Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

PHYSICAL REVIEW E 89, 032401 (2014)

Effect of charge regulation on the stability of electrolyte films

Christiaan Ketelaar* and Vladimir S. Ajaev†


Department of Mathematics, Southern Methodist University, Dallas, Texas 75275, USA
(Received 24 May 2013; revised manuscript received 13 January 2014; published 10 March 2014)
The stability of a thin liquid film of an electrolyte on a solid substrate is investigated. In the framework of
the Debye-Hückel approximation, we show that the commonly used approximation of fixed potential at the
solid-liquid interface does not lead to predictions of film rupture. To reconcile the model with experimental
observations, we consider the constant charge density approximation for the solid substrate and then proceed to
systematically investigate the effects of charge regulation based on a linear relationship between charge density
and potential. Stability criteria are formulated in terms of charge regulation parameters and electrolyte properties,
resulting in different types of stability diagrams. Critical thickness below which the film ruptures is shown to
decrease as the charge regulation at the solid-liquid interface becomes stronger.

DOI: 10.1103/PhysRevE.89.032401 PACS number(s): 68.15.+e, 47.20.Ma, 47.61.Fg

I. INTRODUCTION Davis [11] and Burelbach et al. [12] have examined the
nonlinear evolution of a liquid film on a solid substrate under
Viscous flows in thin liquid layers have been studied both
the action of London–van der Waals forces and found that
experimentally and theoretically by a number of authors,
the nonlinearities of the system accelerate rupture. Zhang and
following the pioneering work of Reynolds on the theory of
Lister [13] have found similarity solutions near the point of van
lubrication [1]. A comprehensive review of the literature on
der Waals rupture of a thin film on a solid substrate and carried
thin-film flows driven by gravity, capillarity, thermocapillarity,
out simulations of nonlinear evolution of the film thickness.
solutocapillarity, and disjoining pressure with or without phase
All three studies have shown that as the instability grows, the
change can be found in Oron et al. [2]. In a more recent review,
nonlinear behavior leads to hole opening in the film in a finite
Craster and Matar [3] discuss the latest developments in the
time.
studies of thin films.
While theoretical studies usually focus on film rupture
Investigations of stability of thin liquid films on solid
driven by the London–van der Waals dispersion forces, an
substrates are important for a number of applications such
alternative mechanism is often encountered in experiments and
as flotation [4], dynamics of the tear film in the eye [5],
is due to electrostatic effects [14]. Many liquids (e.g., aqueous
and microscale heat transfer [6,7]. Film instabilities often
solutions) contain ions and electrical double layers form near
occur in regimes where inertia is negligible and other physical
charged interfaces. We note that the presence of electric
effects such as viscosity, surface tension, electrostatic forces,
charges not only at solid-liquid but also at liquid-air interfaces
and London–van der Waals dispersion forces are significant.
has been established in the studies of Graciaa et al. [15] and
Instabilities can lead to film rupture when the deformable
Li and Somasundaran [16]. Electric charges at interfaces can
interface between the liquid and the gas phase above it touches
be controlled by adding supplementary electrodes within the
the solid substrate, resulting in the formation of dry spots.
solid structures as proposed by Steffes et al. [17].
A driving mechanism of rupture that has received
In order to develop the stability criteria for an electrolyte
widespread attention in the literature is due to the London–van
film on a charged substrate, it is first important to understand
der Waals dispersion forces which are significant only in very
the basic physics of the interaction of two static charged sur-
thin films, typically of thickness below 100 nm. Spontaneous
faces separated by a film of electrolyte. Early models of such
rupture of such thin liquid films on a planar solid wall was
interaction considered both surfaces to be flat and of uniform
studied experimentally by Scheludko [8] and theoretically by
charge densities [18–22]. The result was based on solving the
Ruckenstein and Jain [9]. In the latter study, linear stability
Debye-Hückel equation for the electrostatic potential in the
criteria are established for a liquid layer under the action of
film and usually represented in terms of interaction force or
London–van der Waals forces. Further refinements of the linear
interaction energy between the surfaces. Application of the
theory of [9], e.g., by including the electrostatic effects and
same ideas to films of aqueous solutions with a deforming
viscoelastic properties of the film, are discussed in Chap. 8 of
interface can be found in several studies [23–26]. Both the
[10].
Debye-Hückel and Poisson-Boltzmann equations have been
Nonlinear models of instability and rupture of thin liquid
solved and the results were compared with experiments on
films are often based on the assumption that the film thickness
film drainage [27,28] and films in constrained vapor bubble
is much smaller than the wavelength of the instability [2].
experiments [23]. The results are usually represented as an
This lubrication-type approach has been used by several
additional term in the stress tensor in the film, referred to as the
authors to derive general evolution equations governing the
electrostatic component of disjoining pressure. It is important
spatiotemporal behavior of thin liquid films. Williams and
to note that the electrostatic interaction in all these studies is
assumed to be repulsive and the films are stable.
Since the electric charges at solid-liquid and liquid-gas
*
cketelaarb@smu.edu interfaces usually appear as a result of chemical reactions at
† interfaces, they actually depend on the local electric field, a
ajaev@smu.edu

1539-3755/2014/89(3)/032401(8) 032401-1 ©2014 American Physical Society


CHRISTIAAN KETELAAR AND VLADIMIR S. AJAEV PHYSICAL REVIEW E 89, 032401 (2014)

physical effect known as charge regulation. This effect was


q* y* gas
investigated in the context of interaction of biological cells 2
[29] and amphoteric solid surfaces [30,31]. Carnie and Chan
[32] derived a simplified model of charge regulation based on a
liquid electrolyte d
linear relation between the surface charge density and electric q*
potential and applied it to interaction of flat solid surfaces and 1
x*
spherical particles. Further studies of charge regulation were solid
carried out by Netz [33], Periset-Camara et al. [34], and Gupta
and Sharma [35]. Detailed comparison with experiments by FIG. 1. Sketch of a thin film of liquid electrolyte of viscosity μ
Gupta and Sharma [35] showed that the effects of charge on a flat solid substrate.
regulation are important for achieving reasonable agreement
between experiment and theory for stresses in thin electrolyte where  is the electric permittivity of the liquid, kB is the
films. Boltzmann constant, and T is the temperature. The solid
Several authors applied the formulas for the electrostatic surface shown in Fig. 1 is electrically charged as a result of
component of disjoining pressure to studies of stability of thin the chemical reactions of dissociation of surface groups. A lin-
films of aqueous solutions on solid substrates (see [36] and earized relation between the dimensional charge density q1∗ and
the references therein) and of coating films inside a pipe [37]. the electrostatic potential ψ1∗ at the solid surface is written as
However, in these studies only simple boundary conditions
of either fixed charge or fixed potential at the interfaces are q1∗ = S1 − K1 ψ1∗ , (2)
considered. The objective of the present study is to incorporate
more realistic relations between the electric charge density where S1 and K1 depend on the density of surface groups
and potential at the interfaces into the model of a thin liquid and the dissociation constants, as discussed by Carnie and
film on a substrate and to investigate both linear stability Chan [32]. Here and below the subscript “1” refers to the
and strongly nonlinear evolution of the fluid interface. Of solid surface, while the subscript “2” is reserved for quantities
particular interest are the conditions of film rupture due to describing the liquid-air interface. The applicability of the
the effect of electric charges at the solid-liquid and liquid-gas linearized charge regulation model was discussed in [32] by
interfaces. We also note that the assumption of fixed interfacial comparison with nonlinear models for different electrolyte
potential at a deforming interface, used, e.g., in Conroy et al. concentrations and parameter values corresponding to a range
[37], implies no electrically induced tangential stress at the of solid surfaces such as SiO2 , Fe2 O3 , Al2 O3 , with the general
deforming interface. In the present work, we incorporate the conclusion that the error of the approximation is typically not
effect of such tangential stress. more than a few percent.
The article is organized as follows. In Sec. II, we describe In the limit of a very large film thickness (d  λD ), the
the model of an electrolyte film and in Sec. III we derive electric potential of the solid surface, ψ̄1∗ , can be related to the
the interface shape evolution equation using a lubrication-type charge density using the linearized Grahame equation [39].
approach. The linearized version of the evolution equation Combining the latter with Eq. (2) leads to the formula
is also discussed. In Sec. IV, we present the results of S1
the stability analysis of the model with a fixed potential ψ̄1∗ = . (3)
K1 + λ−1
D
at the solid wall and a constant charge density at the
liquid-gas interface and show that the film does not rupture The origin of electric charges and the effects of charge
under these conditions regardless of its thickness. In Sec. regulation at liquid-air interfaces are not as well understood
V, we present results of the linear and nonlinear stability as for the solid surfaces in contact with liquid. While the
analysis of the model with constant charge densities at both assumption of constant electric charge density q2∗ at the liquid-
interfaces and show that the film can rupture in finite time. In air interface is most likely appropriate for the majority of cases,
Sec. VI we consider a more general model based on linear we keep the formulation more general here by incorporating
charge regulation at the substrate and investigate how the possible effects of linear charge regulation, so that
choice of regulation parameters affects rupture conditions.
Transition to the limiting cases of constant charge density q2∗ = S2 − K2 ψ2∗ . (4)
and constant potential is also discussed. Let us define the nondimensional charge densities and
potentials by
II. FORMULATION
qi∗ (ψ ∗ ,ψi∗ )
Consider a thin liquid film of viscosity μ and thickness d on qi = , (ψ,ψi ) = , i = 1,2. (5)
S1 ψ̄1∗
a flat solid surface, as shown in Fig. 1. The surface tension at
the air-liquid interface is σ . The liquid is an electrolyte with N Then Eq. (2) can be expressed in nondimensional terms as
different types of ions of valencies zk and bulk concentrations
C1 ψ1
n(0)
k (k = 1,2, . . . ,N). Using the standard definition of the q1 = 1 − , (6)
Debye length [38,39], we can write 1 + C1
 where C1 = K1 λD / has the physical meaning of the ratio
kB T of the regulation capacitance of the surface to the electrical
λD = N (0) 2 , (1)
e 2
k=1 nk zk diffuse layer capacitance. A nondimensional version of (4) is

032401-2
EFFECT OF CHARGE REGULATION ON THE STABILITY . . . PHYSICAL REVIEW E 89, 032401 (2014)

written in the form where p0 is the scaled atmospheric pressure. Since we are
q̂(1 + C2 ) − C2 ψ2 interested in situations when film dynamics is dominated by
q2 = , (7) electrostatic effects, other possible contributions to interfacial
1 + C1
stress balances, e.g., due to London–van der Waals dispersion
where forces, are neglected. The liquid shear stress at the interface is
S2 (1 + C1 ) K2 λ D proportional to the tangential electric field component, leading
q̂ = , C2 = . (8) to the second interfacial stress condition of the form
S1 (1 + C2 ) 
To study linear and nonlinear stability of the film, we κ 2 uy + ψy (ψx + ψy hx ) = 0. (17)
will consider perturbations of the base state corresponding The kinematic boundary condition at the liquid-air interface is
to a film of uniform thickness d. The dimensional Cartesian
coordinates along the solid and normal to it are denoted by 1
h
3 t
+ uhx = v. (18)
x ∗ and y ∗ , respectively. Following the standard lubrication-
Application of the divergence theorem for the electric field
type approach [40,41], applicable for the horizontal scale
allows one to obtain relationships between the local charge
of interface deformations being much larger than the film
density and the derivative of the potential at each interface.
thickness, and neglecting the effects of gravity, we introduce
Together with Eqs. (6) and (7), these relationships lead to the
vertical and horizontal length scales defined by d and Ca−1/3 d,
following boundary conditions for the electric potential,
respectively, where Ca = μU/σ is the capillary number. The
velocity and pressure scales, ψy |y=0 = −κ(1 + C1 − C1 ψ1 ), (19)
(d) ψ̄1∗3
3/2
U = , (9) ψy |y=h = κ[q̂(1 + C2 ) − C2 ψ2 ]. (20)
μσ 1/2 λ3D
We note that these equations do not include contributions
 ψ̄ ∗2 from the electric field outside the film since we assume that
P = 21 , (10)
λD relative dielectric permittivity of the electrolyte r is large, i.e.,
effectively take the limit of r → ∞.
are obtained by considering the balance between the electro- The commonly used boundary conditions of constant
static, capillary, and viscous contributions to the stresses in potential and constant charge densities are the special cases
the film as it deforms. The dimensional horizontal and vertical of Eqs. (19) and (20). The constant potential corresponds to
velocity components u∗ and v ∗ , the pressure p∗ , and time t ∗ the limit of Ci → ∞ (i = 1,2), while the constant charge
can now be used to define the corresponding nondimensional density condition corresponds to Ci → 0. Thus, it is the
variables according to relative magnitude of the regulation capacitance to the diffuse
u∗ v∗ p∗ Ut∗ layer capacitance that controls whether an interface behaves
u= , v= , p= , t= . (11) as a constant potential, constant charge density, or charge-
U 1/3
Ca U P 3dCa−1/3
regulating surface. Following Carnie and Chan [32], we
If the capillary number is small, the use of our scaling simpli- characterize the linear charge regulation in terms of the
fies the Navier-Stokes equations to a system of lubrication-type nondimensional parameters −1  i  1 defined by
equations while the electrostatic potential in the liquid satisfies
the Debye-Hückel equation. Thus, the system of governing Ci − 1
i = , i = 1,2. (21)
equation is of the form Ci + 1

px = uyy + ψψx , (12) Note that the constant potential limit then corresponds to
i = 1, while the constant charge limit corresponds to
py = ψψy , (13) i = −1.

ux + vy = 0, (14) III. NONDIMENSIONAL EVOLUTION EQUATION


d
ψyy = κ ψ,
2
κ= . (15) Solving Eq. (15) subject to the boundary conditions (19)
λD and (20) leads to the following expression for the electric
Liquid flow in this model is coupled to the electric field potential in the film,
through the electrostatic body force term in the momentum
ψ = (eκh − 1 2 e−κh )−1 [(eκh − 1 q̂)e−κy
balance as well as through the interfacial boundary conditions
discussed below. The scaled components of this force, ψψx + (q̂ − 2 e−κh )eκy ]. (22)
and ψψy , appear on the right-hand sides of Eqs. (12) and (13),
Note that for a film of uniform thickness (h = 1) this equation
respectively.
reduces to the nondimensional version of the formula obtained
At the solid wall (y = 0) we apply the no-slip and no-
by Carnie and Chan [32] for the case of two flat surfaces with
penetration boundary conditions for the flow. At the interface,
charge regulation.
y = h(x,t), the normal stress boundary condition includes the
We consider the case of constant charge density at the
contribution from the Maxwell stress tensor,
liquid-air interface, i.e., assume 2 = −1, for the rest of
1 ψy2 the present article, since the effects of charge regulation at
p0 − p = hxx − , (16) the deforming interface are not expected to be significant. We
2 κ2

032401-3
CHRISTIAAN KETELAAR AND VLADIMIR S. AJAEV PHYSICAL REVIEW E 89, 032401 (2014)

verified that adding some realistic charge regulation correction parameter 1 :


at that interface does not alter the main conclusions of the
present study, while the formulas become significantly more A = 1 κ(eκ − 1 e−κ ) − 32 1 (eκ + 1 e−κ )
cumbersome. Using Eq. (22) with 2 = −1, the potential ψ2 − 38 (eκ + 1 e−κ )3 , (32)
at the liquid-air interface can be expressed as a function of the
interface shape, 1 
B = −κ 2
(eκ − 1 e−κ )2 − 21
2 + q̂(eκh − 1 e−κh ) + 34 (eκ − 1 e−κ )(eκ + 1 e−κ ) = 0,
ψ2 = . (23) (33)
eκh + 1 e−κh
From Eq. (13), we observe that p − ψ 2 /2 is a function C = −κ(eκ − 1 e−κ ). (34)
independent of y. The horizontal flow velocity is then found
The stability branches in the q̂-κ plane for a fixed 1 are given
by integrating Eq. (12) with the boundary condition (17) and
by
the no-slip condition at the wall, resulting in

u = 12 (px − ψψx )(y 2 − 2yh) − q̂(q̂hx + κ −1 ψ2x )y. (24) −B ± B 2 − 4AC
q̂(κ) = , (35)
2A
Combining the horizontal velocity profile given by (24)
with (14), (16), and the kinematic boundary condition (18), as long as this equation has real solutions; i.e., the discriminant
we obtain B 2 − 4AC is nonnegative.
   
ht + h3 hxx + 12 ψ22 x − 32 q̂h2 (q̂h + κ −1 ψ2 )x x = 0. (25)
IV. FIXED SUBSTRATE POTENTIAL
Note that while this equation only involves the potential at If we consider a constant potential at the solid wall
the deforming interface, the film evolution depends on the (1 = 1), the formula for the potential at the liquid-air
electrical properties of both interfaces since ψ2 comes out of interface, Eq. (23), reduces to
the solution of the Debye-Hückel equation.
Note that while Eq. (25) is nonlinear, it has a simple steady- 1 + q̂ sinh(κh)
state solution corresponding to a uniform film thickness, ψ2 = . (36)
cosh(κh)
h(x,t) = 1. Behavior of a small perturbation ζ (x,t) to this
solution is described by the linearized version of (25), The coefficients A, B, and C defined by Eqs. (32)–(34)
simplify to
ζt + ζxxxx + Gζxx = 0, (26)
A = 2κ sinh κ − 3 cosh κ − 3 cosh3 κ, (37)
where


3q̂ dψ2 3 B = −2κ(sinh2 κ − 1) + 3 sinh κ cosh κ, (38)
G≡ ψ2 − − q̂ 2 (27)
2κ dh 2 h=1
C = −2κ sinh κ. (39)
and ψ2 is now treated as a function of a single variable h, as
defined by Eq. (23). Based on Eq. (26), the growth rate γ of a Remarkably, when these expressions are substituted into
small sinusoidal perturbation of a wave number k is given by the condition (31), we find that the electrolyte film is stable for
all values of the scaled film thickness since the function on the
γ (k) = k 2 (G − k 2 ). (28) left-hand side of (31) is always negative. Thus, even though
Perturbations of all wave numbers decay for G < 0, the assumptions of fixed substrate potential and fixed charge
indicating stability. However, the film is unstable when the density at the fluid interface are often mentioned as a natural
condition G > 0 is satisfied. The growth rate of the fastest framework for studies of electrolyte films on solid substrates
instability mode then corresponds to the maximum value of [36], we find that a model based on these assumptions fails
the function γ (k), equal to to predict film break-up by growth of interfacial perturbations
observed, e.g., in [14]. While film stability at large values of
G2 κ can be explained by the lack of sufficient overlap between
γmax = , (29)
4 the electrical double layers of the two interfaces, it is rather
and is reached at the wavelength puzzling that no rupture threshold is found at small values of
√ κ regardless of the value of q̂. To gain deeper physical insight
2 2π into the behavior of perturbations of the film surface at small
λ= √ . (30) κ, let us consider the limiting case of κ = 0. When the fluid
G
interface is flat, the configuration is that of a capacitor with
Substituting the solution for the electric potential given by the uniform electric field confined between the two interfaces.
(23) into the definition of G given by (27), we can rewrite the The electrostatic energy in the liquid then has constant density,
instability condition in the form denoted by E. Now let us consider the effect of a small
Aq̂ 2 + B q̂ + C > 0, (31) interfacial perturbation ζ . In the framework of the linear
approximation in ζ , the electric field remains uniform in the
where the coefficients A, B, and C are functions of domain 0  y  ζ with the same energy density. The total
the scaled film thickness κ and the charge regulation change in the electrostatic part of the system’s energy due to

032401-4
EFFECT OF CHARGE REGULATION ON THE STABILITY . . . PHYSICAL REVIEW E 89, 032401 (2014)

perturbation is 15
∞ ∞
Eζ dx = E ζ dx = 0, (40) 10 unstable
−∞ −∞

where we used the condition of global mass conservation for


5
the incompressible liquid in the film. Thus, there is no energy
reduction due to electrostatic effect needed to overcome the stable
stabilizing effect of surface tension despite the fact that the 0


two interfaces are oppositely charged.
One of the limitations of our model is the assumption of
fixed charge density at the fluid interface. It is natural to ask −5

whether the conclusions of the present section are still valid unstable
when the effect of interfacial charge transport is incorporated −10
into the model by replacing the fixed value of q̂ with the
evolving scaled charge density q̃(x,t). The standard interfacial
charge transport equation (discussed, e.g., in [37]) for the new −15
0 0.1 0.2 0.3 0.4 0.5
variable q̃ has to be solved together with the evolution equation κ
for film thickness. Thus, the film dynamics is governed by a
system of two coupled partial differential equations. However, FIG. 2. Linear stability diagram for a model with constant charge
when the system is linearized, the leading-order stability crite- densities at both interfaces (1 = 2 = −1).
ria turn out to be unaffected by the interfacial charge transport.
Thus, the conclusions of the present section remain valid.
In order to explain experimentally observed rupture phe-
According to our stability criteria, the film is always
nomena, in the next two sections we modify the assumption of
stable if its scaled thickness κ > 0.530 12, which is consistent
fixed solid substrate potential. It is first replaced by a condition
with experimental results showing that instability is generally
of constant charge density, followed by a discussion of a more
not expected if there is no significant overlap between the
general charge regulation condition.
interfacial electric double layers. Under these conditions, the
film is not destabilized even for large values of the parameter
V. FIXED SUBSTRATE CHARGE q̂; in fact, the asymptotic limit of large q̂ corresponds to the
critical value of the scaled thickness of 0.5051 (the value at
Let us consider another commonly used model of interact-
which the coefficient A passes through zero).
ing charged interfaces in which the effects of charge regulation
The diagram in Fig. 2 indicates that the film can be linearly
are completely neglected by assuming the dimensional charge
unstable at both negative and positive values of the charge
densities to be constant, equal to S1 and S2 at the lower and up-
density ratio parameter, q̂. While instability at negative q̂ can
per boundaries of the film, respectively. In our nondimensional
be easily explained in terms of attraction of oppositely charged
formulation this implies setting 1 = −1 in Eq. (23). The
interfaces, the prediction of linear instability at positive q̂ may
resulting expression for the potential at the liquid-air interface
seem rather surprising. To better understand this result, we
can then be written as
note that it can be interpreted as a manifestation of the well-
q̂ cosh(κh) + 1 known phenomenon of the instability of a charged deformable
ψ2 = . (41)
sinh(κh) interface in an electric field, which in our case is generated by
the charges at the substrate and in the liquid rather than external
Note that q̂ in this case has the simple meaning of the ratio of electrodes (as is often done in experiments showing this
charge densities at the two interfaces. The general coefficients type of instability). This phenomenon has been observed and
defined by Eqs. (32)–(34) simplify to described under a wide range of conditions, including different
electrical conductivities of fluids separated by the charged
A = −2κ cosh κ + 3 sinh κ − 3 sinh3 κ, (42)
interface (see, e.g., [42] and references therein). Another way
B = −2κ(cosh2 κ + 1) + 3 sinh κ cosh κ, (43) to interpret our result is to point out that the deformable
interface interacts not only with the charge at the solid surface
C = −2κ cosh κ. (44) but also with the cloud of charge in the solution, so at least some
of these interactions always involve charges of opposite signs.
The stability diagram based on this equation is shown The linear stability model breaks down as soon as the
in Fig. 2. We note that accounting for tangential stresses at amplitude of the perturbation of the base state is no longer
the interface, as is done in the models of thermocapillary much smaller than unity. To investigate perturbation dynamics
and solutocapillary instabilities, is essential for the correct in this regime and make conclusions about the possibility of
description of film dynamics. For example, in the limit of film rupture, we solved the full nonlinear Eq. (25) with ψ2
κ → 0, the electrostatic contribution to the pressure in the film from (41) numerically using the method of lines with time
has a stabilizing effect, so the instability is entirely due to the stepping by Gear’s BDF method. As is common in film rupture
tangential electric field which arises when a small perturbation studies [13], the size of the computational domain L is dictated
of the initially flat interface is introduced. by the wavelength of the fastest growing perturbation: we

032401-5
CHRISTIAAN KETELAAR AND VLADIMIR S. AJAEV PHYSICAL REVIEW E 89, 032401 (2014)

1.2 1.2
t=0.3
t=0.5
1.15
t=0.6
1

1.1

0.8
1.05

h 1 a 0.6

0.95
0.4

0.9

0.2
0.85

0.8 0
0 0.2 0.4 0.6 0.8 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
x/L t

FIG. 3. Snapshots of the interface shape for κ = 0.4, q̂ = −2. FIG. 4. Evolution of the scaled minimum thickness (denoted
by a) for κ = 0.4, q̂ = −2. Solid line is obtained from the numerical
solution of Eq. (25); dot-dashed line is the prediction of the linear
stability theory.
choose L = λ/2, where λ is given by Eq. (30). The boundary
conditions are
VI. EFFECT OF CHARGE REGULATION
hx (0,t) = hx (L,t) = 0, (45) AT THE SOLID WALL

hxxx (0,t) = hxxx (L,t) = 0, (46) The stability results in the previous two sections clearly
show qualitatively different behavior, but the only difference
between the models is the degree of charge regulation at the
and the initial perturbation is sinusoidal, of the amplitude ζ0 = solid surface, from perfect regulation to no regulation. It is
10−2 . Note that the nonlinear system considered here depends natural to ask how the degree of charge regulation affects the
on two nondimensional parameters, the ratio of the linearized stability of the film. Addressing this issue is the objective of
surface charge densities q̂ and the film thickness scaled by the the present section.
Debye length, κ. If the charge regulation effects are neglected, i.e., 1 = −1,
The plot in Fig. 3 shows snapshots of the interface at the stability diagram is that shown in Fig. 2. As the value of
different times for κ = 0.4 and q̂ = −2. Clearly, the interface the regulation parameter is increased slightly, both stability
shapes are no longer sinusoidal due to the nonlinear effects branches are moving downward in the stability diagram, but
and the film evolution tends to speed up. the overall picture remains qualitatively the same. At the values
In order to better understand the time evolution of the of 1 near −0.5, the lower stability branch starts shifting
liquid-air interface, it is convenient to plot the minimum film downward significantly, eventually disappearing to infinity as
thickness (always reached at x = 0 with our choice of the 1 approaches zero. Typical stages of this process are shown
initial condition) as a function of time, in nondimensional in Fig. 5. The upper stability branch starts deforming so that
terms. The linear stability theory discussed at the end of the region of instability shrinks gradually until in disappears
Sec. III predicts that the scaled minimum thickness, denoted completely in the regime of perfect charge regulation, i.e.,
by a, should evolve according to constant potential. This corresponds to the limiting case of
absolute stability that was discussed in Sec. IV.
a = 1 − ζ0 e G
2
t/4
. (47) Comparison between stability diagrams such as the ones
shown in Fig. 5 shows that for a fixed value of q̂, the
values of κ corresponding to the instability threshold are
This result is represented by a dot-dashed line in Fig. 4. decreasing with increase in 1 . Thus, the critical dimensional
The solid line, corresponding to the numerical solution of thickness for the onset of instability for a given electrolyte
the nonlinear evolution equation (25), follows the predictions properties will decrease as the strength of charge regulation is
of the linear stability analysis initially, but then deviates from increased.
it significantly. The film ruptures in a finite time in a manner
similar to the well-known case of the van der Waals driven
VII. CONCLUSIONS
rupture [12]. We carried out similar studies for several different
values of q̂ and observed the same dynamics. Thus, the model We investigated the stability of a thin electrolyte film on a
described in the present section is indeed capable of describing solid substrate. Both normal and tangential stress conditions at
not only the linear destabilization of the liquid film but also its the film surface include contributions from the electric field. A
rupture in a finite time. linear stability analysis under the assumption of fixed substrate

032401-6
EFFECT OF CHARGE REGULATION ON THE STABILITY . . . PHYSICAL REVIEW E 89, 032401 (2014)

8 8

6 6

4 unstable 4 unstable

2 2

0 0


−2
stable −2
stable

−4 −4

−6 unstable −6
unstable
−8 −8
0 0.1 0.2 0.3 0.4 0.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
κ κ

FIG. 5. Stability regions for 1 = −0.5 (left side) and 1 = −0.3 (right side).

potential shows that the film is always stable. A simple physical electrostatic properties of the two interfaces and κ is the film
interpretation of this conclusion is provided by considering thickness scaled by the Debye length). The region of instability
the change in the electrostatic energy as a result of interfacial shrinks as the value of 1 is increased, in accordance with the
perturbation of an initially flat film in the limiting case when conclusion of stability under constant substrate potential, i.e.,
film thickness is much smaller than the Debye length of the perfect charge regulation.
electrolyte solution. We carried out nonlinear simulations of the film evolution
In order to describe the experimentally observed rupture for the case of constant charge densities and observed that
by electrostatic effects, we considered the approximation of linear destabilization of the interface eventually leads to film
fixed charge density at the solid substrate and then more rupture, in qualitative agreement with experimental data on
general models incorporating the effects of interfacial charge rupture of electrolyte films.
regulation. The latter is characterized by a nondimensional
parameter 1 which depends on the ratio of the solid
surface regulation capacitance to the capacitance of the diffuse
ACKNOWLEDGMENTS
layer formed in the liquid near the solid. We showed that
different types of linear stability diagrams in q̂-κ variables This work was supported by the National Science Foun-
are possible depending on the value of 1 (here q̂ is a dation under Grant No. CBET-0854318. We thank the anony-
nondimensional parameter which characterizes the ratio of mous referees for a number of useful suggestions.

[1] O. Reynolds, Proc. R. Soc. London 40, 191 (1886). [11] M. Williams and S. Davis, J. Colloid Interface Sci. 90, 220
[2] A. Oron, S. H. Davis, and S. G. Bankoff, Rev. Mod. Phys. 69, (1982).
931 (1997). [12] J. Burelbach, S. G. Bankoff, and S. H. Davis, J. Fluid Mech.
[3] R. V. Craster and O. K. Matar, Rev. Mod. Phys. 81, 1131 195, 463 (1988).
(2009). [13] W. Zhang and J. Lister, Phys. Fluids 11, 2454 (1999).
[4] J. Israelachvili, Intermolecular and Surface Forces (Academic [14] H. Schulze, K. Stöckelhuber, and A. Wenger, Colloids and
Press, London, 2011). Surfaces A: Physicochem. Eng. Aspects 192, 61 (2001).
[5] R. Braun, Annu. Rev. Fluid Mech. 44, 267 (2012). [15] A. Graciaa, G. Morel, P. Saulner, J. Lachaise, and R. Schechter,
[6] O. Kabov, D. Zaitsev, V. Cheverda, and A. Bar-Cohen, J. Colloid Interface Sci. 172, 131 (1995).
Exp. Therm. Fluid Sci. 35, 825 (2011). [16] C. Li and P. Somasundaran, J. Colloid Interface Sci. 146, 215
[7] D. Zaitsev, M. Lozano Aviles, H. Auracher, and O. Kabov, (1991).
Microgravity Sci. Technol. 19, 71 (2007). [17] C. Steffes, T. Baier, and S. Hardt, Colloids and Surfaces A 376,
[8] A. Scheludko, Proc. K. Ned. Akad. Wet. B 65, 76 (1962). 85 (2011).
[9] E. Ruckenstein and R. K. Jain, J. Chem. Soc. Faraday Trans. 70, [18] B. Derjaguin, N. Churaev, and V. Muller, Surface Forces
132 (1974). (Springer, New York, 1987).
[10] I. Ivanov, Thin Liquid Films: Fundametals and Applications [19] E. Verwey and J. Overbeek, Theory of the Stability of Lyophobic
(Dekker, New York, 1988). Colloids (Elsevier, Amsterdam, 1948).

032401-7
CHRISTIAAN KETELAAR AND VLADIMIR S. AJAEV PHYSICAL REVIEW E 89, 032401 (2014)

[20] H. Kruyt, Colloid Science (Elsevier, Amsterdam, 1953). [32] S. Carnie and D. Chan, J. Colloid Interface Sci. 161, 260 (1993).
[21] V. A. Parsegian and D. Gingell, Biophys. J. 12, 1192 (1972). [33] R. Netz, J. Phys. Condens. Matter 15, S239 (2003).
[22] J. Gregory, J. Colloid Interface Sci. 51, 44 (1975). [34] R. Pericet-Camara, G. Papastavrou, S. Behrens, and M.
[23] R. Mazzoco and P. Wayner, J. Colloid Interface Sci. 214, 156 Borkovec, J. Phys. Chem. B 108, 19467 (2004).
(1999). [35] A. Gupta and M. Sharma, J. Colloid Interface Sci. 149, 407
[24] D. Chan, R. Dagastine, and L. White, J. Colloid Interface Sci. (1992).
236, 141 (2001). [36] N. Churaev, Adv. Colloid Interface Sci. 103, 197 (2003).
[25] R. Pushkarova and R. Horn, Langmuir 24, 8726 (2008). [37] D. T. Conroy, R. V. Craster, O. K. Matar, and D. T. Papageorgiou,
[26] B. Derjaguin and N. Churaev, J. Colloid Interface Sci. 49, 249 J. Fluid Mech. 656, 481 (2010).
(1974). [38] R. Probstein, Physicochemical Hydrodynamics, 2nd ed. (Wiley,
[27] D. Hewitt, D. Fornasiero, J. Ralston, and L. R. Fisher, J. Chem. New York, 1994).
Soc., Faraday Trans. 89, 817 (1993). [39] B. Kirby, Micro- and Nanoscale Fluid Mechanics: Transport in
[28] R. Manica and D. Chan, Phys. Chem. Chem. Phys. 13, 1434 Microfluidic Devices (Cambridge University Press, Cambridge,
(2011). 2010).
[29] B. Ninham and V. A. Parsegian, J. Theor. Biol. 31, 405 (1971). [40] V. S. Ajaev, Interfacial Fluid Mechanics: A Mathematical
[30] D. Chan, J. Perram, R. White, and T. Healy, J. Chem. Soc., Modeling Approach (Springer, New York, 2012).
Faraday Trans. 71, 1046 (1975). [41] V. S. Ajaev, T. Gambaryan-Roisman, and P. Stephan, J. Colloid
[31] D. Chan, T. Healy, and L. White, J. Chem. Soc., Faraday Trans. Interface Sci. 342, 550 (2010).
72, 2844 (1976). [42] R. V. Craster and O. K. Matar, Phys. Fluids 17, 032104 (2005).

032401-8

You might also like