Вы находитесь на странице: 1из 13

Article

pubs.acs.org/Langmuir

Hydrodynamics of Particles at an OilWater Interface


Archit Dani, Geo Keiser, Mohsen Yeganeh, and Charles Maldarelli*,

The Benjamin Levich Institute for PhysicoChemical Hydrodynamics and Department of Chemical Engineering, The City College of
New York, 140 Convent Avenue, New York, NY 10031, United States

ExxonMobil Research and Engineering Company, Annandale, NJ 08801, United States


S Supporting Information
*

ABSTRACT: This study is a theoretical and experimental investigation of the


hydrodynamics of the mutual approach of two oating spherical particles moving
along an oilwater interface. An analytical expression is obtained for the
(inertialess) Stokes drag for an isolated particle translating on a at interface as a
function of the immersion depth into the water phase for the case in which the
viscosity of the oil is much larger than that of the water. An approximation for the
viscous drag due to the mutual approach of identical spheres is formulated as the
product of the isolated drag multiplied by the resistance of approaching spheres
in an innite medium. Experiments are undertaken on the capillary attraction of
large, millimeter-sized Teon spheres oating at the interface between a very
viscous oil and water. With the use of image visualization and particle tracking,
the separation distance as a function of time [S (t)] is measured along with the immersion depth and predicted by setting the
capillary attraction force equal to the viscous drag resistance. The excellent agreement validates the approximating formula.

interface and create capillary attraction;2427 (iv) electrostatic


interactions: for charged colloids in an aqueous electrolyte
adjoining a low dielectric medium (air or oil), the charge on the
surface and the diuse layer of opposite charge around the
particle in the aqueous phase create surface dipoles giving rise
to interparticle dipolar repulsion screened by the electrolyte,
and a normal electrodipping (image charge) force which
depresses the meniscus and causes capillary attraction on
overlap,2837 and (v) van der Waals attractive interactions.
Gravity-induced capillary attractions typically dominate the selfassembly of colloids on the of order of 10 m or larger at uid

, where
uid interfaces since the capillary length Sc =

INTRODUCTION
The long-standing interest in particles at uid interfaces stems
principally from their application to stabilize foams and
emulsions,1,2 but recent attention has focused on their selforganization into assembled structures. Early experimental
eorts examined particles at an air/water interface in which the
underlying aqueous phase is much deeper than the particle size.
These studies detailed, for uncompressed monolayers, the twodimensional self-organization of colloids into clusters, foamlike
mesostructures and crystalline lattices of hexagonal form in
which separation distances are a few particle diameters, and the
reordering and buckling of these lattices under surface
compression on a Langmuir trough.37 Analogous studies of
colloids on deep layers at the oilwater interface have
demonstrated the eect of compression, particle wettability,
charge and electrolyte,810 particle size and functionality.11 In
addition, particle monolayers of dierent size but with the same
functionality, or the same size but dierent functionalities have
been studied.1113
Self-assembled interfacial congurations of colloids derive
from the interparticle interactions in the monolayer.1418 These
can be divided into ve principal types (Figure 1, panels ad)
including (i) capillary attractive (or otation) interactions:
gravity acting on particles more (less) dense than the
surrounding phase depresses (elevates) the meniscus surrounding the particle, and overlap of local menisci during particle
approach creates an attractive force;1922 (ii) immersion forces:
partially wet colloids resting on substrates in thin layers and
connected by bridging menisci experience a capillary
attraction;1416,23 (iii) irregular wetting: nonspherical particles
or pinning of contact lines along particle surfaces generate
asymmetric contact lines deforming the surrounding uid
XXXX American Chemical Society

is the density dierence between the adjoining uids, g is the


acceleration of gravity and is the interfacial tension becomes
of the order of the particle radius (a). Equivalently the Bond
number B =

a2
Sc2

ga 2
,

which is the ratio of gravitational to

surface tension forces, becomes of order one. Electrostatic


interactions dominate for charged particles on the order of 1
m or smaller, where gravitational forces are unimportant.
Immersion forces operate on all sizes, and asymmetric capillary
attractions dene the interaction for anisotropic particles
oating at an interface.3840
The kinetics of self-organization and the structures of the
assembled states are a balance between the interparticle
interactions (summarized in Figure 1, panels ad) and the
hydrodynamic interactions of the particles, and although the
Received: May 28, 2015
Revised: October 14, 2015

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

Figure 1. Dierent types of interparticle interactions existing on an oilwater interface: (a) gravity induced capillary attraction (Fcap) between two
hydrophobic colloids, (b) electrostatic repulsion force (Frep) between two charged colloids and electrocapillary attractions (Fecap) arising as a
consequence of electrodipping force (Fedip), (c) immersion forces (Fimm) resulting from specic wetting conditions at the three phase contact line,
and (d) capillary attractive forces (Firr) emerging from the pinning of the meniscus to irregularly shaped colloids.

dierent values of the shear and dilatational viscosities. Fisher et


al.45 extend these calculations to the case in which the
dilatational viscosity (or Marangoni number) is innite and the
surface was incompressible (with the shear viscosity nite),
which has been a model used to describe motions of disks
driven by thermal uctuations in Langmuir monolayers of
insoluble surfactants or bilayers.44,46
Some work has been done on experimentally measuring drag
on isolated spheres forced to move along a surface by capillary
or magnetic forces. To compare the drag to the theoretical
predictions, the contact angle or the immersion depth is also
required. For large particles (1 mm), the contact angle can be
measured directly by visualization, while for smaller sizes, other
techniques such as gel trapping, freeze-fracture shadow casting,
atomic force microscopy, and interferometry have been
developed.47 Petkov et al.48 tracked the trajectory and velocity
of single glass particles (few hundreds of microns in radius) as
they moved up an airwater meniscus formed at a Teon
barrier by capillarity. Contact angles were measured directly by
visualization. On the basis of the values of the contact angle and
the generated capillary force, a theoretical particle trajectory
was simulated as a function of time and comparison with the
experimental data yielded the Stokes drag f, which was in very
good agreement with the theoretical predictions of Danov et
al.41,42 Similarly, Vassileva et al.49 experimentally measured f for
single submillimeter-sized glass particles placed at a pentadecane/glycerolwater interface in which the viscosities of the oil
and the aqueous phase were matched. The contact angles were
measured by direct visualization, and the particle pairs
interacted with each other by capillary attractive forces.
Comparison of the experimental trajectories with the
simulations yielded a value of f slightly larger than the
theoretical prediction of Danov et al.42 Ally et al.50 oated
isolated polystyrene particles (a few microns in radius)
embedded with magnetite at a planar airwater and air
silicone oil interface and applied a measured lateral magnetic
eld to move the particles. Contact angles were measured using
lm trapping and interferometry, and from comparisons of the
measured trajectories with the simulations values of the isolated
drag coecient f were obtained in very good agreement with
theory. The fairly good agreement of the theoretical predictions
for the trajectories with the experiments indicates that the eect

forces have been studied at length, the hydrodynamic


interactions are less well understood. The simplest problem
which has been studied is the case of an isolated sphere moving
steadily along a at interface with velocity U between two
immiscible uids [e.g., oil (or air) above with viscosity o and
water (below) with viscosity w] in Stokes ow. For this case,
the hydrodynamic resistance Fdrag is given in terms of a
translational drag resistance coecient f (e.g., for an oil/water
interface, Fdrag = 6oaf(d/a,)U), where d is the immersion
depth into the water phase, = w/o is the viscosity ratio, and
the limit d/a corresponds to the familiar resistance of a
particle in an innite liquid of either adjoining phase. Numerical
calculations of f under these assumptions were undertaken by
Danov et al.41,42 at a at airliquid interface and more generally
on a spherical surface with arbitrary viscosity ratio. These
calculations were extended by Pozrikidis43 to include shear ow
in the bounding phases. Danov et al.41,42 also included the
eect on f of surface rheological forces due to the presence of
monolayer forming surface active species (amphiphilic
molecules). When a particle moves along an interface, it
creates an in-plane shear ow on the interface and a material
area changing ow as surface uid in front of the moving
particle is compressed and surface uid behind the particle is
expanded. The presence of surfactant gives rise to a resistance
to this shear ow in the plane of the interface (surface shear
viscosity, s) and a resistance to the compression or expansion
in the surface area (dilatational viscosity, s). In addition,
gradients in surfactant concentration along the interface create
surface tangential tractions (Marangoni forces). The compression and expansion of surface uid increases the surfactant
concentration at the front and decreases it at the back, creating
a tension gradient which also acts against the surface ow
created by the particle. Hence the presence of surfactant will
increase the drag, and f will depend not only on the immersion
depth of the particle and the viscosity ratio of the bounding
phases but also on the scaled surfactant rheological parameters
[Boussinesq numbers E = s/(ao) and K = s/(ao)] and a
Marangoni number Ma = RT/(oU) where RT is the
thermal energy and is the maximum packing concentration
of surfactant which scales the surface concentration, as well as
nondimensional groups describing the transport of surfactant to
and along the surface.44 Danov et al.41,42 present solutions for
B

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

Figure 2. (a) A sphere translating along an oilwater interface with a velocity U. (b) A two-dimensional illustration of toroidal coordinates when
represented in the plane of the azimuthal angle, = 0. The red circles indicate surfaces of constant , and the blue circles indicate surfaces of
constant .

a single particle wetted area approximation for the isolated drag


[i.e., farea = + (1 )] (where is the wetted area fraction
of the water; see also ref 37) and a mobility function derived
from the bispherical solution for two spheres approaching each
other in a continuous viscous phase. They found reasonable
agreement for uncharged particles, although the experimental
drag was larger than the theoretical prediction. For the charged
particles, comparison of theory and experiment demonstrated
the contribution of the electrostaticinduced capillary attraction
alongside the gravity-induced attraction. For the uncharged
particles, Boneva et al. note that the larger measured drag
relative to the theoretical prediction could be due to the eect
of the meniscus around the particles, although the presence of
the surfactant could also have increased the drag.
This study examines, in more detail, the pairwise hydrodynamic interaction of two colloids trapped at an interface and
moving along their line of centers, and the validity of the
multiplicative approximating formulas for the Stokes drag for
this motion, Fdrag = 6oaf(d/a,)G1(S /a)U. The experimental
validation of this formulation by Vassileva et al.49 and Boneva et
al.29,30 were for an oil/water interface whose viscosity ratio
was one or nearly one. For this case, the hydrodynamic
interaction can principally be accounted for by simply the two
particle mobility function in an innite medium, G(S /a)
because f(d/a, = 1) is between 1 and 1.1 and is not sensitive
to the immersion depth. Here we will experimentally examine
the case of colloids at an oil/water interface in which the oil
viscosity is much larger than that of water ( 0.010.1),
where f becomes a strong function of the immersion depth. To
actuate the particle motion, we will undertake experiments on
uncharged millimeter-sized Teon spheres in which the contact
angle will be measured directly for the evaluation of f, and the
actuation force is capillary attraction without signicant
electrostatic complication. The use of millimeter spheres also
serves the purpose of assessing the importance of the meniscus
drag, since the meniscus depression can be on the order of the
particle radius. To facilitate comparison between experiments
and theory, rather than having to numerically solve the isolated
drag problem, we develop an analytical solution for the drag on
a trapped particle moving along a planar interface at arbitrary
immersion depth and with a very large viscosity stratication
based on a toroidal coordinate solution to the Stokes equations.

of the meniscus around the particles, which is not included in


the theory for the drag, is not very signicant. But it should be
noted that in these experiments, the meniscus deformation was
very small. It was only in the study of Danov et al. (they also
used a heavy metal sphere to produce large deformation) that
the measured drag coecient was found to be much larger than
the prediction, possibly due to the extra drag of the meniscus.
The pairwise hydrodynamic interaction between approaching
particles involves the extra resistance due to the uid drainage
between the particles. While a direct calculation of the drag on
two oating particles approaching each other along an interface
has not been done, experiments on this interaction have been
undertaken by Vassileva et al.49 These authors used an
approximate expression for the hydrodynamic resistance in
which the single particle drag coecient f(d/a, = 1) was
multiplied by the drag coecient for two particles in a
continuous phase approaching one another [i.e., Fdrag =
6oaf(d/a, = 1)G1(S /a)U, where G(S /a) is the mobility
function for two spheres approaching one another in a
continuous phase]. For G(S /a), Vassileva et al. use the
expression of Batchelor51 and by comparing the experiments
to the simulations, they nd a value of f(d/a, = 1) equal to 1.2,
in good agreement with the Danov et al. value for f which lies in
the range of 11.1 for any immersion depth. Similarly, two
particle capillary attraction experiments were undertaken by
Dalbe et al.52 for millimeter-sized polyethylene and nylon
spheres at the air/glycerolwater interface with visual measurement of contact angles. Using the same multiplicative
expression for the drag coecient to simulate the data, these
authors compared theory and experiment by setting f = 1 for all
sphere sizes and wetting angles and found qualitative
agreement, as they did not account for the dependence of f
on the immersion depth. Boneva et al.29,30 considered the
motion of two submillimeter-sized hydrophobized glass colloids
(charged and uncharged) under the inuence of electrostatic
and capillary attractive forces at a tetradecanewater interface
and measured contact angles directly by visualization. A net
attraction was observed to cause the particles to approach one
another, and S (t) was measured as a function of time. They
compared their measured trajectories to theory with the particle
drag formulated as an isolated drag coecient multiplied by the
two particle mobility function in an innite medium. They used
C

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
The study begins rst with this analytical solution and is
followed by the capillary attraction experiments and their
modeling.

unit vector directed along the positive x direction. Resolving


the velocity in r, , and z directions as u, v, and w, respectively,
leads to

u = cos

THEORETICAL CALCULATION OF THE SINGLE


PARTICLE DRAG COEFFICIENT
Formulation. We consider the case of a particle of radius a
straddling a at oilwater interface (Figure 2a) and translating
along the interface with a prescribed low velocity (low
Reynolds number) of magnitude U. The assumption of a at
interface requires the interfacial tension forces to be much
larger than the viscous forces (small capillary number, Ca =
oU/) and that gravitational forces are small so that the
meniscus does not deform appreciably around the particles
(Bond number B < 1). The three phase contact line is assumed
to remain pinned to the particle surface. Pinning the contact
line forbids any rotation along or perpendicular to the axis of
the particle. Also, for the interpretation of the experiments in
which the viscosity of the oil phase is considerably larger (one
or 2 orders of magnitude) than the viscosity of the underlying
aqueous phase, the aqueous phase is assumed inviscid and the
system can be described by the Stokes equation and the
equation of continuity in the oil.
7 = 2 V

(1)

V = 0

(2)

u
= 0,
z

(4b)

w=

z 7 + w sin

cos(0) = 1

r=

c sinh
(cosh cos )

(9)

d
a

(10)

i [A(s)cos(s) + B(s)sin(s)]P1 1/2+s(cosh )

7=

(11a)

ds
i

i [C(s)cos(s) + D(s)sin(s)]P1/2+s(cosh )
(11b)

ds
i

i [E(s)cos(s) + F(s)sin(s)]P21/2+s(cosh ) ds
(11c)

(4c)

w =

m
+
+ 2 2
r r
r 2
z
r

i [G(s)sin(s) + H(s)cos(s)]P1 1/2+s(cosh )


(11d)

ds

where 3 m is a Laplacian-like operator given by


3m =

(8)

and in eq 9, c is dened as c = a sin (0). In order to obtain the


velocity and pressure elds, the governing equations are
transformed into toroidal coordinates and then solved by
using a Mehler-Fock transform.28 On transforming eq 5 into
toroidal coordinates and applying a Mehler-Fock transform, the
general solution satisfying zero ow far from the particle can be
written as

where 7 , , , and w satisfy the following harmonic relations


317 = 0, 3 0 = 0, 3 2 = 0, 31w = 0
(5)
2

w=0

where and are spatial variables in a toroidal coordinate


system. The physical surfaces of interest, namely the surface of
the sphere and the plane interface excluding the part covered by
the sphere, are given by = 0 and = 0, respectively, 0 being
determined from the immersion depth of the sphere. 0 is
related to d and a as given by

1
v = (r 7 + )sin
2

v
= 0,
z

c sin
;
(cosh cos )

z=

where u, v, and w are the components of the velocity eld in the


radial direction (er), azimuthal direction (e), and vertical
direction (ez), respectively. Due to the imposed boundary
conditions (eq 7), the velocities and pressure are rst
harmonics with reference to the azimuthal angle , and
therefore, the other Fourier modes obtained from the general
solution of a Laplacian can be ignored.53 The velocity elds
comprise a homogeneous solution and a pressure-dependent
particular solution given by
(4a)

(7)

Finally, we require the velocity eld to tend to zero far from the
particle in the viscous (oil) phase.
Solution In Toroidal Coordinates. A toroidal coordinate
system (Figure 2b) is an appropriate choice for a spherical
particle straddling a plane interface54 since the coordinate
surfaces coincide with the physical surfaces of the problem. The
transformation relations between cylindrical coordinates and
toroidal coordinates are represented by

(3)

1
u = (r 7 + + )cos
2

w=0

The tangential stress condition and zero normal velocity


conditions on the plane interface can be expressed as

where 7 is the pressure normalized by oU/a, V is the velocity


eld nondimensionalized by U, o is the viscosity of the oil and
distances are scaled by a. V can be resolved into three
components of velocity when represented in cylindrical
coordinates (see Figure 2a).

V = uer + ve + wez

, v = sin ,

(6)

cosh cos
2i

(12)

where
is an associated Legendre function of the rst
kind of integer order m and imaginary degree. The functions
A(s), B(s),..., H(s) are unknown functions of the imaginary
transformed variable s, which are determined from the
boundary conditions imposed. On substituting 11 into 2 and

P m1/2 + s

For the system under consideration, there are three boundary


conditions. They are zero tangential stress and zero normal
velocity at the at, uid interface and no slip on the surface of
the sphere. At the surface of the sphere, V = ex, where ex is the
D

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
transforming into toroidal coordinates, we obtain the following
set of relations between the unknown functions:

In order to solve 13 and 14, it is convenient to dene two


intermediate functions X(s) and Y(s) such that

3
3
c
s + A(s + 1) + 5A(s) s A(s 1)
2
2
a

X(s) =

tan(s)
2

[C(s + 1) 2C(s) + C(s + 1)]

sinh
1
7( , 0)P 1/2
+ s(cosh ) d
(cosh cos 0)3/2

(15)

3
5
3
s + s + E(s + 1) s E(s)

2
2
2

Y (s ) =

3
5
3
s s E(s 1) s + E(s) +

2
2
2

tan(s)
2

d 7( , 0) 1
sinh
P 1/2 + s
3/2
d
(cosh 1)

(cosh ) d

3
3
2s + G(s + 1) 2sG(s) + s G(s 1) = 0

2
2

(16)

The advantage of dening the intermediate functions is that


all the eight unknown functions can be represented in terms of

(13)

3
3
c
s + B(s + 1) + 5B(s) s B(s 1)
2
2
a

the two intermediate functions, and the whole problem reduces

[D(s + 1) 2D(s) + D(s 1)]

eqs 11a and 16, the function B(s) can be expressed in terms of

3
5
3
s + s + F(s + 1) s F(s)

2
2
2

Y(s) by means of a Mehler-Fock transform of order 1.

to obtaining the two intermediate functions. With the use of

B (s ) =

3
5
3
s
s F(s 1) s + F(s) +

2
2
2

3
3
2s + H(s + 1) 2sH(s) + s H(s 1) = 0

2
2

3
3

s + Y (s + 1) 2sY (s) + s Y
s
2
2

(s 1)

(17)

Also, on using eqs 11a and 15, it can be shown that

(14)

3
3
A(s)cos(s0) + B(s)sin(s0) = s + X(s + 1) 2sX(s)cos(0) + s X(s 1)

2
2

On applying a Mehler-Fock transform of order 0, 2, and 1,


respectively, to eq 8, we obtain
D(s) =

3
c
s 2 + 2s + Y (s + 1)

2as
4

3
s 2 2s + Y (s 1)

C(s)cos(s0) + D(s)sin(s0) =

F (s ) =

c
[Y (s + 1) Y (s 1)]
2as

(18)

(20)

H (s ) = 0

(21)

Similarly, on applying a Mehler-Fock transform of order 0, 2,


and 1, respectively, to eq 7, we obtain

(19)

2 2 cos[( 0)s]
2
3
3
c 2
s + 2s + X(s + 1) s 2s + X(s 1) +

2as
4
4
cos(s)

(22)

4
3

s + Y (s + 1) + 2
[14s 2 6s + 8]Y (s)+
s
2
(s 1)
E(s)cos(s0) + F(s)sin(s0) =

4
3
s Y (s 1) = 0
s
2

c
[X(s + 1) X(s 1)]
2a
(23)

G (s ) =

cs sin 0
a sin(s0)

(25)

Since eq 25 is homogeneous, it automatically admits Y(s) = 0


as a solution, which leads to
X (s )

B(s) = 0,

(24)

D(s) = 0,

F (s ) = 0

(26)

On substituting eqs 26, 18, 22, 23, and 24 into eq 13, we


obtain

On substituting 17, 19, 20, and 21 into 14, we obtain


E

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

3
3
s +
Z(s + 1) + 2K (s)Z(s) + s Z(s 1)

2
2
= - (s )

Z (s ) =

K (s ) =

(27)

c[sin(20s) s sin(20)]X (s)


a sin(20s)

(28)

[sin 0 sin(0s) 2s cos 0 cos(0s)]cos(0s)


cos(20s) + cos(20)

3s sin 0 sin(20)sin(20s)
[cos(20s) + cos(20)][sin(20s) s sin(20)]
(29)

- (s ) =

4 2 sin(0s)tan(s)sin 2 0
cos(20s) + cos(20)

(30)

Eq 27 has been solved by Zabarankin for 0 2.247 in the


context of a related problem on the drag of two equal fused
spheres, and its solution is given by
55

Z(s ) =

-(s ) H (s )
2(K (s) s)

Figure 3. Comparison of the obtained analytical solution for f as a


function of d/a with other solutions available in literature. The blue
circles are values reported by Pozrikidis43 by implementing a boundary
integral scheme. The red circles are values obtained by using a
boundary integral scheme developed by Fischer et al.45 for an
incompressible surface, and the green triangles indicate calculated
values using a nite dierence scheme by Danov et al.42

(31)

where H(s) is obtained as the solution of a Fredholm equation


of the second kind.
-(s) = H(s) +

[K (s) s]sin(s)

i s2

1
4

s )cos()H()
d
i (cos(2
) cos(2s)

the shear stresses on the uid surface are equal to zero. The
drag force can then be represented by

(32)

Calculation Of Drag Force. In order to evaluate the drag


force on the particle, it is convenient to sketch a hemispherical
boundary of radius R around the particle as shown in Figure 2a.
Using the divergence theorem, the stresses on the boundaries
Aa, AR, and Az can be related to each other by

A n x dA + A
a

n x dA +

Fdrag =

Fdrag = 2o Ua lim (r 2 7(r , z)|z = 0 )

A n x dA = 0

where is the stress on the surface and n is the unit normal to


the surface. The area integral over the area Az equals zero since
i

(34)

(35)

Substituting and using the asymptotic form of pressure into


eq 35 and nondimensionalizing by using the expression for
Stokes drag, an expression for the dimensionless resistance
coecient f is obtained.

(33)

2 i(sin 0)3

n x dA

When R , the dimensional drag force on the sphere (Fdrag)


is then given by

d
f =
a

A n x dA = A

s sin(20) + 2 sin(20s)
2
1
s X (s )
ds

4
cos(0s)(cos(20s) + cos(20))

Eq 36 allows the dimensionless drag to be computed by


numerical integration after rst solving numerically a Fredholm
equation of the second kind (eq 32) to obtain H(s) and nally
X(s) from eqs 28 and 31. The calculated drag coecient f is
shown in Figure 3 as a function of d/a and also provided in
tabular form in the Supporting Information. When d/a
increases, the particle distributes itself more into a relatively
inviscid aqueous phase and this causes f to decrease. Also, f =
0.5 corresponds to d/a = 1. This illustrates the fact that a
particle immersed halfway into a viscous liquid experiences half
the drag predicted by the Stokes equation. Figure 3 compares
the results obtained from eq 36 to the results reported in the
literature by applying the wetted area approximation [farea = (1
), where is the wetted area fraction of the water] and

(36)

nite dierence and boundary integral techniques to obtain


numerically exact solutions. It is evident that eq 36 accurately
predicts f, and the results match well with the values of f
obtained by nite-dierence42 and boundary integral techniques.43 Figure 3 also shows the boundary integral solution
results of Fisher et al.,45 for the case of an incompressible
surface. Comparing the values of f for a clean interface and an
incompressible one for dierent values of d/a, it is clear that the
eect of the surface incompressibility is to increase the drag by
1020%. Surface shear viscosity can potentially aect the
resistance even more dramatically, and f tends to innity with
innite shear viscosity. The calculations of Danov et al.41 for a
large viscosity contrast between the bounding phases show that
F

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

diameter of the petri dish (184 mm) largely exceeds the capillary
length (S c 5 mm) at the PAO8 and PAO2/water interfaces, and the
experiments were performed near the center of the petri dish to ensure
the atness of the interface in the absence of particles. The rst particle
was carefully placed on the interface by using a pair of tweezers. Once
the particle drift due to deposition at the interface ceases, the second
particle is introduced to the interface at a certain initial separation
distance (S o) which ranges between 0.5lc to 5lc. In order to ensure that
the oilwater interface is devoid of surfactants, the interfacial tension
(IFT) was measured by using a commercial Attesion Theta pendant
drop tensiometer. The value of initial IFT was found to be 46.8 mN/
m, and this value did not change by more than 23 mN/m in 1 h. The
approach of the particles was recorded by using a charged couple
device (CCD) camera at 30 frames per second. The raw data was
analyzed by using NIS-Elements AR tracking software (Nikon), which
extracts the coordinates of both particles from every frame of the video
and obtains the center-to-center separation S as a function of time.
Multiple measurements were carried out for each particle size by
varying the initial separation S o.
Measurement Of Contact Angle and Immersion Depth. To
simulate the experimental particle trajectories S (t), the capillary
attraction force has to be calculated. The capillary attraction force
(eq 41) is a function of the contact angle of the meniscus at the three
phase contact perimeter (as measured through the aqueous phase) and
the inclination angle of the meniscus with the horizontal. Once the
contact angle is known, the inclination angle can be obtained from a
vertical force balance on the particle, given the densities of the particle
and uid phases. The immersion depth d can also be calculated by
geometry from and , so that f, which depends on d/a, can be
calculated and the theoretical simulation of S (t) can be constructed.
The measurement of the contact angle at the particle surface is not
straightforward due to contact angle hysteresis.47 In order to account
for the contact angle hysteresis, the contact and inclination angles and
immersion depths were all directly measured by imaging the meniscus
in a view orthogonal to the plane of the interface using a CCD camera
(see Figure 5a). An oilwater interface was formed in an optical quartz
cuvette and a single particle was deposited on the interface in the same
manner as the capillary attraction experiments. Although the
immersion depths and inclination angle varied with the particle size,
the contact angles in both PAO8 and PAO2 were all approximately the
same (160), most likely due to the reproducibility in the placement of
the particles at the interface. We also measured (using a contact angle
goniometer) the contact angle of millimeter-sized sessile drops of
water in PAO8 and PAO2 and obtained a value of 135, indicating that
pinning of the particle at the interface was signicant.
Additional experiments were performed wherein two particles
placed on an oilwater interface were allowed to come together by
capillary attraction. The particle surface was marked with an oil
insoluble dye in order to detect any possible particle rotation during

for E = K = 1, the drag coecient for a range of d/a is


approximately 20% higher than the clean interface calculation.

EXPERIMENTAL MEASUREMENTS OF CAPILLARY


ATTRACTION

Materials. Ultrapuried water having a conductivity of 18 M cm


was used as the aqueous phase. PAO8 (polyalphaolen-8), which has a
density of 832.6 kg/m3 and PAO2 (polyalphaolen-2), which has a
density of 797 kg/m3 (both at 21 C) were chosen as the oil phase for
two separate sets of experiments and were obtained from ExxonMobil
and used directly without any further purication. All the experiments
were carried out at 21 1 C. The viscosities of the oils were
computed by using a Walther Correlation curve constructed by using
the values of viscosities provided by the manufacturer at two
temperatures of 40 and 100 C. PAO8 has a viscosity of 98.8 cP at
21 C, and PAO2 has a viscosity of 7.19 cP. Teon spheres of dierent
diameters (1.09, 2.2, 3.18, and 3.8 mm) were purchased from
Engineering Laboratories (Oakland, NJ). They were cleaned by
consecutive ultrasonication steps in ethanol, acetone, and nally in DI
water for 5 min each and dried by using nitrogen prior to use. A glass
petri dish having an inner diameter of 184 mm and height of 101 mm
was purchased from VWR and cleaned by immersing in a 36.5% w/v
solution of HCl at 90 C for half an hour, washing thrice times in DI
water and eventually drying using nitrogen.
Capillary Attraction Experiments. An oilwater interface was
formed in the glass petri dish, placed on an antivibration table, by
pouring 150 mL of water followed by 150 mL of oil as shown in Figure
4. The uids in the petri dish were allowed to equilibrate for 1 h. The

Figure 4. A schematic of the experimental setup used to track particle


trajectories on an oil (PAO8/PAO2)water interface. CCD cameras
are used to capture videos orthogonally in projection and top view.

Figure 5. (a) Front view of a 3.8 mm Teon sphere on a PAO8water interface. The darker phase is PAO8, and the lighter phase is water. (b) A
plot of the variation of the particle size corresponding to a particular value of on the PAO8water interface. The experimental data points are
indicated by the data symbols. The simulation results, based on eq 37, are indicated by the solid curves. The red curve corresponds to = 135, and
the green curve corresponds to = 160.
G

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
1 dS

approach. The experimental results indicated the absence of particle


rotation during approach. As rotation is prevented by contact angle
pinning, these experiments lend support to the large values of the
contact angle hysteresis identied between the goniometry measurements and the images of the colloids at the interface.
The uniformly measured value of the contact angle, and the varying
inclination angles and immersion depths with the dierent particle
radii, are also consistent with the vertical force balance on the
particle.56 An equation for this balance is given as

velocity of the colloids 2 dt is qualitatively observed to increase as the


particles come closer. In this range of large separation relative to the
particle size, the increase in velocity is due to the increase in the
capillary attraction with decreasing S , as the meniscus between the
particles continually attens as the particles approach one another.
The hydrodynamic interaction between the particles does become
important as the separation decreases, and we examine this point in
more detail below, where the simulations (the red curve in Figure 6)
are compared to the data.
The regime of closer separation (8 > S /a > 2) is shown in the
graphs in Figure 7 and Figure 8 for the high viscosity and low viscosity
oils, respectively. Each graph reports the results of dierent
experiments with the same particle size and oil viscosity; the
separation distance is scaled with the particle radius a, and time is
plotted dimensionally with t = 0, representing the time in which the
particles are 8 particle radii away and the nal times representing the
point at which the particles appear to contact (S /a = 2). For all sets,
the data at shorter separations (S /a 34) show that the velocity
begins to decrease, as the hydrodynamic resistance overwhelms the
increase in capillary attraction. This behavior was also observed by
Dalbe et al.52 in their study of millimeter-sized particles at the air
water interface, and we further quantify this competition below (the
lines in the graphs), where S (t) is simulated.
In general, it is also clear from the graphs that the average particle
1 dS
velocities ( 2 dt ) range from 3a/20 0.075 mm/sec for the slowest
velocity (more viscous oil, smaller particle size) to 3a/0.5 6.6 mm/
sec for the largest velocity (less viscous oil, larger particle size).
Qualitatively, the data also show expected trends on the eect of oil
viscosity and particle size: the more viscous the oil, the longer the time
required for capillary attraction to bring the particle pair together
(compare Figures 7 and 8). In addition, the larger the particle size the
shorter the time to contact due to the greater capillary attraction as
larger particles create a larger interfacial depression. These eects are
all quantitatively described below, where the separation distance
during capillary attraction is simulated (solid lines in the gures).

2
2
[ + w ] = sin( + )sin() +
B
3 o
1
[cos 3( + )] [cos( + )]
3
h
+ sin 2( + )
a
where o =

s o
w o

and w =

s w
w o

(37)

are nondimensionalized densities

and s is the particle density, o is the oil density, w is the density of


the aqueous phase, and the Bond number (B) as dened earlier is
(w o )ga 2

. In eq 37, h is the deformation of the meniscus at the three

phase contact line, and is obtained by numerically solving the Young


Laplace equation

h2
=
2a 2

sin(+) r(r) ddr [1 + (r((rr)))2 ]1/2 dr

(38)

where (r) being the variation of the meniscus deformation with


distance (r) from the particle center. In Figure 5b, the measured
inclination angles (red data points) are plotted with respect to the
particle radius a for particles oating at the PAO8water interface.
The lines in the gure are the simulations of from the solution of the
vertical force balance for 135 and 160 contact angles, and it is clear
that the angle measured directly from the image of the particle at the
interface (160) ts the measurements very accurately.
Experimental Results on Particle Pair Separation During
Capillary Attraction. Data for the center-to-center particle
separation as a function of time [S (t)] are obtained for particle pairs
at the interfaces between the more viscous oil (PAO8) and water, and
the less viscous oil (PAO2) and water, each for dierent particle sizes.
An illustrative set of data is given rst in Figure 6 for the more viscous
oil and a particle diameter of 1.09 mm for S from approximately 17.5
mm (S /a = 35) to 2 mm (S /a = 4), which represents the trajectory
prior to intimate hydrodynamic interaction and contact. The approach

SIMULATIONS OF CAPILLARY ATTRACTION OF


PARTICLE PAIRS AND COMPARISON TO
EXPERIMENTS
The particle pair trajectories are simulated by considering a
horizontal force balance on each particle equating the drag
(Fdrag) to the capillary attraction force (Fcap). In the
experiments summarized in Figures 7 and 8, the Reynolds
number in the oil phase (Re =

o (a)U
o

) is 0.1 or less for the more

viscous oil experiments, and between 1 and 10 for the less


viscous oil, so inertial eects are relatively unimportant except
possibly for the experiments on the less viscous oilwater
interface. The capillary number (oU/) is less than or equal to
103 for all experiments, and therefore viscous stresses do not
distort the interface from a at conguration. The Bond
number varies between 0.01 (for the smallest particle, 1.09 mm
diameter) to 0.17 for the largest particle, 3.8 mm in diameter).
The largest particles give the greatest capillary attraction and
meniscus height h (Figure 5a). The normalized meniscus height
(h/a) varies from approximately 1 for the largest particle (3.8
mm, see Figure 5a) to less than 0.1 for the smallest (1.09 mm).
Hence, meniscus dragging could be important for the larger
particles (3.8, 3.18, and 2.2 mm). Noting these calculations on
the scaling groups, we rst simulate the hydrodynamic drag on
the particle assuming Stokes ow and a at surface and then
assess the validity through comparison with the experiments.
For the Stokes drag on the particle moving laterally along the
surface, we multiply the hydrodynamic drag on a single particle
as developed earlier [Fdrag = 6oaf(d/a, = 0)U] by G1(S /a),

Figure 6. Interparticle distance (l) between two Teon particles of


1.09 mm diameter each on a PAO8water interface as a function of
approach time. The blue triangles represent experimental data, and the
red curve is the simulation for a value of f = 0.72 which best matches
the experiental data.
H

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

Figure 7. Interparticle separation l/a for particles of (a) 1.09 mm diameter, (b) 3.18 mm diameter, and (c) 3.8 mm diameter plotted as a function of
approach time for PAO8. The blue points are experimental data, and the red curve is the theoretical particle trajectories excluding interparticle
hydrodynamics. The green curve predicts the particle trajectories when interparticle hydrodynamic interactions are taken into account.

where S is the center to center interparticle separation, lc is the


capillary length of the oilwater system, and K1 is a modied
Bessel function of the rst kind and order 1. The Nicolson
approximation requires23 the inclination of the interface to be
small enough such that sin2() < 1, and the capillary length to
be long enough that (a sin( ) /S c) 2 < 1. These inequalities
are satised in our experiments, cf. the values of and in
Figure 5b. Using the vertical force balance to eliminate the
inclination angle,56 the capillary attraction can be written as
S
Fcap = 2aB5/2 2K1
Sc
Figure 8. Interparticle separation l/a for particles of (a) 1.09 mm
diameter and (b) 2.2 mm diameter plotted as a function of approach
time for PAO2. The blue points are experimental data, and the red
curve is the theoretical particle trajectories excluding interparticle
hydrodynamics. The green curve predicts the particle trajectories when
the separation distance dependent interparticle hydrodynamic
interactions are taken into account.

3o af (d /a , = 0)G1(S /a)

a
G1(S /a) = 1 +

2(l 2a)

S
2aB5/2 2K1
Sc

(ln(l 2a) ln a + 0.6789)2


1 + 0.3766 exp

6.297

(39)

Q p2

Q p = a sin sin( )

cos 3

d
S=
dt

(43)

Integration of eq 43 provides the interparticle separation as a


function of time, which can then be compared to the
experiments. In undertaking this comparison, we note rst
that the immersion depth d has to be specied in order that
f(d/a, = 0) can be calculated; all other parameters in eq 43 are
measured. While it is known that particles have a tendency of
sinking deeper into the interface during approach,57 we assume,
in our experiments, that the pinning of the meniscus to the
sphere does not allow any changes in d and inhibits the
propensity of the particles to sink. We therefore will take the
inclination angle and immersion depth to be constant as the
particles approach and calculate d from and as d/a = 1 +
cos( ), where the inclination angle is determined from
the vertical force balance of an isolated particle. We note that
the value of d computed in this way also agrees with the value

As we noted earlier, the mobility function accounts for the


increased hydrodynamic resistance associated with the drainage
of the lm between the colloids as they come in close
proximity.
The capillary attraction force (Fcap) is formulated in the
Nicolson approximation.19,20,23,56
S
K1
Sc Sc

cos

where B is the Bond number, and = 0 3 W 2 + 6 .


This expression makes clear that as the separation distance
decreases, the capillary attraction increases as the Bessel
function increases with a decrease in its argument [K1(S /S c)
S c/S as S /S c 0]. In addition, with increasing particle size a,
the capillary attraction also increases as a6, accounting for the
dependence of the Bond number on a.
With these formulations for the hydrodynamic drag and the
interparticle capillary attraction, the horizontal balance on the
colloids becomes

where G(S /a) is the hydrodynamic mobility of two particles in a


single phase approaching one another at a distance S /a along
their line of centers. For the latter mobility function, we use the
correlation developed from the innite series bispherical
solution as given by29

Fcap = 2

(42)

(40)
(41)
I

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
measured directly from the sideview images of the particles
straddling the interface (Figure 5a). With the value of d
assigned, we integrate eq 43 assuming a value for f(d/a, = 0)
and then change the drag coecient until the simulation of S (t)
agrees with the experiments.
In matching the simulations with the data, we have found
that in experiments where the particles were initially located
very far apart, the data for the largest values of the separation
could not be matched to the simulations. This is illustrated in
Figure 6, where the initial separation is approximately 35 radii.
On adjusting the value of f to match the simulations of the
interparticle separations to the experimental measurements, we
were unable to match the time interval from 0 to approximately
400 s, where S /a changed from approximately 35 to
approximately 25. Our matched simulation shown as the
solid red curve agrees very well with the data for the times in
which the particles are relatively close. We believe that the
reason for this discrepancy is that the capillary attraction forces
at the larger distances (in which the colloids are in fact 2.53.5
capillary lengths away from each other, S c 5 mm) are small
relative to background ow drifts in the experiment as the oil/
water meniscus has relaxed almost completely. Eq 43 that
governs the simulation is valid and determining only for smaller
separations and the larger capillary attractions. (In Figure 6, this
is true for S /a between 25 and 4, or capillary lengths less than
2.5S c and times between 300 and 700 s.) This behavior was
found in our matching to all the experiments reported in
Figures 7 and 8, and we systematically t all data for separation
distances less than approximately 2.5 capillary lengths.
An important issue is the drag correction factor [G(S /a)] to
account for particleparticle interactions. To illustrate the
signicance of this factor, the matching of simulation to theory
was undertaken without the correction (i.e., using only the
single particle drag coecient). This optimization is shown as
the red curves in Figures 7 and 8), which only poorly t the
data as it underestimates the predicted aggregation time for all
particle sizes. However, the green curve, which incorporates the
pairpair hydrodynamics, models the experiment perfectly. In
particular, it is evident by comparing the two simulations that
the inclusion of this eect is responsible for the experimentally
observed reduction in the approach velocities at the closest
separations. Without this eect, the capillary attraction force
remains unchecked, and the particles would increase their
velocity indenitely during approach. Note however that for
larger separations (i.e., larger than S /a > 8), the two simulations
merge, indicating that interparticle hydrodynamic resistance is
unimportant. The merged simulations agree very well with the
data concerning interparticle separations up to 2.5 capillary
lengths, as shown in Figure 6 where the red curve is the merged
solution. For greater separations, the capillary attraction force
becomes weaker than drift currents in the Petri dish, and as we
have noted, the particle trajectories cannot be modeled by
balancing the drag with capillary attraction alone.
In Figure 9, we have plotted (blue data labels) the values of
the single particle drag coecient f obtained by tting the
simulation to the data for each of the experiments on a given
particle size and oil viscosity as summarized in Figures 7 and 8
as a function of the corresponding values for d/a for the
experiment. Error bars are constructed from the dierent
realizations of each of the experiments for a given oil and
particle size. The theoretical curve for the single particle drag f
as a function of d/a, as obtained from the toroidal coordinate
calculations (eq 36), is plotted alongside the data. The

Figure 9. A comparison of the values of theoretical f and experimental


f by using Teon balls on a PAO8water interface and a PAO2water
interface as well as other data available in literature. The blue are the
data obtained by using Teon balls on a PAO8water and PAO2
water interface. The green represent data obtained by Ally et al.50
on an airwater interface and an airsilicone oil interface by using
glass particles. The pink are experimental data points obtained by
Petkov et al.48 by measuring the motion of glass and polystyrene
spheres on an airwater interface.

agreement between the experimentally derived points and the


simulations without any adjustable parameters provides strong
support for the formulation of the hydrodynamic drag of
trapped interfacial particles as a composite product of the single
particle drag and the two-particle in a single phase mobility
function. In addition, the agreement evident in Figure 9 and the
close agreement of the simulations with the measured values of
S (t) for the pair capillary attraction suggests that the eect of
meniscus drag is not important. Note there is no systematic
deviation of the simulations from the measured S (t) or the
optimized f from the theoretical curve for the larger particles
(greater immersion depths) for which the menisci are large and
their eect on the drag would be noticeable. In addition, the
assumption of Stokes ow appears to be validated even for the
low viscosity oil experiments where the Reynolds numbers are
order 110, as no systematic deviations of the experiments
from the theory are observed.
In Figure 9, we have also added the experimentally measured
single particle nondimensional drag coecients obtained by
Ally et al.50 and Petkov et al.,48 which were undertaken at air
liquid interfaces with the large viscosity contrast applicable to
our theory. The contact angle was measured directly, and
hence, the immersion depth was known to allow these
measurements of f to be plotted in Figure 9 as a function of
the known immersion depth. Their data also fall on the
theoretical line computed from our expression (eq 36) further
validating the theory.

CONCLUSIONS
This study has focused on the hydrodynamic motion of
particles oating on a uid interface in the limit of a zero
J

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir

ection coatings) or indirectly as lithography masks or


templates for materials fabrication. The key step in assembling
the particle mesostructures is to tune the interparticle forces
(see Introduction) to achieve the desired assembly. We
envision that the primary use of our study is to implement
accurate simulations of the assembly process in an eort to
understand how the interparticle forces can be integrated in a
rational way to develop a desired lattice. This study shows how
to correctly formulate an expression for the pairwise hydrodynamic interaction between two particles on a surface, in
particular for the case in which the viscosity of one phase
bounding the interface is much larger than the other (as, for
example, the airwater interface). The correct modeling of this
pairwise interaction is a necessary building block for a
multiparticle simulation, and the multiplicative expression for
the two-particle resistance that we have developed here and
veried through experimentation should provide an accurate
rendering of the hydrodynamic forces determining the
assembly.

Boussinesq number. Specically, our interest was in elucidating


pairwise hydrodynamics at an oilwater interface as a rst step
toward understanding the eect of interfacial ow on particle
assembly. We have proposed a formulation for the pairwise
resistance as the product of the resistance of an isolated particle,
f, translating along the surface at a particular immersion depth,
d, and the (well-known) hydrodynamic resistance of two
particles moving toward each other in a single (innite) phase.
To examine the validity of this formulation, we rst obtain a
new analytical solution for the single particle drag coecient by
solving the Stokes equations using toroidal coordinates. We
obtain an integral expression for the drag coecient f, involving
a function which is obtained by solving a Fredholm integral
equation. The solution is valid for the case in which there is a
large contrast between the viscosities of the two phases (e.g.,
airwater or viscous oilswater), and we verify the solution by
comparison with boundary integral and nite dierence
numerical solutions for the single particle drag coecient f in
the literature. In experimental studies, this analytical solution is
valuable since values for the drag coecient can be computed
for the immersion depths relevant to the experiment by
solution of a Fredholm integral, thereby eliminating the
necessity of a full scale numerical simulation. The results
make clear that the large viscosity contrast between the
adjoining uids makes the single particle drag (and hence the
hydrodynamic pairwise interaction) a strong function of the
immersion depth of the particle into the more viscous phase.
Inasmuch as the immersion depth is determined by the three
phase contact angle, this study demonstrates that knowledge of
the contact angle is important for the accurate description of
the collective hydrodynamics of surface colloids on interfaces
separating phases with large viscosity contrast.
We then perform experiments on the capillary attraction of
millimeter-sized hydrophobic particles oating at an oil/water
interface and image their trajectories to determine the
interparticle separation distance as a function of time. The
contact angle of the meniscus at the three phase contact line
between oil, water, and the particle surface and the immersion
depth of the particles are measured separately by directly
imaging oating particles in a view orthogonal to the plane of
the interface. We simulate the interparticle separation by
balancing the capillary attractive force driving the particles
together with the hydrodynamic resistance which is formulated
as the product of the single particle drag f as obtained from our
toroidal analytical solution and the resistance function for two
particles in an innite phase. Since the contact angle is
measured, there are no adjustable parameters, and we nd the
simulations agree with the experimental measurements thus
validating our proposed formulation for the pairwise uid
resistance. Our theoretical prediction for the single particle drag
coecient f is also compared to measurements based on the
experimental data of Petkov et al.48 and Ally et al.50 for the
motion of single particles straddling an interface. Their
experimental data also agrees well with our theoretical
predictions.
Most of the technological applications involving particles at
uid interfaces (apart from stabilizing foams and emulsions) are
related to generating tunable particle mesostructures. Such
assemblies are either formed directly on a solid substrate (e.g.,
convective assembly58) or transferred to a substrate (e.g.,
nanosphere lithography59). The mesostructures are then either
used directly (e.g., as biosensor arrays, photonic crystals,
optoelectronic components, or superhydrophobic or antire-

ASSOCIATED CONTENT

* Supporting Information
S

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acs.langmuir.5b02146.
Dimensionless drag coecient as a function of the
normalized immersion depth of the particle (PDF)
Video of the setup used to track particle trajectories on
an oil (Video)

AUTHOR INFORMATION

Corresponding Author

*E-mail: charles@chemail.engr.ccny.cuny.edu.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The authors would like to acknowledge Elizabeth Knapp for
assisting in the process of particle tracking by using the NISElements AR tracking software and Exxon Mobil Research and
Engineering for funding and NSF (CTS1512458).

REFERENCES

(1) Binks, B. Particles as Surfactants - Similarities and Differences.


Curr. Opin. Colloid Interface Sci. 2002, 7, 2141.
(2) Colloidal Particles at Liquid Interfaces; Binks, B., Horozov, T.,
Eds.; Cambridge University Press: Cambridge, U.K., 2006.
(3) Pieranski, P. Two-Dimensional Interfacial Colloidal Crystals.
Phys. Rev. Lett. 1980, 45, 569572.
(4) Ghezzi, F.; Earnshaw, J.; Finnis, M.; McCluney, M. Pattern
Formation in Colloidal Monolayers at the Air-Water Interface. J.
Colloid Interface Sci. 2001, 238, 433446.
(5) Ruiz-Garcia, J.; Gamez-Corrales, R.; Ivlev, B. I. Formation of twodimensional colloidal voids, soap froths, and clusters. Phys. Rev. E: Stat.
Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 1998, 58, 660663.
(6) Horvolgyi, Z.; Mate, M.; Zrinyi, M. On the universal growth of
two-dimensional aggregates of hydrophobed glass beads formed at the
(aqueous solution of electrolyte)-air interfaces. Colloids Surf., A 1994,
84, 207216.
(7) Chen, W.; Tan, S.; Huang, Z.; Ng, T.-K.; FOrd, W.; Tong, P.
Measured Long-range Attractive Interaction Between Charged
Polystyrene Latex Spheres at a Water-Air Interface. Phys. Rev. E
2006, 74, 021406.
K

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
(8) Horozov, T.; Aveyard, R.; Binks, B. P.; Clint, J. H. Structure and
Stability of Silica Particle Monolayers at Horizontal and Vertical
Octane-Water Interfaces. Langmuir 2005, 21, 74057412.
(9) Reynaert, S.; Moldenaers, P.; Vermant, J. Control Over Colloidal
Aggregation in Monolayers of Latex Particles at the Oil-Water
Interface. Langmuir 2006, 22, 49364945.
(10) Law, A. D.; Auriol, M.; Smith, D.; Horozov, T. S.; Buzza, D. M.
A. Self-Assembly of Two-Dimensional Colloidal Clusters by Tuning
the Hydrophobicity, Composition, and Packing Geometry. Phys. Rev.
Lett. 2013, 110, 138301.
(11) Law, A.; Buzza, D.; Horozov, T. Two Dimensional Colloidal
Alloys. Phys. Rev. Lett. 2011, 106, 128302.
(12) Tarimala, S.; Dai, L. L. Structure of Microparticles in SolidStabilized Emulsions. Langmuir 2004, 20, 34923494.
(13) Ma, H.; Dai, L. L. Structure of Multi-Component Colloidal
Lattices at Oil-Water Interfaces. Langmuir 2009, 25, 1121011215.
(14) Kralchevsky, P. A.; Nagayama, K. Capillary Interactions Between
Particles Bound to Interfaces, Liquid Films and Biomembranes. Adv.
Colloid Interface Sci. 2000, 85, 145192.
(15) Kralchevsky, P. A., Nagayama, K. Particles at Fluid Interfaces and
Membranes: Attachment of Colloid Particles and Proteins to Interfaces and
Formation of Two Dimensional Arrays; Elsevier: Amsterdam, 2001.
(16) Kralchevsky, P. A.; Denkov, N. D. Capillary forces and
structuring in layers of colloid particles. Curr. Opin. Colloid Interface
Sci. 2001, 6, 383401.
(17) Bresme, F.; Oettel, M. Nanoparticles at fluid Interfaces. J. Phys.:
Condens. Matter 2007, 19, 413101.
(18) Bleibel, J.; Dominiguez, A.; Oettel, M. Colloidal Particles at
Fluid Interfaces: Effective Interactions, Dynamics and Gravitation-like
Instability. Eur. Phys. J.: Spec. Top. 2013, 222, 30713087.
(19) Nicolson, M. M.; Evans, R. C. The interaction between floating
particles. Math. Proc. Cambridge Philos. Soc. 1949, 45, 288295.
(20) Chan, D., Jr; Henry, J. D.; White, L. The interaction of colloidal
particles collected at fluid interfaces. J. Colloid Interface Sci. 1981, 79,
410418.
(21) Blanc, C.; Fedorenko, D.; Gross, M.; In, M.; Abkarian, M.;
Gharbi, M. A.; Fournier, J.-B.; Galatola, P.; Nobili, M. Capillary force
on a micrometric sphere trapped at a fluid interface exhibiting arbitrary
curvature gradients. Phys. Rev. Lett. 2013, 111, 058302.
(22) Galatola, P.; Fournier, J. Capillary Force Acting on a Colloidal
Particle Floating on a Deformed Interface. Soft Matter 2014, 10,
21972212.
(23) Kralchevsky, P. A.; Nagayama, K. Capillary forces between
colloidal particles. Langmuir 1994, 10, 2336.
(24) Kralchevsky, P. A.; Denkov, N. D.; Danov, K. D. Particles with
an undulated contact line at a fluid interface: Interaction between
capillary quadrupoles and rheology of particulate monolayers.
Langmuir 2001, 17, 76947705.
(25) Stamou, D.; Duschl, C.; Johannsmann, D. Long-range
Attraction Between Colloidal Spheres at the Air/Water Interface:
The Consequence of an Irregular Meniscus. Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top. 2000, 62, 52635271.
(26) Fournier, J.; Galatola, P. Anisotropic Capillary Interactions and
Jamming of Colloidal Particles Trapped at a Liquid-Fluid Interface.
Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2002,
65, 031601.
(27) Danov, K. D.; Kralchevsky, P. A. Capillary forces between
particles at a liquid interface: General theoretical approach and
interactions between capillary multipoles. Adv. Colloid Interface Sci.
2010, 154, 91103.
(28) Danov, K. D.; Kralchevsky, P. A. Electric forces induced by a
charged colloid particle attached to the water-nonpolar fluid interface.
J. Colloid Interface Sci. 2006, 298, 213231.
(29) Boneva, M. P.; Christov, N. C.; Danov, K. D.; Kralchevsky, P. A.
Effect of electric-field-induced capillary attraction on the motion of
particles at an oil-water interface. Phys. Chem. Chem. Phys. 2007, 9,
63716384.
(30) Boneva, M. P.; Danov, K. D.; Christov, N. C.; Kralchevsky, P. A.
Attraction between Particles at a Liquid Interface Due to the Interplay

of Gravity- and Electric-Field-Induced Interfacial Deformations.


Langmuir 2009, 25, 91299139.
(31) Danov, K.; Kralchevsky, P. A. Interaction Between Like-charged
Particles at a Liquid Surface: Electrostatic Repulsion vs. Capillary
Attraction. J. Colloid Interface Sci. 2010, 345, 505514.
(32) Foret, L.; Wurger, A. Electric-Field Induced Capillary
Interaction of Charged Particles at a Polar Interface. Phys. Rev. Lett.
2004, 92, 058302.
(33) Wurger, A.; Foret, L. Capillary Attraction of Colloid Particles at
an Aqueous Interface. J. Phys. Chem. B 2005, 109, 1643516438.
(34) Oettel, M.; Dominguez, A.; Dietrich, S. Effective Capillary
Interaction of Spherical Particles at Fluid Interfaces. Phys. Rev. E 2005,
71, 051401.
(35) Dominguez, A.; Oettel, M.; Dietrich, S. Theory of CapillaryInduced Interactions Beyond the Superposition Approximation. J.
Chem. Phys. 2007, 127, 204706.
(36) Oettel, M.; Dietrich, S. Colloidal Interactions at Fluid Interfaces.
Langmuir 2008, 24, 14251441.
(37) Park, B. J.; Pantina, J. P.; Furst, E. M.; Oettel, M.; Reynaert, S.;
Vermant, J. Direct Measurements of the Effects of Salt and Surfactant
on Interaction Forces between Colloidal Particles at Water-Oil
Interfaces. Langmuir 2008, 24, 16861694.
(38) Loudet, J.; Alsayed, A.; Zhang, J.; Yodh, A. Capillary Interactions
Between Anisotropic Colloidal Particles. Phys. Rev. Lett. 2005, 94,
018301.
(39) Loudet, J. C.; Yodh, A. G.; Pouligny, B. Wetting and Contact
Lines of Micrometer-Sized Ellipsoids. Phys. Rev. Lett. 2006, 97,
018304.
(40) Botto, L.; Lewandowski, E.; Cavallaro, M.; Stebe, K. J. Capillary
Interactions between Anisotropic Particles. Soft Matter 2012, 8, 9957.
(41) Danov, K.; Aust, R.; Durst, F.; Lange, U. Influence of the surface
viscosity on the hydrodynamic resistance and surface diffusivity of a
large Brownian particle. J. Colloid Interface Sci. 1995, 175, 3645.
(42) Danov, K. D.; Dimova, R.; Pouligny, B. Viscous drag of a solid
sphere straddling a spherical or flat surface. Phys. Fluids 2000, 12,
27112722.
(43) Pozrikidis, C. Particle motion near and inside an interface. J.
Fluid Mech. 2007, 575, 333357.
(44) Dimova, R.; Danov, K.; Pouligny, B.; Ivanov, I. Drag of a Solid
Particle Trapped in a Thin Film or at an Interface: Influence of Surface
Viscosity and Elasticity. J. Colloid Interface Sci. 2000, 226, 3543.
(45) Fischer, T. M.; Dhar, P.; Heinig, P. The viscous drag of spheres
and filaments moving in membranes or monolayers. J. Fluid Mech.
2006, 558, 451475.
(46) Sickert, M.; Rondelez, F.; Stone, H. Single-particle Brownian
dynamics for characterizing the rheology of fluid Langmuir
monolayers. EPL 2007, 79, 66005.
(47) Maestro, A.; Guzman, E.; Ortega, F.; Rubio, R. G. Contact angle
of micro-and nanoparticles at fluid interfaces. Curr. Opin. Colloid
Interface Sci. 2014, 19, 355367.
(48) Petkov, J. T.; Denkov, N. D.; Danov, K. D.; Velev, O. D.; Aust,
R.; Durst, F. Measurement of the drag coefficient of spherical particles
attached to fluid interfaces. J. Colloid Interface Sci. 1995, 172, 147154.
(49) Vassileva, N. D.; van den Ende, D.; Mugele, F.; Mellema, J.
Capillary forces between spherical particles floating at a liquid-liquid
interface. Langmuir 2005, 21, 1119011200.
(50) Ally, J.; Amirfazli, A. Magnetophoretic measurement of the drag
force on partially immersed microparticles at air-liquid interfaces.
Colloids Surf., A 2010, 360, 120128.
(51) Batchelor, G. Brownian diffusion of particles with hydrodynamic
interaction. J. Fluid Mech. 1976, 74, 129.
(52) Dalbe, M.-J.; Cosic, D.; Berhanu, M.; Kudrolli, A. Aggregation of
frictional particles due to capillary attraction. Phys. Rev. E 2011, 83,
051403.
(53) Lee, S.; Leal, L. Motion of a sphere in the presence of a plane
interface. Part 2 An exact solution in bipolar co-ordinates. J. Fluid
Mech. 1980, 98, 193224.
L

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Article

Langmuir
(54) Schneider, J. C.; ONeill, M. E.; Brenner, H. On the slow viscous
rotation of a body straddling the interface between two immiscible
semi-infinite fluids. Mathematika 1973, 20, 175196.
(55) Zabarankin, M. Asymmetric three-dimensional Stokes flows
about two fused equal spheres. Proc. R. Soc. London, Ser. A 2007, 463,
23292350.
(56) Vella, D.; Mahadevan, L. The Cheerios Effect. Am. J. Phys.
2005, 73, 817.
(57) Vella, D.; Metcalfe, P. D.; Whittaker, R. J. Equilibrium
conditions for the floating of multiple interfacial objects. J. Fluid
Mech. 2006, 549, 215224.
(58) Prevo, B.; Kuncicky, D.; Velev, O. Engineered deposition of
coatings From nano- and micro-particles: a brief review of convective
assembly at high volume fraction. Colloids Surf., A 2007, 311, 210.
(59) Burmeister, F.; Schafle, C.; Matthes, T.; Bohmisch, M.;
Boneberg, J.; Leiderer, P. Colloid Monolayers as Versatile Lithographic
Masks. Langmuir 1997, 13, 29832987.

DOI: 10.1021/acs.langmuir.5b02146
Langmuir XXXX, XXX, XXXXXX

Вам также может понравиться