Вы находитесь на странице: 1из 19

Advances in Colloid and Interface Science 161 (2010) 2947

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / c i s

From surfactant adsorption kinetics to asymmetric nanomembrane mechanics:


Pendant drop experiments with subphase exchange
James K. Ferri a,, Csaba Kotsmar b, Reinhard Miller c
a
b
c

Department of Chemical and Biomolecular Engineering, Lafayette College, Easton, PA 18042, USA
Department of Chemical Engineering, The University of Texas at Austin, Austin, TX 78712, USA
Max Planck Institut fr Kolloid and Grenzchenforschung, D-14424 Golm/Potsdam, Germany

a r t i c l e

i n f o

Available online 11 August 2010


Keywords:
Adsorption
Desorption
Kinetics
Convection
Surfactant
Protein
Polymer
Macromolecule
Assembly
Interface
Structure
Nanomechanics
Elasticity
Nanomembrane
Viscoelastic
Rheology
Constitutive
Soft matter

a b s t r a c t
Adsorption equilibrium is the state in which the chemical potential of each species in the interface and bulk
is the same. Dynamic phenomena at uiduid interfaces in the presence of surface active species are often
probed by perturbing an interface or adjoining bulk phase from the equilibrium state. Many methods
designed for studying kinetics at uiduid interfaces focus on removing the system from equilibrium
through dilation or compression of the interface. This modies the surface excess concentration i and allows
the species distribution in the bulk Ci to respond. There are only a few methods available for studying uid
uid interfaces which seek to control Ci and allow the interface to respond with changes to i. Subphase
exchange in pendant drops can be achieved by the injection and withdrawal of liquid into a drop at constant
volumetric ow rate RE during which the interfacial area and drop volume VD are controlled to be
approximately constant. This can be accomplished by forming a pendant drop at the tip of two coaxial
capillary tubes. Although evolution of the subphase concentration Ci(t) is dictated by extrinsic factors such as
RE and VD, complete subphase exchange can always be attained when a sufcient amount of liquid is used.
This provides a means to tailor driving forces for adsorption and desorption in uiduid systems and in
some cases, fabricate interfacial materials of well-dened composition templated at these interfaces. The
coaxial capillary pendant drop (CCPD) method opens a wide variety of experimental possibilities.
Experiments and theoretical frameworks are reviewed for the study of surfactant exchange kinetics,
macromolecular adsorption equilibrium and dynamics, as well as the fabrication of a wide range of soft
surface materials and the characterization of their mechanics. Future directions for new experiments are also
discussed.
2010 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A new single drop methodology . . . . . . . . . . . . . . . . . . .
2.1.
Coaxial capillary pendant drop method overview . . . . . . .
2.2.
Physics of subphase exchange . . . . . . . . . . . . . . . . .
Surfactant and small amphiphile exchange: equilibrium and dynamics .
3.1.
Convection-enhanced adsorption . . . . . . . . . . . . . . . .
3.2.
Desorption kinetics and reversibility . . . . . . . . . . . . . .
3.3.
Sequential surfactant adsorption and replacement . . . . . . .
Polymer and macromolecule adsorption . . . . . . . . . . . . . . . .
4.1.
Thermodynamic equilibrium and convection-enhanced transport
4.2.
Desorption kinetics and reversibility . . . . . . . . . . . . . .
4.3.
Competitive adsorption and displacement . . . . . . . . . . .
Soft interfacial nanocomposites and nanomembranes . . . . . . . . .
5.1.
Prerequisites for interfacial nanocomposite synthesis . . . . . .
5.1.1.
Strong polyelectrolytes . . . . . . . . . . . . . . . .
5.1.2.
Weak polyelectrolytes . . . . . . . . . . . . . . . .

Corresponding author. Tel.: +1 610 330 5820; fax: +1 610 330 5059.
E-mail address: ferrij@lafayette.edu (J.K. Ferri).
0001-8686/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.cis.2010.08.002

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

30
30
30
30
32
32
33
34
35
35
36
37
38
38
38
38

30

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

5.1.3.
Protein and biomacromolecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Adsorption of non-ionic polymers to lipid monolayers . . . . . . . . . . . . . . . . . . . . . . . .
5.3.
Electrostatic assembly of asymmetric nanocomposites of strong polyelectrolytes . . . . . . . . . . .
5.4.
Polysaccharides, peptides and weak polyelectrolyte assemblies . . . . . . . . . . . . . . . . . . . .
5.5.
Proteins and nanobiomembranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.6.
Metrology of nanomembrane mechanics: elasticity and constitutive behavior . . . . . . . . . . . . .
6.
Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.
Physicochemical mechanics of surface materials: surface elastic modulii of surfactants, macromolecules,
nanoparticles and their composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.
Diffusive transport through nanocomposite surface materials . . . . . . . . . . . . . . . . . . . . .
6.3.
Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
The study of equilibrium and exchange kinetics of surfactants,
macromolecules, and their assemblies at immiscible uid and soft
interfaces as well as the rheology of such structures are subjects of
interest in both science and a wide range of technological applications. Substantial interest in particular non-equilibrium properties
comes from all elds of application dealing with foams and emulsions,
as discussed in [13]. As most practical systems are based on multicomponent adsorption layers, studies of adsorption kinetics from
mixed solutions [4], penetration of one component into an established
adsorbed layer [5], and viscoelastic phenomena in well-dened
interfacial nanocomposites are extraordinarily important.
Not many experimental techniques are suitable for the study of
equilibrium and exchange kinetics or rheology and constitutive
behavior of liquid-supported surface material assemblies. Classic
techniques to prepare well-dened interfacial composites employ the
use of a Gibbs or Langmuir monolayer at the interface of a Langmuir
trough followed by the transfer of the monolayer from one reservoir
to another via a translating barrier to exchange the subphase [6,7].
Often studies undertaken with this method are limited because of the
large subphase volume required. Additionally, highly stable surface
layers are required because of the interfacial hydrodynamic shear
which arises from the motion of the adjacent bulk during transfer.
Drop and bubble methods are more suitable for such experiments
because they allow for more stringent control of the environmental
conditions and therefore, more uniform temperature, pressure and
concentration at the interface, smaller amounts of material needed,
and a much higher interface/volume ratio than in conventional
Langmuir troughs.
Svitova et al. describe a modied pendant bubble method which
they term continuous ow tensiometry (CFT) which utilizes a
convection cell to exchange the external bulk liquid [8]. A pendant
bubble is formed in a cuvette and remains during the exchange, and
the interfacial tension is monitored by axisymmetric bubble shape
analysis. Another approach was presented by Wege et al. [911] that
consists of a coaxial double capillary which allows for in-situ internal
subphase exchange in single pendant drops.
The aim of this review is to provide an overview and examples of
the many applications of the coaxial capillary pendant drop (CCPD)
method. A summary of the method and dynamics of subphase
exchange is given in Section 2. Section 3 describes experiments and
theoretical frameworks for surfactants and small amphiphiles dealing
with desorption [12] and reversibility of adsorbed molecules [13], and
sequential adsorption of different surfactants [14]. In Section 4,
experiments concerning macromolecules and biomacromolecules as
well as experiments with interfacial mixtures of small and large
molecules are described. Protein desorption kinetics [15,16], molecular displacement, specically of proteins by surfactants [17] and the
penetration of surface active molecules into an existing surface layer
[18] are detailed. Section 5 outlines fabrication and metrology of soft

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

39
39
40
41
41
42
44

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

44
45
46
46
46

elastic interfacial nanocomposites and nanomembranes using nonspecic adsorption of polymers [19] and polyelectrolytes of opposite
charge with electrostatic layer-by-layer assembly [20] and interfacial
chemistry.
There are obviously many applications not yet realized, such as
investigations of interphase transport phenomena. These and other
more sophisticated experiments, for example studies of the effect of
solvent conditions such as pH and/or ionic strength [21], composition
of mixed solvents [22] on interfacial rheology, and molecular
transport through surface composite layers are described in Section 6.
2. A new single drop methodology
Axisymmetric drop shape analysis (ADSA) has emerged as a
powerful tool for the study of equilibrium and dynamic adsorption at
uid/uid interfaces [23]. Recently a technique was introduced in
which the drop subphase is exchanged using coaxial capillaries to
exchange the volume of the drop interior [911]. For most of the
experiments described in this review, a commercially available
pendant drop tensiometer PAT-1D (Sinterface Technologies, Berlin,
Germany) was used [24].
2.1. Coaxial capillary pendant drop method overview
In the hydrostatic version of the pendant drop method, a drop
(5 b VD b 15 L) of surfactant solution is formed at the tip of a capillary.
A silhouette of the drop is cast onto a CCD camera and digitized. The
digital images of the drop are recorded over time and t to the Young
Laplace equation to accurately (0.1 mN/m) determine surface
tension. Specially constructed tips for the pendant drop apparatus
such as the concentric capillaries described in CabrerizoVilchez et al.
[11] are used for subphase exchange experiments. A schematic is
shown in Fig. 1.
An experiment is performed as follows. A liquid drop of species
concentration C1, is formed using Microinjector 1 (M1) and allowed
to quiesce. The drop subphase is then exchanged by the injection of a
second liquid of species concentration C2, using Microinjector 2
(M2). The drop volume VD or area AD is maintained constant via
feedback control using the drop shape and the withdrawal of liquid
from the droplet interior at the same volumetric ow rate (RE) using
M1. During the exchange, the concentration of each species in the
drop, i.e. C1(t) and C2(t), evolves continuously from the initial
distribution, C1(t = 0) = C1, and C2(t = 0) = 0, to the nal distribution, C1(t = ) = 0 and C2(t = ) = C2,. The surface tension before,
during, and after the exchange can be monitored using drop prole
analysis tensiometry.
2.2. Physics of subphase exchange
The simplest description of the evolution of the subphase
concentration can be derived if the drop subphase is assumed to be

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

31

Fig. 1. Schematic of coaxial capillary pendant drop technique, as in Reference [16]. a) Experimental set-up, and b) schematic of drop subphase exchange.

perfectly mixed. For a perfectly mixed drop subphase, the concentration of species 1 is described by [12]:
C1 t = C;1 expt =

2:1

To visualize spatiotemporal distribution of subphase concentration, experiments were performed using an aqueous solution of a
low molecular weight dye (Brilliant Green, C2, = 3.3 mg/mL,

where is the residence time of the liquid in the drop = VD/RE.


The concentration of species C2, (i.e. C1, = 0)
C2 t = C;2 1 expt =

2:2

Experiments which measure the rate of exchange can be


performed using simple liquids, for example, using water and binary
mixtures of water and ethanol [12] or DMSO and chlorobenzene [25].
See Fig. 2a. In most cases the experimentally measured characteristic
exchange time constant N VD/RE as shown in Fig. 2b indicates that the
drop is not perfectly mixed.
However, the details of the dynamics of the subphase exchange are
important to equilibrium and kinetic experiments. For equilibrium it is
important to quantify the volume of uid required to achieve complete
exchange. Additionally, in order to separate interfacial exchange
kinetics from the dynamics of the subphase exchange process, the
evolution of the subphase concentration must be described as a function
of the extrinsic conditions such as capillary tip geometry, average and
local uid velocity, and subphase concentration.
Direct numerical simulation of the liquid velocity in the drop
provides details of the ow structure on a microscopic level, i.e. uid
streamlines, velocity distribution, and concentration distribution, and
on a macroscopic level, can be used to determine the overall residence
time of liquid in the drop. The spatiotemporal species distribution can
also be calculated to determine the rate at which the drop subphase
attains compositional uniformity. In general, the velocity distribution
is described by the solution of the NavierStokes equation:



v
2
v
+ v
 + v + g
  = P
t

2:3

which describes the velocity vector eld v(t,x,y,z) in terms of the


intrinsic uid properties, viscosity , density , and the pressure P.
The species continuity equation accounts for both convective and
diffusive transport of each species Ci(t,x,y,z) in the drop subphase by
Ci
2
+ v C
 i = Di Ci
t

2:4

where Di is the diffusion coefcient of the species in the liquid phase.


Solution of Eqs. (2.3) and (2.4) subject to appropriate boundary and
initial conditions yields v and Ci.

Fig. 2. Macroscopic characterization of rate of subphase exchange. a) Surface tension


() versus time for water exchange with 2% (v:v) aqueous ethanol. C1, = 0 and
C2, = 2% (v:v) ethanol for 0.2 L/s, solid line is best t to Eq. (2.1), dashed line is the
calculated evolution of bulk concentration of ethanol from Eq. (2.2). b) Time constant
vs. exchange ow rate RE as measured; symbols () and () from Reference [11] and
symbols () from Reference [22]; error bars represent 95% condence and solid line is
for perfect mixing.

32

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

D2 6 10 6 cm2/s); see Reference [26] for details. Fig. 3a shows


the evolution of dye distribution for the exchange ow rate
RE = 0.2 L/s. Direct numerical simulation of the velocity and
concentration distribution using Eqs. (2.3) and (2.4) in cylindrical
axisymmetry and appropriate boundary conditions are shown in
Fig. 3b. Further details of the simulation as well as the impact of
structure formation on the boundary condition at the airwater
interface are provided in [26].
3. Surfactant and small amphiphile exchange: equilibrium and
dynamics
The equilibrium and kinetics of surfactant and small amphiphile
exchange at uid interfaces draw a wide range of interest for scientists
and engineers; see Reviews [2731]. The simplest framework which
relates the equilibrium surface excess concentration eq,1 of small
molecules to their equilibrium bulk concentration C,1 is the Langmuir
adsorption isotherm
x1 =

eq;1
C;1

= 
;1
a1 + C;1

3:1

1
is the equilibrium adsorption constant, ,1 is the
1
maximum packing of the surfactant in a monolayer, and x1 is the
fractional coverage of the interface.
The kinetics of surfactant exchange can be described by a mass
balance at the interface [32]:

surfactant may then be calculated solving Eq. (3.2) if the time


dependence of the Cs,1(t) is known; c.f. Section 2.2 for details.
Eq. (3.2) can be non-dimensionalized to elucidate the competition
between the intrinsic physicochemical kinetics and the rate of
convective exchange of the drop subphase by scaling the surface
excess concentration by the equilibrium surface excess concentration
1 ~ eq,1, the subphase concentration by the equilibrium subphase
concentration, Cs,1 ~ C,1, and time by the convective exchange
timescale, t ~ .
In non-dimensional form, Eq. (3.2) becomes
 
h
 
 
 i
x1 1 t

= 1 k1 Cs;1
t 1x1 1 t x1 1 t

3:3

Cs;1
1
t
where the dimensionless variables 1 =
, C =
, and t = ,

eq;1 s;1
C;1
eq;1
and the x1 =
is the equilibrium fractional coverage of species 1,
;1
C;1
and the adsorption number, k1 = 1
is a scaled bulk concentration
1
of species 1. When desorption kinetics are much faster than convection,
i.e. 1 1, a local equilibrium exists between the bulk and the
interface, and the surface concentration is described by
 
t =

 
t
k1 Cs;1
 

t
1 + kCs;1

where a1 =

x1 1



1 t
= 1 Cs;1 t ;1 1 t 1 1 t
t

In this case, the surface excess concentration evolution is


determined only by the rate of convective exchange.
The equilibrium and dynamic surface tensions eq and (t)
corresponding to Langmuir adsorption are described by the surface
equation of state,

3:2

Eq. (3.2) describes the rate of change of the surface excess


concentration (t) as the difference between the adsorptive ux
which is the product of the adsorption kinetic rate constant 1, the
instantaneous subphase concentration Cs,1(t), and the space available on the interface, i.e. the difference between the maximum
surface concentration ,1 and the surface concentration 1(t). The
desorptive ux is equal to the product of the desorption kinetic
constant 1 and 1(t). The evolution of the surface concentration of

3:4

eq = o + RT ;1 ln 1x1 

3:5

and
 

 
t = o + RT ;1 ln 1x1 1 t

3:6

where o is the surface tension of the surfactant-free interface and RT


is the product of the gas constant and the temperature.
In this section, experiments studying the adsorption dynamics of
surfactants and small amphiphiles using the CCPD method are
illustrated using a homologous series of phosphine oxide surfactants
which are reasonably described by the Langmuir framework. See
Table 1 and Reference [33] for details of equilibrium data and model
constants.
3.1. Convection-enhanced adsorption
Convective exchange adsorption experiments are performed by
formation of a pendant drop of surfactant solution followed by
continuous injection of the same surfactant solution into the drop
subphase and withdrawal of uid from the drop at the same
volumetric ow rate. This convection can enhance the rate and extent
to which an apparent equilibrium state is reached. A discussion of
these two effects and representative experiments follows.

Table 1
Langmuir isotherm constants for CnDMPO surfactants.

Fig. 3. Microscopic characterization of subphase exchange dynamics: experiment and


simulation. a) Subphase exchange with Brillant Green (C 2, = 3.3 mg/mL) at
RE= 0.2 L/s, and b) direct numerical simulation of species distribution and velocity
contours using Eqs. (2.3) and (2.4) for the same experimental conditions as in a) and
constant surface tension boundary condition. See Reference [26] for details.

CnDMPO
n

a (mol/L)

(mol/m2)

8
10
12
14

4.12 104
4.39 105
5.88 106
5.85 107

3.40 106
3.64 106
4.39 106
4.44 106

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

33

Consider a semi-innite surfactant solution immediately after the


formation of an interface. Surfactant adsorbs at the interface depleting
the local concentration in the bulk and a concentration gradient from
the bulk to the interface is initiated. Often bulk diffusion is the
mechanism controlling the rate of surfactant adsorption. The
characteristic diffusion time scale D for surfactant adsorption
depends on how effective the surface depletes the bulk of surfactant
and on the diffusion coefcient of surfactant in the bulk solution.
2

h1
D1

D =

3:7

where the adsorption depth, h1, can be dened from the ratio of the
equilibrium surface concentration of a surfactant to the bulk
concentration, i.e.
h1 =

eq;1
C;1

3:8

which can be calculated from an equilibrium model such as Eq. (3.1)


[31]. Eqs. (3.7) and (3.8) show that the smaller is the concentration;
the larger is the adsorption depth and the diffusion timescale.
Therefore, experiments which study surfactants at low concentrations
can require long equilibration times. Low concentration pendant drop
experiments also present an additional difculty; surfactant adsorption can deplete the bulk concentration shifting the apparent
equilibrium. This occurs because at low concentration, the bulk
phase does not act as an innite reservoir of surfactant. This effect can
be qualitatively predicted by the depletion number, Sb,
Sb;1 =

AD eq;1
VD C;1

3:9

where AD is the total interfacial area and VD is the total volume of the
pendant drop. The depletion number is a ratio of the equilibrium mass
of surfactant adsorbed at the interface to the total amount of
surfactant available in the bulk. When Sb 1, the pendant drop can
be treated as an innite reservoir. The larger is Sb, the greater the shift
of the apparent adsorption equilibrium. Convective exchange of the
drop subphase during adsorption experiments mitigates both of these
effects.
The following experiments for the surfactant C10DMPO illustrate.
At a bulk concentration C = 1 10 4 mol/L, the adsorption depth, h, is
2.5 10 5 m. The diffusion timescale for surface tension equilibration
is about 6 s. Under typical experimental conditions (VD = 15 L,
AD = 25 mm2), the depletion number is about 5 10 2. This predicts
a rapid equilibration with respect to the timescales in the experiment,
and the depletion number suggests that the pendant drop will behave
as an innite reservoir of surfactant. These expectations are conrmed
by the data shown in Fig. 4. The surface tension is unchanging over all
observed times and agrees with the equilibrium value as predicted by
the Langmuir isotherm.
For the surfactant C 14 DMPO at a bulk concentration of
C = 6 10 6 mol/L, the adsorption depth as predicted by the constants
of the Langmuir model given in Table 1 is about 6 10 4 m suggesting
that the diffusion timescale for adsorption equilibration is on the order
of hours. The depletion number for the experiment Sb is nearly unity. In
Fig. 5, the surface tension relaxation is shown for both static and
convection-enhanced adsorption equilibrations. In the case of no
exchange (i.e. static pendant drop), bulk depletion alters the apparent
equilibrium, however continuous exchange accelerates the approach to
equilibrium and the extent to which equilibrium is reached; the long
time asymptote for the exchanged subphase is well described by the
Langmuir model.
Acceleration of adsorption kinetics can be accomplished in
situations where there is no bulk depletion effect. These situations

Fig. 4. Dynamic surface tension of C10DMPO () at a bulk concentration of


C = 1 10 4 mol/L: rapid adsorption, no bulk depletion. Solid line is the drop
interfacial area; dashed line is the equilibrium surface tension as predicted with
Langmuir model.

arise typically for macromolecules which have long timescales at low


(to moderate) concentration, cf. Section 4.1 for examples.

3.2. Desorption kinetics and reversibility


Surfactant desorption kinetics can be measured in experiments
where a pendant drop of surfactant solution is formed, and then
subsequently the drop subphase is replaced with an aqueous phase,
free of surfactant. Experiments may be designed to measure both the
rate of desorption (and therefore desorption kinetic constants) and
the extent to which desorption occurs. A discussion of theory and
experiments follows.
Consider an interface of a pendant drop having an equilibrium
distribution of the surfactant adsorbed from the adjacent bulk phase.
If the bulk concentration of the surfactant in the drop is depleted by
replacing its subphase with water, a driving force for surfactant
desorption is established. Eq. (3.3) subject to the initial condition,


1 t = 0 = 1, describes the dynamic surface concentration and can
be readily integrated via the RungeKutta method. The desorption
process is a function of the dimensionless groups, 1, which is the

Fig. 5. Dynamic surface tension of C14DMPO at a bulk concentration of C = 6 10 6 mol/L


for static () adsorption and convection exchange () experiments: adsorption depletion
and shift in apparent equilibrium. Solid line is the drop area, dashed line is the equilibrium
surface tension as predicted with the Langmuir model.

34

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

ratio of the convection and desorption timescales, the adsorption


number, k, and the fractional interfacial coverage, x1. When the
convective timescale is much smaller than the desorption timescale,
the subsurface concentration during the exchange is negligible and
the surface concentration is desorption-controlled [34], i.e.
 


1 t = exp 1 t

3:10

As desorption proceeds, the surface tension increases. The rate of


this increase can be recast in dimensionless form by scaling the
surface tension by its equilibrium lowering according to:

 
 
ln 1x1 1 t
 
t eq
= 1
t =
ln1x1
o eq

3:11

Fig. 6 shows the dimensionless time evolution of the surface


tension as a function of the ratio of convection to adsorption
timescales (1) for a xed adsorption number k1, representing the
effect of varying either the surfactant physical chemistry (i.e. the
desorption coefcient 1) or the exchange ow rate RE.
Consider the variation of surfactant physical chemistry at a xed
exchange rate. As the desorption coefcient increases, the increase in
becomes more rapid. However this effect has an upper bound. When
desorption is essentially instantaneous, the interface maintains a local
equilibrium with the bulk phase as described by Eq. (3.4); this is the
dashed curve in Fig. 6. This suggests that the study of a homologous
series of surfactants, viz. n-alkanes conjugates of the same hydrophilic
moiety, at a xed exchange rate should be able to identify the
presence of a desorption barrier and demonstrate its dependence on
hydrocarbon chain length. Finally, consider the effect of the exchange
ow rate. The greater is RE the more rapid is the exchange forcing the
bulk concentration towards zero. In this case, the solutions to
Eqs. (3.3) and (3.10) converge, i.e. the process is desorptioncontrolled and independent of exchange rate. The effect of adsorption
number on the surface tension increase for a xed surfactant physical
chemistry and exchange ow rate have also been reported.
Data is presented in Fig. 7 which shows the surface tension and
drop area versus time for C8, C10, C12, and C14DMPO at bulk
concentrations having an equilibrium surface tension of 45 mN/m.
Here, the temporal axis was rescaled so that (t = 0) occurs at the
initiation of exchange. In these experiments, the exchange rate and
drop volume for each experiment were constant so the convection
timescale is the same in each experiment so that the only adjustable

Fig. 7. Desorption kinetics of surfactants: CnDMPO. Scaled surface tension C8 (),


C10 (), C12 (), and C14 () versus time.

parameter is the desorption coefcient 1. Eq. (3.3) was integrated


using the form of C1(t) given in Eq. (2.2) with the best t value for
from the experiment. The desorption coefcients for each surfactant
were reported [12]; these values (s 1) are 4.1 10 3 (C8), 2.8 10 3
(C10), 2.1 10 3 (C12), and 5.5 10 4 (C14).
3.3. Sequential surfactant adsorption and replacement
Finite surfactant sorption kinetics during sequential subphase
exchange can give rise to non-equilibrium surface concentrations and
thereby non-equilibrium surface tensions. Consider an interface
having an equilibrium distribution of the rst surfactant (component
1) adsorbed from the adjacent bulk phase. A mass balance for each
surfactant can also be written at the interface which describes the rate
of change of the surface concentration, i, of each component as a
difference between adsorptive and desorptive kinetic uxes. The
adsorptive ux of each species is proportional to the bulk concentration, Ci, of that species and the space available on the interface; i.e. the
difference between the maximum surface concentration, , and the
total instantaneous surface concentration, 1 + 2. Assuming negligible interactions among adsorbed species, the desorptive ux of each
component is linearly proportional to its surface concentration. For a
binary system, the sorption kinetic equations are:
1 t
= 1 Cs;1 t 1 t 2 t 1 1 t
t

3:12

2 t
= 2 Cs;2 t 1 t 2 t 2 2 t
t

3:13

where i and i are the adsorption and desorption kinetic rate


constants of each component respectively. Note that in Eqs. (3.12) and
(3.13), the maximum surface concentration is assumed to be the same
for each component, which is reasonable for homologous surfactants
possessing the same polar moiety.
The evolution of the surface concentration of surfactant may then
be calculated solving Eqs. (3.12) and (3.13) together with Eqs. (2.2)
and (2.3) to describe the evolution of the bulk concentration of each
component. These may be rearranged, scaling time by the convection
timescale, ,
Fig. 6. Desorption kinetics for different surfactant physical chemistries and exchange
ow rates at a xed adsorption number (k = 10). (1) = 1.0, (2) = 0.1 and (3)
= 0.01. Desorption-control shown for ( = 0.01).

 
h
 
 
 
 i
1 t

= 1 k1 Cs;1
t 1 t 2 t 1 t
t

3:14

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

 
h 
 
 
 
 i
2 t

= 2 k2 Cs;2
t 1 t 2 t 2 t

3:15

t
where t = . Eqs. (3.14) and (3.15) show that the evolution of the

surface concentration of each species in a binary system depends on


i C;i
four dimensionless groups: the adsorption numbers ki =
which
i
are scaled bulk concentrations and the desorption coefcients i.
When these kinetic constants are sufciently large, Eqs. (3.14) and
(3.15) reduce to the LangmuirHinshelwood equilibrium adsorption
isotherms [35] for a binary system in which adsorption follows a local
equilibrium with convective exchange, as in Eq. (3.4) for a unary
system:
 
 

1 t
k1 C1 t
 
 
x1 t =
=

1 + k1 C1 t + k2 C2 t
 
 
k2 C2 t
 
2 t
 
 
=
x2 t =

1 + k1 C1 t + k2 C2 t

3:16

3:17

The surface tension evolution is dictated by the surface concentration via the equation of state which relates the dynamic surface
tension (t) to the surface concentrations 1(t) and 2(t).
 
 3
2

 
1 t
2 t

4
5

t = o + RT ln 1

3:18

where o is the surface tension of the surfactant-free interface and RT


is the product of the gas constant and the temperature.
When the adsorption number for each component is the same (i.e.
ki = const.), the equilibrium interfacial coverage and therefore the
surface tension are the same prior and subsequent to the exchange.
This same behavior is displayed when the sorption kinetic constants
of both species are identical. However, the surface tension during a
sequential exchange of surfactants with different desorption coefcients manifests a temporal dependence that depends on both
magnitude of the difference and the sequence in which the exchange
occurs. Representative experiments demonstrating minima and
maxima in the dynamic surface tension during the exchange are
shown in Fig. 8. This was demonstrated to be consistent with
interfacial over- and under-population during the exchange process
using the theory described in this section and assuming species

Fig. 8. Sequential adsorption of surfactants: C10DMPO and C14DMPO. Surface tension


() versus time.

35

concentrations of the form given in Eqs. (2.1) and (2.2). Further


details were reported earlier by Gorevski [14].
4. Polymer and macromolecule adsorption
Adsorption of proteins and their interaction with other biocompatible/biodegradable macromolecules or low molecular-weight
surfactants at liquid interfaces are of high relevance to food,
pharmaceutical and cosmetic industrial applications. Central problems associated with polymers and macromolecules at liquid
interfaces are those of adsorption and thermodynamic equilibrium,
dynamic surface tension, adsorption and desorption kinetics, and the
interactions in interfacial mixtures of surfactants and macromolecules
[3644].
4.1. Thermodynamic equilibrium and convection-enhanced transport
Proteins and macromolecules desorb from interfaces into pure
solvents very slowly [4548]. The reason is likely the high Gibbs free
energy of adsorption. One can assume that the adsorption kinetics of
high molecular weight species at surfaces is an irreversible process, in
analogy to kinetically irreversible or unidirectional chemical reactions, neglecting the reverse ux of material from the surface into the
bulk [49,50].
The adsorption of macromolecules at solid or liquid interfaces is in
many cases kinetically irreversible, i.e. the adsorption and interfacial
tension of a solution calculated from the extrapolation to innite time
depends signicantly on the conditions under which the adsorption
layer was formed, rather than solely on the macromolecule concentration. Convective transfer of the molecules from the bulk to the
solution interface can enhance the adsorption process, and can lead to
increased adsorbed amounts of the molecules in equilibrium at the
same bulk concentration caused by the shorter time needed for the
unfolding of adsorbed macromolecules [50]. Therefore, the adsorption
isotherm and surface tension isotherm for protein solutions derived
from thermodynamic or molecular statistical considerations can fail to
describe such a system.
Fainerman et al. [50] investigated the adsorption kinetics of three
different proteins, -casein, -lactoglobulin and human serum
albumin (HSA) at the at the air/water interface with and without
forced convection. (These experiments were performed using a
bubble prole analysis tensiometer and CFT similar to Svitova [8] as
shown in Fig. 9a.) The adsorption experiments were performed at
different solution concentrations with the three different proteins
both with and without forced convection. The data indicate a
signicant difference between the adsorption rates in case of all
proteins at all concentrations in the two different types of experiments. With forced convection, both the rate of surface tension
decrease and consequently the adsorption rate were approximately
one order of magnitude larger. Fig. 9b represents the differences in the
case of 10 7 mol/L BLG solutions. It was also shown that the
equilibrium surface tensions, i.e. the values extrapolated to innite
time, are independent of the adsorption rate. Fig. 9c shows that for the
two experiments with 5 10 9 mol/L -casein solutions the limiting
surface tension values are the same to within the experimental error.
Similar results were found for HSA and BLG solutions.
It can be seen that the rate of protein adsorption, even though
varied in a wide range, does not affect the equilibrium surface tension
of the solution, and consequently the adsorption. Therefore, despite
the essential irreversibility of the protein adsorption kinetics at liquid
interfaces [51], this process is thermodynamically reversible. This
behavior may be caused by the fact that the characteristic time for
conformational change is essentially shorter than the time necessary
to attain the equilibrium state in experiments with forced convection.
The adsorption equilibrium of poly(ethylene oxide)poly(propylene
oxide)poly(ethylene oxide), hereafter PEO/PPO/PEO, was studied in

36

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

Fig. 10. Dynamic surface tension of PEO/PPO/PEO. a) C1, = 6.4 107 mol/L; exchange
RE = 0.5 L/s () and diffusion only (). b) C1, = 1.6 107 mol/L; exchange
RE = 0.5 L/s and diffusion only.

equilibrium state; see Fig. 10b However, in both cases, the long time
asymptotes of the dynamic surface tension between both convective
exchange and diffusion alone are the same suggesting that a
thermodynamic equilibrium is reached in either case. Furthermore,
Fig. 10b provides an illustration of acceleration of adsorption kinetics
even when adsorption equilibrium is accessible by diffusion as discussed
in Section 3.1. Although polymers are traditionally thought to have
frozen, unresponsive structures at interfaces, these studies suggest that
the polymer is highly mobile in the adsorbed layer and readily
rearranges at the interface. Similar results have been reported
elsewhere; for additional discussion see Reference [52] and Section 5.2.

Fig. 9. a) Pendant bubble external convective exchange as described in References [8,37];


b) dependence of surface tension on time for BLG solution (concentration 10 7 mol/l)
with () and without () forced convection; RE = 25 L/s; c) surface tension as function of
inverse time for 5 10 9 mol/L -casein solutions with () and without () forced
convection; solid lines represent approximations to innite time.

the presence and absence of convective exchange [19]. At high


concentration, the interface reaches equilibrium under convective
exchange under comparable time scales to diffusion alone, see
Fig. 10a. However, at lower concentration the diffusion timescale is
much longer and convection greatly enhances the rate of approach to an

4.2. Desorption kinetics and reversibility


The study of the desorption kinetics of proteins and macromolecules can also be used to elucidate the energy of the adsorbed layer.
For example, the desorption of PEO/PPO/PEO was studied [53,54]; it
was shown that these macromolecules are kinetically irreversible on
the timescale of available experiments; see Fig. 11 for representative
data of dynamic surface tension versus time during subphase washout
using the CCPD method.
Desorption kinetics of different proteins, -casein and -lactoglobulin were studied as a function of the initial surface concentration
of protein. Adsorption layers were prepared by adsorption from

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

37

the competition between desorption and degradation at longer


timescales.
4.3. Competitive adsorption and displacement

Fig. 11. Desorption of PEO/PPO/PEO; C1, = 9.5 107 mol/L. Surface tension () and
drop area (solid line) vs. time. It is assumed to be kinetically irreversible on
experimental timescales observed.

protein solutions of different concentrations and the kinetics and


extent of desorption were studied [34]. It was found that desorption of
proteins from liquid interfaces depends on the conditions under
which they have been adsorbed. At low concentrations, the
adsorption process takes a relatively long time and the molecules
have enough space and time to adsorb and unfold at the interface. In
the case of higher bulk concentrations the adsorption is faster and the
adsorbing molecules strongly compete with each other from the
beginning of the process. Fig. 12 shows representative data for casein supporting that rate and extent of desorption for protein
solutions are a function of surface coverage. It was observed that at all
concentrations and for both proteins, only a small fraction of the
adsorbed molecules desorbs. Additionally, the surface tension change
is little, which suggests irreversible adsorption. This nding was
contradictory to other results in literature [55], therefore, additional
theoretical analyses were performed. These showed that the relative
desorption for proteins is 104108 times slower than that for usual
surfactants. Hence, experiments performed on timescales O(104 s) are
unable to discriminate between the reversibility and the irreversibility of adsorbed protein molecules; additional complications arise from

Fig. 12. Desorption of -casein with varying monolayer concentrations: comparison of


desorption kinetics from adsorption layers formed from C1, = 5 108 () 1 107
() 5 107 () mol/L; RE = 0.2 L/s.

Mixtures of proteins and low molecular weight surfactants are


applied in many practical systems due to favorable properties
achieved by the synergist effects [4]. Small surfactants adsorb rapidly
and provide interfaces low interfacial tension for emulsication or
foaming processes, and the proteins adsorb more slowly and form a
rather stable layer preventing systems from coalescence.
A number of studies on mixed proteinsurfactant systems to
which a broad spectrum of experimental characterization techniques
were applied are summarized in Reference [4]. The CCPD method
allows for a quantitative comparison of the properties of adsorption
layers fabricated via sequential (component by component) and
simultaneous adsorption (from a mixed solution).
Experimental protocol and results from mixed -casein/C12DMPO
(non-ionic surfactant) and mixed -casein/DoTAB (cationic surfactant) systems are in detail in References [56,57]. Although the
adsorption isotherms of these mixed layers obtained in the two
different ways are very similar, after subphase exchange their
equilibrium states are substantially different due to differences in
the composition. Figs. 13 and 14 present the measured dynamic
surface tensions during the bulk exchange with buffer solution
(washing-out) of a pendant drop with mixed -casein/C12DMPO
adsorption layers (increasing concentrations of C12DMPO) built up
with sequential and simultaneous adsorption, respectively. Surfactants are able to displace/replace proteins from the surface because of
their high surface activity, high concentration, and ability to modify
biomacromolecules, especially in the adsorbed state. Note that by
subphase exchange, surfactants and small amphiphiles can be
completely removed from the airwater interface due to their low
adsorption energy. Additionally, protein/surfactant complexes of low
surface activity, i.e. anchored at the interface via only a few adsorbed
segments, also desorb easily. However, free protein molecules not
displaced by the surfactant, remain in the surface layer. Increasing
concentration of C12DMPO used during the assembly of the mixed
layers leads to increasing amounts of displaced/replaced proteins,
shown by signicantly lower surface tensions after the adsorption
process and much higher values after the subsequently performed

Fig. 13. Dynamic surface tensions during the dropbulk exchange process with buffer
solution (washing-out) after simultaneous adsorption experiments with different
C12DMPO concentrations (A: 5 106 mol/L; B: 105 mol/L; C: 3 105 mol/L;
D: 4 105 mol/L; E: 5 105 mol/L; F: 8 105 mol/L) and a xed -casein
concentration of 106 mol/L.

38

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

nanocomposites with well-dened composition, architecture, and


processing conditions is possible. There are two essential prerequisites: 1) a driving force for interfacial assembly and 2) a kinetic
hindrance to disassembly; i.e. slow desorption kinetics. There also
exists an ancillary requirement; 3) the constituents of the composite
should partition relatively weakly to the solid interfaces contacted by
the exchange liquid in the experimental set-up to prevent large scale
fouling. These requirements are addressed for representative systems
of strong polyelectrolytes, weak polyelectrolytes, and protein-based
nanomembranes.

Fig. 14. Dynamic surface tensions during the dropbulk exchange process with buffer
solution (washing-out) after simultaneous adsorption experiments with different
C12DMPO concentrations (A: 5 106 mol/L; B: 105 mol/L; C: 3 105 mol/L;
D: 4 105 mol/L; E: 5 105 mol/L; F: 8 105 mol/L) and a xed -casein
concentration of 106 mol/L.

washing-out process. At the highest surfactant concentrations, the


surface tension of solution does not reach that of the pure buffer
because of the (small amount) of protein remaining at the airwater
interface. The exchange experiments demonstrate that in sequential
adsorption, proteins and surfactants interact only in the adsorption
layer. In this case, the surface tension reaches much higher values in
Fig. 13, i.e. protein displacement by surfactant was more effective as
compared to the replacement from a layer formed via simultaneous
adsorption in Fig. 14. Surface dilational rheological experiments with
mixed -casein/C12DMPO and -casein/DoTAB adsorption layers
conrmed this nding.
5. Soft interfacial nanocomposites and nanomembranes
Soft surface nanocomposites have scientic and practical applications spanning elds from electro-optical [5861] and mechanosensitive materials [6264] to biofunctional surfaces for the stimulation of cell proliferation, differentiation, and gene expression [6571].
Ultrathin surface materials can be categorized according to the
nature and strength of bonding forces by which they are formed, and
strategies ranging from physical forces such as electrostatic and
hydrophobic interactions to interfacial chemistry can be employed,
see References [72,73] for examples. In this section we discuss
prerequisites for nanomembrane fabrication using the coaxial
capillary pendant drop method and describe experiments including
adsorption to insoluble monolayers by non-ionic macromolecules,
electrostatic templating and layer-by-layer assembly of polyelectrolytes, and representative interfacial covalent crosslinking chemistry of
polysaccharides, peptides, and proteins. Particularly, we review
results for polyelectrolyte multilayer assemblies including poly(styrene sulfonate) (PSS) with poly-(allylamine hydrochloride)
(PAH) or poly-(acrylic acid) (PAA), hyaluronic acid (HA)poly(Llysine) (PLL), and brin-based nanomembranes. Strategies for metrology and materials characterization are also described. In most cases,
electrostatic complexation, hydrophobic association, or covalent crosslinking at the airwater interface leads to the formation of supramolecular networks which confers mechanical rigidity that is outside the
description provided by equilibrium surface thermodynamics; i.e. Gibbs
elasticity. A continuum framework for data analysis is described.

5.1.1. Strong polyelectrolytes


Strong polyelectrolytes possess high degree of counter-ion
dissociation and are therefore highly soluble in aqueous systems.
This satises 3) but necessitates an electrostatic template at the air
water interface. For this, a monolayer of insoluble molecules
possessing a charged headgroup must be spread on the surface of
the drop, and the lm is brought to the desired state of compression
and therefore surface charge density by varying the drop volume
using microsyringe (M1). A number of different phospholipid
monolayers [9,19,74] have been demonstrated to endure the process
under a wide range of experimental conditions, viz. monolayer lm
pressure, drop volume, and exchange ow rate see Fig. 15.
After the monolayer is deposited and brought to the desired state
of compression, the subphase of the drop is then exchanged with a
subphase containing a polyelectrolyte of opposite charge to the lipid.
Alternate cycles of polycation (PC) and polyanion (PA) and intermittent rinsing with aqueous monovalent electrolyte result in a
freestanding polymeric nanocomposite with a thickness dened by
the number of adsorbed layers; see Fig. 16.
This sequential assembly of oppositely charged strong polyelectrolytes at the airwater interface results in nanocomposites of welldened composition and transverse dimension. The free energy
changes as measured by ADSA during the adsorption cycles can be
used to provide information on the dynamics of structure formation in
the nanocomposite, as shown in Fig. 17. Further details of adsorption
dynamics are reported in Reference [20].
5.1.2. Weak polyelectrolytes
Weak polyelectrolytes have a solution chemistry-dependent
degree of dissociation; therefore under some conditions aqueous
solubility is considerably less than for strong polyelectrolytes. This
presents an experimental difculty associated with 3). Additionally,
changing solution conditions between alternate layers can result in

5.1. Prerequisites for interfacial nanocomposite synthesis


The CCPD method permits subphase exchange with a relatively
low convective disturbance to the interface, therefore fabrication of

Fig. 15. Surface pressure isotherm of l-DPPC: monolayer surface pressure versus area
per molecule; effect of surface coverage and rate of exchange.

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

Fig. 16. Schematic of freestanding asymmetric polyelectrolyte nanomembrane


fabrication at the airwater interface using lipid templating.

desorption and simplex formation as shown below in Fig. 18. The


surface tension increases concomitant with PAH (pH = 6.0) adsorption as in Fig. 17. However, after exchange with PAA (pH = 9.0), there
is a decrease in surface tension that is consistent with PAA/PAH
simplex formation and desorption. Because the surface equation of
the state of the insoluble layer is very sensitive to small impurities, the
surface tension of the interface during compression and expansion
can serve as an indicator of polymer adsorption. Fig. 19 shows the
expansion and compression of the lipid monolayer before and after
contact with the polyelectrolyte solutions. The complete reversal of
PAH adsorption underscores the notion of the reversible nature of the
adsorbed polymer layer as well as the importance of maintaining
constancy of solution chemistry, which is well known in any layer-bylayer assembly of weak polyions [75,76].
5.1.3. Protein and biomacromolecules
Protein, virus, or reactive nanoparticle-based nanomembranes
[77,78] can be synthesized by covalent network formation through
non-specic, viz. using divalent complexing agents such as Ca++,
polycondensation reactions, or biochemically specic reactions, such
as the cleavage of brinogen by thrombin to form brin bers [79].
Because most proteins have large surface afnity and low rates of
desorption, the uid interface can be used to template the transverse
connement required to produce a nanomembrane. When the
kinetics of desorption are a priori unknown, they can be measured
using methodology outlined in Section 3.2; see for example

Fig. 17. Surface tension versus time for the assembly of strong polyelectrolyte bilayer:
DMPG(PAH/PSS)1. CPAH, = 1 mg/mL, CPSS, = 1 mg/mL; RE = 0.2 L/s; pH = 6.0;
CNaCl = 0.5 mol/L. Surface tension during DMPGPAH adsorption (), NaCl wash
(), and DMPGPAHPSS adsorption (). Solid line is the drop area.

39

Fig. 18. Desorption of polyelectrolyte monolayer and simplex formation. Surface tension
versus time for (dis)assembly of weak polyelectrolyte bilayer: DMPG(PAH/PAA)1.
CPAH, = 1 mg/mL, CPAA, = 1 mg/mL; RE = 0.2 L/s; pH= 6.0 for PAH; pH= 9.0 for PAA;
CNaCl = 0.5 mol/L. Surface tension during DMPGPAH adsorption (), NaCl wash (), and
DMPGPAHPAA adsorption (). Solid line is the drop area.

representative data for brinogen in Fig. 20. In this case, adsorption


is driven by the hydrophobicity of the reactants satisfying 1); however
in some situations, condition 3) is problematic, particularly when the
subphase is not completely exchanged.

5.2. Adsorption of non-ionic polymers to lipid monolayers


Non-specic adsorption of macromolecules at soft and solid
interfaces represents a wide range of technologically important
problems ranging from immunoassay specicity [80] to reverse
osmosis membrane ltration fouling [81,82]. In the eld of surface
science, understanding the kinetics of adsorption and equilibration at
soft interfaces remains a largely unsolved problem. An insoluble lipid
monolayer spread at the airwater interface can be used as a
representative system to study both the kinetics and structure of
macromolecular adsorption at soft surfaces. One of the difculties in
studying lipidpolymer interactions by adsorption onto a monolayer
at the airwater interface is the ambiguity in the initial condition of
the mixed monolayer. Previous studies, for example as in Reference
[83], investigated the mixed monolayer by rst spreading the lipid
onto a Langmuir trough containing a subphase free of polymer and

Fig. 19. Phospholipid monolayer expansion/compression isotherm and equation of


state before and after disassembly of strong polyelectrolyte (PAH) monolayer by
sequential adsorption of (PAA). Surface tension during dilation of DMPG only () and
after DMPGPAHPAA adsorption (). Solid line is the drop area.

40

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

Fig. 20. Kinetically irreversible adsorption of brinogen at airwater interface. Surface


tension () vs. time for brinogen (C1, = 1 mg/mL; 10 mM Ca++). Desorption driven
by complete subphase exchange with pure water. Solid line is the drop area.

Fig. 21. Adsorption of the triblock polymer onto the lipid monolayer. Surface pressure
versus lipid area per molecule for l-DPPC and PEO/PPO/PEO (C1, = 2.4 105 mol/L).
Exchange rate RE = 0.2 L/s. PEO/PPO/PEO (); PEO/PPO/PEO adsorption onto l-DPPC
(). Solid line is the l-DPPC alone.

subsequently injecting a concentrated polymer solution into the


subphase below the lipid monolayer. In this case, the polymer must
diffuse to uniformity in the bulk to reach adsorption equilibrium.
Spreading the lipid on a subphase of the polymer solution is also
precluded because the preexisting adsorbed phase lowers the surface
tension and attenuates the Marangoni effect by which homogeneous
spreading of the lipid is accomplished.
The CCPD method can be used to study the equilibrium penetration
of a wide variety of polymers and copolymers into phospholipid
monolayers as a function of polymer bulk concentration. Uniform
adsorption of polymer onto the lipid is achieved by rst spreading the
insoluble monolayer onto the pendant water drop and replacing the
original subphase for a polymer solution. Compression isotherms can
be used to gain insight into the large deformation response of the
mixed system and controlled studies of the response of the interfacial
tension to periodic area perturbations can be used to obtain the
dilational rheological parameters of the interfacial layer and provides
information about the rate of exchange of matter between the surface
and the subphase and the relaxation kinetics in the monolayer.
Fig. 21 shows representative data for the adsorption of the PEO
PPOPEO copolymer onto a DPPC monolayer. Here, it can be seen that
the surface pressure of the penetrated monolayer is approximately
the same as for adsorption of the polymer only. This is due to the
relatively low surface pressure of the preexisting monolayer; the lipid
is in the liquid expanded state. However, as the lipid molecular area is
decreased by compression, the surface pressure increases in a way
that is consistent with squeeze-out of the polymer from the lipid
monolayer. This implies reversible adsorption of the polymer and
suggests that the desorption timescale for the same copolymer at the
airwater interface shown in Fig. 10 is much longer than the
experimental observation. It is also possible to measure the
viscoelasticity of the interfacial nanocomposites prepared using this
method; data for the PEO/PPO/PEODPPC system suggests some
interaction between the lipid and the copolymer in the adsorbed
layer; see Reference [19] for details.

transport, glass transition, and stress to failure, although the


mechanisms responsible for these phenomena are still not fully
understood.
Recently, the CCPD method was used to fabricate and characterize
the mechanics of strong polyelectrolyte multilayers as described in
Section 5.1.1. Fig. 22 shows the mechanical response of strong
polyelectrolyte multilayers as a function of thickness and the surface
tension response of strong polyelectrolyte nanomembranes DMPG
(PAH/PSS)n under dilation and compression. It is shown that strong
polyelectrolyte nanocomposites are likely to be elastomeric structures
that stretch semi-reversibly upon large deformation with an increasing dependence of the lm surface elastic modulus on lm thickness
and template charge density; see Reference [87] for details.
The effect of strong polyelectrolyte molecular weight and solvent
ionic strength on the elastic modulus of the nanomembrane was also
studied by Cramer and Ferri [88]. Fig. 23 shows the dependence of
surface and bulk elastic moduli on anion (13 kDa and 70 kDa)
molecular weight and salt concentration. For either molecular weight,
there is a transition corresponding to a decrease in elastic modulus
demonstrating a saloplastic effect [89]. A lowering of the yield stress
and a decrease in the surface elastic modulus were also observed for

5.3. Electrostatic assembly of asymmetric nanocomposites of strong


polyelectrolytes
Ultrathin freestanding polyelectrolyte lms have recently received
increasing attention due to their potential use as micromechanical
sensors, actuators, and barrier materials [8486]. An increasing
number of experimental results have shown that thin lms exhibit a
signicantly different behavior as compared to bulk materials, i.e.

Fig. 22. Mechanics of strong polyelectrolyte asymmetric membranes at airwater


interface. Synthesis conditions: C1, = 1 mg/mL; CNaCl = 0.5 mol/L; exchange rate
RE = 0.2 L/s. Surface tension and drop areas versus time for DMPG(PSS/PAH)n;
n = 1 (), 2 (), 3 (). Solid line is the drop area.

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

41

increases in ionic strength. The bulk elastic modulus E is related to the


surface elastic modulus Es via E = Es/h where h is the nanomembrane
thickness. Additionally, it was reported that for strong polyelectrolyte
multilayers out of plane coupling is limited to about two layers. The
magnitude of the elastic modulii found for strong polyelectrolytes is
consistent with elastomeric rubbers. For such materials, some degree
of a reversibility and viscoelastic response would be expected; both
were experimentally observed [87]. Studies of other polyelectrolyte
nanomembranes using CFT at the airwater interface are reported in
References [90,91].
5.4. Polysaccharides, peptides and weak polyelectrolyte assemblies
Strong polyelectrolytes have a high degree of dissociation and
therefore interlayer interactions are stronger; lm growth is linear in
the number of layers. Weak polyelectrolytes have a lower degree of
dissociation and weaker interlayer interactions, consequently the lm
growth is exponential in the number of layers [92]. The mechanical
properties of weak polyelectrolyte multilayers have been studied by a
variety of techniques [9396].
Differences in the mechanics of strong and weak polyelectrolyte
lms as studied by the CCPD method are manifold. Consider the data
shown in Fig. 24 which compares the surface tension response to
dilation for two bilayers (n= 2) for strong (PSS/PAH) and nanocomposites for weak polyelectrolytes hyaluronic acid (HA) and poly(L-lysine)
(PLL). The strong PE pair exhibits a linear (elastic) relationship between
stress and area for up to 10% areal dilation. The weak polyelectrolyte pair
shows plastic ow over a relatively low deformation.
The surface modulus of DMPG(HA/PLL)n as a function of lm
growth was shown in Reference [97] to be approximately constant for
(2 b n b 6) signifying no increase in the surface elastic modulus for an
increase in membrane thickness. A constant surface modulus which
accompanies increasing lm thickness signies a decreasing bulk
modulus. The decrease was also conrmed independently for (HA/
PLL)n using the colloidal probe AFM technique in Reference [67].
The mechanical strength of the membrane can be modied via
polycondensation of HA and PLL to form an interlayer amide. Fig. 25
shows surface tension as a function of deformation for DMPG(HA/
PLL)2 as assembled and after covalent crosslinking.
Fig. 25 demonstrates that the surface tension (i.e. stress response)
of the interface during dilation exceeds the limit as dened for a Gibbs
surface layer. That is, when the surface tension is a function of the
surface density of adsorbed species only, the surface stress has an
upper bound of the surface tension of the pure solvent subphase;

Fig. 23. Effect of solvent strength on the elasticity of strong polyelectrolyte


nanomembranes. Surface modulus Es (N/m) () and bulk modulus E (MPa) () of
DMPG(PSS/PAH)3 as a function of aqueous sodium chloride concentration.

Fig. 24. Surface tension (symbols) and drop area (solid line) versus time for DMPG
(PAH/PSS)2 () and DMPG(PLL/HA)2 (). Solid line is the drop area.

o = 72 mN/m for airwater. As can be seen in the surface stress


response for the crosslinked and non-crosslinked HA/PLL (n = 2), the
crosslinked membrane displays elastic behavior over a signicantly
wider range of strain and also demonstrates fracture. Additionally,
Fig. 26 shows the increase in the deviation from the YoungLaplace
equation with increasing dilation which occurs for the crosslinked
nanomembrane. This systematic deviation from the YoungLaplace
equation suggests that a fundamental assumption of the Laplace
equation is violated as deformation increases; see Section 5.6 for
discussion and alternative framework for drop shape analysis for
elastic nanomembranes.
5.5. Proteins and nanobiomembranes
Recent developments in producing biocompatible materials have
enabled numerous demonstrations that cells can be extremely
sensitive to changes in the mechanical properties of their substrates
under chemostatic conditions [98]. Proteins perform a diverse array of
tasks in living cells, including signal transduction, metabolic and
catalytic functions, and mechanical support [99]. Because mechanical
forces can lead to protein domain deformation and unfolding, this is
an important subject in molecular biomechanics.
The viscoelasticity of the protein brin is unique and remarkable
among biopolymers [79,100,101]. The surface stress response to

Fig. 25. Surface tension and drop area versus time for DMPG(PLL/HA)2 () and
DMPG(PLL/HA)2 after crosslinking () for 12 h under constant drop area.

42

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

and technology of mesoscopic materials transversely constrained to


molecular and supramolecular dimensions.
Although there have been advances in nanomaterials characterization based on compressive loading [102,103], this section is
restricted to a review of the theoretical framework for a twodimensional (i.e. ultrathin) nanomembrane of a linearly elastic
continuum of arbitrary, axisymmetric curvature under tension. See
for example Reference [104] for further details.
Consider an axisymmetric curved elastic membrane subject to an
internal pressure, for example a pendant drop. A point on the
membrane is characterized by the cylindrical coordinates (r, , z) and
the surface coordinates (, ). The undistorted state is given by
axisymmetric shape, which can be expressed as r0(z0) or 0(z0). When
subjected to a tensile load such as ination, this shape deforms. The
motion of a point during a rotationally symmetric ination can be
expressed as r(z0) and z(z0) or (z0) so that knowledge of these
functions species the shape of the membrane in the deformed state.
The surface coordinate
0 is related to the cylindrical coordinates (r0,
q
Fig. 26. Deviation from the YoungLaplace versus drop area for the deformation of
DMPG(PLL/HA)2 () and DMPG(PLL/HA)2 after crosslinking ().

ination of brin nanomembranes was investigated. Fig. 27 compares


the surface tension during deformation of a brinogen monolayer and
a brin nanomembrane at the airwater interface. As in Fig. 25, the
surface tension exceeds that of the pure solvent o which demonstrates the impact of covalent structure on interfacial stress response.
The non-linear nature of the surface stress as a function of strain and
the absence of fracture also illustrates the role of internal structure on
material response.

5.6. Metrology of nanomembrane mechanics: elasticity and constitutive


behavior
Progress in the understanding of the mechanical performance of
nanomaterials fabricated at mesoscopic size has been limited due to the
lack of suitable instrumentation and methods; particularly because the
metrology of nanomaterials at intermediate (10 4 b l b 10 2 m) length
scales is difcult to access. Therefore, the elaboration of both
experimental methodology and theoretical framework for their
interpretation is of signicant importance in advancing both the science

z0) by d0 =

r0 z0 + 1dz0 , where the prime denotes differentia-

tion with respect to z0. Similarly, in the deformed state, the surface
coordinate
0) is related to the cylindrical coordinates r and z by
q(z

dz0 = r z0 + z 2 z0 dz0 . Therefore, the stretches of the membrane in the directions of the surface coordinate lines are:
dz0
=
1 =
d0 z0

q
2
2
r z0 + z z0
q
2
r0 z0 + 1

5:1

r z0
2 =
r0 z0
where the subscript (1) denotes an extension in the surface direction
and the subscript (2) denotes an extension in the surface direction .
As the dilation of the membrane proceeds, surface stresses develop
in the membrane in response to its stretching. For isotropic materials,
the stretches and stress are collinear. In linear elasticity, the physical
components for the stress tensor T11 = T1 of the membrane are
T1 =


i
Gs h 2
2
1 1 + s 2 1
1s

5:2

where surface stress in the orthogonal direction (2) is obtained by


exchanging of indices (1) and (2).
The local force balance in the membrane requires that the
divergence of the stress in the membrane equals the jump of pressure
across it. The force balance has two tangential components and one
normal component. The normal and the tangential () components of
the force balance are
s
 
 2
1
1 r
1
r

T1 s
1
+
T
= p
2
 2
r

r
1

5:3

T1 1 r

T T1 = 0
z0 r z0 2

Fig. 27. Comparison of the interfacial stress response of native brinogen () before
and after biochemical crosslinking with thrombin () to form a brin nanomembrane.
Synthesis conditions are Cbrinogen, = 1 mg/mL; Cthrombin, = 20 units/mL; 10 mmol/L
CaCl2. Surface tension of native brinogen () and after biochemical crosslinking with
thrombin () during drop dilation. Solid line is the drop area.

and the tangential () component is identically satised. The pressure


jump [p] across the membrane arises from hydrostatics, i.e.
p  = p 0 + g zz0 .
The geometric and constitutive relationships can be substituted into
the equations of equilibrium to form a coupled system of non-linear
partial differential equations for the deformed shape of the droplet;
namely r(z0) and z(z0). For purely elastic surface materials, Eqs. (5.1)
through (5.3) form coupled two point boundary value equations which
can be converted to initial value problems by appropriate substitution.

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

Solutions of these equations yield p(z0), 1(z0), 2(z0), T1(z0), T2(z0),


and most importantly the deformed conguration r(z0) and z(z0) as a
function of the constitutive parameters, c.f. Gs and s in Eq. (5.2). Note
that as a consequence of elasticity, all dependent variables are a
function of the initial conguration, i.e. z0. Comparison of r(z) to
measured ination proles form the basis of an inverse method to
determine the constitutive parameters of the nanomembrane.
For purely viscous interfaces, the surface stress is isotropic and
independent of history. Now, the shape of the drop depends only on
the thermodynamic surface tension and the pressure jump across the
interface, i.e. r = r(z) only not z0.
In this case, the momentum balance reduces to the YoungLaplace
equation:
"
#"



 #3

2 r z
r z 2
r z 2 2
r z

1
1
+
= p
r z
z
z
z2

5:4

where is the surface tension, r(z) is the shape of the drop in the
deformed conguration, and [p] is the pressure jump across the
interface. Solution of Eq. (5.4) has been extensively documented [24].
Accordingly, there are different mechanisms of interfacial elasticity. For a liquid interface, the isotropic surface stress is a function of
surface excess concentration, []. Dilation of the interface results in a
local or global dilution of the surface excess concentration . The Gibbs
adsorption equation species that as surface excess concentration
decreases, surface tension increases. For liquid interfaces, the Gibbs

elasticity EG =
relates changes in surface tension to changes in
lnA
area, i.e. deformations which dilute the surface excess concentration
are resisted. Controlled perturbation of the liquid interface can be
used to measure the complex modulus E, which contains both
storage (surface equation of state) and dissipation (interfacial
transport and relaxation kinetics) contributions [105109].
For a solid interface, the surface stress is related to the strain
energy which describes the intrinsic capacity of the material to store
energy and the deformation history of the material. The isotropic
membrane stress = 12 T1 1 ;2 + T2 1 ;2  is the average of the
two principle stresses from Eq. (5.2).
Fluid interfaces which support ultrathin soft nanocomposites can
display liquid and solid-like mechanical responses. The total interfacial stress in surface direction 1, S1, can be written as the sum of the
thermodynamic surface tension and the membrane contribution, i.e.
S1 = + T1 1 ;2 . This approach was recently used to separate
surface tension effects from nanomembrane contributions in Pickering emulsion deformation; see Reference [77] for details.
The total isotropic surface stress measured by ADSA is the result of
both contributions as shown in Eq. (5.5).

S = + 1 ;2

43

properties such as constitutive behavior and structureproperty


relationships in soft surface materials. Comparable methods which
rely on small amplitude perturbation to extract parameters such as
elasticity, see for example Reference [24], are incapable of describing
behavior observed under large deformation; typically surface strains


Eis = 12 2i 1 b0:1. Although the strain eld of the pendant drop
ination is not strictly isotropic, this approximation signicantly
simplies data reduction; the constitutive
 2law for the extra surface
+ s
stress in linear elasticity is = Gs 11
1 or simply = K A
A0 .
s

ADSA measures the total surface stress S; the extra surface stress
can be found by assuming () = eq, or the surface tension of the
interface in the undeformed state, and subtracting.
First consider the constitutive behavior of the nanocomposite
DMPG(PAH/PSS)2 for tensile loading and unloading shown is
compared to DMPG alone in Fig. 28. In the gure, an initial region of
linear elasticity (constant K) can be seen followed by yielding and a
transition to plasticity. The dilational modulus of the lipid alone is
constant for the entire deformation. In each experiment, the drop area
is held constant after reaching an areal dilation of 60%. It can also be
noted that a stress relaxation is observable for the polyelectrolyte
nanocomposite but not for the lipid alone. By rescaling the surface
t 0
stress according to the equilibrium relaxation,
, the relaxation
eq 0
time constant for a simple Maxwell model, E = E0e( t/) can be
bounded by the experimental data; see Fig. 29. These results suggest
that more complicated expressions for linear or non-linear viscoelasticity may apply.
The recoverable energy during deformation can be calculated from
the loading and unloading curves as measured during ination and
deation. Fig. 30 shows the evolution of average extra surface stress
during loading and unloading for the DMPG(PSS/PAH)3 as a function
BC
, can be used to
of surface dilation. The plasticity index, = AC
AC

describe the fraction of the deformation that is irreversible and can


give insight into the internal structure of the nanomaterial. In this
case, the plasticity index is 0.30, indicating that 30% of the
deformation is unrecoverable. The impact of processing conditions,
for example solvent annealing, and polyelectrolyte molecular weight
on elastic modulus, yield stress, and plasticity as discussed further in
Reference [88].
The extra surface stress as a function of surface strain for a brin
nanomembrane is shown in Fig. 31. It can be noted that the strainstiffening effect gives rise to a dilational modulus of the form K = K0
exp[(/)/0]; for these data K0 4.5 mN/m and 0 0.14. The
hyperelastic material response can also be noted in this case; b 0.05.
See Reference [97] for further experimental details.

5:5

For isotropic materials, when the principle stretches are approximately equal (1 2), the principle stresses are equal; the
application of ADSA for interfacial stress measurement is reasonable.
It should be noted that because all materials are viscoelastic, the
macroscopic behavior depends on the ratio of the timescale for
internal relaxation R and the timescale of an experimental observation O known as the Deborah number, De = R/O. For De 1
macroscopic behavior is elastic, for De 1 it is viscous, and for De ~ 1
it is viscoelastic. Therefore, the difference of applicability between
Eqs. (5.1)(5.3) and Eq. (5.4) to describe the shape and state of stress
of the interface arises not from the difference between solids and
liquids, but rather the capacity of the interface to support a non-zero
deviatoric stress = S1 S2.
The CCPD method can be used to prepare liquid-supported elastic
nanomembranes (i.e. asymmetric nanocomposites), and this set-up in
conjunction with ADSA can be used as a test frame to study material

Fig. 28. Extra surface stress versus areal dilation for DMPG(PAH/PSS)2 () and DMPG
only ().

44

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

Fig. 29. Surface stress relation versus time for DMPG(PAH/PSS)2 (): experiment and
theory (simple Maxwell model) for = 10, 25, and 40 s. Surface stress during relaxation
() vs. time. Solid lines are the theory for differing characteristic relaxation times
( = 10, 25, and 40 s).

Fig. 31. Constitutive behavior for protein-based nanocomposite: hyperelasticity and


strain-stiffening in extra surface stress versus areal dilation for brin nanomesh;
h 70 nm; preparation conditions are the same as in Fig. 27. Loading () and
unloading ().

The effect of internal molecular structure on nanomembrane


stress response and constitutive behavior is apparent. Fig. 32
compares the extra surface stress and deviation from the Gauss
Laplace equation for the crosslinked linear biopolymers and the
protein-based membrane. Interestingly, the low deformation
modulus of both nanomembranes is approximately the same;
only large deformation is able to discriminate between the different
materials. Although the specic origin of the strain-stiffening effect
is unclear (either due to protein tertiary molecular structure or
superstructure of the brous mesh itself), the difference is
signicant.

using the CCPD method. A brief outline of both types of experiments


is also provided.

6. Future directions
Continuation of investigations of surfactant and protein adsorption kinetics and reversibility is relatively straightforward, however
the CCPD method provides access to a fairly wide range of systematic
investigations on the mechanics of well-dened nanocomposites
and structureproperty relationships for materials beyond those
described in Section 5. Additionally, studies of diffusion and
molecular transport in these materials should also be accessible

Fig. 30. Plasticity in the strong polyelectrolyte nanomembrane DMPG(PSS/PAH)3


under large deformation: loading (AC) and unloading (CB) hystereses in extra surface
stress versus areal dilation.

6.1. Physicochemical mechanics of surface materials: surface elastic


modulii of surfactants, macromolecules, nanoparticles and their
composites
Both sequential and co-adsorption provide means to fabricate
surface materials of well-dened composition and morphology;
Fig. 33 illustrates permutations between some of the basic building
blocks available for surface modication. In many cases, the interface
can be designed to be a kinetically irreversible structure which can
provide engineered functionality for a variety of applications, many of
which are referenced in the preceding sections. Because the subphase
can be effectively exchanged using the CCPD method, it is possible to
measure dilational modulii of interfacial nanocomposite materials
and separate intrinsic surface mechanics from the kinetics effects that
arise due to the dynamics of adsorption between the interface and the
bulk that occurs at the elevated bulk concentrations typically required

Fig. 32. Comparison of the impact of covalent intramembrane nanostructure:


nanomembrane of crosslinked linear polyelectrolyte DMPG(HA/PLL)2 () and brin
(): extra surface stress and deviation from GaussLaplace equation versus areal
dilation. Solid lines are the sum of the error between the observed shape and Gauss
Laplace equations.

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

45

Fig. 33. Surface mechanical properties of interfacial nanocomposite materials: surfactant surface equations and state and transport; polymer and soft nanocomposites; nanoparticles
and nanoparticle composites.

for fabrication. A wide variety of constitutive behavior such as those


described in Section 5 would be directly accessible.
6.2. Diffusive transport through nanocomposite surface materials
Interphase transport phenomena remain a relatively unexplored area in interface science [110112]. Fick's law is the
simplest constitutive law that describes molecular transport; it
relates the ux of a material through a continuous medium to the
concentration gradient of that material and its transport coefcient. For the transport of small solutes through the nanocomposite
materials shown in Fig. 33, the continuum approximation may be
reasonable; this requires the characteristic length scale of the
solute l(A) to be much smaller than the transverse dimension of
the interface h.
The rate of transport of A per interfacial area (or simply ux of A)
through the interface (i) in dilute limit is given by
NA = DA;i

CA
r

6:1

where NA is the ux with respect to a xed frame of reference, i.e.


the drop interfacial area S, DA,i is the diffusion coefcient of species
CA
A through the interfacial composite, and
is the concentration
r
gradient normal to the interface. The gradient can be written in
terms of the concentration in the () phase CA, the concentration
C C
A
in the () phase CA, and interface thickness, h, using C
B h A.
r
This approximation attributes all resistance to mass transfer to the
interface; however boundary layer effects in both phases may be
easily included.
For interphase transport, the driving force must be adjusted to
reect the
 difference in solute partitioning between phases;
;
CA* = f CA CA where CA*, is the () phase concentration
in
 
equilibrium with the () phase concentration and f CA is the
thermodynamic function that describes equilibrium partitioning

Fig. 34. Interphase transport through interfacial nanocomposite materials: a) solvent


permeation, b) solute permeation and interphase transport, and c) nanomembrane
permeability and dialysis.

46

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947

of species A between the () and () phases. Substitution into


Eq. (6.1) yields an expression for the permeability of the
interface p:
p=

NA
;

CA CA*

6:2

where p is the product of DA,i and h. For a given interfacial composite, p


is an intrinsic material property. Therefore, it provides additional
information concerning the internal structure of the material, as well
as itself being a technologically relevant parameter.
Fig. 34 summarizes a variety of experiments which can be
designed to assess the transport of solvents and solutes with relatively
large gas phase solubilities, interphase transport of solutes which are
mutually soluble in immiscible phases, and solutes which are soluble
only in a single (for example aqueous) phase.
6.3. Closure
Subphase exchange in pendant drops provides a means to tailor
driving forces for adsorption and desorption in uiduid systems
and in some cases, fabricate surface materials of well-dened
composition at these interfaces. Since a complete exchange can
always be attained, the CCPD method opens a wide variety of
experimental possibilities, previously not achievable with more
traditional methods, including a new framework for assessing
mechanics and transport of soft surface materials.
Acknowledgments
This work was nancially supported by projects of the DLR (50WM
0640 and 0941), the DFG SPP 1273 (Mi418/16-2), and the NSF
(CMMI) Award 0729403.
References
[1] Dickinson E, Miller R, editors. Food colloids fundamentals of formulation,
Special Publication No 258Royal Society of Chemistry; 2001.
[2] Bos MA, van Vliet T. Adv Colloid Interface Sci 2001;91:437.
[3] Fainerman VB, Leser ME, Michel M, Lucassen-Reynders EH, Miller R. J Phys Chem
B 2005;109:9672.
[4] Kotsmar C, Pradines V, Alahverdjieva VS, Aksenenko EV, Fainerman VB,
Kovalchuk VI, et al. Adv Colloid Interface Sci 2009;150:41.
[5] Zhao J, Vollhardt D, Wu J, Miller R, Siegel S, Li JB. Colloids Surf, A 2000;166:235.
[6] Vollhardt D, Wittig M. Colloids Surf 1990;47:233.
[7] Sundaram S, Ferri JK, Vollhardt D, Stebe KJ. Langmuir 1998;14:1208.
[8] Svitova TF, Wetherbee MJ, Radke CJ. J Colloid Interface Sci 2003;261:170.
[9] Wege HA, Holgado-Terriza JA, Cabrerizo-Vilchez MA. J Colloid Interface Sci
2002;249:263.
[10] Wege HA, Holgado-Terriza JA, Neumann AW, Cabrerizo-Vilchez MA. Colloids
Surf, A 1999;156:509.
[11] Cabrerizo-Vilchez MA, Wege HA, Holgado-Terriza JA, Neumann AW. Rev Sci
Instrum 1999;70:2438.
[12] Ferri JK, Gorevski N, Kotsmar C, Leser ME, Miller R. Colloids Surf, A 2008;319:13.
[13] Fainennan VB, Miller R, Ferri JK, Watzke H, Leser ME, Michel M. Adv Colloid
Interface Sci 2006;123:163.
[14] Gorevski N, Miller R, Ferri JK. Colloids Surf, A 2008;323:12.
[15] Maldonado-Valderrama J, Wege HA, Rodriguez-Valverde MA, Galvez-Ruiz MJ,
Cabrerizo-Vilchez MA. Langmuir 2003;19:8436.
[16] Maldonado-Valderrama J, Galvez-Ruiz MJ, Martin-Rodriguez A, CabrerizoVilchez MA. Langmuir 2004;20:6093.
[17] Kotsmar C, Kragel J, Kovalchuk VI, Aksenenko EV, Fainerman VB, Miller R. J Phys
Chem B 2009;113:103.
[18] Kotsmar C, Grigoriev DO, Xu F, Aksenenko EV, Fainerman VB, Leser ME, et al.
Langmuir 2008;24:13977.
[19] Ferri JK, Miller R, Makievski AV. Colloids Surf, A 2005;261:39.
[20] Ferri JK, Dong WF, Miller R. J Phys Chem B 2005;109:14764.
[21] Caro AL, Nino MRR, Patino JMR. Colloids Surf, A 2009;332:180.
[22] Li P, Xiu GH, Rodrigues AE. AlChE J 2007;53:2419.
[23] Rotenberg Y, Boruvka L, Neumann AW. J Colloid Interface Sci 1983;93:169.
[24] Loglio G, Pandolni P, Miller R, Makievski AV, Ravera F, Ferrari M, et al. Drop and
bubble shape analysis as tool for dilational rheology studies of interfacial layers.
In: Mobius D, Miller R, editors. Novel methods to study interfacial layers. Studies
in Interface ScienceAmsterdam: Elsevier; 2001. p. 439.

[25] Wege HA, Holgado-Terriza JA, Galvez-Ruiz MJ, Cabrerizo-Vilchez MA. Colloids
Surf B Biointerfaces 1999;12:339.
[26] Javadi A, Ferri JK, Karapantsiosc ThD, Miller R. Colloids and Surfaces A:
Physicochemical and Engineering Aspects 2010;365(13):14553.
[27] Noskov BA. Adv Colloid Interface Sci 1996;69:63.
[28] Chang CH, Franses EI. Colloids Surf, A 1995;100:1.
[29] Langevin D. Curr Opin Colloid Interface Sci 1998;3:600.
[30] Eastoe J, Dalton JS. Adv Colloid Interface Sci 2000;85:103.
[31] Ferri JK, Stebe KJ. Adv Colloid Interface Sci 2000;85:61.
[32] Baret JF. J Colloid Interface Sci 1969;30:1.
[33] Aksenenko EV, Makievski AV, Miller R, Fainerman VB. Colloids Surf, A 1998;143:
311.
[34] Miller R, Grigoriev DO, Kragel J, Makievski A, Maldonado-Valderrama J, Leser M,
et al. Food Hydrocolloids 2005;19:479.
[35] Chang CH, Wang NHL, Franses EI. Colloids Surf 1992;62:321.
[36] Fainerman VB, Lucassen-Reynders EH, Miller R. Adv Colloid Interface Sci
2003;106:237.
[37] Goddard ED. J Colloid Interface Sci 2002;256:228.
[38] Miller R, Fainerman VB, Makievski AV, Kragel J, Grigoriev DO, Kazakov VN, et al.
Adv Colloid Interface Sci 2000;86:39.
[39] Dickinson E. Colloids Surf B Biointerfaces 1999;15:161.
[40] Dickinson E. J Chem Soc, Faraday Trans 1998;94:1657.
[41] Makievski AV, Fainerman VB, Bree M, Wustneck R, Kragel J, Miller R. J Phys Chem
B 1998;102:417.
[42] Tripp BC, Magda JJ, Andrade JD. J Colloid Interface Sci 1995;173:16.
[43] Miller R, Joos P, Fainerman VB. Adv Colloid Interface Sci 1994;49:249.
[44] Walstra P, Deroos AL. Food Rev Int 1993;9:503.
[45] Ramsden JJ, Roush DJ, Gill DS, Kurrat RG, Willson RC. J Am Chem Soc 1995;117:
85116.
[46] Cohen Stuart MA. Biopolymers at interfaces. Marcel Dekker; 1999.
[47] Norde W, Haynes CA. Interfacial phenomena and bioproducts. Marcel Dekker;
1999.
[48] Ramsden JJ. Biopolymers at interfaces. Marcel Dekker; 1999.
[49] Adamczyk Z. J Colloid Interface Sci 2000;229:477.
[50] Fainerman VB, Lylyk SV, Ferri JK, Miller R, Watzke H, Leser ME, et al. Colloids Surf,
A 2006;282283:21721.
[51] Fainerman VB, Leser ME, Michel M, Lucassen-Reynders EH, Miller R. J Phys Chem
B 2005;109:9672.
[52] Xir AF, Granick S. Nat Mater 2002;1:129.
[53] Svitova TF, Radke CJ. Ind Eng Chem Res 2005;44:1129.
[54] Ferri JK, unpublished data.
[55] Mac Ritchie F. Vol. 7. Elsevier; 1998.
[56] Kotsmar C, Grigoriev DO, Makievski AV, Ferri JK, Krgel J, Miller R, et al. Colloid
Polym Sci 2008;286:1071.
[57] Kotsmar C, Krgel J, Kovalchuk VI, Aksenenko EV, Fainerman VB, Miller R. J Phys
Chem B 2009;113:103.
[58] Kang SJ, Kocabas C, Kim HS, Cao Q, Meitl MA, Khang DY, et al. Nano Lett 2007;7:
3343.
[59] Ravirajan P, Haque SA, Durrant JR, Bradley DDC, Nelson J. Adv Funct Mater
2005;15:609.
[60] Pardo-Yissar V, Katz E, Lioubashevski O, Willner I. Langmuir 2001;17:1110.
[61] He JA, Samuelson L, Li L, Kumar J, Tripathy SK. Langmuir 1998;14:1674.
[62] Rogach A, Susha A, Caruso F, Sukhorukov G, Kornowski A, Kershaw S, et al. Adv
Mater 2000;12:333.
[63] Caruso F. Adv Mater 2001;13:11.
[64] Schmidt DJ, Cebeci FC, Kalcioglu ZI, Wyman SG, Ortiz C, Van Vliet KJ, et al. ACS
Nano 2009;3:2207.
[65] Li WJ, Mauck RL, Cooper JA, Yuan XN, Tuan RS. J Biomech 2007;40:1686.
[66] Wittmer CR, Phelps JA, Lepus CM, Saltzman WM, Harding MJ, Van Tassel PR.
Biomaterials 2008;29:4082.
[67] Picart C, Schneider A, Etienne O, Mutterer J, Schaaf P, Egles C, et al. Adv Funct
Mater 2005;15:1771.
[68] Hoffman AS, Stayton PS, Press O, Murthy N, Lackey CA, Cheung C, et al. Polym Adv
Technol 2002;13:992.
[69] Discher DE, Janmey P, Wang YL. Science 2005;310:1139.
[70] Georges PC, Janmey PA. J Appl Physiol 2005;98:1547.
[71] Mertz D, Vogt C, Hemmerle J, Mutterer J, Ball V, Voegel JC, et al. Nat Mater 2009;8:
731.
[72] Rehage H, Husmann M, Walter A. Rheologica Acta 2002;41:292.
[73] Burger A, Rehage H. Angew Makromol Chem 1992;202:31.
[74] Ruths J, Essler F, Decher G, Riegler H. Langmuir 2000;16:8871.
[75] Ball V, Hubsch E, Schweiss R, Voegel JC, Schaaf P, Knoll W. Langmuir 2005;21:
8526.
[76] Hammond PT. Curr Opin Colloid Interface Sci 1999;4:430.
[77] Ferri JK, Carl P, Gorevski N, Russell TP, Wang Q, Boker A, et al. Soft Matter 2008;4:
2259.
[78] Russell JT, Lin Y, Boker A, Su L, Carl P, Zettl H, et al. Angew Chem Int Ed 2005;44:
2420.
[79] Collet JP, Shuman H, Ledger RE, Lee ST, Weisel JW. Proc Natl Acad Sci USA
2005;102:9133.
[80] Shen HY, Watanabe J, Akashi M. Anal Chem 2009;81:6923.
[81] Ulbricht M. Polymer 2006;47:2217.
[82] Jin WQ, Toutianoush A, Tieke B. Langmuir 2003;19:2550.
[83] Maskarinec SA, Hannig J, Lee RC, Lee KYC. Biophys J 2002;82:1453.
[84] Krogman KC, Lowery JL, Zacharia NS, Rutledge GC, Hammond PT. Nat Mater
2009;8:512.

J.K. Ferri et al. / Advances in Colloid and Interface Science 161 (2010) 2947
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]

Jiang CY, Tsukruk VV. Adv Mater 2006;18:829.


Nolte M, Donch I, Fery A. Chemphyschem 2006;7:1985.
Ferri JK, Dong WF, Miller R, Mohwald H. Macromolecules 2006;39:1532.
Cramer AD, Ferri JK, in preparation.
Porcel CH, Schlenoff JB. Biomacromolecules 2009;10:2968.
Guzman E, Ritacco H, Ortega F, Svitova T, Radke CJ, Rubio RG. J Phys Chem B
2009;113:7128.
Guzman E, Ritacco H, Rubio JEF, Rubio RG, Ortega F. Soft Matter 2009;5:2130.
Lavalle P, Gergely C, Cuisinier FJG, Decher G, Schaaf P, Voegel JC, et al.
Macromolecules 2002;35:4458.
Richert L, Boulmedais F, Lavalle P, Mutterer J, Ferreux E, Decher G, et al.
Biomacromolecules 2004;5:284.
Nolte AJ, Rubner MF, Cohen RE. Macromolecules 2005;38:5367.
Richert L, Engler AJ, Discher DE, Picart C. Biomacromolecules 2004;5:1908.
Markutsya S, Jiang CY, Pikus Y, Tsukruk VV. Adv Funct Mater 2005;15:771.
McRuiz JT, Adjei N, Ferri JK, in preparation.
Chicurel ME, Chen CS, Ingber DE. Curr Opin Cell Biol 1998;10:232.
Bao G, Suresh S. Nat Mater 2003;2:715.
Guthold M, Liu W, Sparks EA, Jawerth LM, Peng L, Falvo M, et al. Cell Biochem
Biophys 2007;49:165.

47

[101] Liu W, Jawerth LM, Sparks EA, Falvo MR, Hantgan RR, Superne R, et al. Science
2006;313:634.
[102] Stafford CM, Harrison C, Beers KL, Karim A, Amis EJ, Vanlandingham MR, et al. Nat
Mater 2004;3:545.
[103] Withers JR, Aston DE. Adv Colloid Interface Sci 2006;120:57.
[104] Mueller I, Strehlow P. Rubber and rubber balloons, paradigms of thermodynamics. Springer-Verlag; 2004.
[105] Ravera F, Ferrari M, Santini E, Liggier L. Adv Colloid Interface Sci 2005;117:75.
[106] Liggieri L, Ferrari M, Mondelli D, Ravera F. Faraday Discuss 2005;129:125.
[107] Miller R, Wustneck R, Kragel J, Kretzschmar G. Colloids Surf, A 1996;111:75.
[108] Noskov BA, Akentiev AV, Bilibin AY, Zorin IM, Miller R. Adv Colloid Interface Sci
2003;104:245.
[109] Noskov BA, Loglio G. Colloids Surf, A 1998;143:167.
[110] Ravera F, Ferrari M, Liggieri L, Miller R, Passerone A. Langmuir 1997;13:4817.
[111] Ferrari M, Liggieri L, Ravera F, Amodio C, Miller R. J Colloid Interface Sci 1997;186:
40.
[112] Liggieri L, Ravera F, Ferrari M, Passerone A, Miller R. J Colloid Interface Sci
1997;186:46.

Вам также может понравиться