You are on page 1of 15

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/226295719

Rheology of interfacial layers


Article in Colloid and Polymer Science June 2010
DOI: 10.1007/s00396-010-2227-5

CITATIONS

READS

62

70

6 authors, including:
James K Ferri

Aliyar Javadi

Lafayette College

Technische Universitt Berlin

47 PUBLICATIONS 856 CITATIONS

78 PUBLICATIONS 495 CITATIONS

SEE PROFILE

SEE PROFILE

Nenad Mucic

Rainer Wstneck

Max Planck Institute of Colloids and Interfaces

Max Planck Institute of Colloids and Interfaces

22 PUBLICATIONS 255 CITATIONS

95 PUBLICATIONS 1,927 CITATIONS

SEE PROFILE

SEE PROFILE

All content following this page was uploaded by Aliyar Javadi on 27 July 2016.
The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.

Colloid Polym Sci (2010) 288:937950


DOI 10.1007/s00396-010-2227-5

INVITED REVIEW

Rheology of interfacial layers


Reinhard Miller & James K. Ferri & Aliyar Javadi &
Jrgen Krgel & Nenad Mucic & Rainer Wstneck

Received: 3 February 2010 / Revised: 14 April 2010 / Accepted: 14 April 2010 / Published online: 1 May 2010
# Springer-Verlag 2010

Abstract The response of interfacial layers to deformations


in size and shape depends on their composition. The
corresponding main mechanical quantities are elasticity
and viscosity of dilation and shear, respectively. Hence, the
interfacial rheology represents a kind of two-dimensional
equivalent to the traditional bulk rheology. Due to growing
interest in the quantitative understanding of foams and
emulsions, more works are dedicated to studies on
interfacial rheology. This overview presents the theoretical
basis for traditional and recently developed experimental
tools and discusses their application to different interfacial
systems. While dilational rheology provides information on
the composition of mixed interfacial layers, the shear
rheology gives answers essentially on structures formed at
an interface. The most frequently used methods at present
are the oscillating drop and bubble tensiometry methods for
dilational deformations and oscillating ring/bicone rheometers for shear deformations.
Keywords Interfacial rheology . Dilational elasticity and
viscosity . Shear elasticity and viscosity . Surfactant
adsorption layers . Mixed protein surfactant layers

R. Miller (*) : A. Javadi : J. Krgel : N. Mucic : R. Wstneck


Max Planck Institute of Colloids and Interfaces,
Potsdam/Golm, Germany
e-mail: miller@mpikg.mpg.de
J. K. Ferri
Department of Chemical and Biomolecular Engineering,
Lafayette College,
Easton, PA 18042, USA

Introduction
Interfacial rheology deals with the response of interfaces
against deformations and is relevant in many technical
applications, such as mass transfer, monolayers, foaming,
emulsification, oil recovery or high-speed coating. Its
history started only in the nineteenth century when the
Paris Academy of Sciences published a paper by Ascherson
[1] on the behaviour of a skin formed at the interface
between an aqueous protein solution and oil. Later in 1845,
Hagen [2] postulated a viscosity of the interfacial region
different from that of the adjoining bulk phase. And the
possibly first experiment in interfacial rheology was
performed 25 years later by Plateau [3], who compared
the damping of an oscillating magnetic needle immersed in
a liquid with one floating at its surface, although not
knowing that the observed effect was caused by the
adsorption layer of surface active impurities. This fact was
recognised by Marangoni [4] who explained the effects
observed by Plateau via a surface compression in front of
the needle and a dilation on the opposite side and the
resulting surface tension gradient influenced the movement
of the needle. Gibbs [5] gave a first interpretation for the
interfacial layers response, which was improved then by
Lord Rayleigh [6].
The pioneers of the twentieth century, driving this
scientific field further were Boussinesq [7], who proposed
a two-dimensional analogue of the Newtonian fluid as the
terminology of modern interfacial rheology, which was then
extended by Levich [8], Dorrestein [9], Ericksen [10],
Oldroyd [11], Scriven [12] and others.
Starting from Levichs proposal to discuss surface
tension gradients in terms of surface dilational, Hansen
[13] and the Dutch School [14, 15] developed the modern
view of the dilational rheology.

938

Since recently, the theoretical background of interfacial


rheology and the available experimental tools were only
described in some review articles and book chapters, for
example, by Joly [16], Lucassen [17], Edwards et al. [18] or
Miller et al. [19]. The possibly first book devoted to
interfacial rheology was published only in 2009 [20]. It
contains detailed descriptions of experimental methods of
dilational and shear rheology together with selected
application to various interfacial layers formed by surfactants, polymers, proteins and their mixtures.
In recent time, the role of the mechanical interfacial
behaviour for the stabilisation of foams and emulsions
becomes more and more emphasised, However, there are
various stabilisation mechanisms for foams or emulsions,
including the rupture of liquid films as their building
blocks, liquid drainage and Ostwald ripening; therefore, it
is a very complex problem to define the real impact of
interfacial rheology for these systems [2125].
The aim of the present work is a brief summary of the
state of art of the two-dimensional rheology, studied under
dilation/compression and shear, respectively. This includes
a discussion of the experimental tools as well as a selection
of experimental examples, which can be only very
subjective due to the large number of publications in this
field. As very recently, quite extensive reviews have been
published on shear rheology of adsorption layers and liquid
films [2628], this topic will be kept rather short here.

Survey of interfacial rheology


The main kinds of deformations of an interfacial layer are
expansions, compressions (constant shape but change in
area) and shear (constant area but change in shape). This
principle is demonstrated in Fig. 1 schematically with two
concentric circles.
As mentioned above, the proposal of Boussinesq to treat
the two-dimensional rheology in analogy to the bulk
rheology lead to a similar tool box, i.e. the behaviour of
interfacial layers is discussed in terms of springs and

Fig. 1 Interfacial deformations: compression/expansion (left) and


shear (right)

Colloid Polym Sci (2010) 288:937950

dashpots. A spring is the equivalent to an ideal elastic


behaviour and described by the Hook law, while a dashpot
is the symbol for an ideal viscous behaviour described by
the Newton law (Figs. 2 and 3).
The serial combination of a spring and a dashpot, known
as Maxwell model, represents the typical visco-elastic
behaviour of interfacial layers.
The spring and dashpot approach, its combination in
form of the Maxwell model, and in form of many other
models, was discussed in general in [29] and applies to
shear as well as dilational deformations.
Interfacial shear deformations
In interfacial shear rheology, we have essentially two types
of experimental methods, the indirect and direct methods.
The indirect methods determine surface velocity profiles by
using visible inert particles from which the interfacial shear
properties can be determined. These methods are mainly
restricted to gas/liquid interfaces. In contrast, the direct
methods determine the torsional stress of a liquid interface
stressed by a body just touching it. The geometry of some
measuring bodies is rather sophisticated and also the
techniques to detect the resulting motion or stress is often
difficult. Reviews on shear rheology methods were published very recently, and experimental and theoretical
details can be found therein [27].
The Channel surface viscometer is a two-dimensional
analogue of the HagenPoiseuille law and used for
measurements of the surface shear viscosity S of spread
insoluble monolayers. It is based on the determination of
the flow rate of a film through a narrow channel or slit
under an applied two-dimensional pressure difference
and was first proposed in 1937 [30, 31].
A second group of channel viscometer, also applicable to
adsorbed layers, is the deep channel surface viscometer.
This technique has been first described by Davies and
Mayers [32] and is based on viscous-traction forces to
produce flow without surface pressure gradients. It consists
of a channel formed by two concentric sharp-edged cylinders
which just touch the liquid surface. While the cylinders are
held stationary and concentrically, the dish with the solution is
rotated and from the differences between the surface velocities
of talc particles on a clean and film-covered substrate provide

Fig. 2 Spring (left) and dashpot (right) as symbols for the main
elements of the mechanical behaviour of an interfacial layer

Colloid Polym Sci (2010) 288:937950

939

Fig. 3 Two-element model of a Maxwell (left) and KelvinVoigt (right) liquid interfacial layer

the surface viscosity. For measurements at liquid/liquid


interfaces, Wasan et al. [33] proposed modifications of the
deep channel viscometer.
The rotating knife-edge in wall surface viscometer introduced by Goodrich et al. [34] determines the surface velocity
and the angular velocity of a rotating outer ring, from which
the interfacial shear viscosity can be calculated. The method
has a high sensitivity 105 mN s m1 hs 0:1 mN s m1 . As in
the deep-channel viscometer, a small particle is placed within
the surface to determine the rotational speed of the fluid
interface.
As a fourth method, the Transient rotating cylinder
apparatus was proposed by Krieg et al. [35] for surface
layers and later by Hassager and Westborg [36] for liquid
liquid interfaces. It is based on the registration of the
decreased surface motion after a sudden cessation of a
rotating cylinder. The generation of oscillatory shear
deformations yields the interfacial visco-elastic response,
for which the theoretical basis was elaborated by Prieditis et
al. [37].
The direct methods, in contrast to the indirect ones,
determine the torsional stress of the liquid interface stressed
by a touching body. The techniques can be classified into
steady rotation, transient deformation or oscillating mode
(forced or damped oscillation) and consist classically of a
circular measuring body suspended from a thin torsion wire
such that it just touches the interface. However, also
oscillating needles are known as measuring bodies.
The Torsion pendulum rheometer is one of the oldest
methods in interfacial rheology [38]. The schematic of the
Fig. 4 Schematic of a single
knife-edge (left) and biconus
(right) torsion pendulum
rheometer

surface viscometer of Brown et al. [39] using a sharp knifeedge body is shown in Fig. 4. The cup is rotated and the
torsional stress measured to determine the interfacial shear
viscosity once the steady state has been reached. Most
recently it was shown that the use of a biconical measuring
body is advantageous in particular for liquid/liquid interfaces.
For measurement at very low shear rates, a torsion
pendulum was described by Krgel et al. [40], which allows
experiments with very small mechanical deformations of
the adsorbed layer.
The use of an oscillating needle in an interfacial
rheometer has been proposed first by Shahin [41]. Here, a
magnetic field forces the rod, floating at the interface, to
move within the interface in order to produce a shear filed.
The principal set-up of an oscillating needle experiment
used by Brooks et al. [42] is shown in Fig. 5 (for more
details, see for example [26]).
Oscillating torsion pendulum or needle rheometers are
available as commercial instruments and do not need to be
built individually. The main request thereby is to measure
data which are characteristics of the studied surface layer
rather than depend on the measuring principle or instrument. It was shown recently in [43] that the use of two
rheometers (ISR1 by SINTERFACE and MCR 301-ISR by
Physica) of different sensitivity provide complementary
results, as demonstrated in Fig. 6 for four spread polymer
layers. Note that the MCR is an example of a classical bulk
rheometer, which can be equipped with an interfacial took
to be applied for the characterisation of interfacial layers.
There are more providers of similar equipments; however,

940

Fig. 5 Principle oscillating needle set-up; according to Brooks et al.


[42]

all still have sensitivities significantly less than that of a


torsion wire rheometer. We can expect that it will still take a
bit of time until these much more robust instruments can
replace the rather fragile torsion wire instruments like the
ISR1.
The classical rheometers with an interfacial measuring
set as extra equipment are also suitable for studies of crosslinked monolayers, multi-layers or films [44, 45]. However,
these systems will not be further discussed here in detail as
the rheology of these materials would represent a separate
topic.

Colloid Polym Sci (2010) 288:937950

influence of surfactants on water waves and proposed the


first explanation of the observed effect [48]. He, however,
thought that the oil film acted as a lubricant between the
water and moving air. Lord Rayleigh used capillary waves
to measure the surface tension of pure liquids [49], but the
application to surfactant solutions failed due to a missing
inadequate theory. The foundations of the corresponding
modern theory were formulated only in the middle of the
twentieth century by works of Lamb [50], Levich [8, 51]
and Dorrestein [9, 52]. From his theoretical analysis,
Lucassen [15] predicted the existence of longitudinal
capillary waves, which were then immediately proven by
experiments discovery [53].
The great advantage of the capillary wave damping method
over others is that it is suitable in a frequency range from
25 Hz to very high frequencies of even 4 kHz [14, 54].
Although many designs were developed in the past, so far,
there is no commercial instrument available for the capillary
and longitudinal wave damping technique. In Fig. 7, an
example is given for a capillary wave damping instrument,
as described in more detail elsewhere [46, 55].

Theoretical models of dilational rheology


Expansions and compressions of interfacial layers
The damping of capillary waves is the oldest method in
dilational surface rheology, and the effect of surfactants on
their properties is probably one of the oldest scientific
problems ever discussed in the literature [46]. The first
reference to capillary waves, generated by oil drops when
falling onto the surface of water, were written in clay and
date back to the eighteenth century B.C. [47]. Much more
recently, Benjamin Franklin performed experiments on the

Fig. 6 Interfacial shear elasticity G as a function of the polymer


surface concentration for polymers of different molecular weight: a
2.708105, b 1.59105, c 1.043105, d 8.0104 g/mol; measurements were performed with a knife-edge torsion pendulum technique
(unfilled circles) and a bicone equipped rheometer (filled squares),
respectively; according to [43]

As mentioned above, dilational rheology described the


response of interfacial layers to expansions and compressions. It is based on parameters which are closely linked to
the equilibrium properties of the relevant adsorption layers
and also to the mechanisms controlling their formation.
Only for purely elastic interfaces, the dynamic interfacial
tension response g(t) follows immediately the area change
A(t) without any phase lag. In general, however, the

Fig. 7 Scheme of a capillary wave damping apparatus, according to


[55]

Colloid Polym Sci (2010) 288:937950

941

interfacial tension follows the area change with a certain


delay due to the various relaxation processes within the
interfacial layer and between the interface and the adjacent
bulk phases. For area changes with small amplitudes, the
corresponding surface tension change is described by the
convolution integral [56]:
Zt
dgt

Et  ld ln Aldl

1

where t is the time, dgt gt  g eq is the deviation of


interfacial tension
g(t) from
its equilibrium value g eq and


d ln At  At  Aeq =Aeq is the relative area change.
E (tl) is the weight factor, i.e. the surface relaxation
function on the time difference tl. The Fourier transformation applied to Eq. 1 yields Eiw dg iw=d ln Aiw,
where dg iw F fdgtg, dln Aiw F fdln Atg and
E iw F fEtg are the Fourier transformations of the
respective time functions. The function E(iw) is the
dynamic surface elasticity [56], presented as a complex
quantity. Separated into real and imaginary part,
E iw Er w i  Ei w

We obtain the surface dilatation elasticity Er(w) and the


surface dilatation viscosity hd w Ei w=w, respectively,
at the oscillation frequency f=w/2 with w being the
circular frequency [57].
The dynamic surface elasticity is a function of the
deformation frequency and can be measured via the
interfacial tension response to deformations of the interface,
most frequently harmonic perturbations. For adsorption
layers of a single surfactant controlled by a diffusional
adsorption mechanism, the dynamic surface elasticity is
given by [58, 59]:
E iw E0 1 1  iz 1 E0

1 z iz
1 2z 2z 2

E0 is the high frequency limit of the elasticity and wD the


characteristic frequency.
 
dg
D dc 2
E0 
; wD
4
d ln *
2 d*
p
The parameter z wD =w is a dimensionless parameter. As one can see easily, E0 and wD are given by the
adsorption isotherm as a function of the surfactant bulk
concentration c, and by the surfactants diffusion coefficient
D. From Eq. 3, we can determine the modulus and phase
angle of the dynamic surface elasticity

1=2
jEj E0 1 2z 2z 2
; f arctanz=1 z 

For a more general case of a mixed adsorption mechanism,


including the transport by diffusion and an adsorption step

governed by a kinetic barrier, the dynamic surface elasticity


is given by [60, 61]:
E E0

iwt 1 i=2z
1 iwt 1 i=2z

The relaxation time depends on the adsorption model. For a


Langmuir adsorption model, we have t ak bk c=*1 1
with the kinetic coefficients of adsorption and desorption, k
and bk, respectively, and the maximum adsorbed amount
*1 ; c is the surfactants bulk concentration. For small
relaxation times, Eq. 6 simplifies into Eq. 3.
For surfactant solutions above the critical micelle
concentration (CMC), the diffusion of micelles and their
kinetics of formation and disintegration have a significant
effect on the dynamic surface elasticity. There are several
theoretical approaches for describing the dynamic surface
elasticity of micellar solutions [6265]. In the simplest case,
for a slow relaxation of the micelles, the dynamic surface
elasticity can be described by an equation derived by
Lucassen [62]

1=2 1
2
E E0 1 1  iz 1 nb1=2 1 nD b

where D DM =D is the ratio of the diffusion coefficients


of micelles DM and monomers D, n is the aggregation
number of the micelles, b c  CMC=CMC, c is the
total surfactant concentration in the bulk. Equation 7
simplifies to Eq. 3 when we substitute D by the
effective
2
diffusion coefficient Deff D1 nb 1 nD b .
Joos and van Hunsel obtained the same Eq. 7 for a fast
relaxation of micelles, however, with n=1 [63].
The corresponding expressions for the dynamic surface
elasticity for mixed adsorption layers are much more
complex [66, 67]. For a mixture of two surfactants under
the condition of a diffusional exchange of matter mechanism, the respective expression were discussed very
recently in [68] and also successfully applied to protein
surfactant mixtures [69].

Modern instrumental tools


As mentioned earlier, instruments for studies of interfacial
rheology were special designs in a few laboratories. In many
cases, the efficiency of such apparatuses was not very high,
and a special knowledge was required to use them properly.
Since about 1995, commercial devices are available, and
by this, the studies on the mechanical properties of
interfacial layers are strongly increasing. It can also be the
other way round, i.e. the demand in studies of mechanical
interfacial properties induced the development of professional instrumentation.

942

Equipment for shear deformations


The first commercial shear rheometer was obviously the
CIR from Camtel as briefly mentioned in [70]. At about the
same time, the interfacial shear rheometer ISR1 was
introduced and described in detail [40, 71]. While the CIR
works with forced oscillations, the ISR1 is based on a free
damped oscillation. The main principle of this apparatus is
shown in Fig. 8.
The principle of this rheometer is based on a ring with a
sharp edge hanging at a torsion wire. By applying an
impulsive torque by an instantaneous movement of the
torsion head (motor), the pendulum performs free damped
oscillations. Experiments involving free damped oscillations are simpler to be performed as forced oscillations
because only the displacement of the measuring body has to
be monitored, done by the computer-controlled laser-mirror
position sensitive diode system. The rheological model
behind the theory for data analysis, elaborated a long time
ago by Tschoegl [72], is based on a parallel action of the
shear viscosity and elasticity (KelvinVoigt model) combined with the rheological elements of the experimental setup, including the elasticity of the torsion wire, the moment
of inertia of the measuring body and the friction of the
clean solvent interface.
During the last few years, additional instruments
appeared on the market, mainly bulk rheometers with a
sensitivity high enough to be applied to interfacial layers as
well [73]. These instruments work with rings or biconical
discs and can be applied to liquid/gas and liquid/liquid
interfaces.
Interfacial dilational experiments
The methodology of wave damping is the classical way of
measuring the dilational rheology of interfacial layers.
Elaborated by groups at Unilever [1315, 17, 53, 57] and
Fig. 8 Scheme of the torsion
shear rheometer ISR1
(SINTERFACE Technologies)

Colloid Polym Sci (2010) 288:937950

applied to various interfacial layers, even to interfaces


between two liquids [74], these methods never got designed
in the form of a commercial instrument. The principal set-up
of a capillary wave instrument was shown above in Fig. 7,
while a longitudinal wave apparatus is shown here in Fig. 9.
The way how to run the two experimental wave damping
apparatuses has been described in detail elsewhere [46, 55].
The longitudinal waves complement the capillary waves in
the available frequency window and in this way span over a
rather broad range from about 0.1 Hz (lower limit of
longitudinal waves) to about 4,000 Hz (upper limit of
capillary waves) [75].
During the last two decades, the oscillating drop and
bubble method was developed systematically and became
the most frequently used method. It splits into two
branches, using the shape of a drop/bubble or measuring
the capillary pressure inside it. The first relaxation
experiment based on transient changes of a pendent drop
was performed in 1993 [76], while first harmonic perturbations were published in 1996 [77]. In both cases, the
volume of a drop (or bubble) is changed, and the surface
tension response is determined from the shape, assumed to
be in a hydrodynamic equilibrium. The principle is shown
in Fig. 10. From the image of a drop, the shape coordinates
are extracted and then compared to those calculated via the
GaussLaplace equation of capillarity:


1
1
g

8
P0 rgz
R1 R2
where R1 and R2 are the two principal radii of interface
curvature, g is the interfacial tension, P0 is the pressure
difference at a reference plane, is the density difference,
g is the local gravitational constant and z is the vertical
height measured from the reference plane.
This methodology is generally rather old; however, it
became a suitable method for routine studies only after the
use of electronic cameras and digitising boards for the taken

Colloid Polym Sci (2010) 288:937950

943

Fig. 9 Schematic of a longitudinal wave instrument;


according to [55]

images. The pioneer of this now most widely used methodology is Neumann [78], who applied this technique for
various surface tension and contact angle measurements.
To easier solve Eq. 8, it can be transformed into a set of
first-order differential equations expressed by the geometric
parameters of the drop/bubble profile, i.e. the arc length s
and the normal angle f between the drop radius and the zaxis (i.e. the vertical direction) [79]:
dx
cosf
ds

dz
sinf
ds

10

df 1
2 sinf
bz 
ds R
b
x

11

Fig. 10 Pendent drop (left) and principle of calculated profiles in the


process of least square fitting (right) of theoretical drop shapes (black
shapes) via the Laplace equation to the drop coordinates (red shape)

When the apex of the bubble/drop is located in the point


(0,0), then x is the abscissa of the profile point, z is the
ordinate of the profile point, R is the radius of curvature in
the point (x,z), b is the radius at the apex and b rg=g.
The sign in Eq. 11 stands for sessile drops or captive
bubbles (+) and for pendant drops or emerging bubbles (),
respectively.
As the result of such measurements, we obtain the
surface tension as a function of time in parallel to the
change of the surface area, both harmonic oscillations. A
Fourier analysis allows, then, to extract the dilational
rheology in terms of pairs of parameters, such as highfrequency elasticity E0 and characteristic frequency wD, or
real and imaginary part of the visco-elasticity Er and Ei, or
elasticity module and phase angle |E| and f.
The frequency limit for drop and bubble oscillations is
essentially given by the time a liquid meniscus needs to
assume its equilibrium shape. This depends on its total size,
the surface tension and the viscosity of the liquid. For
liquid/gas interfaces and drops of a size of few millimetres,
it was found that the oscillating drop method is applicable
up to about 0.2 Hz [80]. Although commercial instruments
allow technically higher frequencies, the gained data can
provide erroneous results, such as a dilational elasticity for
the pure water/air interface.
The first attempt to use small spherical drops and
bubbles for studies of the dilational rheology was made
already in 1970 [81], although at this time, electric pressure
sensors were not available yet. The presently used capillary
pressure method for oscillating drops and bubbles goes
back to pioneering work of Passerone et al. [82, 83]. In
Fig. 11, the principle of a capillary pressure method is

944

Colloid Polym Sci (2010) 288:937950

Fig. 11 Schematic set-up of a capillary pressure method for oscillating drop and bubble experiments, using a closed cell (left) or open cell (right)
geometry; according to [84]

shown. On the left, the bubble or drop is immersed into the


measuring cell so that we speak about a closed cell
geometry. In contrast, on the right of Fig. 11, we see the
geometry of an open cell as the drop is formed outside the
measuring cell. The second type of cells has the advantage
that is can be used also for experiments with continuously
growing drops, for which the closed cell geometry cannot
provide sufficient space for a reasonable number of drops
formed and detached from the capillary tip.
The use of the oscillating drop/bubble method based on
the measurement of capillary pressure depends on a reliable
theory that allows extracting the rheological quantities. For
this, the pressure variations inside the cell have to be
carefully analysed. These pressure changes result from the
compression/expansion of the liquid
dPA

1 dP
B
dVA 
dVA
VA d ln V
VA

12

where VA is the volume of the liquid in the cell (Phase A),


and B=dP/dlnV is the bulk elasticity of the liquid (for
water, we have B=2.04109 Pa for an isotherm regime and
B=2.22109 Pa for an adiabatic regime). For our oscillation
experiments, we have to use the adiabatic regime.
For the variation of the liquid volume VA the cell is
given by variation of the piezoelectric actuator, VDrv, of
the volume of liquid passing though the capillary, VCap,
and the inner cell volume variations due to wall deformations, VWall, which leads to the volume balance
dVA dVDrv  dVCap dVWall  dVGas

A careful analysis, as described for example in [85], leads


finally to the real and imaginary part of the visco-elasticity,
which can be calculated from the known geometry of the
drop (aC, h0) and the measured pressure changes:


3 "

#


p h20 a2C
8g 0 h20  a2C
dPA0
dPA0
2

dVDrv0 cos
Er
w G2 
3
Bef =VA
16h20
dVm0 2
p h2 a2
0

Ei 

 "
2 3

p h20 aC
16h20

dVm0 2

dVDrv0 sin wG1

14

[58, 59], we obtain another set of equations for the real and
imaginary part of the visco-elasticity:

#
dPA0

13

15

where Bef is the effective elasticity of the measuring cell,


and G1 and G2 are the viscous and inertial coefficients,
respectively (see details in [85]).
From an interfacial rheological model of a visco-elastic
surface, including a diffusion-controlled exchange of matter
mechanism, proposed by Lucassen and van den Tempel

Er E0

1z
1 2z 2z 2

16

z
17
1 2z 2z 2
p
where z wD =2w, E0 c dg=d ln* is the limiting
(high frequency) elasticity, and wD c Ddc=d*2 is the
Ei E0

Colloid Polym Sci (2010) 288:937950

945

characteristic diffusion relaxation frequency. The comparison of Eqs. 14 and 15 with Eqs. 16 and 17 gives access to
the important quantities of the rheological model.
Also, capillary pressure instruments have a strongly
limited frequency range. Therefore, studies were even
performed under microgravity conditions for extending
the range of application of this methodology and improving
the quality of measured data, as summarised in [85]. The
large advantage of these experimental conditions is the
ideal sphericity of drops and bubbles of any dimension,
which, therefore, allows for oscillations at frequencies
much higher than on ground due to the absence of any
disturbing wobbling modes. The limitations given by the
hydrodynamics of the liquid system of course also hold in
microgravity.

Experimental examples
There are various studies published in literature, and we can
present only few well-selected systems here. The experimental data shown are mainly obtained from investigations
with oscillating drops and bubbles, at low frequencies via
the shape of the menisci, and at higher frequencies via
capillary pressure measurements.
Surfactant solutions
The dilational visco-elastic properties of many surfactants
are systematically investigated. Classical studies were done
with surfactants like fatty acids, fatty alcohols or alkyl
sulphates using capillary wave [58, 59] and oscillating
bubble techniques [81]. In the meantime, since oscillating
drop and bubble instruments became easier accessible,
many surfactant systems were investigated. In a recent
study, the dilational rheology of C12 dodecyl dimethyl
phosphine oxide (DMPO) was measured at various concentrations at low frequencies using the oscillating drop
profile analysis methods [86] and compared with data
obtained from oscillating bubble experiments [87, 88]. The
consideration of the internal compressibility discussed in
[89] allowed to reach good agreement between the
experimental and theoretical dependencies for E0. Some
data are presented in Fig. 12 measured with the drop profile
analysis tensiometer PAT-1 (SINTERFACE Technologies,
Germany) at an oscillation frequency of 0.1 Hz. Both real
and imaginary parts of the surface dilational modulus show
a maximum at a certain C12DMPO concentration, well
reproduced by the theoretical model (for details, see [86]).
The visco-elastic modulus as a function of concentration
|E|(c) measured for the technical non-ionic surfactant Triton
X-165 (commercial polyethylene glycol octylphenyl ethers)
solutions shows two distinct maxima (see Fig. 13). These

Fig. 12 Real and imaginary part of the complex visco-elasticity as a


function of C12DMPO concentrations at an oscillation frequencies of
0.1 Hz; Ei (filled squares) and Er (filled circles); curves are
calculations; according to [86]

maxima were predicted by the theory based on a combined


model including molecular reorientation and internal
molecular compressibility in the surface layer [90, 91].
While the first maximum is caused by the transition from
an expanded state of the adsorbed Triton molecules (where
the ethylene oxide groups are adsorbed at the interface [92])
to a more compact state, the second maximum is caused by
the internal compressibility.
As mentioned above, surfactants form micelles beyond a
critical concentration (CMC). The presence of micelles in
the solution has a strong impact on the dynamic properties
of the interfacial layer. This effect is caused by the fact that
micelles can disintegrate or be formed, depending on the local
bulk concentration. Experiments of adsorption kinetics and

Fig. 13 Dependence of the visco-elastic modulus |E| on the Triton X165 concentration measured by the buoyant bubble method (PAT1) for
the oscillation frequency of 0.1 Hz; experimental data (filled circles)
and theoretical lines according to [90, 91]

946

Colloid Polym Sci (2010) 288:937950

70
60

|E| [mN/m]

50
40
30
20
10
0
1E-03

1E-02

1E-01

1E+00

f [Hz]

Fig. 14 Visco-elasticity module |E| as a function of frequency f for


various 148 concentrations at the CMC=7 mol/l (filled circles),
and above the CMC at 12 (filled triangles), 15 (filled diamonds), 20
(ex symbols) and 30 mol/l (filled squares). Thin and thick dotted
lines are calculated for the CMC and 5 mol/l, respectively; thin solid
lines are calculated for c > CMC; the numbers correspond to
concentrations given in mol/l, according to [93]

dilational rheology have been performed to quantify this effect


[93]. The dependencies presented in Fig. 14 show the viscoelasticity modulus |E| as a function of frequency f for the
non-ionic surfactant 148 (octaethylene tetradecyl ether)
at the CMC=7 mol/l (the highest possible monomer
concentration) and at several concentrations above the
CMC (12, 15, 20 and 30 mol/l). With the increasing
number of micelles the visco-elasticity modulus |E| is
decreased caused by the contribution of the micelle kinetics
to the diffusion-controlled exchange of surfactants between
the surface layer and the solution bulk. The theoretical
curves were calculated from the model proposed by Joos as
given above by Eq. 8, and good agreement between the
theory [94] experimental data is observed.
Protein and polymer layers
There are many experiments performed for protein adsorption layers, using shear as well as dilational rheology. For
dilational rheology, the measured dependencies are more or
less independent of frequency as the relaxation of adsorbed
protein molecules is rather slow. As a first example in
Fig. 15, the values of the visco-elastic modulus |E| are
shown for solutions of -casein (BCS) and bovine serum
albumin (BSA) measured at a fixed oscillation frequency of
0.13 Hz using the bubble profile method [77]. The solid
lines are theoretical calculations obtained for the
corresponding model described in detail in [95]. The
agreement between the theory and experimental data in
Fig. 15 is quite good. The theory does not only predict the
values of |E| for the two proteins correctly but also
reproduces the shape of the |E|() curve: for BSA, this
dependence is monotonous and approaches a limiting

Fig. 15 Dependence of |E| on the surface pressure measured at a


frequency f=0.13 Hz for BSA (filled diamonds) and BCS (filled
triangles); the solid curves are calculated from a model discussed in
[95]; data from [99]

value, while for BCS, the curve is non-monotonous. The


maximum in the elasticity is linked with a conformational
change within the interfacial layer [96], a phenomenon first
mentioned by Graham and Phillips [97].
Another example is shown in Fig. 16, for ovalbumin at
three different water/fluid interfaces. The differences
between the three curves, as discussed in [99], are caused
by differences in the interaction enthalpy parameter. At the
air/water interface, the highest values exist, in line with the
experimental findings. The dotted lines are estimations
from a model proposed also in [98].
As mentioned above, the measured elasticities are almost
independent of the frequency of oscillations, in large
contrast to surfactant adsorption layers. An excellent proof
is shown in Fig. 17, where the elasticity module E0 is
plotted as a function of the surface pressure , determined
from measurements at different concentrations. Due to a
diffusion model for the frequency dependence, one could
expect huge differences when experiments are made in a

Fig. 16 Elasticity module E0() for ovalbumin at three interfaces at


f=0.1 Hz; water/air (filled squares), water/tetradecane (filled circles)
and water/TriAcyl-Glycerol (filled triangles); according to [98]

Colloid Polym Sci (2010) 288:937950

947

The theory given by Eq. 3 can be generalised to mixtures of


different adsorbing species. This diffusion controlled model
for the visco-elasticity was first elaborated in 1995 [66] and
independently later by Joos [67]. Its first application to
experimental data was described in [68] for mixtures of
homologous alkyl dimethyl phosphine oxides and for the
mixture of sodium dodecyl sulphate (SDS) with dodecanol.
The same general model was also adapted to mixtures of
proteins and surfactants [100]. The dilational viscoelasticity modulus as a function of frequency |E|(f) of
adsorption layers of mixed solutions of BLG/C10DMPO at
a fixed amount of BLG (10-6 mol/l) and various amounts of
surfactant is shown in Fig. 18 [101]. The elasticity modulus
of the mixture decreases significantly with increasing
C10DMPO concentration. The theoretical dependencies

(curves shown in Fig. 18) were calculated from a simplified


model earlier proposed by Joos and Garrett [102] and the
obtained agreement with the experimental data is satisfactory, in particular at low C10DMPO concentrations.
For concentrations of C10DMPO above 0.1 mmol/l, the
theoretical predictions become worse, and for a better
agreement, higher diffusion coefficients for the protein
would have to be assumed. Possibly, the application of the
more general model presented in [100] would improve the
agreement significantly.
In a recent study, mixtures of BLG with SDS were
studied at the water/air and also at the water/hexane
interfaces [103]. Although the adsorption behaviour is
rather different at the two interfaces, the general features
are comparable. However, while the surface pressure for
BLG at the water/hexane interface reaches significantly
larger values as compared to the water/air interface, the
dilational elasticities are lower.
The real and imaginary parts of the dilational viscoelastic modulus, Er and Ei, determined for mixed BCS/
C12DMPO adsorption layers in dependence of the concentration of the non-ionic surfactant C12DMPO at different
frequencies, are presented for an oscillation frequency of
0.1 Hz in Fig. 19 [86]. Both dependencies Er(c) and Ei(c)
show a well-pronounced maximum at a concentration of
104 mol/l. The curves calculated from the corresponding
theory provide a satisfactory description of the experimental data over the whole concentration range.
Mixtures of lysozyme with different surfactants were
also systematically investigated, despite the fact that
lysozyme is a protein difficult to study due to its
enormously high surface activity (adsorbs at very low bulk
concentrations). Results for mixtures with a fixed amount
of lysozyme (710-7 mol/l) and various concentrations of

Fig. 18 Dependencies of |E| on frequency f for BLG/C10DMPO


mixtures at a fixed amount of BLG (10-6 mol/l) and various C10DMPO
concentrations of filled triangles 0.02; filled circles 0.04; filled
diamonds 0.1; filled squares 0.2; ex symbols 0.4 mmol/l; theoretical
curves were calculated from a simplified model discussed in [101]

Fig. 19 Real and imaginary parts of the complex visco-elasticity, Er


and Ei, for BCS/C12DMPO mixtures in dependence of the C12DMPO
concentration at a fixed oscillation frequency of f=0.1 Hz; Er (filled
squares), Ei (filled circles); curves are theoretical calculations
according to [86]

Fig. 17 Elasticity module E0() for BLG at the water/tetradecane


interfaces and f=0.1 Hz, adsorbed from aqueous solutions of different
concentration: 3 mg/l (filled triangles), 10 mg/l (filled diamonds),
10 mg/l (filled squares), 1000 mg/l (filled circles); according to [98]

concentrations interval larger than two orders of magnitude.


However, the results collapse into one dependence.
Mixed interfacial layers

948

the added anionic surfactant SDS is shown in Fig. 20. As


one can see, there is a surfactant concentration interval, in
which the elasticity of the mixed layer is significantly
increased. The arrows point at transitions from a nonassociated state of the lysozyme (AB) to a more hydrophobic
protein/surfactant complex (BCD). Thereafter, due to
increasing hydrophobic interactions, the complex becomes
hydrophilised and the behaviour goes back to the behaviour of
the non-associated lysozyme (DE). Of course, the lysozyme
is now not really a native globular protein but surrounded by a
kind of double corona, as it was discussed in [104, 105]. The
parameters for calculating the curves in Fig. 20 were given
[105]. While the thin solid line corresponds to a model
without any complex formation, the bold solid line corresponds to complexes formed by one protein molecule and
eight SDS molecules.

Colloid Polym Sci (2010) 288:937950

Fig. 21 Real (filled circles) and imaginary (filled inverted triangles) part,
Er and Ei, of the visco-elasticity of the interfacial layer between an aqueous
solution of 1 wt.% silica particles in presence of 2104 mol/l CTAB and
1 mM NaCl, against pure hexane; curves are calculated according to [106]

Interfacial layers containing particles


Also for nano-particle dispersions, the surface and interfacial dilational rheology has been investigated by oscillating
drop and bubble methods. It is known that particles selfassemble at interfaces when they have a respective balance
between hydrophobicity and hydrophilicity (surface energy),
as it is the case for all surfactants. A quantitative understanding of the influence of such layers on the interfacial properties
is particularly important for the stabilisation of so-called
Pickering emulsions and foams. Recent studies were dedicated to colloidal dispersions of silica particles, the surface
energy of which was tuned by the adsorption of the cationic
surfactant cetyl trimethyl ammonium bromide (CTAB) [106,
107]. Measurements of the dynamic interfacial tension and
dilational visco-elasticity of dispersions at different particle
CTAB concentration ratios were performed by using DS and
CP tensiometry. The results shown in Fig. 21 demonstrate
that several relaxation processes can be observed in such

Fig. 20 Surface dilational modulus of 710-7 mol/l lysozyme/SDS


versus SDS concentration at a fixed oscillation frequency of f=0.08 Hz;
experimental data (filled squares), lines are theoretical calculations
(see text), according to [104, 105]

particle layers and a specific model for dilational viscoelasticity is needed to account for diffusional exchange as
well as for an additional relaxation process characterised by a
respective kinetic rate constants [108, 109].
The studies also demonstrated that the dilational viscoelasticity changes with the age of the interfacial layer,
becoming more and more similar to the behaviour of an
insoluble monolayer. By direct microscopic visualisation a
skin-like structure at the droplet surface is observed, in
agreement with the measured rheological properties.

Summary and conclusions


The main aim of this manuscript is to demonstrate how
sensitive dilational and shear visco-elasticities of different
types of interfacial layers are in respect to the adsorbed
molecules and their composition. It was demonstrated that
the interfacial composition of mixed protein/surfactant
systems can successfully be analysed by the dilational
rheology approach. More sensitive to structure formation
rather than surface coverage is the shear rheology which
can be used for indication whether polymer molecules are
present at an interface or not. In summary, we can state that
this methodology has reached presently a rather high
quantitative level and gives access to information which is
not accessible by other techniques.
There are, however, still a few phenomena in the
mechanical behaviour of interfacial layers not understood
yet. One is the fact that under certain conditions, negative
dilational viscosities are reported in literature, mainly
measured by capillary wave and oscillating barrier techniques. In 1995, it was reported in [110] that the capillary
wave damping was reduced for some surfactant solutions as

Colloid Polym Sci (2010) 288:937950

compared to a pure water surface. This can be misinterpreted as negative dilatational surface viscosity. The authors
analysed the experimental data and postulated a coupling
between the capillary and dilatational modes as possible
reason. Also with the oscillating barrier method, negative
phase angles were experimentally observed and discussed
in terms of negative viscosities. It is important to state here
that negative dilational viscosities are physically unreasonable and obviously experimental or theoretical artefact.
Also, any use of the terminus effective negative viscosities should be avoided as it is either induced by wrong
experimental conditions or by data interpretations using
incomplete theoretical approaches [75].
Recent publications in literature demonstrate a number
developments that provide new opportunities for the
characterisation of the mechanical behaviour of interfacial
monomolecular layers. One of these new possibilities is the
optical tweezers which allow for looking into local
properties in contrast to the classical 2D rheology which
provides an integral behaviour [111]. An overview of this
so-called micro-rheology was given by Fischer recently
[112]. There are many further ideas to approach the
mechanical properties of interfacial, and even the resistance
of small hollow sphere against deformation can be probed
by respective experimental techniques [113, 114].
Acknowledgements The work was financially supported by projects
of the DFG SPP 1273 (Mi418/16-2 and Wu187/12-1), the German Space
Agency (DLR 50WM0941), and the COST actions P21 and D43.

References
1. Ascherson FM (1840) Archiv Anat Physiol und wiss Med 44
2. Hagen GHL (1845) Abhandlungen der Kniglichen Akademie
der Wissenschaften zu Berlin (Phys Math Kl), Berlin 41
3. Plateau JAF (1869) Phil Mag Ser 4:38445
4. Marangoni C (1870) Nuovo Cim 3:50
5. Gibbs JW (1931) The collected work of JW Gibbs, vol 1.
Longmans Green, New York
6. Lord Rayleigh (1890) Proc Roy Soc (London) 47:281364
7. Boussinesq MJ (1913) Ann Chim Phys Ser 8(29):349364
8. Levich VG (1941) Acta Physicochim 14:307
9. Dorrestein R (1951) Koninkl Ned Akad Wetenschap Proc B
54:260
10. Erickson JL (1952) J Ration Mech Anal 1:521
11. Oldroyd JG (1955) Proc Roy Soc (London) A 232:567
12. Scriven LE (1960) Chem Eng Sci 12:98
13. Hansen RS (1964) J Appl Phys 35:1983
14. van den Tempel M, van de Riet RP (1965) J Chem Phys 42:2769
15. Lucassen J (1968) Trans Faraday Soc 64:2221
16. Joly M (1964) Surface viscosity. In: Danielli JF, Pankhurst KGA,
Riddiford AC (eds) Recent progress in surface science, vol 1.
Academic, New York, pp 148
17. Lucassen J (1981) In: Lucassen-Reynders EH (ed) Surfactant
science series, vol 11. Marcel Dekker, Basel, pp 217265
18. Edwards DA, Brenner H, Wasan DT (1991) Interfacial transport
processes and rheology. Butterworth-Heinemann, Boston

949
19. Miller R, Wstneck R, Krgel J, Kretzschmar G (1996) Colloids
Surf A111:75
20. Miller R, Liggieri L (eds) (2009) Progress in colloid and
interface science series: interfacial rheology, vol 1. Brill, Leiden
21. Dickinson E, Murray BS, Stainsby G (1988) J Chem Soc
Faraday Trans 1(84):871
22. Bhattacharyya A, Monroy F, Langevin D, Argillier JF (2000)
Langmuir 16:8727
23. Stubenrauch C, Miller R (2004) J Phys Chem 108:6412
24. Koelsch P, Motschmann H (2005) Langmuir 21:6265
25. Murray BS, Dickinson E, Wang Y (2009) Food Hydrocoll
23:1198
26. Krgel J, Derkatch SR, Miller R (2008) Adv Colloid Interface
Sci 144:38
27. Krgel J, Derkatch SR (2009) Interfacial shear rheologyan
overview of measuring techniques and their applications. In:
Miller R, Liggieri L (eds) Progress in colloid and interface
science: interfacial rheology, vol 1. Brill, Leiden, pp 372428
28. Krgel J, Derkatch SR (2010) Current Opinion Colloid Interface
Sci 15. doi:10.1016/j.cocis.2010.02.001
29. Krotov VV (2009) In: Miller R, Liggieri L (eds) Progress in
colloid and interface science: interfacial rheology, vol 1. Brill,
Leiden, pp 337
30. Myers RJ, Harkins WD (1937) J Chem Phys 5:601
31. Dervichian DG, Joly M (1937) Comptes rendus 2004:1318
32. Davies JT, Mayers GRA (1960) Trans Faraday Soc 56:691
33. Wasan DT, Gupta L, Vora MK (1971) AIChE J 17:1287
34. Goodrich FC, Allen LH, Poskanzer A (1975) J Colloid Interface
Sci 52:201
35. Krieg RD, Son JE, Flumerfeld RW (1981) J Colloid Interface Sci
79:14
36. Hassager O, Westborg H (1987) J Colloid Interface Sci 119:524
37. Perieditis J, Amundson NR, Flumerfelt RW (1987) J Colloid
Interface Sci 119:303
38. Langmuir I (1936) Science 84:378
39. Brown AG, Thuman WC, McBain JW (1953) J Colloid Sci 8:491
40. Krgel J, Siegel S, Miller R, Born M, Schano K-H (1994)
Colloids Surfaces A 91:169
41. Shahin GT (1986) PhD Thesis, University of Pennsylvania,
Philadelphia
42. Brooks CF, Fuller GG, Frank CW, Robertson CR (1999)
Langmuir 15:2450
43. Maestro A, Ortega F, Monroy F, Krgel J, Miller R (2009)
Langmuir 25:7393
44. Piazza L, Drr-Auster N, Gigli J, Windhab EJ, Fischer P (2009)
Food Hydrocoll 23:2125
45. Vandebril S, Franck A, Fuller GG, Moldenaers P, Vermant J
(2010) Rheol Acta 49:131
46. Noskov BA (2009) Capillary waves in interfacial rheology. In:
Miller R, Liggieri L (eds) Progress in colloid and interface
science: interfacial rheology, vol 1. Brill, Leiden, pp 103136
47. Tabor D (1980) J Colloid Interface Sci 75:240
48. Franklin B (1774) Philos Trans R Soc 64:445
49. Rayleigh L (1774) Phil Mag 30:386
50. Lamb H (1932) Hydrodynamics. Dover, New York
51. Levich VG (1940) Soviet Phys JETP 10:1296
52. Dorrestein R (1951) Proc Koninkl Ned Akad Wet B54:350
53. Lucassen J (1968) Trans Faraday Soc 64:2230
54. Lucassen-Reynders EH, Lucassen J (1969) Adv Colloid Interface Sci 2:347
55. Miller R, Fainerman VB (2004) Emulsions: structure, stability
and interactions. In: Petsev DN (ed) Interface Science and
Technology Series, vol 4. Elsevier, Amsterdam, pp 6190
56. Noskov BA (1989) Izv AN SSSR Ser Fluid Dynamics 2:105
57. Lucassen-Reynders EH, Lucassen J (1994) Colloids Surf A
85:211

950
58. Lucassen J, van den Tempel M (1972) Chem Eng Sci 27:1283
59. Lucassen J, van den Tempel M (1972) J Colloid Interface Sci
41:491
60. Noskov BA (1982) Kolloid Zh (in Russian) 44:492
61. van den Tempel M, Lucassen-Reyders EH (1983) Adv Colloid
Interface Sci 18:281
62. Lucassen J (1975) Faraday Discuss Chem Soc 59:76
63. Joos P, van Hunsel J (1988) Colloids Surf 33:99
64. Noskov BA (2002) Adv Colloid Interface Sci 95:237
65. Ivanov IB, Danov KD, Ananthapadmanabhan KP, Lips A (2005)
Adv Colloid Interface Sci 114115:61
66. Jiang Q, Valentini JE, Chiew YC (1995) J Colloid Interface Sci
174:268
67. Joos P (1999) Dynamic surface phenomena. VSP, Dordrecht
68. Aksenenko EV, Kovalchuk VI, Fainerman VB, Miller R (2007) J
Phys Chem C 111:14713
69. Kovalchuk VI, Aksenenko EV, Miller R, Fainerman VB (2009)
In: Miller R, Liggieri L (eds) Progress in colloid and interface
science: interfacial rheology, vol 1. Brill, Leiden, pp 332371
70. Warburton B (1996) Curr Opin Colloid Interface Sci 1:481486
71. Krgel J, Li JB, Miller R, Bree M, Kretzschmar G, Mhwald H
(1996) Colloid Polym Sci 274:1183
72. Tschoegl NW (1961) Kolloid-Z 181:19
73. Erni P, Fischer P, Windhab EJ (2003) Rev Sci Instrum 74:4916
74. Bonfillon A, Langevin D (1993) Langmuir 9:2172
75. Noskov BA (2010) Curr Opin Coll Interf Sci 15. doi:10.1016/j.
cocis.2010.01.006
76. Miller R, Sedev R, Schano K-H, Ng Ch, Neumann AW (1993)
Colloids Surf A 69:209
77. Benjamins J, Cagna A, Lucassen-Reynders EH (1996) Colloids
Surf A 114:245
78. Rotenberg Y, Boruvka L, Neumann AW (1983) J Colloid
Interface Sci 93:169
79. Maze C, Burnet G (1969) Surf Sci 13:451
80. Leser ME, Acquistapace S, Cagna A, Makievski AV, Miller R
(2005) Colloids Surf A 261:25
81. Kretzschmar G, Lunkenheimer K (1970) Ber Bunsenges Phys
Chem 74:1064
82. Passerone A, Liggieri L, Rando N, Ravera F, Ricci E (1991) J
Colloid Interface Sci 146:152
83. Liggieri L, Ravera F, Passerone A (1995) J Colloid Interface Sci
169:226
84. Kovalchuk VI, Zholkovskij EK, Krgel J, Miller R, Fainerman
VB, Wstneck R, Loglio G, Dukhin SS (2000) J Colloid
Interface Sci 224:245
85. Kovalchuk VI, Ravera F, Liggieri L, Loglio G, Pandolfini P,
Makievski AV, Vincent-Bonnieu S, Krgel J, Javadi A, Miller R
(2010) Adv Colloid Interface Sci. doi:10.1016/j.cis.2010.02.012
86. Kotsmar CS, Krgel J, Kovalchuk VI, Aksenenko EV, Fainerman VB, Miller R (2009) J Phys Chem B 113:103
87. Wantke K-D, Fruhner H (2001) J Colloid Interface Sci 237:185
88. Kovalchuk VI, Krgel J, Makievski AV, Ravera F, Liggieri L,
Loglio G, Fainerman VB, Miller R (2004) J Colloid Interface Sci
280:498

Colloid Polym Sci (2010) 288:937950


89. Fainerman VB, Kovalchuk VI, Aksenenko EV, Michel M, Leser
ME, Miller R (2004) J Phys Chem B 108:13700
90. Fainerman VB, Lylyk SV, Aksenenko EV, Makievski AV, Petkov
JT, Yorke J, Miller R (2009) Colloids Surf A 334:1
91. Fainerman VB, Lylyk SV, Aksenenko EV, Makievski AV,
Ravera F, Petkov JT, Yorke J, Miller R (2009) Colloids Surf
A 334:16
92. Gilnyi T, Varga I, Gilnyi M, Mszros R (2006) J Colloid
Interface Sci 301:428
93. Fainerman VB, Petkov JT, Miller R (2008) Langmuir 24:6447
94. Danov KD, Kralchevsky PA, Denkov ND, Ananthapadmanabhan KP, Lips A (2006) Adv Colloid Interface Sci 119:17
95. Lucassen-Reynders EH, Fainerman VB, Miller R (2004) J Phys
Chem B 108:9173
96. Noskov BA, Latnikova AV, Lin S-Y, Loglio G, Miller R (2007) J
Phys Chem C 111:16895
97. Graham DE, Phillips MC (1980) J Colloid Interface Sci 76:227
98. Benjamins J, Lucassen-Reynders EH (2009) Interfacial rheology
of adsorbed protein layers. In: Miller R, Liggieri L (eds) Progress
in colloid and interface science: interfacial rheology, vol 1. Brill,
Leiden, pp 253302
99. Benjamins J, Lucassen-Reynders EH (1998) In: Mbius D,
Miller R (eds) Studies in Interface Science, vol 7. Elsevier,
Amsterdam, pp 341384
100. Aksenenko EV, Kovalchuk VI, Fainerman VB, Miller R (2006)
Adv Colloid Interface Sci 122:57
101. Miller R, Leser ME, Michel M, Fainerman VB (2005) J Phys
Chem 109:13327
102. Garrett PR, Joos P (1976) J Chem Soc Faraday Trans 1(72):2161
103. Pradines V, Krgel J, Fainerman VB, Miller R (2009) J Phys
Chem B 113:745
104. Miller R, Alahverdjieva VS, Fainerman VB (2008) Soft Matter
4:1141
105. Alahverdjieva VS, Grigoriev DO, Fainerman VB, Aksenenko
EV, Miller R, Mhwald H (2008) J Phys Chem C 112:2136
106. Ravera F, Santini E, Loglio G, Ferrari M, Liggieri L (2006) J
Phys Chem B 110:19543
107. Ravera F, Ferrari M, Liggieri L, Loglio G, Santini E, Zanobini A
(2008) Colloids Surf A 323:99
108. Noskov BA, Loglio G (1998) Colloids Surf A 143:167
109. Liggieri L, Ferrari M, Ravera F (2007) Colloid stability: the role
of surface forcespart II. In: Tadros Th (ed) Colloids and
Interface Science Series, vol 2. Wiley, Hoboken, pp 313344
110. Earnshaw JC, McCoo E (1995) Langmuir 1:1087
111. Dhar P, Cao YY, Fischer TM, Zasadzinski JA (2010) Phys Rev
Lett 104:016001
112. Fischer TM (2009) Optical tweezers for 2D micro rheology.
In: Miller R, Liggieri L (eds) Progress in colloid and
interface science: interfacial rheology, vol 1. Brill, Leiden,
p 654
113. Zhang L, DAcunzi M, Kappl M, Auernhammer GK, Vollmer D,
van Kats CM, van Blaaderen A (2009) Langmuir 25:2711
114. Kluge D, Abraham F, Schmidt S, Schmidt HW, Fery A (2010)
Langmuir 26:3020

You might also like