You are on page 1of 27

Advances in Colloid and Interface Science

104 (2003) 245271

Dilational surface viscoelasticity of polymer


solutions
B.A. Noskova,*, A.V. Akentieva, A.Yu. Bilibina, I.M. Zorina,
R. Millerb
a

Chemical Department, St. Petersburg State University, Universitetsky pr. 2, 198904 St. Petersburg,
Russia
b
Kolloid- und Grenzflachenforschung,

MPI fur
Forschungcampus Golm, D14476 Golm, Germany

Abstract
A review of recent results on the dilational surface viscoelastic properties of aqueous
solutions of non-ionic polymers is given. In the frequency range from 0.001 up to 1000 Hz
the methods of transverse and longitudinal surface waves and the oscillating barrier method
were applied. Viscoelastic behavior of adsorbed polymer films significantly differs from the
behavior of films formed by only conventional surfactants of low molecular weight. For
example, the dynamic surface elasticity of the former systems is low and almost constant in
a broad concentration range. One can observe the increase of the surface elasticity only at
extremely low concentrations andyor in the range of semi-dilute solutions. If the surface
stress relaxation in conventional surfactant solutions is usually determined by the diffusional
exchange between the surface layer and the bulk phase, the relaxation processes in the
polymer systems proceed mainly inside the surface layer. Possible mechanism of the latter
relaxation is discussed.
2003 Elsevier Science B.V. All rights reserved.
Keywords: Polymer solutions; Polymer films; Dynamic surface properties; Capillary waves; Surface
viscoelasticity

1. Introduction
Surface properties of aqueous polymer solutions have been investigated for more
than half a century w1,2x. However, the progress in this field was rather slow for
*Corresponding author. Tel: q7-812-428-4093; fax: q7-812-428-6939.
E-mail address: boris@BN1664.spb.edu (B.A. Noskov).
0001-8686/03/$ - see front matter 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0001-868603.00045-9

246

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

many years, mainly due to lack of suitable experimental techniques. Traditional


methods of surface science give usually too scarce information when applied to
polymer systems. For example, the application of the Gibbs adsorption equation
usually leads to large errors when used to determine the polymer adsorption w35x.
Only recently our understanding of the structure of adsorbed polymer films has
increased significantly mainly due to the application of neutron reflection, which
allows to determine not only the adsorption but also the distribution of monomers
normal to the liquid surface w2,58x. At the same time any information on the
dynamics of polymer chains at liquid surfaces is still quite limited w9,10x.
Non-equilibrium surface properties of polymer solutions are mainly investigated
via measurements of the dynamic surface tension w2,1117x and by comparison of
the results with the solution of the corresponding boundary problem for the diffusion
equation w12,13x or with empirical equations, such as that given by Hua and Rosen
w18x. However, more recent studies have shown that surface tension can be
insensitive to the state of the surface layer and is not a good predictor of the total
amount of polymer in the surface layer w6,7,19x. Thus, for example, conformational
transitions are possible in adsorption films of poly(ethylene oxide) (PEO) and
poly(ethylene glycol) (PEG) although the surface tension is almost constant w19x.
On the other hand, the creation of a new liquid surface is usually accompanied by
non-linear hydrodynamic processes, which can influence the measured dynamic
surface tension at short times w20x. For long times natural convection and liquid
evaporation complicate the interpretation of experimental results.
The methods of mechanical relaxation spectrometry of liquid surface layers are
essentially free from these drawbacks and, in our opinion, most suitable for
investigation of the mechanical relaxation in polymer surface films. All these
methods are based on subtle perturbations of the equilibrium state and the subsequent
measurement of the system response w2022x. One usually induces small harmonical
changes of the area of the liquid surface dS and measures the resulting changes in
surface tension dg. For small perturbations linear relationships describe the system
and the surface tension perturbation is proportional to the area perturbation w22x
dgs(v)dS

(1)

The oscillations of surface tension and surface area can proceed out of phase and
the parameter (the dilational dynamic surface elasticity) is in general a complex
quantity. It can depend on the angular frequency v of the perturbations. The dynamic
surface elasticity is a fundamental surface property and can be related to both
equilibrium surface properties and kinetic coefficients of the relaxation processes
w22x. If the perturbations are small, one can substitute dA and dg by their Fourier
transforms and apply Eq. (1) to aperiodic processes too.
First studies of the dilational dynamic surface properties go back to the 1960s
w23,24x. However, the crucial influence of the complex dynamic surface elasticity
on the dynamics of various colloid systems, for example, on the stability of foams
and emulsions has been realized much later w2528x. It is rather surprising, but
until now the dynamic surface elasticity of polymer solutions, unlike the elasticity

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

247

of insoluble polymer monolayers, has been measured rather seldom. Most of the
results correspond to high frequencies ()10 kHz) where the method of surface
quasielastic light scattering (SQELS) can be applied w4,10,12,29,30x. In this
frequency range, the viscoelastic behavior of adsorbed polymer films is analogous
to insoluble monolayers, indicating that the characteristic times of the main relaxation
processes correspond to lower frequencies w4,12x.
It is noteworthy that the surface dilational viscosity Im()yv has been obtained
negative by the SQELS method for both solutions of conventional surfactants
w31,32x and polymer films w3336x. For surfactant solutions, this effect was
originally connected with an adsorption barrier w31x. However, it was shown later
that the surface viscosity must be positive for small deviations from equilibrium
regardless of the adsorption mechanism w22x. Buzza et al. recently characterized a
negative surface viscosity as an effective quantity and related it to the inadequacy
of the theoretical foundation of the SQELS method w33x. For example, the apparent
negative surface viscosity can be explained, if one assumes a sublayer of special
properties of 1 mm thickness below the interface w33x. However, the neutron
reflection method does not confirm such a thick layer for dilute polymer solutions
w3,58x. More recently this effect was explained by a missing non-linear term in
the equation of convective diffusion in the derivation of the dispersion relation for
capillary waves w32,37x. At present, two questions still remain: what is the reason
of the influence of non-linear terms in the case of small surface deformations of an
equilibrium system and what is the reason for a negative surface viscosity for
insoluble monolayers, where there is no bulk diffusion?
Although one can expect that the application of low-frequency experimental
techniques will give more information on the relaxation processes in adsorbed nonionic homopolymer films, these methods have been applied mainly to spread
polymer films w3843x. According to our knowledge only Scott and Stephens have
measured the characteristics of capillary waves for solutions of poly(ethylene glycol)
at a single concentration and a single frequency w44x and Myrvold and Hansen used
the oscillating bubble method to determine the dynamic surface elasticity of
ethylhydrooxyethyl cellulose solutions at a few concentrations in the frequency
range from 0.2 to 2 Hz w45x. Zhang and Pelton applied the oscillating drop method
to solutions of poly(N-isopropylacrylamide) (PNIPAM), but the surface deformations were large and they did not determine the dynamic surface elasticity w17x.
In this paper we present a brief review of some recent results on the viscoelasticity
of adsorbed non-ionic homopolymer films at the liquidgas interface w4649x. The
dynamic surface elasticity was measured in broad frequency (0.0011000 Hz) and
concentration (0.000011 wt.%) ranges using a combination of three different
methods of relaxation spectrometry of the surface layer.
The subsequent section will be devoted to the theory of surface dilational
viscoelasticity of linear flexible polymer solutions. After that the methods of
transverse and longitudinal surface waves, and the oscillating barrier method, will
be described. In the final section we will discuss experimental data for solutions of
non-ionic polymers of different chemical nature wPEO, PEG, PNIPAM,
poly(vinylpyrrolidone) (PVP)x.

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

248

2. Theory
The dilational surface stresses in solutions of conventional surfactants usually
relax due to the diffusional exchange between the surface layer and the bulk phase
and the dynamic surface elasticity is determined by the characteristic time of
surfactant diffusion to and from the surface w20,24,50x. The latter quantity increases
with the molecular weight of the solute. Consequently, for macromolecules one can
neglect the influence of bulk diffusion on the dynamic surface elasticity at
frequencies characteristic for surface waves (G1 Hz). In this case one can consider
the surface layer of surfactant solutions as an insoluble film. If N independent
relaxation processes proceed in the surface layer, the dynamic surface elasticity
takes the following form w22,51x
B

'

N
B g E
dg
F y8
sC
d ln S D ln S Gz is1

g Ej B g E
F yC
F
D ln S G
D ln S GAi,jj

1qivti

(2)

where ti is the isothermic relaxation time of the normal process i in the surface
layer, ji is the corresponding chemical variable (structural parameter for the
processes of the structure rearrangement), Ai is the chemical affinity. The subscript
ji means that the derivative corresponds to a non-equilibrium process (the process
i is frozen) but is taken at the equilibrium values of the thermodynamic variables.
The subscript Ai indicates the chemical equilibrium for the process i.
One can apply Eq. (2) to surface films only if the bulk phase concentration is
independent of the distance from the interface. Account of the diffusion in the bulk
phase complicates essentially the expression for the dynamic surface elasticity.
However, if we have a single surfactant and a single relaxation process in the
surface layer with the characteristic time t, for example adsorptionydesorption, the
surface elasticity takes a simple form w22,51,52x
B

g
F
D ln S Gj,c

sC

y 2Dv Gc
v G
1qivtq1qi.y
2D c
ivtq1qi.

(3)

where c is the bulk concentration, G is the adsorption, D is the bulk diffusion


coefficient.
If the intrinsic adsorption processes is fast enough and t0, Eq. (3) reduces to
the equation of Lucassen and Hansen w24x
v G
y
2D
c
g
sy
ln G
v G
1q1qi.y
2D c

1qi.

(4)

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

249

One can apply Eq. (3) and Eq. (4) to polymer solutions only in the limiting
cases of low molecular weight and extremely low frequency, which are usually out
of the scope of most experimental studies. The thermodynamic Eq. (2) is more
suitable but gives only the frequency dependence of the surface elasticity and
contains at least 2 N undetermined parameters. Therefore, more detailed investigations require a model of the surface layer of polymer solutions.
The dynamic surface properties of polymer solutions have been considered until
recently only in the framework of the dumbbell model w53,54x. This model leads to
reasonable results only for dilute solutions and, consequently, is hardly applicable
to the surface layer of solutions of surface-active macromolecules. A more general
theory for solutions of linear flexible polymers based on the Rouse model and
reptation model by de Genes w55,56x has been proposed a few years ago w57x. The
main conclusions of the theory have been corroborated for insoluble polymer films
w42,58x. We will present below only relationships relevant to adsorbed polymer
films.
Although we neglect the exchange of macromolecules between the surface layer
and the bulk phase, the exchange of monomers between different regions of the
surface layer has to be taken into account. Two main parts of the surface film are
considered: a relatively narrow concentrated region (I) contiguous to the gas phase
(the proximal region), and region (II) of tails and loops protruding into the bulk
of the liquid where the global concentration of monomers is essentially lower (the
distal region). It is essential that the width of region (II) exceeds that of region (I)
and the monomer concentration in region I is essentially higher. The second
important assumption is that the surface tension depends first of all on the
concentration in region (I) and the relaxation of surface stresses at a surface dilation
proceeds at the expense of drawing chains up to the surface wtransition of monomers
from (II) to (I)x or squeezing chains out of the surface (the reverse transition). If
we consider a linear theory of surface viscoelasticity (the case of small deformations), the departure of a polymer coil from the surface (disentanglement of the
coil) or the reverse process of entanglement of a free coil to the surface turn out to
be relatively rare events and can be neglected. Then the exchange of monomers
between the regions (I) and (II) under expansion (compression) of the surface can
take place as a result of two processes: the relaxation of inner strains of a polymer
chain or the drawing up (squeezing in) of the chain as a whole. The first process
can be described mathematically by means of a solution of the corresponding
boundary problem for the Rouse equation
B
rj
3kBT 2rj E
F
sBCfjq 2
D
t
b j2 G

(5)

Here the real polymer chain is assumed to consist of a large number of frictional
units (Rouse segments). These segments have the mobility constant B, and are
connected by harmonic springs with the mean square separation b 2 . rj is the position
of the j-th Rouse segment at time t, f j is an external force, kB is the Bolzmann
constant, and T is the temperature.

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

250

Note that Eq. (5) corresponds to ideal chains and does not take into account
excluded volume effects and hydrodynamic interactions. However, in the proximal
region, where the local concentration is high and the chains are unfolded, this
assumption can be satisfactory.
On the other hand, the high monomer concentration in the region (I) leads to
interactions between the adsorbed polymer trains. Then the motion of a train as a
whole can be considered in the framework of the two-dimensional reptation model
by de Genes w55x. The neighboring adsorbed polymer trains form a two-dimensional
network of entanglements. The real nature of the entanglements is not important. In
the two-dimensional case the entanglements can be formed, for example, by the
transitional parts of the chains where a polymer chain transits from region (I) to
region (II). Then the adsorbed train wriggles (reptates) inside a network of
entanglements, which form a two-dimensional tube. The velocity of a train with
one transitional region from (I) to (II) (a tail lying on the surface) can be

proportional to the acting force Dpa

Dpas

N 2 d L2
B d t

(6)

where a is the mean distance between the walls of the two-dimensional tube of
entanglements, is the deviation of the surface pressure from the equilibrium value,
N2 is the number of monomers in the train under consideration, L2 is the length of
the train.
These simple assumptions allow us to describe mathematically the motion of
polymer trains in the surface layer under local surface dilation. If the number of
loops and tails (in region II) per chain n is large (n41), the following expression
for the complex dynamic surface elasticity can be obtained w57x

sy

B
g w 2G3
8itBv
2G3 E ivt2 z
C
F
9 2 2
q
1y
x
|
8
ln Gm y pGm
p p 1qivtB yp2. D
pGm G 1qivt2 ~

(7)

where p changes from 1 to infinity and the sum includes only the odd terms, Gm is
the number of moles of monomers per unit surface area in region (I), G3 is the
number of moles of those monomers per unit area in region (I) that belong to trains
with two transitional regions from (I) to (II), tB is the relaxation time of inner
stresses of a polymer chain
tBs

b2N02
3p2kBTB

(8)

and the relaxation time t2 corresponds to the motion of a train in region (I) as a
whole
t 2s

N 2L 2

BG2apyG
m.

(9)

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

251

G2 is the number of moles of monomers per a unit area in region (I), which
belong to trains confined by one transitional region from (I) to (II) (one loop or
tail), N0 is the number of monomers inside the two-dimensional tube of
entanglements.
It follows from Eq. (9) that the relaxation time decreases with G2 and,
consequently, with the number of tails protruding into the bulk phase. Actually, each
additional tail creates a new path for the relaxation of stresses in region (I) of the
surface layer and decreases the resistance to this relaxation. Similarly, the increase
of the number of loops leads to the decrease of N0 and, consequently, to the decrease
of tB according to Eq. (8).
If the chains are short enough and the number of adsorbed chains with more than
one loop (train) in region (II) is small (n-1, G3<G2-Gm), the relaxation of
surface stresses at frequencies comparable with 1yt2 will mainly take place at the
expense of squeezing separate trains out of region (I) wdrawing into the region (I)x.
In this case the relaxation of local surface dilational stresses can proceed mainly at
the expense of reptation of trains with a tail in the bulk phase. Then

sy

g
ivt2
ln Gm 1qivt2

(10)

and the rheological Maxwell model describes the surface dilational viscoelasticity.
Measurements of the frequency dependency of the dynamic surface elasticity give
a possibility to determine the important characteristics of the surface layer
sylnGm and t2. The former quantity is an analog of the Gibbs elasticity of
solutions of conventional surfactants w59x and characterizes the interaction of
polymer chains in the proximal region. The relaxation time t2 can be connected
with the dynamic characteristics of macromolecules wcf. Eq. (9)x. This relationship
contains a number of parameters and cannot be applied for the determination of
molecular properties without additional information on the packing of the macromolecules in the surface layer. The calculation of the dynamic surface elasticity
from molecular parameters is probably a difficult problem. Nowadays, even reliable
estimations of the main relaxation times in adsorbed films of the most important
polymers are still lacking. For example, even the data for PEO films from different
authors are not always in agreement w11,12,38,44,46,60x. Therefore, in the remaining
part of this paper we will use Eq. (7) only for qualitative comparison with
experimental results.
It was shown also in w57x that the conventional dispersion relation that relates the
characteristics of surface waves with surface properties, holds at least for dilute
polymer solutions. For waves on liquids of finite depth it takes the following form
w61x
rv2wyrv2ygk3qrgk.tanhkh.z~q4irmv3k2q4m2v2k3wymtanhkh.ykz~
x

srqiis

mk2wyrv2ygk3qrgk.tanhkh.z~qk3gk3qrgk.
x

(11)

252

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

where ks2 pylqia is the complex wave number, a is the damping coefficient, l
is the wavelength, g is the gravitational acceleration, m 2 sk 2 yivrym, m is the
viscosity, r is the density, h is the depth of the liquid in the trough.
Eq. (11) can be considered to be a complex algebraic equation relative to k. The
two main roots of this equation correspond to transversal and longitudinal surface
waves. Both wave modes arise as a result of spontaneous thermal fluctuations at
the liquid surface or can be excited by a wave generator. The longitudinal waves
disappear if the complex dynamic surface elasticity tends to zero. The detection of
longitudinal waves is a more difficult problem in comparison with transversal waves
because they almost do not change the geometrical shape of the surface.
Note that although the specific energy dissipation in the surface layer, where the
monomer concentration is higher, exceeds the specific dissipation in the bulk
solution, the energy of waves dissipates mainly in the bulk solution because the
wavelength (the penetration depth of the motion into the bulk liquid) exceeds the
thickness of the surface layer by approximately 6 orders of magnitude. The surface
rheology influences the wave characteristics only as a result of alterations of the
boundary conditions for the equations of liquid dynamics w61x. Therefore the
damping coefficient of surface waves is a function of both components of the
dynamic surface elasticity. Even if the imaginary part of the surface elasticity is
zero (pure elastic film), the damping coefficient can be a complicated (nonmonotonous) function of the elasticity modulus.
3. Experimental methods
Most of the mentioned methods of mechanical relaxation spectrometry of liquid
surface layers relate to frequencies less than approximately 0.2 Hz. In this case one
can easily create homogenous deformations of the liquid surface in a Langmuir
trough and measure the corresponding changes of the surface tension, for example,
by using a Wilhelmy plate. The dynamic surface elasticity can be calculated directly
according to Eq. (1). Numerous authors used different modes of area variations
w21,29,42,49,6265x. In the set-up shown in Fig. 1 a mechanical generator of lowfrequency oscillations transforms the rotation of an electromotor to a translational
motion with reversion and gives the possibility to control frequency and amplitude
w49x. The moving part of the generator is connected to a Teflon barrier by a hollow
steel rod. In operation the barrier glides back and forth along the polished brims of
the Langmuir trough and produces the surface area oscillations of the liquid in the
trough. The amplitude of the barrier oscillations varies from 0.5 to 5 mm. The
corresponding surface tension oscillations are measured by the Wilhelmy plate
method using a roughened glass plate. The elasticity modulus can be determined
from the amplitude ratio of the oscillations of the surface tension and surface area,
while the phase shift between the oscillations of the two parameters (surface tension
and surface area) determines the phase angle of the dynamic surface elasticity.
At oscillation frequencies less than approximately 0.2 Hz the length of surface
longitudinal waves far exceeds the length of the Langmuir trough and the surface
tension oscillations in the trough are homogeneous. At higher frequencies the surface

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

253

Fig. 1. Schematic illustration of the experimental set-up for measurements of the dynamic surface elasticity by the oscillating barrier method.

tension oscillations become inhomogeneous and one has to apply other methods.
The number of experimental techniques applicable at frequencies exceeding 0.2 Hz
is rather limited. Except for the SQELS method discussed above, one can use only
the surface wave methods and the oscillating bubble (or drop) method w20,50,66,67x.
Note that in spite of enormous efforts to develop the latter technique over the last
years, an agreement between experimental results and the theory for simple model
systems is still unsatisfactorily for frequencies higher than 10 Hz and large surface
coverage. Therefore, only the methods of longitudinal and transverse surface waves
have been applied to polymer solutions in this frequency range.
Two different experimental set-ups have been used to measure the characteristics
of capillary waves w20,68,69x. The first one is based on the principles of electrocapillary generation and optical detection of the waves (cf. Fig. 2) w70x. Capillary
waves are excited by applying a sinusoidal voltage from an electric generator to a
thin metallic blade placed above the surface of the investigated solution. A grounded
reference platinum electrode was immersed into the solution. The distance between
the blade and the surface determines the initial amplitude of the transverse waves
and is less than a millimeter. To ensure the coincidence of the frequencies of the
waves and the electric generator a frequency divider (1:2) is included in the circuit.
The laser beam goes through a focusing optical system, is reflected from the liquid
surface and directed by an optical fiber to the position-sensitive photo detector. The
main part of the optical system is placed on a translation stage, which can be moved
via a microscrew. The alternating current appearing as result of the oscillations of
the laser beam incident on the photo detector is amplified and compared in phase
with the signal from the electric generator. Simultaneous measurements of the

254

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 2. Schematic representation of the experimental setup consisting of an electro-capillary wave generator and an optical detection system.

displacement of the point where the laser beam is reflected from the liquid surface
and the corresponding changes of the amplitude and phase of the electric signal
allow the determination of the damping coefficient and the length of the capillary
waves.

Fig. 3. Schematic representation of the electromechanical capillary wave generation and detection system.

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

255

In the second set-up a capacity wave detector is used (cf. Fig. 3) w68,69x. A
mechanical generator made from thin capillary tubes generates the surface waves.
The lower tube of the elliptical cross-section and a mean diameter of approximately
1 mm is coated by paraffin to diminish wetting. The mechanical generator oscillates
perpendicular to the liquid surface under the action of the electrodynamic vibrator,
which is connected with a generator of low-frequency sinusoidal voltage. The
dynamic condenser is used as receiving probe and formed by a metal plate and the
liquid surface. The metal plate is placed parallel to the liquid surface and
perpendicular to the direction of the wave motion. The width of the plate is less
than half of the wavelength and the oscillations of the liquid surface lead to
periodical changes of the capacity of the condenser and consequently induce the
alternating electric current in the circuit. The electric signal is amplified and
compared in phase with the current from the electric generator. The air electrode
(metallic plate) is connected with a translation stage that allows movement of the
system perpendicular and in parallel to the liquid surface. The mechanical generator
is placed on a similar translation stage and can be moved by a microscrew. The
electric circuit contains also a grounded reference platinum electrode immersed into
the liquid in the Langmuir trough. This apparatus allows determining the amplitude
and the phase of the electric signal too, and therefore, the length and the damping
coefficient of capillary waves. The wave characteristics measured by both set-ups
coincide within the error limits.
To create the longitudinal surface waves in the Langmuir trough we apply almost
the same mechanical generator as in the oscillating barrier method. The only
difference relates to the construction of the barrier. Instead of the Teflon barrier we
use a Teflon frame with a metal wire tightened inside the frame. When the frame
lays on the polished brims of the Langmuir trough the wire of the diameter 0.05
mm is wetted by the liquid and in parallel to the liquid surface. In operation the
oscillations of the wire produce the longitudinal surface waves. Their characteristics
are determined by means of transverse surface waves (cf. Fig. 4) w71x. The
transverse waves have the constant frequency of 180 Hz and are excited by the
electro-capillary method (see Fig. 2). Under the action of the strong alternating
electric field in the gap between the blade and the liquid transverse surface waves
are excited and propagate in a direction perpendicular to the propagation of
longitudinal waves. The transverse surface waves are detected by an optical method
(Fig. 4). The alternating electric current appearing as a result of the oscillations of
the laser beam incident on the photodetector is amplified and applied to a scope.
Another signal to the scope is applied from the electrical generator. A comparison
of the changes of the phase difference of these two signals at simultaneous
measurements of the displacements of the point where the laser beam is reflected
from the liquid surface allows us to determine the length of the transverse waves.
Propagation of longitudinal waves leads to oscillations of the surface concentration
and thereby to oscillations of the surface tension. In turn, this induces oscillations
of the length of transverse waves according to the dispersion Eq. (11). Moreover,
this is the only observable effect connected to the propagation of longitudinal waves

256

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 4. Schematic illustration of the longitudinal wave setup.

because the length of longitudinal waves exceeds the length of transverse waves by
more than an order of magnitude and only linear waves are used. In other words
we can ignore the non-linear interaction of surface waves. The oscillations of the
length of transverse waves are observed as oscillations of the phase of the electric
signal (relative to the phase of the signal from the electric generator). A phase
difference gauge converts the phase oscillations to a low-frequency signal with a
frequency equal to that of the longitudinal surface waves and with amplitude
proportional to the amplitude of these waves. This signal is observed on the screen
of the scope and recorded in the memory of a computer. The recording is
synchronized with the motion of the Teflon frame.
The change of the distance between the mechanical generator of longitudinal
waves and the region where the laser beam is reflected from the liquid surface leads
to corresponding changes of the amplitude and the phase of the signal applied to
the scope. This allows determining the length and the damping coefficient of
longitudinal waves by the same method as used for the measurements of the
characteristics of transverse capillary waves. Between ten and fifteen patterns of the
signal in numerical form are recorded for each position of the mechanical generator
and averaged afterwards to determine the amplitude and the phase shift relative to
oscillations of the generator by means of a non-linear regression analysis. The
derivative of the dependency of the logarithm of amplitude on the distance between
the oscillating wire and the region where the laser beam is reflected is equal to the

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

257

Fig. 5. Concentration dependency of the surface tension of solutions of PEG4 (triangles pointing down),
PEG20 (circles), PEO (open triangles pointing up), PVP10 (solid triangles pointing up), PVP55 (crosses) and PNIPAM (squares).

damping coefficient of longitudinal waves and the derivative of the phase shift over
this distance allows determining the wavelength.
PEG (Merck) of two molecular weights Ms4000 (PEG4) and 20 000 (PEG20),
PEO with Ms100 000 (Aldrich), PVP (Aldrich) of two molecular weights Ms
10 000 (PVP10) and Ms55 000 (PVP55) were used as received. PNIPAM with
the average molecular weight 300 000 was prepared using the methodology of
Schild and Tirrell w72x.
4. Results and discussion
4.1. Surface tension
Fig. 5 shows that the static surface tensions of PEO, PEG and PVP solutions do
not depend on the molecular weight in a broad concentration range and change not
more than by 1 mNym when the concentration changes by more than 3 orders of
magnitude. Obviously any estimation of the surface excess according to the Gibbs
equation will be too rough for these systems.
The variation of surface tension with concentration has to be very gradual because
of the low molar surface excess (large area per macromolecule). Since the adsorption
can only decrease with the decrease in bulk concentration, the change of the surface
tension with concentration must be even more gradual for more dilute solutions.
These considerations lead to a paradox as discussed by An et al. w5x: the g vs. c
plot must be concave with respect to c or ln c axis and it becomes impossible for
the surface tension to approach the low concentration limit where g must be close
to the value for water. In practice one usually observes not a concave but even a
convex concentration dependence of the surface tension. A possible explanation of
this effect was given by Linse and Hatton who noted that adsorption at low

258

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 6. Kinetic dependence of the surface tension of PVP10 solutions at concentrations 0.00001 wt.%
(triangles), 0.000059 wt.% (circles). Curves are the guides for an eye.

concentrations will deplete the bulk phase and thus leads to a compression of the
c-axis and consequently to a change of the shape of the g vs. c plot w73x.
However, another reason of the convex plot of g vs. c seems to be more important.
If the concentration is low enough, the surface tension depends on time and one
measures not static but a dynamic surface tension, which can significantly exceed
the static value w5x. Indeed, we observed slow changes in surface tension with time
at low concentrations (c-2=10y4 wt.%) for all the polymers. For cF3=10y5
wt.% it was impossible to obtain equilibrium values within one day. Many authors
studied the dynamic surface tension of polymer solutions at low concentrations and
usually connected the time effects with slow diffusion of the polymer coils from
the depth of the bulk phase to the interface followed by adsorption, which can be
sometimes accompanied by conformational transitions in the surface layer w11
17,74x. Fig. 6 shows, as an example, some kinetic dependencies of the surface
tension for PVP solutions. The comparison with results calculated according to the
equation of molecular diffusion shows that the adsorption is a little faster than
predicted. This can be explained by the influence of convection in the bulk phase
w49x.
Another peculiarity of the data represented in Fig. 5 consists in the more abrupt
changes of the surface tension in the range of relatively high PVP and PEG20
concentrations where the surface tension depends on the molecular weight. The
abrupt surface tension drop has been observed also by Kim and Cao in the range
of semi-dilute PEG solutions w75x and by Huang and Wang for dilute PVP solutions
w4x. Lou et al. also described similar concentration dependencies of the surface
tension for solutions of poly(vinylcaprolactam) (PV-CAP) w76x. The reason of this
effect is probably different for different polymers. Kim and Cao assumed that the
change of solvent quality in the surface layer of relatively concentrated aqueous
PEO solutions could lead to surface aggregation and to additional reduction of the

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

259

surface tension w75x. However, to the best of our knowledge there are no data
indicating similar changes of the solvent quality for PVP solutions. Besides, the
changes of the surface tension at c)0.1 wt.% for PVP solutions is accompanied by
slow time effects, which were also observed for PV-CAP w76x but not for PEG and
PEO solutions.
The data for PVP solutions in Fig. 5 represent static values only for c-0.1 wt.%.
At higher concentrations the figure shows only dynamic surface tensions at a surface
age of approximately 500 min for PVP10 and more than 1200 min for PVP55. The
attainment of equilibrium surface tension is difficult because of the required length
of time. While the interpretation of the time effects at low PVP concentrations (Fig.
6) is rather traditional, the time effects at high concentration are new findings w76x
and can hardly be explained by the same arguments. One can expect that the
adsorption rate increases with concentration and the equilibrium is reached beyond
the time accessible to experimental observation in this study (tG10 s). Indeed, we
did not observe any changes of the surface tension with time in the PVP
concentration range from 2=10y4 wt.% up to 0.1 wt.%. More probable that the
anomalous behavior of surface properties at high PVP concentrations is caused by
traces of impurities of higher surface activity. It is well known that their influence
can be overwhelming in studies of adsorption kinetics w20,7780x. To the best of
our knowledge polymer adsorption films have not been investigated yet in this
respect.
Fig. 5 shows that the equilibrium surface pressure of PVP is rather low in a
broad concentration range (approx. 3 mNym). This means that if the surface activity
of a substance is high, even small traces can be sufficient to exceed the PVP surface
pressure. In this case the impurity will displace the polymer from the surface layer.
The displacement of polymers including proteins by conventional surfactants has
been extensively studied for the last years w8184x. At low and moderate PVP
concentrations the total admixture concentration is insufficient to create a surface
pressure that exceeds 3 mNym. While the PVP surface pressure is almost independent of concentration, the impurity equilibrium surface pressure can increase with
concentration and at approximately c)0.1 wt.% for PVP10 and at c)0.5 wt.% for
PVP55 exceeds this threshold. Then PVP is gradually displaced from the surface
by a low concentrated and hence slowly adsorbing impurity. At small admixture
concentration, its adsorption rate can be extremely slow w7780x.
The proposed mechanism of slow processes observed at high PVP concentrations
has been confirmed by the method of Lunkenheimer and Miller w78x. Expansion
and subsequent compression of adsorbed surface films during different stages of the
equilibration process showed that it is hardly possible to explain all the observed
time effects if one assumes only the possibility of conformational transitions in the
surface layer without any influence of admixture desorptionadsorption w49x. The
data on the dynamic surface elasticity of PVP solutions (cf. next subsection) also
confirm this interpretation. Thus, only the data (Fig. 5) at low and medium
concentrations (c-0.1 wt.% for PVP10 and at c-0.2 wt.% for PVP55) correspond
more or less to a pure PVP adsorption film. At higher concentration an admixture
probably determines the surface properties.

260

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 7. Concentration dependency of the wavelength of capillary waves for solutions of PVP10 at the
frequencies 120 (squares), 180 (circles), 220 (triangles pointing up) and 270 (triangles pointing down).

4.2. Dilational surface viscoelasticity


Most of the results on the dynamic surface elasticity of polymer solutions have
been obtained by means of capillary waves. The changes of the wavelength of
transverse surface waves with polymer concentration have the same features at all
the frequencies. As an example, Fig. 7 shows the l(c) dependency for solutions of
PVP55 at three different frequencies. The wavelength almost does not change with
concentration except at high concentrations, similar to the corresponding dependency
of the surface tension. The similarity of the dependencies l(c) and g(c) follows
from the dispersion Eq. (11). Numerical estimations show that the wavelength is
determined first of all by the surface tension and the surface elasticity gives only a
small correction (some per cent) to l w20,31,59,69x. In contrast, the surface
dilational viscoelasticity determines the damping coefficient of capillary waves.
Unlike the surface tension, the dynamic surface properties of polymer solutions
are non-monotonic functions of concentration and depend on the molecular weight
in a broad concentration range. Fig. 8 shows, as an example, the experimental data
of the damping coefficient of transverse capillary waves for solutions of PEG and
PEO at a frequency of 200 Hz. Non-monotonic dependencies a (c) have been
observed also for PVP solutions (Fig. 9) w49x.
At a first sight, the concentration dependence for PEG4 solutions resembles the
corresponding curves for conventional surfactants where a local maximum is also
followed by a local minimum w20,24,56,69x. However, a careful inspection allows
detecting significant differences. First, the damping coefficient of the polymer
solutions changes slower with concentration. The concentration range in Fig. 8 and
Fig. 9 spans over 5 orders of magnitude while the corresponding range is
approximately 5 times shorter for solutions of conventional surfactants. Second, in
the latter case the damping coefficient in the region of the local minimum

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

261

Fig. 8. Concentration dependency of the damping coefficient of capillary waves at the frequency 200
Hz for solutions of PEG4 (triangles), PEG20 (asterisks) and PEO (squares).

corresponds to high surface elasticity (approximation of the infinite surface elasticity


in the theory of capillary waves) and is equal to approximately 3a0, where a0 is
the damping coefficient at zero surface elasticity. For polymers the obtained a
values in the region of the local minimum are essentially lower. Finally, the rate of
increase of the damping coefficient at high concentrations also exceeds the
corresponding rate for solutions of conventional surfactants.
For solutions of PEG20, PEO and PVP the decrease in damping for dilute
solutions proceeds more slowly and the local maximum disappears. One can assume
that this maximum shifts towards lower concentrations where measurements of the
surface properties are difficult because of the slow equilibration. For PEO solutions

Fig. 9. Concentration dependence of the damping coefficient of transverse capillary waves for PVP55
solutions.

262

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 10. Real part of the dynamic surface elasticity of solutions of PEG4 vs. concentration at the frequencies 120 Hz (squares), 180 Hz (circles), 220 Hz (triangles pointing up) and 270 Hz (asterisks).

the damping coefficient is almost constant. Similar results have been obtained also
for PNIPAM solutions.
Fig. 10 shows the real part of the dynamic surface elasticity of PEG4000 solutions
calculated with the dispersion Eq. (11) from the characteristics of capillary waves
at different frequencies. The imaginary part of the elasticity was close to zero at
frequencies exceeding 100 Hz. At low concentrations the real part goes though a
local maximum and gradually decreases after that up to the range of semi-dilute
solutions, where an abrupt increase of the surface elasticity accompanies the drop
of surface tension. When the polymer molecular weight decreases the local maximum
disappears and the decrease in elasticity becomes more gradual. For PEO solutions
the elasticity is almost independent of concentration. Similar results have been
obtained at other frequencies using other experimental methods and other polymers.
Fig. 11 shows, as an example, the results for PVP solutions obtained by three
different methods. The abrupt increase in elasticity at high concentrations can be
the consequence of the displacement of PVP from the surface layer by a stronger
surface-active admixture. High elasticities in this concentration range can indicate
the non-polymeric nature of the impurity. At lower concentrations PVP chains in
the surface layer determine the surface viscoelastic properties. One can see that r
gradually decreases from approximately 5 mNym to approximately 1 mNym for
both polymers when the concentration increases up to 0.01 wt.%. Fig. 12 shows an
example of the frequency dependence of both components of the surface viscoelasticity for PVP55 solutions at two concentrations. One can see that the surface
elasticity does not depend on frequency in a broad range of the spectrum. In general,
all the results obtained for PVP and PNIPAM solutions at small deviations from the
equilibrium give no distinct evidence of any relaxation processes in the frequency
range from approximately 0.001 Hz up to 1000 Hz. However, this can be only a
consequence of the low dynamic surface elasticity. Indeed in the concentration range

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

263

Fig. 11. Concentration dependence of the real (solid symbols) and imaginary (open symbols) components of the dynamic surface elasticity of PVP10 and PVP55 (only real part, asterisks) solutions at the
frequencies 0.05 Hz (squares), 0.5 Hz (triangles, crosses) and 270 Hz (circles).

from 0.0001 wt.% up to 0.1 wt.% the real part of the elasticity does not exceed 5
mNym for PVP solutions and 3 mNym for PNIPAM solutions, and the imaginary
part changes only from 0 up to 2 mNym. Thus, any frequency changes of the
elasticity could lie within the error limits, which are approximately "0.5 mNym
and "1 mNym for r and i, respectively, in the medium concentration range.
At this stage we can make some preliminary conclusions and describe the generic
features of the surface viscoelasticity, which are probably related to most of the
aqueous polymer solutions w4649x. Note that we consider only the range of medium

Fig. 12. Frequency dependence of the real (solid symbols) and imaginary (open symbols) components
of the dynamic surface elasticity of PVP55 solutions at the concentrations 0.0158 wt.% (triangles),
0.0001 wt.% (squares).

264

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

concentration from approximately 0.0001 wt.% up to 0.1 wt.%, or for PEG solutions
up to 1 wt.%. If the molecular weight is high enough (G100 000), the real
component of the elasticity is low (approx. 12 mNym) and independent of
concentration. The imaginary part is close to zero. The adsorbed film is purely
elastic at fist approximation. If the molecular weight decreases, the surface elasticity
increases at low concentrations. In other words, the dynamic surface elasticity begins
to decrease gradually with concentration. If the molecular weight is low enough
(PEG4), a local maximum appears in the concentration dependency of the surface
elasticity real part and the imaginary part deviates from zero at low frequencies.
The adsorbed film becomes viscoelastic at these concentrations and frequencies.
This behavior is entirely different from the viscoelasticity of conventional
surfactant solutions of low molecular weight when the dynamic surface elasticity
can be one or two orders of magnitude higher w71,85,86x. Although the decrease of
the dynamic surface elasticity with increasing concentration at a given perturbation
frequency is also observed for solutions of conventional surfactants w85,86x. Then,
the rate of the decrease is significantly higher, which can be easily explained by the
fast decrease of the diffusion relaxation time wthe coefficient Gyc.2 y2D in Eq.
(3) and Eq. (4)x for surfactants with concentration. When this characteristic time
approaches the period of perturbations the dynamic surface elasticity begins to drop.
Obviously this explanation cannot be applied to polymer solutions, where the
decrease of r is much slower and does not depend on frequency in a broad range.
For most of the frequencies and concentrations under investigation the period of
perturbations significantly exceeds the characteristic diffusion time.
The surface elasticity of pure water equals zero in the frequency range accessible
to experimental investigations. Therefore, the increase of dynamic surface elasticity
at dilution (cf. Fig. 10 and Fig. 11) can indicate that the elasticity of polymer
solutions goes through a local maximum at sufficiently low concentrations. The rate
of the elasticity change decreases with the increase of polymer molecular weight.
We can assume that the slope maximum of the dependency (c) is a generic
phenomenon for polymer solutions and this maximum shifts to more dilute solutions
with increasing molecular weight.
It is impossible to check up this assumption directly because at concentrations
lower than approximately 3=10y5 wt.% the equilibration process requires too much
time and only the surface elasticity of the system far from equilibrium can be
measured. However, one can examine the states of local equilibrium between the
surface and the subsurface layers when the system slowly approaches the adsorption
equilibrium. Indeed the adsorption kinetics of non-ionic polymers is controlled by
diffusion in the bulk phase w12,49x and the system meets the conditions of local
equilibrium. Every point of the dynamic surface elasticity must correspond to a
definite equilibrium subsurface concentration and the transition from (t) to (c)
requires only a (non-linear) transformation of the abscissa.
Fig. 13 and Fig. 14 show that the kinetic dependencies of the surface elasticity
for PVP and PNIPAM solutions are non-monotonical indeed. Similar results have
been obtained for several polymer concentrations and corroborate our assumption.

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

265

Fig. 13. Kinetic dependence of the real (solid symbols) and imaginary (open symbols) components of
the dynamic surface elasticity of PVP55 solutions at the concentration 0.00003 wt.%.

Moreover, in the range of local maximum of the elasticity real part the imaginary
part deviates from zero as in the case of PEG4 solutions at low frequencies w47x.
This means that the adsorbed film becomes viscoelastic there and the characteristic
time of the processes in the surface layer is comparable with the period of
oscillations. For solutions of PEG4 this allows us to determine the frequency
dependence of both dynamic surface elasticity components (Fig. 15). In this case
the polyoxyethylene chain is relatively short and we can assume that the number of
loops and tails per adsorbed chain is small (n-1). Therefore, the Maxwell
rheological model (10) can be applied to describe the frequency dependencies in

Fig. 14. Kinetic dependence of the real (solid symbols) and imaginary (open symbols) components of
the dynamic surface elasticity of PNIPAM solutions at the concentration 0.0001 wt.%.

266

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 15. Frequency dependency of the real (solid squares) and imaginary (empty squares) parts of the
dynamic surface elasticity for solutions of PEG4 at the concentration 0.001 wt.%. The lines are the best
fit to equation (11) for the real (solid line) and imaginary (dotted line) parts of the surface elasticity.

Fig. 15. The parameters gylnG and t2 have been determined by a non-linear
regression analysis and the curves in Fig. 15 represent the results of calculations
according to Eq. (10). The deviation of the real part of the surface elasticity from
the curve at low frequencies is probably caused by other relaxation processes. Note
that Eq. (10) describes only the main contribution to the surface elasticity. The
quantity sylnG, which in the present case is equal to 7.2 mNym, is an important
characteristic of the surface layer. It depends on the molecular interactions in the
surface layer and is one of the most important parameters in physicochemical
hydrodynamics of capillary systems w59x. The relaxation time (t2 s0.06 s for 0.001
wt.% PEG solutions) is related to both properties of the polymer chain a, B and the
characteristics of the chain packing in the surface layer N2, L2 wcf. Eq. (9)x and
cannot probably be estimated independently.
The obtained results give evidence that the dynamic surface elasticity of polymer
solutions is a complex function of concentration. The reason is obviously some
subtle differences in the structure of the adsorbed films at different polymer
concentrations. These differences lead to the distinctions in the relaxation times and
relaxation strengths, and consequently, to the distinctions in surface elasticity and
damping coefficient. The possibility of comparison between the results for adsorbed
w4649x and spread w48,87x films of the same polymer facilitates interpretation of
the observed non-monotonic dependencies. One can indeed obtain PEO surface
films by spreading the polymer from an organic solvent onto the aqueous surface
w38x. In this case one can obtain the dynamic surface elasticity directly as a function
of the surface excess (Fig. 16) w87x. The corresponding curve for the elasticity real
part of PEO films also has a local maximum in the range of low surface

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

267

Fig. 16. Dependence of the real (squares) and imaginary (circles) parts of the dynamic surface elasticity
on the surface concentration of PEO at the frequency 1 Hz. The solid line represents the static surface
elasticity calculated from the surface pressure isotherm.

concentrations Gs and the maximal value is close to the corresponding value for
adsorbed PEG films (Fig. 10). At even lower surface concentrations the film is
purely elastic and the polymer chains in the surface layer have a flat, almost twodimensional conformation without long loops and tails into the bulk phase. The
elasticity is determined by mutual repulsion of monomers in the surface film.
Beyond the maximum the surface elasticity deviates from static values (solid line
in Fig. 16) and drops fast, thus indicating the formation of a self-similar structure
with a large number of loops and tails protruding into the substrate w88,89x. Although
the imaginary part of the surface elasticity is close to zero, the deviation of the real
part form the static values can indicate slow relaxation processes with characteristic
times much lower than the period of applied surface oscillations. This process can
be connected with the global reconstruction of the self-similar structure w87x.
Similar conformational changes probably proceed in the adsorbed polymer films
too. At the same time there are also some differences in the viscoelastic behavior
between the adsorbed and spread films. First, the static surface elasticity of the
adsorbed films is zero and thus the dynamic surface elasticity always deviates from
the static values. Second, the imaginary part of the surface elasticity for adsorbed
films deviates from zero in the range of the local maximum unlike the case of
spread films at the same frequencies. One can assume that these differences are
connected with more gradual conformational transitions in adsorbed films. At very
low bulk concentrations the adsorbed macromolecules also have an almost flat twodimensional conformation and do not form long loops and tails (Fig. 17). Then the
film is elastic and the elasticity increases with concentration at the expense of
repulsion of neighboring monomers in the surface layer. At further increase of the
bulk concentration the surface excess increases and hinders complete unfolding of
some of the adsorbing polymer coils in the surface layer. Gradually some loops and
tails appear in the distal region of the surface layer. This leads to a relaxation of

268

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

Fig. 17. Schematic illustration of the structural changes with concentration in the surface layer of polymer
solutions.

surface stresses at the expense of monomer exchange between the proximal and
distal regions of the surface layer. The imaginary part of the surface elasticity
increases and the film becomes viscoelastic. Subsequent increase of the number of
loops and tails leads to a faster exchange between the two regions of the surface
layer and, consequently, to a decrease of both dynamic surface elasticity components;
the surface elasticity goes through a maximum. At higher concentrations, beyond
the elasticity maximum, the number of loops and tails increases. This leads to a fast
exchange of monomers between the proximal and distal regions, and the reverse
characteristic time of this process exceeds the frequencies accessible to experimental
techniques. At the same time, a slow co-operative process involving the reorganization or the whole surface layer can accompany the fast monomer exchange
between the two surface layer regions. The latter process can be responsible for the
low but non-zero values of the real part of the dynamic surface elasticity for most
of the studied systems This picture is in a qualitative agreement with the theory
described in the second section of this work. A quantitative comparison with the
theory is a future task.
5. Conclusions
The comparison of the dilational surface viscoelasticity of polymer solutions and
solutions of conventional surfactants yields significant distinctions. For polymers,
the elasticity changes more gradually and never reaches high values in the
concentration range accessible for experimental investigation. However, for some
systems (PNIPAM solutions) the surface elasticity increases up to the higher values
at the fist stages of the adsorption process corresponding to extremely low bulk
concentrations. Thereafter, it decreases again. Unlike the surface tension, the dynamic
surface elasticity of dilute polymer solutions depends on the molecular weight and
the concentration. If the molecular weight is low enough one can observe a local
maximum in the concentration dependence of the real part of the surface elasticity.
In the region of maximum the adsorbed film can be viscoelastic. With increasing
molecular weight this maximum shifts to the range of extremely low concentrations
inaccessible to experimental investigation. However, one can observe a local
maximum in the kinetic dependence of the surface elasticity. The main features of

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

269

the dilational surface viscoelasticity can be explained, at least qualitatively, by the


theory based on the Rouse equation and de Gennes reptation model.
Acknowledgments
This work was financially supported by the Russian Foundation of Fundamental
Research (Project No 03-03-32366). A.V.A. is grateful to INTAS for the fellowship
grant for young scientists (No YSF 00-166).
References
w1 x
w2 x
w3 x
w4 x
w5 x
w6 x
w7 x
w8 x
w9 x
w10x
w11x
w12x
w13x
w14x
w15x
w16x
w17x
w18x
w19x
w20x
w21x
w22x
w23x
w24x
w25x
w26x
w27x
w28x
w29x
w30x
w31x
w32x
w33x
w34x
w35x
w36x

E.L. Lovell, H. Hibbert, J. Am. Chem. Soc. 62 (1940) 2144.


A. Couper, D.D. Elley, J. Polym. Sci. 3 (1948) 345.
J.R. Lu, T.J. Su, R.K. Thomas, J. Penfold, R.W. Richards, Polymer 37 (1996) 109.
Q.R. Huang, C.H. Wang, J. Chem. Phys. 105 (1996) 6546.
S.W. An, R.K. Thomas, C. Forder, N.C. Billingham, S.P. Armes, J. Penfold Langmuir 18 (2002)
5064.
L.T. Lee, B. Jean, A. Menelle, Langmuir 15 (1999) 3267.
R.M. Richardson, R. Pelton, T. Cosgrove, J.u Zhang, Macromolecules 33 (2000) 6269.
J. Penfold, Curr. Opinion Coll. Interface Sci. 7 (2002) 139.
N. Sato, K. Sugiura, S. Ito, M. Yamamoto, Langmuir 13 (1997) 5685.
C. Booth, R.W. Richards, M.R. Taylor, G.-E. Yu, J. Phys. Chem. Part B 102 (1998) 2001.
J.E. Glass, J. Phys. Chem. 72 (1968) 4459.
B.B. Sauer, H. Yu, Macromolecules 22 (1989) 786.
I. Nahringbauer, J. Coll. Interface Sci. 176 (1995) 318.
J.u Zhang, R. Pelton, Langmuir 12 (1996) 2611.
S.-U. Um, E. Poptoschev, R. Pugh, J. Coll. Interface Sci. 193 (1997) 41.
J.u Zhang, R. Pelton, Langmuir 15 (1999) 8032.
J.u Zhang, R. Pelton, Coll. Surf. Part A 156 (1999) 111.
X.Y. Hua, M.J. Rosen, J. Coll. Interface Sci. 124 (1988) 652.
B.A. Noskov, D.A. Alexandrov, G. Loglio, R. Miller, Coll. Surf. Part A 156 (1999) 307.
B.A. Noskov, Adv. Coll. Interface Sci. 69 (1996) 63.

R. Miller, R. Wustneck,
J. Kragel,
G. Kretzschmar, Coll. Surf. Part A 55 (1996) 75.
B.A. Noskov, G. Loglio, Coll. Surf. Part A 143 (1998) 167.
F.C. Goodrich, J. Phys. Chem. 66 (1962) 1858.
J. Lucassen, R.S. Hansen, J. Coll. Interface Sci. 23 (1967) 319.
D. Langevin, Curr. Opinion Coll. Interface Sci. 3 (1998) 600.
D. Langevin, Adv. Coll. Interface Sci. 88 (2000) 209.
P.J. Wilde, Curr. Opinion Coll. Interface Sci. 5 (2000) 176.
M.A. Bos, T. van Vliet, Adv. Coll. Interface Sci. 91 (2001) 437.
E.J. McNally, J. Zografi, J.J. Coll. Interface Sci. 138 (1990) 61.
B.H. Cao, M.W. Kim, H.Z. Cummins, J. Chem. Phys. 102 (1995) 9375.
J.C. Earnshaw, E. McCoo, Langmuir 11 (1995) 1087.
F. Monroy, M.G. Munoz, J.E.F. Rubio, F. Ortega, R.G. Rubio, J. Phys. Chem. Part B 106 (2002)
5636.
D.M.A. Buzza, J.L. Jones, T.C.B. McLeish, R.W. Richards, J. Chem. Phys. 109 (1998) 5008.
S.K. Peace, R.W. Richards, N. Williams, Langmuir 14 (1998) 667.
A.J. Milling, L.H. Hutchings, R.W. Richards, Langmuir 17 (2001) 5927.
A.J. Milling, R.W. Richards, F.L. Baines, S.P. Armes, N.C. Billingham, Macromolecules 34
(2001) 4173.

270
w37x
w38x
w39x
w40x
w41x
w42x
w43x
w44x
w45x
w46x
w47x
w48x

w49x
w50x
w51x
w52x
w53x
w54x
w55x
w56x
w57x
w58x
w59x
w60x
w61x
w62x
w63x
w64x
w65x
w66x
w67x
w68x
w69x
w70x
w71x
w72x
w73x
w74x
w75x
w76x
w77x
w78x
w79x
w80x

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271
M. Hennenberg, S. Slavtchev, B. Weyssow, J.-C. Legros, J. Coll. Interface Sci. 230 (2000) 216.
R.L. Shuler, W.A. Zisman, J. Phys. Chem. 74 (1970) 1523.
K. Tamada, K. Miyano, Jpn. J. Appl. Phys. 33 (1994) 5012.
C.-S. Gau, H. Yu, G. Zografi, Macromolecules 26 (1993) 2524.
R. Skarlupka, Y. Seo, H. Yu, Polymer 39 (1998) 387.
F. Monroy, F. Ortega, R.G. Rubio, Phys. Rev. Part E 58 (1998) 7629.
F. Monroy, S. Rivillon, F. Ortega, R.G. Rubio, J. Chem. Phys. 115 (2001) 530.
J.C. Scott, R.W.B. Stephens, J. Acoust. Soc. Am. 52 (1972) 871.
R. Myrvold, F.K. Hansen, J. Coll. Interface Sci. 207 (1998) 97.
B.A. Noskov, A.V, Akentiev, G. Loglio, R. Miller, Mendeleev Commun. N2 (1998) 190.
B.A. Noskov, A.V. Akentiev, G. Loglio, R. Miller, J. Phys. Chem. Part B 104 (2000) 7923.
B.A. Noskov, A.V. Akentiev, D.A. Alexandrov, G. Loglio, R. Miller, in: E. Dickinson, R. Miller
(Eds.), Food Colloids. Fundamentals of Formulation, Royal Society of Chemistry, Cambridge,
2001, p. 191.
B.A. Noskov, A.V. Akentiev, R. Miller, J. Colloid Interface Sci. 255 (2002) 417.
B.A. Noskov, Adv. Coll. Interface Sci. 95 (2002) 237.
B.A. Noskov, Kolloid Zh. (in Russian) 44 (1982) 492.
M. van den Tempel, E.H. Lucassen-Reyders, Adv. Coll. Interface Sci. 18 (1983) 281.
J.H. Aubert, J. Coll. Interface Sci. 96 (1983) 135.
P.O. Brunn, et al., J. Coll. Interface Sci. 128 (1989) 328.
P.G. de Gennes, Scaling Concepts in Polymer Physics, London, 1979.
M. Doi, S.F. Edwards, The Theory of Polymer Dynamics, Clarendon Press, Oxford, UK, 1986.
B.A. Noskov, Coll. Polymer Sci. 273 (1995) 263.
F. Monroy, F. Ortega, R.G. Rubio, J. Phys. Chem. 103 (1998) 2061.
V.V. Krotov, A.I. Rusanov, Physicochemical hydrodynamics of capillary systems, Imperial College
Press, London, 1999.
R.W. Richards, M.R. Taylor, J. Chem. Soc. Faraday Trans. 92 (1996) 601.
Q. Jiang, Y.C. Chew, J.E. Valentini, J. Coll. Interface Sci. 159 (1993) 477.
G. Loglio, U. Tesei, R. Cini, Rev. Sci. Instrum. 59 (1988) 2045.
J.J. Kokelaar, A. Prins, M. de Gee, J. Coll. Interface Sci. 146 (1991) 507.
B.S. Murray, P.V. Nelson, Langmuir 12 (1996) 5953.
Q. Jiang, Y.C. Chiew, Coll. Surf. Part B 20 (2001) 303.

V.I. Kovalchuk, J. Kragel,


E.V. Aksenenko, G. Loglio, L. Liggieri, In: Studies in Interface Sci.

Vol.11, D. Mobius
and R. Miller (Editors), Elsevier, Amsterdam, 2000, p. 485.
K.-D. Wantke, H. Fruhner, J. Coll. Interface Sci. 237 (2001) 185.
B.A. Noskov, N.N. Kochurova, A.I. Rusanov, USSR patent N 1017999, 1983.
B.A. Noskov, A.A. Vasiliev, Kolloid Zh. (in Russian) 50 (1988) 909.
C.H. Sohl, K. Miyano, J.B. Keterson, Rev. Sci. Instrum. 49 (1978) 1464.
B.A. Noskov, D.A. Alexandrov, R. Miller, J. Coll. Interface Sci. 219 (1999) 250.
H.G. Schild, D.A. Tirrell, Langmuir 7 (1991) 665.
P. Linse, T.A. Hatton, Langmuir 13 (1997) 4066.
J.M.G. Lankveld, J. Lyklema, J. Coll. Interface Sci. 41 (1972) 454.
M.V. Kim, B.H. Cao, Europhys. Lett. 24 (1993) 229.
A. Lou, B.A. Pethica, P. Somasundaran, Langmuir 16 (2000) 7691.
K.J. Mysels, A.T. Florence, J. Coll. Interface Sci. 43 (1973) 577.
K. Lunkenheimer, R. Miller, Tenside 16 (1979) 6.
K. Lunkenheimer, R. Miller, H. Fruhner, Coll. Polymer Sci. 269 (1982) 599.
K.J. Mysels, Langmuir 2 (1986) 423.

B.A. Noskov et al. / Advances in Colloid and Interface Science 104 (2003) 245271

271

w81x D.J. Cook, J.A.K. Blondel, J. Lu, R.K. Thomas, E.A. Simister, J. Penfold Langmuir 14 (1998)
1990.
w82x B. Jean, L.-T. Lee, B. Cabane, Langmuir 15 (1999) 7590.
w83x A.R. Mackie, A.P. Gunning, P.J. Wilde, V.J. Morris, J. Coll. Interface Sci. 210 (1999) 157.
w84x R. Miller, V.B. Fainerman, A.V. Makievski, J. Kragel,

D.O. Grigoriev, V.N. Kazakov, O.V.


Sinyachenko, Adv. Coll. Interface Sci. 86 (2000) 39.
w85x B.A. Noskov, D.O. Grigoriev, R. Miller, J. Coll. Interface Sci. 188 (1997) 9.
w86x B.A. Noskov, D.O. Grigoriev, Langmuir 12 (1996) 3399.
w87x A.V. Akentiev, B.A. Noskov, Coll. J. 64 (2002) 129.
w88x C. Ligoure, J. Phys. II 3 (1993) 1607.
w89x M. Auboy, O. Guiselin, E. Raphael, Macromolecules 29 (1996) 7261.

You might also like