201,202], different surface dilational
methods [203-205], or the modified pendent drop
experiment [206,207]. By moving a barrier on the
trough, changing the shape of the elastic ring,
lifting the funnel or the strip, or increasing/decreasing the volume of a pendent drop, a variety of area
changes can be performed, such as jumps, square
waves, ramp type, trapezoidal and harmonic area
changes or continuous linear and non-linear expansions. The whole theoretical treatment of the deft-

t > 2q + t2

(84)

Here, the relative area change is denoted by f2 =


(d In A / d t ) = ( 1 / q ) In [ 1 - (AA/Ao)]. tl and t2 are
the characteristic times of the trapezoidal perturbation. O)o denotes the characteristic frequency. For
a diffusion-controlled exchange of matter COo is
defined by Eq. (73), as it was defined for a hatmonic area change. The system theory used to
derive the surface tension response functions is
explained for example by Miller et al. [149], who
also summarised response functions to other types
of area deformation functions.
The derivation of A y ( t ) for surfactant mixtures
and the consideration of specific peculiarities of
liquid/liquid interfaces were described by Miller
et al. [149]. For a surfactant mixture, consisting
of m compounds, the functions AT(t) for a square
pulse result in

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

102

AA ,..,~ ~ e x p ( 2 o g i o t ) e r f c ( ~ )
A71(t) = E A-0o i=1 ~
A72(t) = A71(t) - AVl(t - t2)

at 0 < t < t 2
at t > t2

(85)
(86)

of adsorbed polymer molecules. The special case


s = 1 is known as the Langmuir reaction mechanism. Based on the rate equations a model for an
adsorption layer relaxation can be derived by
considering the time-dependent area A(t). For the
case of Eq. (90) this modification leads to

with the total adsorption FT defined by


FT =

Fi

(87)

A dt - kadc

1 --

-- kdes

i=1

and the characteristic frequencies, defined for each


of the components, by
( d c i ~ z Di

(88)

After some rearrangement and using the solution


to the problem in the general form [185,198]
F = Fo + AFexp(iogt)

(92)

the exchange of matter equation reads,


The surface tension response function A?(t) of
relaxations at the liquid/liquid interface has exactly
the same form as the one at the liquid/air interface,
except that the characteristic frequency is defined
differently, to take into account the peculiarity of
solubility of the surfactant in both adjacent phases.
As mentioned above, besides the diffusion-controlled model other models exist to describe the
adsorption kinetics and exchange of matter. De
Feijter et al. [211] have developed a relation taking
into consideration simultaneous adsorption of proteins and surfactants at an interface. As a special
case a relation results which describes the equilibrium state of adsorption of polymer molecules at
a liquid interface
F
(

i~
Kx + ko

(93)

with
Kx = skad c o (1 --~-!F"~ts-1) + kdes
F~ \
1"oo/
Foo

(89)

1 -

with adF the concentration characterising the surface activity of the polymer, and s the number of
adsorption sites per adsorbing molecule.
The following kinetic relation
1 --

- kde, ~

(94)

By applying the inverse Fourier transform, according to the calculus discussed in Miller et al. [ 149],
the interracial tension response can be calculated.
Under the assumption of a step-type disturbance
the response function reads
AA
AT(t) = E Too exp(-Kxt)

~___~) --c/adv

- k,dCo

E(ko) --- E

(95)

The surface dilational modulus E defined by


Eq. (70), is related to the surface coverage by the
equation
[
E = R Tf'~

Fio 1
Fo
F
( 1 - s) ~ + s -1 -

(96)

(90)

turns into the isotherm Eq. (89) in the equilibrium


state. Here k,d and kd~ are the rate constants of
adsorption and desorption, respectively. This equation can be used to describe the relaxation process

Loglio et al. [203] have demonstrated that the


most useful disturbance for interracial relaxation
experiments is the trapezoidal area change. For
the time regimes of most of the transient relaxation
experiments the trapezoidal area change can be

R. Miller et aL/Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

approximated adequately by a square pulse. For


the square pulse area change we obtain
AA
ATe(t) = E T o exp(- Kx t)

at 0 < t <~tz
(97)

Ay2(t) =

zJ71(t)

A y l ( t - tl)

at

> t2

(98)

where t2 is the duration of the pulse and AA/Ao is


the relative area change,
Eqs. (95), (97) and (98) can be used to interpret
experimental data by use of a fitting procedure.
The resulting values for E and K~ are obtained
independently of any knowledge of the parameters
of the equilibrium adsorption isotherm. On the
other hand both parameters contain specific constants of the adsorption isotherm and are therefore
of interest from this point of view too.
A very simple relation for the description of
relaxation processes has been given by Graham
and Phillips [,,212-216]
dF
/
H A A ~|
dt = kadco exp

RT J
IIz]A\

--kdesF(t ) exp

/|

(99)

This relation can be used instead of the above


derived response functions.

4.3. Interfacial relaxation methods

There are many experimental techniques for


studying interfacial relaxations of soluble adsorption layers. Except for wave damping techniques,
these methods have been developed and only used
by single research groups. Up to now, no commercial set-up exists and relaxation experiments are
not therefore very common. New developments in
this field, especially the transient methods, are
simple to construct and data are easy to handle,
so will probably increase the number of investigators studying the dynamic and mechanical properties of adsorption layers. In this section, wave
damping and other harmonic methods as well as
transient relaxation techniques will be described.

103

4.3.1. Relaxations to harmonic interfacial


disturbances

The oldest relaxation techniques, developed to


measure dilational elasticities and exchange of
matter properties, are the methods of wave damping. If a wave is generated by mechanical or other
means, it propagates along the surface and is
damped by hydrodynamic and surface mechanical
properties [-217].
The construction of experimental set-ups differs
in the way of wave generation and wave propagation and damping detection. Yasunaga and Sasaki
[218] for example produce the capillary waves by
a vibrator attached to a drive unit of a loudspeaker.
The wave damping is determined via a microscope
and stroboscope. The set-up works in a frequency
range from 25 Hz to 4 kHz. Other designs use, for
example, the condenser principle to record the
wave damping and phase lag [219-223].
A new type of capillary wave, cylindrical waves,
is used by Jiang et al. [-164]. A sharp needle as a
point source is used to excite cylindrical waves
through electrocapillarity. The wavelength and the
damping coefficient are detected from specular
reflection of a laser beam. Unlike plane waves,
cylindrical waves propagate isotropically in the
radial direction and no shear deformation caused
by edge effects is to be expected. The schematic
diagram of an experimental set-up designed by
Jiang et al. [, 164] is shown in Fig. 15. The motor
allows the surface wave profiler to move along the
surface to scan the distance to the wave generation
origin. Both wave generation and detection avoid
direct contact with the interface. The technique is
applicable in a frequency range from 50 to 1000 Hz.
To perform measurements of longitudinal wave
propagation and damping characteristics, a device
designed by Lemaire and Langevin [ 194] can be
used. While the longitudinal wave is generated by
a piezo element, its propagation is measured by
analysing the propagation of high-frequency capillary waves, generated by electrocapillarity and
analysed via the specular reflection of a laser beam
from the interface [ 194,218]. The apparatus, using
two types of waves, capillary wave damping to
analyse the propagation of longitudinal waves, is
shown schematically in Fig. 16.
The set-ups designed by Wielebinski and

R. Miller et al./Colloids Surfaces A. Physicochem. Eng. Aspects 111 (1996) 75-118

104

11
SURFACEWAVEPROFILER I
1". . . .
"1
I M2 . ~ ,

_,

LASER
He-N

~;;~ BEAMSPLITTER , . _ ~

PREAMPLIFIER ]

I
Fig. 15. Schematic diagram of the apparatus for the generation and detection of propagating cylindrical waves; according to Jiang
et al. [164].

Capillarywnws
excitationblade

T
Lon~tu,41.~! waves
excitation

Lon~tt)dlnRI waves
prop~don
CapiUaty waves
propagation

Fig. 16. Schematic diagram of the apparatus for the generation and detection of longitudinal waves; according to Lemaire &
Langevin [194].

Findenegg [219] and Lemaire and Langevin [ 194]


are also applicable to liquid/liquid interfaces without significant modification,
Wielebinski and Findenegg 1-219] used a light
scattering method to measure the characteristic of
very small amplitude, thermally induced capillary
waves. These ever-present waves are of very high
frequency and usually exchange of matter can be

ignored. As the accuracy of these measurements is


very high, a light scattering technique is an attractive method of studying dynamic interfacial
properties.
Besides the capillary wave techniques, the pulsaring bubble method was also used in early experiments for measuring the surface dilational elasticity
[172,173,199]. For soluble adsorption layers it

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

allows the determination of the exchange of matter


at a harmonically deformed bubble surface,
The principle of the oscillating bubble experiment as modified by Fruhner and coworkers
[224,225] is shown in Fig. 17. A small air bubble
is formed at the tip of a capillary which is immersed
in the solution. A piezoelectric excitation system
immersed in the solution volume generates directly
harmonic oscillations of the bubble surface. A
pressure sensor also mounted on the system reads
the pressure changes due to bubble area, and
therefore adsorption layer dilations and expansions, in dependence on frequency, while keeping
the bubble oscillation amplitude constant. From
the harmonic surface tension response the dilational elasticity and the exchange of matter can be
calculated. The comparatively complex theory for
data interpretation has recently been described by
Wantke and coworkers [ 173,199]. The method
can be applied in a frequency interval from few
hertz up to several hundred hertz,
Modified experimental set-ups for pulsating
bubble studies have been described by other
authors as well, for example by Johnson and Stebe
[200] and Chang and Franses [178,179]. The
main difference between the new instruments and
the apparatus described earlier by Lunkenheimer
and Kretzschmar [ 172] is the direct measurement
of the pressure variation in the bubble by pressure
transducers,
Studies of dilational rheology of monolayers and

II

j-

Fig. 17. Oscillating bubble set-up to measure the dilational


elasticity and the exchange of matter of a surfactant solution
as a function of frequency; 1, capillary;2, brass block with lid;
3, solution; 4, piezoelectric driver; 5, low pressure quartz
transducer; 6, amplifier and measuring instrument for signal
amplitude and phase lag; 7, frequencygenerator,

105

adsorption layers can also be performed by using


an oscillating barrier method as described for
example by Lucassen and Barnes [226], Lucassen
and Giles [182], and Kretzschmar [227]. This
method permits a direct measurement of the surface pressure oscillation and the phase angle
between generation of the area oscillation and the
resulting pressure oscillations. Due to hydrodynamic limitations of the method the frequency
range is limited to less than one hertz.
4.3.2. Relaxations to transient interfacial
disturbances

An experimental set-up for the study of surface


relaxation processes in soluble adsorption layers
after transient surface area disturbances was developed by Loglio et al. [150,201-203]. The main
feature of the apparatus is the elastic circular ring,
which confines the surface area in the sample vessel
and replaces the traditional barrier and trough. By
changing the shape of the ring, which is immersed
in the solution, small area changes of the solution surface can be applied. The surface tension
response after such deformations is registered via
force measurements using a Wilhelmy plate. The
software-driven apparatus allows different types of
area changes: step and ramp type, square pulse
and trapezoidal as well as sinusoidal area deformations. The construction ensures that area changes
are almost isotropic. Area changes used in transient
and harmonic relaxation experiments are of the
order of 1-5%. The surface tension response measured via the Wilhelmy balance has an accuracy
of better than +0.1 mN m -1.
A recently developed modification of the pendent
drop method allows definite area changes of the
drop surface, which can be used to initiate transient
relaxation processes [206,207]. A metering system
consisting of two syringes (see Fig. 18) is used to
form a drop with a definite volume (syringe 1) and
then to change this volume by small increments
by using the highly precise syringe 2.
In this way transient relaxation experiments with
any area disturbance are possible. The high resolution and excellent accuracy of the interfacial
tension determination [228,229], better than
+0.1 mN m -~, provides a useful tool for studies
of dynamics of soluble adsorption layers. The

106

IE Miller et al./Colloids Surfaces A." Physicochem. Eng. Aspects 111 (1996) 75-118
motor

syringe 2

syringe 1
monitor

light
source

drop

microscope

Fig. 18. Video-enhanced pendent drop method modified for relaxation experiments according to Miller et al. [206].

method can be applied to liquid/gas as well as


liquid/liquid interfaces and is easily temperature
controlled. Experimental data obtained by this
technique on aqueous solutions of surfactants and
proteins are given below [206,207].
The drop pressure experiment is a modification
of the pendent drop technique, which also allows
relaxation studies on different area deformations.
The directly coupled pressure transducer can register the change of the pressure inside the drop (or
bubble) and therefore, the interfacial tension with
time. The time resolution is very high and fast
computers are able to acquire data in time scales
less than 1 ~ts. Liggieri et al. [230] and Passarone
et al. [231] and later Horozov et al. [180],
MacLeod and Radke [174,175] and Nagarajan
and Wasan [181] developed different drop pressure experiments. Wasan et al. [177] demonstrated
different types of surface area changes of surfactant
solution drops and the respective surface tension
responses. A typical set-up of a drop pressure
experiment is shown in Fig. 19, designed for studies
at a liquid/liquid interface,
The high accuracy of modern pressure transducer and metering systems and the high speed of
simple desk computers will make this type of

1-.._~
"
2---__..

6
~

Fig. 19. Principle of a drop pressure experiment; 1, drop;

2, capillary; 3, second liquid; 4, solution under study;


5, differentialpressure sensor;6, syringe;accordingto Passarone
etal.[231].
experiment very successful. Its design is easy and
the application of the method is very general.
A very recently developed method for transient
and low-frequency harmonic relaxation experiments has been published by Kokelaar et al. [232],
which appears as a modification of the elastic ring
principle of Loglio et al. [ 150,201-203 ]. A cylindrical ring is placed vertically in the liquid surface
(see Fig. 20). The area inside the ring can be
changed by moving it up and down. The resulting

R. Miller et aL/Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

1 II
=
_

~
2

Fig. 20. Perspective view of the cylindrical ring method;


1, Wilhelmy plate; 2, appliance to move the ring up and down;
3, cylindrical ring; 4, vessel; according to Kokelaar et al. [232].

changes in surface tension can be measured by


using a Wilhelmy plate, kept in permanent contact
with the surface in the centre of the ring. The
amount of relative surface area increase or decrease
is in the range of up to 15%, which is sufficiently
large. Usual relaxation experiments need only very
small disturbances to ensure the validity of the
elaborated linearised theories,
Another experimental set-up also applicable for
relaxation studies is the overflowing cylinder discussed by Bergink-Martens et al. [233] and Prins
[234]. Similar set-ups were used for the measurement of dilational properties by de Rijcker and
Defay [235] and Joos and de Kayser [236], or
for desorption kinetics studies by Kretzschmar and
Vollhardt [237].

4.4. Experimental results of relaxation studies


The present section is dedicated to experimental
results obtained by different techniques. So far
there are only a few experiments performed to
compare results of different methods for one and
the same surfactant solution, as Dukhin et al. 1-12]
also point out. Each set of data has to be seen in
close relation to the specific experiment,
This section aims to demonstrate the variety of
relaxation studies performed with surfactant solutions. Results for single surfactants and surfactant
mixtures as well as micellar solutions will be shown.
Newly available techniques for liquid/liquid interfaces are also discussed here. The presentation is

107

far from being complete and should be seen as just


a selection.
As mentioned above, only a few attempts have
been made to compare different techniques. For
example, Jiang et al. [164] studied the dilational
elasticity of Triton X-100 adsorption layers with
plane and cylindrical capillary waves. The results
of the two measurements are given in Fig. 21.
Although the experimental conditions and the theories differ from each other, the agreement between
the results is excellent.
Lucassen and Hansen [193] were one of the
first to investigate the damping of capillary waves
of surfactant solutions. In Fig. 22 their results for
n-octanoic acid and dodecylamine hydrochloride,
both in 0.005 N HC1 (nonionic surfactants), are
shown in the form of effective dilational elasticity
as a function of concentration. These experimental
data are used as standard and many new experiments are compared with the results given forty
years ago.
The curves have a typical shape with a maximum
at a certain concentration which corresponds to a
surface concentration of about 50% of the maximum surface coverage. The maximum is caused by
the competitive effect of two phenomena: increase
of the dilational elasticity modulus E with concentration and increase of the exchange of matter with
increasing concentration, diminishing the effective
elasticity.
The first studies of the effect of micelles on the
exchange of matter to harmonical disturbances
of the surface area of a surfactant solution were
performed by Lucassen [ 186]. He used an aqueous
solution of hexadecyldimethylammonium propanesulphonate (HDPS) below and above the
CMC. The exchange of matter, shown by the
effective dilational elasticity c, is greatly affected
by the presence of micelles, as was discussed above.
The lower the frequency of disturbance, the more
pronounced is the influence of micelle kinetics on
the exchange of matter, see Fig. 23 where the line
marks the CMC of HDPS. Lucassen was able to
describe the behaviour by using the theory given
by Eqs. (79) and (80).
The surface relaxation behaviour of mixed surfactant solutions was studied by Garrett [238]
using the longitudinal wave technique. By keeping

R. Miller et al./ Colloids Surfaces A." Physicochem. Eng. Aspects 111 (1996) 75-118

108

1,5

z
E
,,,
o
--

[]
C3

0,5
-10

-9

-8

-7

-6

log c [mol/cm 3]

Fig. 21. Comparison of surface dilation elasticities determined from two different wave damping experiments with Triton X-100
solutions at 150 Hz; plane waves, []; cylindrical waves, m; according to Jiang et al. [164].
45

4O
35

O
[]

3O

~ 25
z

20

,,,

15

[]

10
5
0

-7,5

r-

-7

---~

-6,5

--

-6

-5,5

-5

-4,5

log c [mol/cm 3]
Fig. 22. Effective dilational elasticity of n-octanoic acid, i , and dodecylamine hydrochloride, D, at 200 Hz, measured with a
capillary wave technique; according to Lucassen and Hansen [193].

the total surfactant concentration constant at


10 -s tool cm -3 different compositions of a mixture
ofnonionic dodecyl tri(ethyleneglycol)(C12E3)and
dodecyl hexa(ethyleneglycol) (C12E6) were investigated. Although the surface activity of the two
components do not deviate very much from each
other, the frequency dependence of the effective
dilational elasticity changes remarkably with the

composition. The theory for the exchange of matter


for mixed surfactant solutions (see Eqs. (77) and
(78)) agrees well with his experimental data.
Dilational experiments with insoluble surfactant
monolayers can be performed using the capillary
wave damping technique [134,239]. When the
frequency applied is very high only the dilational
elasticity of the monolayer can be measured. For

109

R. Miller et aL/Colloids Surfaces A: Physicoehem. Eng. Aspects 111 (1996) 75-118


18

16

14

IICMC

12

z
E

* ?o

.1.

YD
ip

08

0,6

O~
O

--

O
,

[3

0,4

[3

0,2
0

I
9

I
7

I
6

log c [mol/cm "~]

Fig. 23. Concentration dependenceof the effectivedilational elasticityof hexadecyldimethylammoniumpropanesulphonatesolutions


at differentfrequencies;f = 10, A; 5, O; 2, ~1,;1, [3; 0.5, c.p.m.,vertical line, CMC; according to Lucassen [186].
measuring the dilational rheological parameters
of a monolayer an oscillating barrier method can
be used [ 182,226,227]. As the oscillating barrier
method permits a direct measurement of the surface pressure oscillation, the dilational elasticity
and the phase angle between the generated area
oscillation and the resulting surface pressure oscillations can be easily determined,
The dilational elasticity and the phase angle,
expressed as dilational viscosity, of D P P C is given
in Fig. 24. The measurements are performed under
the same conditions as the shear rheological studies
reported in section 3.3.1. which allows a direct
comparison of the results,
The discussion of the rheology was done in
close relation to the monolayer morphology
[137,138,240-243]. The dilational elasticity and
viscosity of both monolayers show a strong dependence on surface pressure with a minimum in the
coexistence region. This behaviour was explained
as follows. As long as two phases coexist a monolayer compression does not result in a significant
pressure increase but in a transfer of the liquid
expanded (LE) into the LC phase. During expansion of the monolayer the opposite process hinders
a surface pressure change. Thus almost no phase
shift and no elasticity are observed. Only below
and above the coexistence region can surface pres-

sure changes be generated by area changes so that


a measurable dilational elasticity and an increasing
dilational viscosity are observed. Both parameters
increase and level off at a surface pressure above
the end of the coexistence region. From the present
experiments it is not possible yet to specify the
mechanism of the relaxation process in the monolayer which is responsible for the observed dilational viscosity (phase angle). Probably both a
lateral relaxation along the monolayer and an
orientational relaxation happen simultaneously.
Further experiments with different distances
between the oscillating barrier and the pressure
balance and variation in tile oscillation frequency
are necessary to enable us to distinguish between
these two, and possibly still other, relaxation
processes 1-241-244].
Besides rheological studies of surfactant adsorption layers increasing interest can be observed in
experiments with protein adsorption layers. Serrien
et al. [245] measured the damping of planar
longitudinal waves generated with a barrier on a
Langmuir trough and registration with a Wilhelmy
balance. The longitudinal waves correspond to
slow periodic compressions/dilations of the protein
adsorption layer (BSA, casein). Further experiments using a stress relaxation technique were
performed with BSA and buttermilk. The results

110

R Miller et al./Colloids Surfaces A: Physicoehem. Eng. Aspects 111 (1996) 75-118


140

600

120

5.00
100
-~.
Z

/
80,

;'~

"

.~

60 -

300

o
"~

'~
"~

400

~__~o~/

200

"-"

40.

20.

100

0
0

lJ0

I'2

14

pressure (mN/m)
Fig. 24. Dilational elasticity, O, and viscosity, D, of a D P P C rnonolayer as a function of surface pressure, T = 20.5C, L = 7.5 cm,
m = 0.628 s - 1.

show that the characteristic parameters of the


models used depend on the age of the adsorption
layer. The discussion of the results comprises explanations with respect to the established structure,
to the distance from the equilibrium state, and the
amount of denatured molecules.
The dynamics of mixtures of surfactants with
proteins is of great importance for many practical
processes, such as coating of photographic films,
where gelatine in mixtures with surfactants and
surface active dyes adsorb at the interface. Hempt
et al. [246] studied the relaxation behaviour of
gelatine solutions in both the presence and absence
of surfactants (SDS, tetradecyldimethylphosphine
oxide, cetyltrimethylammonium bromide, n-decanoic acid, perfluoro-octanoic acid tetraethylammonium salt). Adsorbed gelatine molecules alone
do not show a frequency dependence of surface
elasticity, which refers to the behaviour of an
insoluble monolayer. The presence of surfactants
changes the elastic and relaxation behaviours dramatically. With increasing SDS concentration the
elasticity modulus (frequency independent plateau
value of the elasticity) first increases and then
decreases. The dynamic behaviour of the mixed

adsorption layer changes from one completely


formed by gelatine molecules to an adsorption
layer completely controlled by surfactant molecules
(Fig. 25). Similar behaviour can be observed for
CTAB and a perfluorinated surfactant [246].
Transient relaxation experiments of protein
adsorption layers were published by Miller and
coworkers [206,207]. The experiments were performed using a modified pendent drop technique
as described above. The surface tension response
to three subsequent square pulse perturbations
of 0.1mgm1-1 HA adsorbed at the aqueous
solution/air interface [206] are shown in Fig. 26.
If the results are discussed on the basis of a
diffusion-controlled exchange of matter model, the
resulting diffusion coefficients were found to be
two to three orders of magnitude higher than those
expected from the Stokes-Einstein relation. The
same experimental data have been interpreted on
the basis of Eqs. (97) and (98). Using a leastsquare fitting algorithm, values of E and Kx are
obtained. The quality of agreement between experimental data and the fitting curve is also demonstrated in Fig. 26.
The calculated values of E and Kx for the

R. Miller et al./Colloids Surfaces A." Physicochem. Eng. Aspects 111 (1996) 75-118

111

120

I O0

O
0

80

<>
0

&

60

[]

c
O

40

o []

20

. . . .

GD

20

40

60

[]

80

1 O0

120

140

160

f [Hz]

Fig. 25. Effective dilational elasticity determined from oscillating bubble experiments of gelatine/SDS mixtures at 0.5 wt% gelatine;
SDS concentration co = 2 x 10 -7, A, 5 x 10 -7, ~ ; 1 x 10 -6, R; 1.5 x 10 6, 0 ; 3 x 10 -6, mol cm -3, according to Hempt et al. [246].

1,5

o,

1,

0,5[

-2,5 ~1t
-3 ~
Time [s]

~
,

-4
(a)

Time [s]

On
c

-2

-2,5

500

1000

-3 ~
(b)

Time [s]

(c)

Fig. 26. Surface tension relaxation to a surface area square pulse of an aqueous 0.1 mg m1-1 HA solution with the modified pendent
drop method at the water/air interface; symbols, experimental data; solid lines, theory Eqs. (97) and (98); AA/Ao = (a) 0.041; (b) 0.084;
(c) 0.091; according to Miller et al. [206].

individual runs differ from each other more than


expected from the experimental accuracy. This
inconsistency is probably caused by different experimental conditions. Although the relative area

adsorption isotherm leading to non-linearity effects


even at very small deviations from equilibrium.
These effects are not considered in the present
theory.

c h a n g e A A / A o s h o u l d n o t affect the r e l a x a t i o n
b e h a v i o u r of a n a d s o r p t i o n layer, a n a d s o r b e d H A
layer seems to be very sensitive to it. T h i s p h e n o m e n o n c a n be e x p l a i n e d b y the v e r y high slope of the

I n t e r f a c i a l r e l a x a t i o n s of H A at the w a t e r / d e c a n e
interface were p e r f o r m e d w i t h the s a m e t e c h n i q u e
[ 2 4 7 ] . T h e i n t e r p r e t a t i o n of the e x p e r i m e n t s b y
fitting different m o d e l s to the d a t a has s h o w n t h a t

112

1~ Miller et al. /Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75 118

the kinetic model yields values of the parameters


which are rather consistent while those obtained
from the diffusion model scatter more. In addition,
the resulting diffusion coefficients are found to be
several orders of magnitude too high, hence diffusion is not the controlling mechanism of the albumin relaxation process studied,

ideas about dilational properties of composite


adsorption layers were published by Lucassen
[252]. These ideas can be helpful for understanding the rheological behaviour of monolayers of
insoluble surfactants. A theory has also been developed by Johnson and Stebe 1-200] to differentiate
exchange of matter from dilational viscosity. This
calls for new experimental developments.

5. Summary and outlook


Acknowledgements
This paper describes theoretical and experimental aspects of interfacial shear and dilational rheology. Besides the description of experimental details
and the theoretical basis of some methods, several
examples are discussed to demonstrate the capability of various experiments.
Interfacial shear experiments are useful for learning about interactions of molecules in adsorption
layers or monolayers. Especially in mixed layers
the interactions change dramatically with the composition. For example, mixed polymer-polymer or
polymer-surfactant systems form very complex
adsorption layers and the time dependence of the
interfacial shear properties can provide an understanding of structure formation and molecular
exchange processes.
Transient and harmonic relaxation techniques
are used to investigate the dilational rheology of
liquid interfaces. New developments have been
made, such as the elastic ring, the modified pendent
drop or different drop-bubble pressure experiments. New instrfiments based on oscillating bubbles and drops have been designed as well to study
interfacial properties at medium frequencies. Stress
relaxation experiments provide simultaneously
information about exchange of matter and dilational rheological parameters of freshly formed
adsorption layers.
So far there is lack of sufficiently systematic
experimental studies. Only on the basis of a large
data base can further improvements in the theories
be made. Thus, the more effective instruments and
newly developed techniques should enable surface
scientists to produce more systematic data. There
is also a deficiency in respect to relaxation theories
for polymers and mixed polymer-surfactant systems [12] and micellar solutions [248-251]. First

This work was financially supported by the


DFG (Mi 418/5-1, Wu 187/3-2, WU 187/5-1) and
the Fonds der Chemischen Industrie (RM 400429).

Appendix A. List of symbols


a
aL
adF
A
Bo

c
Co
D
Def
t

Dn
E
E=E'+E"

Er
E'
E"
f
AF
Fr
g
G
Gs
h
h

width of a canal
constant of the Langmuir isotherm
(mol cm -2)
constant of the de Feijter isotherm
area of the interface (cm2)
Boussinesq number
bulk concentration (molcm -3)
equilibrium
bulk
concentration
(mol cm -3)
diffusion coefficient (cm2 s -a)
effective
diffusion
coefficient
(cmz s - 1)
diffusion coefficient of a micelle
(cm2 s -1)
thermodynamic surface dilational
modulus (mN m- 1)
complex surface elasticity
elasticity of the torsion wire
real part of the complex modulus
imaginary part of the complex
modulus
frequency of oscillations (Hz)
free energy of activation
friction of the clean solvent interface
acceleration constant (cm s-2)
interfacial Gibbs free energy
interfacial shear modulus of rigidity
depth of liquid
Planck constant

R. Miller et al./Colloids Surfaces A: Physicochern. Eng. Aspects I l l (1996) 75-118

H
H,
AH
Ir
k
kad
kde s

K
K.
Kx
L
m
n
P
Q
r
R
Rr
s
S
t
tad
T
ux(x, y, t)
uy(x, y, t)
ff
ff
t~a
U
V
x
y

enthalpy
geometric constant
amplitude of oscillations (cm)
moment of inertia of the measuring
system
Boltzmann's
constant
(1.38 x 10 -23 J mo1-1 K -1)
rate constant of adsorption (cm s-1)
rate constant of desorption (s- 1)
distribution coefficient
number of moving kinetic units
rate of relaxation (s- 1)
length of a canal (cm)
number of components
micellar aggregation number
pressure (N m -2)
film flow rate
radius of a molecule (nm)
gas law constant (8.314gcm 2
(s2 mol K - l ) )
ring radius (cm)
number of adsorption sites per
adsorbed molecule
entropy
time (s)
characteristic adsorption time (s)
absolute temperature (K)
component of ff in x direction
component of ~ in y direction
tensor of deflection
vector of deflection
vector of deflection velocity
anti-symmetric tensor of deflection
energy of activation
volume (cm 3)
direction normal to the interface (cm)
direction tangential to the interface
(cm)

Greek letters

fl
//8
~i,j
~o

amplitude of an oscillation
damping coefficient
damping coefficient of shearing
interfacial tension (mN m - 1)
components of the stress tensor
interfacial tension of the pure
solvents (mN m - 1)

~
Fd
F~
FT
F0
F~

F = F(t)/Fo
6
~
~/
r/d
~/8
O(t)
~:
2
#
Pi
v
~
~
H= ~o-~
p
zi
zd
Zo
~0
~

~i
m
~oo
g2
2p
2~

113

stress tensor
adsorption at time t = 0 (mol cm -2)
adsorption
of
component
i
(mol cm -2)
total
surface
concentration
(mol cm- 2)
equilibrium surface concentration
(tool cm -2)
maximum surface concentration
(tool cm -2)
dimensionless adsorption
average amplitude of thermal
oscillation
two-dimensional unit tensor
bulk viscosity
surface
dilational
viscosity
(mN s m - ~)
surface shear viscosity (mN s m - l )
stimulation of oscillation
complex wave number
wavelength (cm)
chemical potential
chemical potential of component i
wave frequency
wave number
displacement of a surface element
surface pressure
density (g cm -3)
relaxation time of dilation/compression deformation
relaxation time of shear deformation
period of oscillation
loss angle
strain tensor
the trace of the tensor
angular frequency (rad- 1)
relaxation frequency (rad -~)
area disturbance (d In A/dO
angular rotation of the particle
angular rotation of the ring

Appendix B. Abbreviations
CMC
CTAB
FRAP

critical concentration of micellisation


cetyltrimethylammonium bromide
fluorescence recovery after photobleaching

114

HA
SDS

R. Miller et al./Colloids Surfaces A: Physicochem. Eng. Aspects I l l (1996) 75-118

human serum albumin


sodium dodecylsulphate

References
[1] G. Kretzschmar, J. Inf. Rec. Mater., 21 (1994) 335.
[2] H. Fruhner, K.-D. Wantke, J. Kr~igel and
G. Kretzschmar, J. Inf. Rec. Mats., 22 (1994) 29.
[3] R.B. Dorshow and R.L. Swofford, J. Appl. Phys., 65
(1989) 3756; Colloids Surfaces, 43 (1990) 133.
[4] A.R. Kovscek, H. Wong and C.J. Radke, AIChE J., 39
(1993) 1072.
[5] E. Dickinson, J.A. Hunt and D.G. Dalgleish, Food
Hydrocolloids, 4 (1991)403.
[6] E. Dickinson and J.L. Gelin, Colloids Surfaces, 63
(1992) 329.
[7] E. Dickinson, in C. Gallegos (Ed.) Interactions in
Protein-Stabilized Emulsions. Progress and Trends in
Rheology IV, Proceedings of the Fourth European
Rheology Conference, Seville, 1194, p. 227.
[8] D.C. Clark, P.J. Wilde, D.R. Wilson and R. Wtistneck,
Food Hydrocolloids, 6 (1992) 173.
[9] D.C. Clark, A.R. Mackie, P.J. Wilde and D.R. Wilson,
in K.D. Schwenke and R. Mothes (Eds.), Food
Proteins - - Structure and Functionality, VCH,
Weinheim, 1993, p. 263.
[10] D.C. Clark, P.J. Wilde, D.J.M. Bergink-Martens,
A.J.J. Kokelaar and A. Prins, in E. Dickinson and
P. Walstra (Eds.), Food Colloids and Polymers: Stability
and Mechanical Properties, 1194, p. 354.
[11] P.J. Wilde and D.C. Clark, J. Colloid Interface Sci., 155
(1993) 48.
[12] S.S. Dukhin, G. Kretzschmar and R. Miller, Dynamics
of adsorption at liquid interfaces, in D. M6bius and
R. Miller (Eds.), Studies of Interface Science, Vol. 1,
Elsevier, Amsterdam, 1995.
[13] G. Loglio, U. Tesei, G. Mori, R. Cini and F. Pantani,
Nuovo Cimento, 8 (1985) 704.
[14] G. Loglio, U. Tesei and R. Cini, Boll. Oceanologia
Teorica Applicata, 3 (1987) 195.
[15] G. Loglio, N. Degli-Innocenti, U. Tesei, A.M. Stortini
and R. Cini, Ann. Chim. (Rome), 79 (1989) 571.
[16] G. Loglio, N. Degli-Innocenti, U. Tesei, R. Cini and
Wang Qi-Shan, Nuovo Cimento, 12 (1989) 289.
[17] M. van den Tempel and E.H. Lucassen-Reynders, Adv.
Colloid Interface Sci., 18 (1983) 281.
[18] P.R. Garrett and P. Joos, J. Chem. Soc. Faraday Trans.
1, 69 (1976) 2161.
[19] E.H. Lucassen-Reynders and K.A. Kuijpers, Colloids
Surfaces, 65 (1992)175.
[20] D.T. Wasan, A.D. Nikolov, L.A. Lobo, K. Koczo and
D.A. Edwards, Prog. Surface Sci., 39 (1992) 119.
[21] D.A. Edwards, H. Brenner and D.T. Wasan, Interfacial
Transport Processes and Rheology, ButterworthHeineman Publishers, 1991.

[22] K. Malysa, Adv. Colloid Interface Sci., 40 (1992) 37.


[23] P.R. Garrett and P.R. Moore, J. Colloid Interface Sci.,
159 (1993) 214.
[24] K. Lunkenheimer and K.-D. Wantke, Colloid Polym.
Sci., 259 (1981) 354.
[25] J. Boussinesq, Ann. Chim. Phys. (Ser.), 29 (1913) 349.
[26] J. Boussinesq, C.R. Acad. Sci., 156 (1913) 1124.
[27] V.V. Krotov, in Theory of Interfacial Phenomena,
University of Leningrad, 1979, p. 146 (in Russian).
[28] M. Joly, Surface Colloid Sci., 5 (1964) 1.
[29] B. Stuke, Fortschr. Kolloide Polym., 55 (1971) 106.
E30] J. Frenkel, Trans. Faraday Soc., 33 (1937) 58.
[31] R.H. Ewell and H. Eyring, J. Chem. Phys., 5 (1937) 726.
[32] W.J. Moorew and H. Eyring, J. Chem. Phys., 6
(1938) 391.
[33] E. Boyd and W.D. Harkins, J. Am. Chem. Soc., 61
(1939) 1188.
[34] L Prandtl, Z. Angew. Math. Mech., 8 (1928) 85.
[35] Andrade, Philos. Mag., 17 (1934)705.
[36] S. Glasstone, K.J. Laidler and H. Eyring, The Theory
of Rate Processes, McGraw-Hill, New York, 1941.
[37] R. Merigoux, C.R. Acad. Sci., 202 (1936) 2049; 203
(1936) 848; 205 (1937) 115.
[38] J.H. Schulman and T. Teorell, Trans. Faraday Soc., 34
(1938) 1337.
[39] M. Joly, J. Phys. Radium, 8 (1937) 471.
[40] M. Joly, J. Phys. Radium, 9 (1938) 345.
[41] M. Joly, Kolloidzschr., 89 (1939) 26.
[42] P.A. Rehbinder, Kolloid-Z., 8 (1946) 157.
[43] G.M. Bartenev, UspechiKolloidnoj ChimiilV. Reologija
i Fiziko-Chimiceskoje Mechanika, Nauka, Moscow,
1973, p. 174 (in Russian).
[44] Z.A. Rogovin, Uspechi Khimii i Fizika Polimerov,
Moscow, Khimia, 1970.
[45] G.M. Bartenev and Yu.V. Zelenev, Physik der Polymere,
Verlag der Grundstoffindustrie, Leipzig, 1978.
[46] R.J. Myers and W.D. Harkins, J. Chem. Phys., 5
(1937) 601.
[47] D.G. Dervichian and M. Joly, J. Phys. Radium, 10
(1939) 375.
[48] W.D. Harkins and J.G. Kirkwood, J. Chem. Phys., 6
(1938) 53.
[49] H. H0hnerfuss, J. Colloid Interface Sci., 107 (1985) 84.
[50] J.T. Davies, Proceedings 2nd International Congress on
Surface Activity, 1 (1957) 220.
[51] R.J. Mannheimer and R.S. Schechter, J. Colloid Interface
Sci., 32 (1970) 195.
[52] R.J. Mannheimer and R.S. Schechter, J. Colloid Interface
Sci., 32 (1970) 212.
[53] R.J. Mannheimer and R.S. Schechter, J. Colloid Interface
Sci., 32 (1970)225.
[54] A.J. Pintar, A.B. Israel and D.T. Wasan, J. Colloid
Interface Sci., 37 (1971) 52.
[55] D.T. Wasan, L. Gupta and M.K. Vora, AIChE J., 17
(1971) 1287.
[56] L. Gupta and D.T. Wasan, Ind. Eng. Chem. Fundam.,
13 (1974) 26.

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118
[57] V. Mohan, L. Gupta and D.T. Wasan, J. Colloid
Interface Sci., 57 (1976) 496.
[58] A.R. Deemer, J.D. Chen, M.G. Hedge and J.C. Slattery,
J. Colloid Interface Sci., 78 (1980) 87.
[59] A.K. Chattopadhyay, L. Ghaicha, S.G. Oh and
S.G. Shah, J. Phys. Chem., 96 (1992) 6509.
[60] I. Blute, M. Jansson, S.G. Oh and S.G. Shah, 36.
Hauptversammlung der Deutschen Kolloidgesellschaft,
J~ilich, 1993.
[61] F.C. Goodrich, L.H. Allen and A. Poskanzer, J. Colloid
Interface Sci., 52 (1975) 201.
[62] A. Poskanzer and F.C. Goodrich, J. Colloid Interface
Sci., 52 (1975) 213.
[63] A. Poskanzer and F.C. Goodrich, J. Phys. Chem., 79
(1975) 2122.
[64] R.D. Krieg, J.E. Son and R.W. Flumerfelt, J. Colloid
Interface Sci., 79 (1981)14.
[65] O. Hassager and H. Westborg, J. Colloid Interface Sci.,
119 (1987) 524.
[66] J. Prieditis, N.R. Amundson and R.W. Flumerfelt,
J. Colloid Interface Sci., 119 (1987) 303.
[67] I. Langmuir, Science, 84 (1936) 378.
[68] A.G. Brown, W.C. Tbuman and J.W. McBain, J. Colloid
Sci., 8 (1853) 491.
[69] R.J. Mannheimer and R.A. Burton, J. Colloid Interface
Sci., 32 (1970) 73.
[70] N. Lifschutz, M.G. Hedge and J.C. Slattery, J. Colloid
Interface Sci., 37 (1971)73.
[71] F.C. Goodrich and L.H. Allen, J. Colloid Interface Sci.,
40 (1972) 329.
[72] P.B. Briley, A.R. Deemer and J.C. Slattery, J. Colloid
Interface Sci., 56 (1976) 1.
[73] F.C. Goodrich and A.K. Chatterjee, J. Colloid Interface
Sci., 34 (1970) 36.
[74] R. Shail, J. Eng. Math., 12 (1978) 59.
[75] S.G. Oh and J.C. Slattery, J. Colloid Interface Sci., 67
(1978) 516.
[76] A.M. Davis and M.E. O'Neill, Int. J. Multiphase Flow,
5 (1979)413.
[77] J. Ross, J. Phys. Chem., 62 (1957) 531.
[78] N.W. Tschoegl, Kolloid-Z., 181 (1962) 19.
[79] K. Inokuchi, Bull. Chem. Soc. Jpn., 26 (1953) 471.
[80] K. Inokuchi, Bull. Chem. Soc. Jpn., 27 (1954) 203.
[81] K. Inokuchi, Bull. Chem. Soc. Jpn., 27 (1954) 432.
[82] K. Inokuchi, Bull. Chem. Soc. Jpn., 28 (1955) 453.
[83] K. Inokuchi, Bull. Chem. Soc. Jpn., 26 (1953) 471.
[84] T. Tachibana, K. Inokuchi and T. Inokuchi, Kolloid-Z.,
167 (1959) 141.
[85] K. Motomura and R. Matuura, J. Colloid Sci., 18
(1963) 295.
[86] B. Biswas and D.A. Haydon, Proc. R. Soc. London Ser.
A, 271 (1963) 296.
[87] B. Biswas and D.A. Haydon, Proc. R. Soc. London Ser.
A, 271 (1963) 317.
[88] G.S. Patil, S.S. Katti and B. Biswas, J. Colloid Interface
Sci., 25 (1967) 462.

115

[89] G.S. Patil, S.S. Katti, J. Colloid Interface Sci., 30


(1969) 219.
[90] B.M. Abraham, K. Miyano, S.Q. Xu and J.B. Ketterson,
Rev. Sci. Instrum., 54 (1983) 213.
[91] B.M. Abraham, K. Miyano, S.Q. Xu andJ.B. Ketterson,
Phys. Rev. Lett., 49 (1982) 1643.
[92] B.M. Abraham, K. Miyano and J.B. Ketterson, Ind.
Eng. Chem. Prod. Res. Dev., 23 (1984) 245.
[93] S.S. Feng, R.C. MacDonald and B.M. Abraham,
Langmuir, 7 (1991) 572.
[94] B.M. Abraham and J.B. Ketterson, Langmuir, 1 (1985)
461.
[95] B.M. Abraham and J.B. Ketterson, Langmuir, 2 (1986)
801.
[96] J.B. Peng, G.T. Barnes and B.M. Abraham, Langmuir,
9 (1993)3574.
[97] T.S. Jiang, J.D. Chen and J.C. Slattery, J. Colloid
Interface Sci., 96 (1983) 7.
[98] R. Shail, Int. J. Multiphase Flow, 5 (1979) 169.
[99] R. Shail and D.K. Gooden, Int. J. Multiphase Flow, 7
(1981) 245.
[100] Y.C. Ray, H.O. Lee, T.L. Jiang and T.S. Jiang, J. Colloid
Interface Sci., 119 (1987) 81.
[ 101 ] J. Kr/igel, S. Siegel, R. Miller, M. Born and K.-H. Schano,
Colloids Surfaces A, 91 (1994) 169.
[102] J. Kr~igel, S. Siegel and R. Miller, Prog. Colloid Polym.
Sci., 97 (1994) 183.
[103] J. Kr/igel, R. Wiistneck, D.C. Clark, P.J. Wilde and
R. Miller, Colloids Surfaces A, 98 (1995) 127.
[104] J. Kr~gel, A.M. Stortini, N. Degli-Innocenti, G. Loglio
and R. Miller, Colloids Surfaces A, 101 (1995) 129.
[105] J. Kr/igel, G. Kretzschmar, J.B. Li, G. Loglio, R. Miller
and H. M6hwald, Thin Solid Films, (1995), in press.
[106] G.T. Shahin, Ph.D. Thesis, University of Pennsylvania,
1986.
[107] H.E. Gaub and H.M. McConnell, J. Phys. Chem., 90
(1986) 6830.
[108] K.V. Zotova and A.A. Trapeznikov, Kolloidn. Zh., 26
(1964) 190.
[109] A.A. Trapeznikov and E.S. Dokukina, Dokl. Akad.
Nauk USSR, 231 (1976) 405.
[ 110] A.A. Trapeznikov and E.S. Dokukina, Kolloidn. Zh., 40
(1978) 92.
[111] A.A. Trapeznikov and L.A. Korsunova, Kolloidn. Zh.,
40 (1978) 100.
[112] E.S. Dokukina and A.A. Trapeznikov, Kolloidn. Zh., 44
(1982) 667.
[113] V.I. Izamilova, G.P. Jampolskaja, L.E. Bobrova and
Z.D. Tulovskaja, in A.A. Abramzon and E.D. Shchukin
(Eds.), Poverchnostnye Javlenija i PoverchnostnoAktivnye Vescestva, Chimija, Leningrad, 1984 (in
Russian).
[114] A.A. Trapeznikov and E.S. Dokukina, Kolloidn. Zh., 39
(1977) 1209.
[115] R. Wtistneck and H. Fruhner, Colloid Polym. Sci., 259
(1982) 1228.
[116] R. W0stneck, Colloid Polym. Sci., 262 (1984) 821.

116

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

[,117] R. Wiastneck, H. Hermel and G. Kretzschmar, Colloid


Polym. Sci., 262 (1984) 827.
[118] R. WQstneck, V.V. KrotovandM. Ziller, ColloidPolym.
Sci., 262 (1984) 67.
[ 119] R. Wflstneck and L. Zastrow, Colloid Polym. Sci., 263
(1985) 778.
[ 120] R. W0stneck, L. Zastrow and G. Kretzschmar, Kolloidn.
Zh., 47 (1985) 462.
[,121] R. W0stneck and L. Zastrow, Colloid Polym. Sci., 263
(1985) 749.
[ 122 ] R. Wtistneck, L. Zastrow and G. Kretzschmar, Kolloidn.
Zh., 49 (1987) 10.
[,123] R. W0stneck and G. Kretzschmar, Kolloidn. Zh., 49
(1987) 555.
[ 124] R. WQstneck, G. Kretzschmar and L. Zastrow, Kolloidn.
Zh., 49 (1987) 239.
[ 125] R. Wtistneck, N.P. W0stneck, H. Hermel and L. Zastrow,
Kolloidn. Zh., 49 (1987) 244.
[,126] V.G. Vins and A.A. Trapeznikov, Kolloidn. Zh., 41
(1979) 392.
[,127] E.J. Vernon-Carter and P. Sherman, J. Dispersion Sci.
Technol., 2 (1981) 399.
[,128] V.D. Kiosseoglou and P. Sherman, Colloid Polym. Sci.,
261 (1983)520.

[,146] R. W0stneck, E. Buder, R. Wetzel and H. Hermel,


Colloid Polym. Sci., 267 (1989) 429.
[,147] R. Wiistneck, E. Buder, R. Wetzel, H. Hermel and
K. Lunkenheimer, Colloid Polym. Sci., 267 (1989) 516.
[ 148] R. W0stneck and T. W/irnheim, Colloid Polym. Sci., 266
(1988) 926.
[,149] R. Miller, G. Loglio, U. Tesei and K.-H. Schano, Adv.
Colloid Interface Sci., 37 (1991) 73.
[,150] G. Loglio, U. Tesei, R. Miller and R. Cini, Colloids
Surfaces, 61 (1991) 219.
[,151 ] K. Lunkenheimer and R. Miller, Tenside Detergents, 16
(1979) 312.
[,152] K. Lunkenheimer and R. Miller, Material Sci. Forum,
25-26 (1988) 351.
[-153] E.H. Lucassen-Reynders, J. Colloid Interface Sci., 42
(1973) 573.
[-154] E.H. Lucassen-Reynders, J. Colloid Interface Sci., 42
(1973) 573.
[,155] E.H. Lucassen-Reynders, J. Lucassen, P.R. Garrett,
D. Giles and F. Hollway, in E.D. Goddard (Ed.), Adv.
Chem. Ser., 144 (1975) 272.
[,156] G. Kretzschmar and R. Miller, Adv. Colloid Interface
Sci., 36 (1991) 65.
[157] F. de Voeght and P. Joos, J. Colloid Interface Sci., 98

[,129] H.J. Rivas and P. Sherman, Colloids Surfaces, 11


(1984) 155.

(1984) 20.
[,158] S. Hhrd and R.D. Neuman, J. Colloid Interface Sci., 115

[,130] G. Doxastakis and P. Sherman, Colloid Polym. Sci.,


264 (1986) 254.
[,131] E. Dickinson, B.S. Murray and G. Stainsby, J. Colloid
Interface Sci., 106 (1985) 259.
[,132] E. Dickinson, B.S. Murray and G. Stainsby, J. Chem.

(1987) 73.
[,159] C. Stenvot and D. Langevin, Langmuir, 4 (1988) 1179.
[,160] J.C. Earnshaw and C.J. Hughes, Langmuir, 7 (1991)
2419.
[,161] J.C. Earnshaw, R.C. McGivern, A.C. McLaughlin and
P.J. Winch, Langmuir, 6 (1990) 649.
[-162] J.C. Earnshaw, P.J. Winch, J. Phys. Condensed Matter,
2 (1990) 8499.
[,163] M. Hennenberg, X.-L. Chu, A. Sanfeld and M.G.
Velarde, J. Colloid Interface Sci., 150 (1992) 7.
[,164] Q. Jiang, Y.C. Chiew and J.E. Valentini, Langmuir, 8
(1992) 2747.
[,165] V. Romero-Rochin, C. Varea and A. Robledo, Physica
A, 184 (1992) 367.
[166] S.M. Sun and M.C. Shen, J. Math. Anal. Appl., 172
(1993) 533.
[,167] V. Thominet, C. Stenvot and D. Langevin, J. Colloid
Interface Sci., 126(1988)54.
[,168] R.C. McGivern and J.C. Earnshaw, Langmuir, 5 (1989)
545.
[,169] J.C. Earnshaw and D.J. Robinson, J. Phys. Condensed
Matter, 2(1990) 9199.
[170] K. Sakai, P.-K. Choi, H. Tanaka and K. Takagi, Rev.
Sci. Instrum., 62(1991)1192.
[171] K. Sakai, H. Kikuchi and K. Takagi, Rev. Sci. Instrum.,
63 (1992)5377.
[,172] K. Lunkenheimer and G. Kretzschmar, Z. Phys. Chem.
(Leipzig), 256 (1975) 593.
[,173] K.-D. Wantke, R. Miller and K. Lunkenheimer, Z. Phys.
Chem. (Leipzig), 261 (1980)1177.
[-174] C.A. MacLeod and C.J. Radke, 9th International

Soc. Faraday Trans., 84 (1988) 871.


[,133] E. Dickinson, S.E. Rolfe and D.G. Dalgleish, Int. J. Biol.
Macromol., 12(1990)189.
[134] H. H0hnerfuss, P.A. Lange and W. Walter, J. Colloid
Interface Sci., 108 (1985) 442.
[ 135] H. H0hnerfuss, J. Colloid Interface Sci., 120 (1987) 281.
[,136] H. Hiihnerfuss, J. Colloid Interface Sci., 126 (1988) 384.
[,137] G. Weidemann and D. Vollhardt, Colloids Surfaces A,
100 (1995)187.
[,138] G. Weidemann and D. Vollhardt, Thin Solid Films, 264
(1995) 94.
[139] H. Rehage and M. Veyessie, Ber. Bunsenges. Phys.
Chem., 89 (1985) 1166.
[,140] H. Rehage and M. Veyessie, Progress and Trends in
Rheology II, Supplement to Rheol. Acta, 26 (1988) 200.
[,,141] H. Rehage and M. Veyessie, Angew. Chem., 102
(1990) 497.
[,142] K. Miyano and M. Veyessie, Phys. Rev. Lett., 52
(1984) 1318.
[,143] K. Lunkenheimer and R. Miller, J. Colloid Interface
Sci., 120 (1987) 176.
[,144] R. Wtistneck and H.-J. Miiller, Colloid Polym. Sci., 264
(1986) 97.
[,145] R. Wi~stneck, R. Wetzel, E. Buder and H. Hermel,
Colloid Polym. Sci., 266 (1988) 1061.

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

[175]
[176]
[177]

[178]
[179]
[180]

[181]
[182]
[183]

[184]
[185]
[186]
[ 187]
[188]
[ 189]
[190]
[191]
[192]
[193]
[194]
[ 195]
[196]
[197]
[ 198]
[199]
[200]
[201]
[202]
[203]
[204]

Symposium Surfactants in Solution, Varna, 1992,


T2.A3.2.
C.A. MacLeod and C.J. Radke, J. Colloid Interface Sci.,
160 (1993) 435.
K.J. Stebe arid D.O. Johnson, 67th Annual Colloid and
Surface Science Symposium, Toronto, 1993, p. 404.
D.T. Wasan, K. Koszo and R. Nagarajan, 67th Annual
Colloid and Surface Science Symposium, Toronto,
1993, p. 285.
C.H. Chang and E.I. Franses, Chem. Eng. Sci., 49
(1994) 313.
C.H. Chang and E.I. Franses, J. Colloid Interface Sci.,
164 (1994) 107.
T. Horozov, K. Danov, P. Kralschewsky, I.B. Ivanov
and R. Borwankar, 1st World Congress on Emulsions,
Paris, Vol. 2, 1993, 3-20-137.
R. Nagarajan and D.T. Wasan, J. Colloid Interface Sci.,
159 (1993) 164.
J. Lucassen and D. Giles, J. Chem. Soc. Faraday Trans.
1, 71 (1975) 217.
D.S. Dimitrov, I. Panaiotov, P. Richmond and L. TerMinassian-Saraga, J. Colloid Interface Sci., 65 (1978)
483.
G. Kretzschmar and K. K6nig, J. SAM, 9 (1981) 203.
J. Lucassen and M. van den Tempel, Chem. Eng. Sci.,
27 (1972)1283.
J. Lucassen, Faraday Discuss. Chem. Soc., 59 (1976) 76.
V.G. Levich, Acta Physicochim. URSS, 14 (1941) 307.
V.G. Levich, Acta Physicochim. URSS, 14 (1941) 322.
V.G. Levich, Physicochemical Hydrodynamics, Prentice
Hall, Englewood Cliffs, NJ, 1962.
R.S. Hansen, J. Colloid Sci., 16 (1961) 549.
R.S. Hansen and J.A. Mann, J. Appl. Phys., 35
(1964) 152.
M. van den Tempel and R.P. van de Riet, J. Chem.
Phys., 42 (1965) 2769.
J. Lucassen and R.S. Hansen, J. Colloid Interface Sci.,
22 (1966) 32.
C. Lemaire and D. Langevin, Colloids Surfaces, 65
(1992) 101.
J. Lucassen, Trans. Faraday Soc., 64 (1968) 2221.
J. Lucassen, Trans. Faraday Soc., 64 (1968)2230,
M. van den Tempel, Chem. Ing. Tech., 43 (1971) 1260.
J. Lucassen and M. van den Tempel, J. Colloid Interface
Sci., 41 (1972) 491.
K.-D. Wantke, K. Lunkenheimer and C. Hempt,
J. Colloid Interface Sci., 159 (1993) 28.
D.O. Johnson and K.J. Stebe, J. Colloid Interface Sci.,
168 (1994) 21.
G. Loglio, U. Tesei and R. Cini, Colloid Polym. Sci.,
264 (1986) 712.
G. Loglio, U. Tesei and R. Cini, Rev. Sci. Instrum., 59
(1988) 2045.
G. Loglio, U. Tesei, N. Degli-Innocenti, R. Miller and
R. Cini, Colloids Surfaces, 57 (1991) 335.
G. Loglio, R. Miller, A. Stortini, N. Degli-Innocenti,
U. Tesei and R. Cini, Colloids Surfaces A, 90 (1994) 215.

117

[205] M. van Uffelen and P. Joos, J. Colloid Interface Sci.,


158 (1993) 452.
[206] R. Miller, Z. Policova, R. Sedev and A.W. Neumann,
Colloids Surfaces, 76 (1993) 179.
[207] R. Miller, R. Sedev, K.-H. Schano, C. Ng and
A.W. Neumann, Colloids Surfaces, 69 (1993) 209.
[208] P. Joos and M, van Uffelen, J. Colloid Interface Sci.,
155(1993) 271.
[209] P. Joos and M. van Uffelen, Colloids Surfaces, 75
(1993) 273.
[210] G. Loglio, R. Miller, A. Stortini, N. Degli-Innocenti,
U. Tesei and R. Cini, Colloids Surfaces A, 95 (1995) 63.
[211] J. de Feijter, J. Benjamins and M. Tamboer, Colloids
Surfaces, 27 (1987)243.
[212] D.E. Graham and M.C. Phillips, J. Colloid Interface
Sci., 70 (1979) 403.
[213] D.E. Graham and M.C. Phillips, J. Colloid Interface
Sci., 70 (1979) 415.
[214] D.E. Graham and M.C. Phillips, J. Colloid Interface
Sci., 70 (1979) 427.
[215] D.E. Graham and M.C. Phillips, J. Colloid Interface
Sci., 76 (1980)227.
[216] D.E. Graham and M.C. Phillips, J. Colloid Interface
Sci., 76 (1980) 240.
[217] F.C. Goodrich, J. Phys. Chem., 66 (1962) 1858.
[218] T. Yasunaga and M. Sasaki, in W.J. Gettins and E.
Wyn-Jones (Eds.), Techniques and Application of Fast
Reactions in Solutions, D. Reidel, 1979, p. 579.
[219] D. Wielebinski and G.H. Findenegg, Prog. Colloid
Polym. Sci., 77 (1988) 100.
[220] B.A. Noskov and A.A. Vasilev, Kolloidn. Zh., 50
(1988) 909.
[221] B.A. Noskov and M.A. Schinova, Kolloidn. Zh., 51
(1989) 69.
[222] B.A. Noskov, O.A. Anikieva and N.V. Makarova,
Kolloidn. Zh., 52 (1990) 1091.
[223] B.A. Noskov, Kolloidn. Zh., 45 (1983) 689; Mech. ~idk.
Gasa, 1 (1991) 129.
[224] H. Fruhner and K.-D. Wantke, Colloids Surfaces A,
in press.
[225] H. Fruhner, K. Lunkenheimer and R. Miller, in L. Ratke
and B. Feuerback (Eds.), Lecture Notes in Physics,
Vol. 464, Springer-Verlag, Berlin, 1995, p. 41.
[226] J. Lucassen and G.T. Barnes, J. Chem. Soc. Faraday
Trans. 1, 68 (1972) 2129.
[227] G. Kretzschmar, Prog. Colloid Polym. Sci., 77 (1988) 72.
[228] P. Cheng, Ph.D. Thesis, University of Toronto, 1990.
[229] P. Cheng, D. Li, L. Boruvka, Y. Rotenberg and A.W.
Neumann, Colloids Surfaces, 43 (1990) 151.
[230] L. Liggieri, F. Ravera and A. Passerone, J. Colloid
Interface Sci., 140 (1990) 436.
[231] A. Passerone, L. Liggieri, N. Rando, F. Ravera and
E. Ricci, J. Colloid Interface Sci., 146 (1991) 152.
[232] A.J.J. Kokelaar, A. Prins and M. de Gee, J. Colloid
Interface Sci., 146 (1991) 507.
[233] D.J.M. Bergink-Martens, H.J. Bos, A. Prins and B.C.
Schulte, J. Colloid Interface Sci., 138 (1990) 1.

118

R. Miller et al./ Colloids Surfaces A: Physicochem. Eng. Aspects 111 (1996) 75-118

[234] A. Prins, Chem. Ing. Tech., 64 (1992)73.


[235] M. de Rijcker and R. Defay, Bull. Soc. Chim. Belg., 65
(1956) 794.
[236] P. Joos and P. de Kayser, The overflowing funnel as a
method for measuring surface dilational properties,
Levich Birthday Conference, Madrid, 1980.
[237] G. Kretzschmar and D. Vollhardt, Ber. Bunsenges. Phys.
Chem., 71 (1967) 410.
[238] P.R. Garrett, J. Chem. Soc. Faraday Trans. 1, 69
(1976) 2174.
[239] I.N. Kremlev and V.I. Lobyshev, Biophys., 30 (1985) 74.
[240] J.B. Li, R. Miller and H. MOhwald, Thin Solid Films,
(1996) in press.
[241] G. Kretzschmar, J.B. Li, R. Miller, H. Motschmann and
H. M6hwald, Colloids Surfaces A, in press,
[242] H. MOhwald, Annu. Rev. Phys. Chem., 41 (1990) 441.
[243] J. Daillant, L. Bosio, B. Harzallah and J.J. Benattar,
J. Phys. II, 1 (1991) 149.

[244] P. Joos, M. van Uffelen and G. Serrien, J. Colloid


Interface Sci., 152 (1992) 521.
[245] G. Serrein, G. Geeraerts, L. Ghosh and P. Joos, Colloids
Surfaces, 68 (1992) 219.
[246] C. Hempt, K. Lunkenheimer and R. Miller, Z. Phys.
Chem. (Leipzig), 266 (1985) 713.
[247] R. Miller, J. Kr~igel, G. Loglio and A.W. Neumann,
Proceedings of the 1st World Congress on Emulsions,
Paris, Vol. 2, 1993, 3 20-143.
[248] D.O. Grigoriev, B.A. Noskov and S.I. Semchenko,
Kolloidn, Zh., 55 (1993) 45.
[249] B.A. Noskov and D.O. Grigoriev, Prog. Colloid Polym.
Sci., 97 (1994) 1.
[250] B.A. Noskov and D.O. Grigoriev, Langmuir, submitted
for publication.
[251] B.A. Noskov, D.O. Grigoriev and R. Miller, J. Colloid
Interface Sci., submitted for publication.
[252] J. Lucassen, Colloids Surfaces, 65 (1992) 139.

You might also like