Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

International Journal of Multiphase Flow 107 (2018) 82–103

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


journal homepage: www.elsevier.com/locate/ijmulflow

The low surface Péclet number regime for surfactant-laden viscous


droplets: Influence of surfactant concentration, interfacial slip effects
and cross migration
V. Sharanya a, G.P. Raja Sekhar a,∗, Christian Rohde b
a
Department of Mathematics, Indian Institute of Technology, Kharagpur 721302, India
b
Institute of Applied Analysis and Numerical Simulation, University of Stuttgart, Pfaffenwaldring 57, Stuttgart 70569, Germany

a r t i c l e i n f o a b s t r a c t

Article history: We consider the motion of a viscous drop in an arbitrary unsteady Stokes flow such that the surface of
Received 5 September 2017 the drop is fully covered with a stagnant surfactant layer. In particular the regime of low surface Péclet
Revised 26 February 2018
number is analyzed and we account for the effect interfacial slip on the overall behavior of the flow
Accepted 5 May 2018
field. The hydrodynamic problem is solved by the solenoidal decomposition method and the drag force is
Available online 14 May 2018
computed in terms of Faxen’s laws, using a perturbation ansatz in powers of the surface Péclet number.
MSC: The surface equation of the deformed sphere has been determined by an iterative method up to the first
76D05 order approximation. Analytical expressions for the migration velocity of the drop are likewise given.
76D45 Based on this analysis we can completely characterize various flow situations including a given ambient
76R50 flow as uniform flow, Couette flow and Poiseuille flow. Moreover, we find out that a surfactant-induced
82D15
cross-stream migration of the drop occurs towards the center-line in both, Couette and Poiseuille flow
Keywords: cases. The variation of the drag force and the migration velocity is computed for different parameters
Unsteady Stokes flow such as the Péclet number and the Marangoni number. Finally, the theoretical findings are validated on
Surfactant concentration with available experimental data.
Faxen’s laws
© 2018 Elsevier Ltd. All rights reserved.
Liquid drop
Migration velocity

1. Introduction variation in the interfacial tension gradient can be driven by


thermal (temperature), chemical (solute), electrical or electro
Understanding the motion of drops and bubbles is important chemical pressures. The study of droplet migration under the
for many industrial and chemical applications. Deformability, influence of any of the above is often referred as “active droplets”
inertia, external (thermal and chemical) gradients influence the and in the absence of any of these factors driven by an external
global flow field. The variation of temperature or the pres- hydrodynamic velocity or pressure is referred as “passive droplet”.
ence of surfactants causes variations in the interfacial gradient. Ramachandran et al. (2012) discussed the impact of interfacial
Young et al. (1959) were the first to study flow past a drop slip on the dynamics of a passive drop in a Stokes flow by using
by considering thermal effects. Subramanian and Balasubrama- a numerical approach based on the boundary integral method.
niam (2001) have computed the drag force in terms of Faxen’s Ramachandran and Gary Leal (2012) studied the effect of inter-
laws by considering the thermal effects in an axisymmetric Stokes facial slip on the drop deformation in a steady Stokes flow by
flow. Subramanian (1983) calculated the settling velocity of a using Navier slip boundary terms of Faxen’s laws for an oscil-
drop by considering thermal effects in a steady axisymmetric latory Stokes flow past a passive drop. The solution is obtained
flow. The unsteady motion of a vertically falling liquid drop in an using a double curl representation for the velocity in terms of
axisymmetric flow has been analyzed by Chisnell (1987). Dill and two scalars. This method is shown to be very useful while han-
Balasubramaniam (1992) have studied the thermocapillary migra- dling arbitrary flows past spherical geometries (see Choudhuri and
tion of a drop in an axisymmetric unsteady Stokes flow. In general Raja Sekhar, 2013; Choudhuri and Padmavati, 2014). Choudhuri and
Raja Sekhar (2013) have obtained the drift of a spherical active
∗ conditions. Choudhuri and Padmavati (2014) have calculated
Corresponding author.
E-mail addresses: sharanya@iitkgp.ac.in (V. Sharanya), the drag and torque in drop in a steady arbitrary Stokes flow
rajas@maths.iitkgp.ernet.in, rajas@iitkgp.ac.in (G.P.R. Sekhar), where the surface activity is due to the thermocapillary effects.
crohde@mathematik.uni-stuttgart.de (C. Rohde).

https://doi.org/10.1016/j.ijmultiphaseflow.2018.05.008
0301-9322/© 2018 Elsevier Ltd. All rights reserved.
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 83

Mandal et al. (2015) computed the shape of a drop by considering form similar to Faxen’s laws. For example, the recent study by
the interfacial slip effect in an arbitrary steady Stokes flow by Choudhuri and Raja Sekhar (2013) discussed thermocapillary mi-
using Lamb’s solution. The above literature indicates that the gration of a viscous spherical drop and obtained the corresponding
thermocapillary migration of spherical drops in a viscous fluid has Faxen’s laws. Consequently, Sharanya and Raja Sekhar (2015) have
been studied extensively taking into account of unsteady nature addressed thermocapillary migration of a spherical drop in an ar-
as well as internal effects (see Choudhuri and Raja Sekhar, 2013; bitrary unsteady Stokes flow. We are motivated by these studies
Sharanya and Raja Sekhar, 2015). Moreover, spherical drops in an and consider the motion of a viscous spherical drop whose in-
unbounded viscous fluid governed by arbitrary Stokes flow have terface is covered with a stagnant layer of surfactant in an ar-
also been studied. However, surfactant-coated spherical drops in bitrary unsteady Stokes flow. The arbitrary Stokes flow case is
arbitrary viscous flows have not been explored completely. considered by Pak et al. (2014), where they restrict the flow to
Surfactants are surface active agents that are adsorbed at a be steady, and the surfactant to coat the whole interface. The
fluid-fluid interface or at a liquid-gas interface, where they typ- slip reduces the deformation of a drop in a shear-type flow (see
ically lower the interfacial tension and cause a Marangoni ef- Ramachandran et al., 2012; Ramachandran and Gary Leal, 2012).
fect. It is observed that even a small amount of surfactant can Also, it is noted that due to this slip condition the disturbance
reduce the terminal velocity of a drop. For example, Leven and flow produced by a drop is expected to be weakened in magni-
Newman (1976) studied the effect of surfactants on the termi- tude. It may be noted that very limited experimental work is avail-
nal velocity of a drop in an axisymmetric flow. There it is as- able on the solutocapillary caused Marangoni migration of a drop
sumed that along the interface, the surfactant is governed by a at low surface Péclet number (see Bratukhin et al., 2005; Zuev and
convection-diffusion equation. Holbrook and Levan (1983a) and Kostarev, 2008). Bratukhin et al. (2005) were the first who con-
Holbrook and Levan (1983b) have used a collocation method to ducted an experiment to observe the effect of concentration of the
solve the convection-diffusion problem for high Péclet numbers dissolved components (surfactants or impurities) on the migration
and studied the retardation of drop motion in presence of surfac- of a drop. They have analyzed the effect of temperature and sur-
tant. Sadhal and Johnson (1983) studied the flow past a drop which factants separately and compared both thermocapillary and solu-
is partially coated with a stagnant layer of surfactant for large sur- tocapillary migrations. The study of active droplets has drawn the
face Péclet number. Many authors have examined the effect of sol- interest of researchers recently due to potential applications in bi-
uble and insoluble surfactants on the motion of drops using var- ology like locomotion of microorganisms and bacteria. Hence, un-
ious numerical techniques (see Oguz and Sadhal, 1988; Alke and derstanding self-propelling droplets under the influence of exter-
Bothe, 2007). Stone and Leal (1990) derived a convection-diffusion nal factors is very important. To this extent, in our present study,
equation for the surfactant transport along a deforming interface. we attempt a more generalized problem of an arbitrary transient
Stone and Leal (1990) used a numerical approach to analyze the Stokes flow past a drop for low surface Péclet number. Also, we
effect of surfactants on the deformation and breakup of a drop. take into account the effect of interfacial slip on the flow. We
Hanna and Vlahovska (2010) discussed the surfactant-induced mi- solve the problem for any given ambient flow and consider some
gration of a drop in an unbounded Poiseuille flow for large Péclet special cases to validate our results. We also show a comparison
numbers. A simplified CFD simulation was performed to study the with the experimental results of Bratukhin et al. (2005); Zuev and
influence of surfactants on the rise of bubbles by Fleckenstein and Kostarev (2008).
Bothe (2013). Recently, Pak et al. (2014) calculated the migration The objective of our present paper is to analyze the behav-
of a drop in a steady Poiseuille flow at low surface Péclet num- ior of the flow when the interfacial slip effect and the surfactant
bers. Das et al. (2017) investigated the axisymmetric motion of a concentration effect occurs for low surface Péclet numbers. We
surfactant-laden droplet in combined presence of linearly varying use the solenoidal decomposition method (double curl method)
temperature field and imposed Poiseuille flow. They have extended to solve the unsteady Stokes equations (see Venkatalaxmi et al.,
the work by considering arbitrary Poiseuille flow instead of ax- 2004). We use slip boundary conditions to see the effect of in-
isymmetric flow (see Das et al., 2018). terfacial slip on the flow behavior which has been previously
It was observed that a clean viscous drop at zero Reynolds used by Ramachandran et al. (2012) and Ramachandran and
number in a pressure driven flow moves only along the flow di- Gary Leal (2012). If we denote the surfactant concentration as  ,
rection without experiencing any cross migration in the absence we assume that  is governed by a convection-diffusion equation
of inertia and deformation (see Hetsroni and Haber, 1970). On the (see Sadhal and Johnson, 1983; Stone and Leal, 1990; Wong et al.,
other hand it was also observed that a viscous drop experiences 1996). We find the surfactant concentration up to second order for
cross migration due to inertia. It is experimentally observed that, an arbitrary Stokes flow, i.e., up to O(Pe2s ) (see Pak et al., 2014). We
for three dimensional Poiseuille flow and for Couette flow, the mi- observe area-specific surfactant distribution on the interface of the
gration due to deformation occurs towards the center line (see drop. We also solve for the flow fields and obtain the settling ve-
Goldsmith and Mason, 1962; Karnis and Mason, 1967; Chan and locity of the drop. We compute migration velocity corresponding
Leal, 1979). The cross migration due to inertial effects is also stud- to surfactant coated drop in Poiseuille flow and Couette flow and
ied by many authors (see Cox and Brenner, 1968; Ho and Leal, make some observations on the cross migration velocity. Further,
1974). It is also found that the surfactant redistribution can also the droplet is assumed as spherical as a first approximation and
cause cross stream migration of drops (see Hanna and Vlahovska, then the equation of the interface is obtained by using the itera-
2010; Stan et al., 2013; Pak et al., 2014). However, these studies tive method as in the work of Hetsroni and Haber (1970).
are restricted to the steady flows in ambient Poiseuille flow. The
literature cited above on surfactant coated viscous spherical drops 2. Problem statement and mathematical formulation
indicates that a generalized version of migration of viscous spher-
ical drop considering surfactant effects in an arbitrary Stokes flow We consider the motion of a liquid drop of radius a and vis-
is not explored. Hence, we generalize the problem to an unsteady cosity μi in an unsteady Stokes flow, suspended in another un-
arbitrary ambient flow, by considering the effects of interfacial slip bounded Newtonian fluid of viscosity μe (see Fig. 1). Let the ve-
as well as surfactant concentration effects. locity of the fluid inside the drop be vi and the velocity of the
We are interested in the case of arbitrary Stokes flow past drops fluid outside the drop be ve . We assume that the settling veloc-
which is challenging due to its three dimensional nature. Note that ity of the drop is U, which we determine later. The presence of
the corresponding drag and torque can be obtained in a compact a small amount of surface-active agents (surfactants) causes the
84 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

Fig. 1. Geometry of the problem.

variation in interfacial tension which influences the migration of which together with some pressure field p∞ satisfies the unsteady
the drop. We analyze the problem if effects caused by the surfac- Stokes and continuity equations.
tant’s concentration and interfacial slip become important. Surfac- We assume that the surfactant is confined to the interface. It
tants are surface-active agents that lower the interfacial tension cannot diffuse into the drop or bulk fluid. The surfactant trans-
between two liquids. We neglect the inertial terms under negli- port is governed by an unsteady convection-diffusion equation (see
gible Reynolds number assumption, and assume that the Capillary Sadhal and Johnson, 1983; Stone and Leal, 1990), which is given in
number μσUc is small enough for the drop to maintain a spherical the non dimensional form as follows
shape as the first approximation. We assume a low surface Péclet ∂  
P rs + P es ∇ s .(vs ) + (v.nˆ )∇
 s .nˆ = ∇ 2  , (5)
number Pes . Further, we assume that the (dimensional) interfacial ∂t s

tension, σ ∗ , depends in an affine way on the (dimensional) surfac- where vs is the velocity component tangential to the surface of
tant concentration,  ∗ , i.e., the drop and Pes = aU c
Ds is the surface Péclet number which mea-
σ ∗ = σ − RT  ∗ , (1) sures the importance of convection relative to diffusion. Here Ds is
the dimensional surface-diffusion constant. P rs = Dνs is the Prandtl
where σ is the interfacial tension when the interface is clean, R number which is dimensionless and is defined as the ratio of mo-
is the gas constant and T is the absolute temperature (Pak et al., mentum diffusivity to surfactant diffusivity, and ∇ s is the surface
2014). We non-dimensionalize the lengths by the drop radius a, ve- gradient operator, i.e.,
locities by the characteristic velocity scale of the background flow,
Uc , time by its characteristic time scale tc and the surfactant con- ∂ φˆ ∂

s = ∇
 − nˆ (nˆ .∇
 ) = θˆ + .
centration by its equilibrium value when the distribution is uni- ∂θ sin θ ∂φ
form,  eq . The pressure is non-dimensionalized by μaUc . Eq. (5) includes the convective and diffusive contribution to the
We assume that the flow inside and outside the drop is gov- surfactant transport and a source-like contribution accounting for
erned by the unsteady Stokes equations and the continuity equa- the variation of surfactant concentration resulting from the local
tions which are given in the non-dimensional form as follows changes in the interfacial area (see Stone and Leal, 1990).
(see Fig. 1 for the geometrical setting with r 2 = x2 + y2 + z2 , x = We solve the problem in a reference frame which is moving
r sin θ cos φ , y = r sin θ sin φ and z = r cos θ ): for r < 1 with the velocity of the drop, U, in which the drop appears to be
∂vi stationary (see Fig. 1). In this moving frame, the velocity fields in-
βi = −∇
 pi + ∇ 2vi ; ∇
 .vi = 0, (2) side and outside the drop are given respectively by
∂t
and for r > 1
 i = vi − U,
u
∂ve
βe = −∇
 pe + ∇ 2ve ; ∇
 .ve = 0. (3)
∂t  e = ve − U.
u
2 2
In the above equations, βe = νae tc and βi = νatc represent the un- One can observe that, these velocity fields also satisfy the un-
i
steadiness parameters corresponding to the flow inside and outside steady Stokes and continuity equations given by for r < 1
2
the drop respectively, which we assume to be unity, i.e., tc = aν . ∂ u i dU
j + = −∇
 pi + ∇ 2 u
i; ∇
 .u
 i = 0, (6)
We assume that the velocity field far from the drop approaches ∂t dt
the undisturbed background flow, v∞ , i.e., and for r > 1
∂ u e dU
ve → v∞ as r → ∞, (4) + = −∇
 pe + ∇ 2 u
e; ∇
 .u
 e = 0. (7)
∂t dt
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 85

 e is expected to meet the following far field


The external velocity u We have from the properties of Associated Legendre polynomials
condition in the reference frame that the function
n 
∞ 
 
e → u
u  ∞ = v∞ − U as r → ∞. (8) X (θ , φ ) = 0
Enm cos mφ + Fnm
0
sin mφ Pnm (cos θ ),
We follow the physical interpretations discussed by various au- n=0 m=0

thors (e.g. Subramanian and Balasubramaniam, 2001; Happel and will satisfy Eq. (15) provided
Brenner, 2012; Clift et al., 2005; Polyanin et al., 2001) and adopt ∞ 
 n  0 
the following kinematic boundary conditions on the surface of the (l02 − n(n + 1 )) Enm cos mφ + Fnm
0
sin mφ Pnm (cos θ ) = 0,
drop in non-dimensional form: n=0 m=0
Vanishing normal component of the velocities, i.e., (16)
 e .nˆ = 0; u
u  i .nˆ = 0, (9) This enables us to write the following relations
2π π
Slip in the tangential component of velocities, i.e., 2 ( n + m )! −1
0
Enm π = (0 )Pnm (cos θ)
 i .tˆ = ατnˆetˆ ,
 e .tˆ − u
u (10) (2n + 1 )(n − m )! n(n + 1 ) − l02 φ =0 θ =0
× cos mφ sin θ dθ dφ , (17)
where nˆ and tˆ denote unit normal and unit tangential vectors
respectively. τ eˆ denotes the tangential stress component exter-
nˆt 2π π
nal to the drop and similarly we denote the same quantity inter- 2 ( n + m )! −1
0
Fnm π = (0 )Pnm (cos θ)
nal to the drop by τ i ˆ . This boundary condition is obtained from (2n + 1 )(n − m )! n(n + 1 ) − l02 φ =0 θ =0
nˆt
the definition of the Navier slip boundary condition and is used × sin mφ sin θ dθ dφ , (18)
by Leal (2007) and Ramachandran and Gary Leal (2012). Finally,

α = aμαe is the dimensionless slip coefficient where α  is the di- except for n = 0, m = 0. Further, we have from Eq. (16) − n(n + l02
mensional slip coefficient. 1 ) = 0 for n = 0, m = 0. Therefore, we have that  0 is a constant,
Tangential stress balance, i.e., which we take as unity, i.e., 0 = 1.
We have solved the leading order arbitrary steady Stokes
τnˆetˆ − μτnˆi tˆ = Ma∇
 s  .tˆ, (11) equations through a representation via solenoidal decomposition
method (Venkatalaxmi et al., 2004). The details are given in
where  denotes the dimensionless surfactant concentration, μ =
μi RT eq
Appendix A. Accordingly, the zeroth order drag force experienced
μe is the viscosity ratio, Ma = μeUc is the Marangoni number by a spherical drop can be computed using the formula
which measures the relative importance of stress induced by sur- π 2π
factant concentration gradient to the viscous stress. Since the  =
D τ̄¯ .nˆ dS, (19)
stress fields and the surfactant concentration on the surface of the θ =0 φ =0
drop remain the same in both the laboratory frame and the mov- where dS represents the surface element, nˆ is the unit normal to
ing frame, the tangential stress balance takes the same form as in the boundary of the drop, and τ̄¯ is the stress tensor. We have com-
both reference frames. We note that the surfactant transport equa- puted the zeroth order drift in case of a transient Stokes flow past
tion given in Eq. (5) simplifies in the moving reference frame to a viscous drop. It is expressed in terms of Faxen’s law, given by
 

∂  0 = 4π λ2 αˆ 0 A0 iˆ + B0 jˆ + A0 kˆ eλ2e t ,
P rs + P es ∇  s ) = ∇s2  .
 s .( u (12) D (20)
∂t e 1 11 11 10

Here u
 s is the velocity tangential to the surface of the drop in the where iˆ, jˆ, kˆ are the unit vectors in the Cartesian coordinate sys-
moving frame. tem. Note that the above structure in terms of the known vec-
tor (A011 , B011 , A010 ) is due to the spherical harmonics Sn0 (θ , φ ) given
3. Method of solution in Eq. (A.21). Corresponding to a given ambient flow, one can de-
termine the coefficient αˆ 10 . For example, in case of uniform ambi-
We expand the velocity and pressure fields, surfactant concen- ent flow, we have n = 1 and the corresponding expression for αˆ 10
tration and migration velocity as a regular perturbation expansion can be obtained using αˆ n0 given in Appendix A. Consequently from
for low surface Péclet number (Pes  1), i.e., Eq. (20), we have the following expression for the drag force
   
 i , pe , pi ,  , U = u
e, u
u  i0 , pe0 , pi0 , 0 , U0
 e0 , u  0 = 2 π Y + μX + α P [ u
D  0∞ ]0 +
V + μU + α H
[∇ 2 u
 0∞ ]0 . (21)
  W + μZ + α G W + μZ + α G
 i1 , pe1 , pi1 , 1 , U1
 e1 , u
+ P es u
μi
  The above quantity depends on μ = μ e , the ratio of the viscosi-
+ Pe2s  i2 , pe2 , pi2 , 2 , U2 + O(Pe3s ). (13)
 e2 , u
u
ties, and α the dimensionless slip coefficient. Since  = 1, ∇  s
0 0
vanishes and the tangential stress becomes continuous. Hence at
3.1. Leading order problem leading order, we do not observe any influence of the surfactant.
The expanded form of the quantities X, Y, P, G, Z, W, U, V, H etc.,
The zeroth order surfactant transport equation corresponding to are given in Appendix B. It may be noted that the above compact
the general case given in Eq. (12) is form is due to the following relations
∂ 0
2


= ∇s2 0 .  0∞ ]0 = 2α10 + λe β10 A011 iˆ + B011 jˆ + A010 kˆ eλe t ,
2
P rs (14) [u
∂t 3
We follow Carslaw and Jaeger (1959) to obtain the solution of the
above equation. Correspondingly, we put 0 = e−l0 t/Prs X (θ , φ ), η =
2
2 3 0 0 ˆ
[∇ 2 u λ β (A i + B011 jˆ + A010 kˆ )eλe t ,
2
 0∞ ]0 =
cos θ to get 3 e 1 11
 
∂ 2 ∂X 1 ∂ 2X 2λe 0 0 ˆ
(1 − η ) + + l02 X = 0. (15) [∇
 ×u
 0∞ ]0 = 0 ˆ λ2e t
γ (C i + D011 jˆ + C10 k )e .
∂η ∂η 1 − η2 ∂φ 2 3 1 11
86 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103



One may observe that, when the slip coefficient in the zeroth order
1 = R1n (θ , φ )e−l1 t/Prs ,
2
drag force is equal to zero (i.e., α = 0), then the drag force reduces (30)
to n=0

 0 = 2 π Y + μX [ u V + μU where
D  0∞ ]0 + [∇ 2 u
 0∞ ]0 . (22) n 
W + μZ W + μZ  
R1n (θ , φ ) = 1
Enm cos mφ + Fnm
1
sin mφ Pnm (cos θ ), (31)
In the context of thermocapillary migration of a spherical drop, m=0
Sharanya and Raja Sekhar (2015) obtained an expression for the 1 , F 1 have to be determined
drag force exerted on a spherical drop. The above expression in are the spherical harmonics, and Enm nm
Eq. (22) agrees with their results when the thermocapillary ef- such that  1 satisfies the Eq. (28). Since ∇s2 R1n (θ , φ ) = −n(n +
∞ 2
fects are neglected. Note that the zeroth order drag force given in 1 )R1n (θ , φ ), we observe that ∇s2 1 = − n(n + 1 )R1n (θ , φ )el1 t/Prs .
Eq. (21) is with respect to a reference frame which is moving with n=0

a velocity U0 . Therefore the drag force in the laboratory reference The coefficients in R1n (θ , φ ) can be determined as follows:
frame in terms of a given ambient hydrodynamic field is given by ∞ 
 n  1 
Y + μX + α P (−n(n + 1 ) + l12 ) Enm cos mφ + Fnm
1
sin mφ
 0 = 2π V + μU + α H
D ([v0∞ ]0 −U0 ) + [∇ 2v0∞ ]0 , n=0 m=0
W + μZ+ α G W + μZ+ α G
×Pnm (cos θ )e−l1 t/Prs = ∇
2
 s .u
 0s . (32)
(23)
This enables us to write the following relations
where U0 is the zeroth order migration velocity which is yet to be
determined. 2 ( k + j )!
Ek1j π e−l1 t/Prs
2

The force balance in the absence of gravity when the flow (2k + 1 )(k − j )!
is transient is given by (see Subramanian and Balasubramaniam, 2π π
−1
2001; Chisnell, 1987), = (∇  0s )Pkj (cos
 s .u θ)
k(k + 1 ) − l φ =0 θ =0
2

dU × cos jφ sin θ dθ dφ , (33)


M =D
, (24)
dt
where M = 43 π ρi is the mass of the drop with unit radius. Here, ρ i
2 ( k + j )!
Fk1j π e−l1 t/Prs
2
is the density of the drop. From the above Eq. (24), we have the
leading order force balance as follows (2k + 1 )(k − j )!
2π π
−1
d U0 = (∇  0s )Pkj (cos
 s .u θ)
M = D0 . (25) k(k + 1 ) − l 2 φ =0 θ =0
dt
On using the expression for the drag given in Eq. (23) (for any × sin jφ sin θ dθ dφ , (34)
given ambient flow), this would enable us to obtain the following which implies −l12 /P rs = λ2e (< 0 ) and
expression for the migration velocity of the drop 

1
Enm = (n + 1 )αn0 + βn0 (λe fn+1 (λe ) + (n + 1 ) fn (λe ) )
3 Y + μX + α P V + μU + α H
U0 = [v0∞ ]0 + [∇ 2v0∞ ]0
2 ρi + ρe W + μ Z + α G W + μZ + α G − nαˆ n0 + βˆn0 ( (n + 1 )gn (λe ) − λe gn+1 (λe ) )

3 Y + μX + α P
−1  
+ λ2e . (26) n (n + 1 )
2 ρi + ρe W + μ Z + α G × A0nm , (35)
n(n + 1 ) + λ2e P rs
We observe that, when the slip coefficient α is zero, the above ze-
roth order migration velocity reduces to the one that is obtained 
1
Fnm = (n + 1 )αn0 + βn0 (λe fn+1 (λe ) + (n + 1 ) fn (λe ) )
by Sharanya and Raja Sekhar (2015) provided thermal effects are
neglected. In this case, we have
− nαˆ n0 + βˆn0 ( (n + 1 )gn (λe ) − λe gn+1 (λe ) )

Y + μX V + μU

U0 =
3
[v0∞ ]0 + [∇ 2v0∞ ]0
 
2 ρi + ρe W + μ Z W + μZ n (n + 1 )
× B0nm . (36)

3 Y + μX
−1 n(n + 1 ) + λ2e P rs
+ λ2e . (27)
2 ρi + ρe W + μ Z The first order hydrodynamic problem is solved in Appendix C. Fol-
lowing a similar approach that is used to solve the leading order
3.2. First-order correction problem, we compute the first order drag given by

 1 = 2π − Y + μX + α P U1 + 2Maλe f2 (λi )g1 (λe )
2
The first order surfactant transport equation due to the expan- D
W + μZ + α G (W + μZ + α G )
sion in Eqs. (13) and (12) is given by
1 ˆ λe t
× (E11 k )e
2
1 ˆ 1 ˆ
∂ 1 i + F11 j + E10 . (37)
−∇ 1 + P rs
2
= −∇
 s .u
 0s , (28)
s
∂t dU
The force balance M dt1 = D  1 together with the expression for D
1
where u 0s is the zeroth order tangential velocity vector on the sur-
given in Eq. (37) leads to the first order migration velocity of the
face of the drop. Since the non homogeneous term on the right
drop
hand side of the above equation is of the form g(θ , φ )eλe t , we seek
2
 
2Maλ2e f2 (λi )g1 (λe ) 1 ˆ
2
surfactant concentration of the form: 1 (θ , φ , t ) = 1 (θ , φ )e−l1 t/Prs 3 1 ˆ λ2e t
U1 = (E11 i + F111 jˆ + E10 k )e
(l12 is amplification factor), so that Eq. (28) reduces to the 2 ρi + ρe (W + μZ + α G )
(∇s2 + l12 )1 = ∇
 s .u
 0s . (29)

3 Y + μX + α P
−1
× + λ2e , (38)
In order to obtain the first order surfactant concentration  1 , we 2 ρi + ρe W + μ Z + α G
express  1 in terms of spherical harmonics, i.e., where
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 87

 
2A011  k (k + 1 ) (2k + 1 )(k − j )!
1
E11 = 2α10 + β10 (λe f2 (λe ) + 2 f1 (λe ) ) Ek1j −
(2 + λ2e Prs ) k(k + 1 ) + λ2e P rs 2 π ( k + j )!
2π π
−λ2e t
− αˆ 10 + βˆ10 (2g1 (λe ) − λe g2 (λe ) ) , (39) e
× (∇ s .(1 u
 0s ))
k(k + 1 ) + λ2e P rs φ =0 θ =0

2B010  ×Pkj (cos θ ) cos jφ sin θ dθ dφ , (48)


1
F11 = 2α10 + β10 (λe f2 (λe ) + 2 f1 (λe ) )
(2 + λ 2 Pr
s )
e

− αˆ 10 + βˆ10 (2g1 (λe ) − λe g2 (λe ) ) , (40) Fk2j = −kαˆ k2 + βˆk2 ( (k + 1 )gk (λe ) − λe gk+1 (λe ) )
 
and k (k + 1 ) (2k + 1 )(k − j )!
× Fk1j −
2A010  k(k + 1 ) + λ2e P rs 2 π ( k + j )!
1
E10 = 2α10 + β10 (λe f2 (λe ) + 2 f1 (λe ) )
(2 + λ2e Prs ) e−λe t
2 2π π
× (∇
 s .(1 u
 0s ))
− αˆ 10 + βˆ10 (2g1 (λe ) − λe g2 (λe ) ) , (41) k(k + 1 ) + λe P rs φ =0 θ =0
2

×Pkj (cos θ ) sin jφ sin θ dθ dφ . (49)


Here we observe that, only three modes of concentration 1 ,
E11 1
F11
1 are contributing to the drag and migration velocity.
and E10 Evaluating the double integral on the right hand side is difficult for
any arbitrary flow. However, these can be evaluated for a specific
ambient flow so that we have the second order concentration. Ac-
3.3. Second-order correction
cordingly, we compute these double integrals for specific cases like
uniform flow, Couette flow etc.
The second order surfactant transport equation is given by
Once we obtain the second order concentration for a given flow,
∂ 2 one can solve the above equations by following similar procedure
−∇s2 2 + P rs = −∇
 s .(0 u
 1s + 1 u
 0s ), (42)
∂t that is used to solve the zeroth and first order equations. The sec-
ond order drag is given by
where u  0s , u
 1s are the zeroth order and first order tangential

velocity components on the drop surface, respectively. Following
 2 = 2π Y + μX + α P 2Maλ2e f2 (λi )g1 (λe )
the similar arguments as given for the first order correction, we D (−U2 ) +
W + μZ + α G (W + μZ + α G )
can assume that the second order concentration is of the form
2

2 (θ , φ , t ) = 2 (θ , φ )e−l2 t/Prs . Then, the Eq. (42) reduces to × (E11
2 ˆ 2 ˆ
i + F11 2 ˆ λ2e t
j + E10 k )e . (50)
(∇s2 + l22 )2 = ∇
 s .(0 u
 1s + 1 u
 0s ). (43)
dU2
The force balance M dt
= D2 leads to
In order to obtain the second order surfactant concentration,  2 ,  
we adopt a similar procedure that is used in Section 3.2. We ex- 3 2Maλ2e f2 (λi )g1 (λe ) 2 ˆ 2 ˆ −λ2e t
press  2 in terms of spherical harmonics, i.e., U2 = (E11 i + F112 jˆ + E10 k )e
2 ρi + ρe (W + μZ + α G )


3 Y + μX + α P
−1
2 = R2n (θ , φ )e−l2 t/Prs ,
2
(44) + λ2e , (51)
n=0 2 ρi + ρe W + μ Z + α G
where where X, Y, P, G, Z, W, U, V, H etc., are given in the Appendix B.
We therefore, conclude that the second order migration velocity

n  
R2n (θ , φ ) = 2
Enm cos mφ + Fnm
2
sin mφ Pnm (cos θ ), (45) and drag depend only on three modes of the concentration namely,
2 , F 2 and E 2 .
E11
m=0 11 10
2 , F 2 have to be determined such that  satisfies the
and Enm nm 2
Eq. (43). Correspondingly, the coefficients R2n (θ , φ ) can be deter- 4. Migration of a surfactant coated spherical drop with
mined as follows: interfacial slip in steady Stokes flow

2 ( k + j )!
Ek2j π e−l2 t/Prs If we consider the special case of steady flow past a drop, i.e.,
2

(2k + 1 )(k − j )! λe = λi = 0, the leading order migration velocity Eq. (26) reduces
2π π
−1 to
= (∇
 s .(0 u
 1s + 1 u
 0s ))
μ
k(k + 1 ) − l22 φ =0 θ =0 U0 = [v0∞ ]0 + [∇ 2v0∞ ]0 . (52)
(4 + 6μ + 12αμ )
×Pkj (cos θ ) cos jφ sin θ dθ dφ , (46)
Up to our knowledge the above migration velocity is obtained for
the first time in the literature. If we consider a zero slip coefficient,
2 ( k + j )! i.e., α = 0, then the zeroth order terminal velocity reduces to
Fk2j π e−l2 t/Prs
2

(2k + 1 )(k − j )! μ
2π π U0 = [v0∞ ]0 + [∇ 2v0∞ ]0 , (53)
−1 4 + 6μ
= (∇
 s .(0 u
 1s + 1 u
 0s ))
k(k + 1 ) − l22 φ =0 θ =0 which is exactly matching with the one that is obtained by
×Pkj (cos θ ) sin jφ sin θ dθ dφ , (47) Pak et al. (2014).
If we consider the special case of steady flow past a droplet, i.e.,
which implies −l22 /P rs = λ2e , and λe = λi = 0, the first order migration velocity reduces to

Ek2j = −kαˆ k2 + βˆk2 ( (k + 1 )gk (λe ) − λe gk+1 (λe ) ) 2Ma
U1 = (e1 iˆ + f11
1 ˆ
j + e110 kˆ ), (54)
3(2 + 3μ + 6αμ ) 11
88 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

where We find the variation δ F in the area when f changes as follows


2π π π 2π    
2 ( k + j )! −1 1 ∂ ∂f
e1k j π = (∇  0s )
 s .u
δF = 2a ( 1 + f ) + a − sin θ
(2k + 1 )(k − j )! k(k + 1 ) φ =0 θ =0 θ =0 φ =0 sin θ ∂θ ∂θ

×Pkj (cos θ ) cos jφ sin θ dθ dφ , (55) ∂2 f 1
− δ (a f ) sin θ dθ dφ . (62)
∂φ 2 sin2 θ
2π π
2 ( k + j )! −1 Thus we get
fk1j π = (∇  0s )
 s .u
  
(2k + 1 )(k − j )! k(k + 1 ) φ =0 θ =0
1 1 2 2f 1 1 ∂2 f 1 ∂ ∂f
×Pkj (cos θ ) sin jφ sin θ dθ dφ . + = − − + sin θ ,
(56) R1 R2 a a a sin θ ∂φ
2 2 sin θ ∂θ ∂θ
It may be noted that the structure of the above expressions (63)
matches with the corresponding expressions (Eqs. (4.10) and
where R1 , R2 are the principle radii of curvature. We expand f(θ , φ ,
(4.12)) of Pak et al. (2014). Hence, if the slip coefficient α =
t), which denotes the deformation of the sphere, as a regular per-
0, the expression for the first order migration velocity given
turbation expansion for low surface Péclet number (Pes  1), i.e.,
in Eq. (54) agrees with the corresponding one obtained by
Pak et al. (2014) (Eqs. (4.10) and (4.12)). f (θ , φ , t ) = f0 (θ , φ , t ) + Pes f1 (θ , φ , t ) + Pe2s f2 (θ , φ , t ) + O(Pe3s ).
If we consider the special case of steady flow past a drop, i.e., (64)
λe = λi = 0, the second order migration velocity reduces to
We expand each fj (θ , φ , t) in terms of spherical harmonics as
2Ma
U2 = (e2 iˆ + f11
2 ˆ
j + e210 kˆ ), (57) ∞

6 + 9μ + 18αμ 11
f j (θ , φ , t ) = Lnj (t )Snj (θ , φ ), (65)
where n=1
2π π
−1 j
e2nm = (∇
 s .(0 u
 1s + 1 u
 0s )) where Ln ’s are constants to be determined. Hence the shape of the
n(n + 1 ) φ =0 θ =0 surface is given by
×Pnm (cos θ ) cos mφ sin θ dθ dφ , (58)  

 ∞

r = a 1+ L0n (t )Sn0 (θ , φ ) + Pes L1n (t )Sn1 (θ , φ ) + O(Pe2s ).
2π π n=1 n=1
−1
2
fnm = (∇
 s .(0 u
 1s + 1 u
 0s )) (66)
n(n + 1 ) φ =0 θ =0
×Pnm (cos θ ) sin mφ sin θ dθ dφ . (59) The normal stress balance in the dimensional form is given by

1 1

It may be noted that the structure of the above expressions resem- μi τnˆi nˆ − μe τnˆenˆ = σ ∗ + . (67)
bles with the corresponding expressions (Eqs. (4.17) and (4.18)) of R1 R2
Pak et al. (2014). Hence, if the slip coefficient α = 0, the expression We non-dimensionalize the above equation to get
for the first order migration velocity given in Eq. (57) agrees with
1
1 1

the corresponding one obtained by Pak et al. (2014) (Eqs. (4.17) and μτnˆi nˆ − τnˆenˆ = − Ma + . (68)
Ca R1 R2
(4.18)).
Using the Eqs. (63) and (66), we get
5. The equation of the interface 2 1


(μτnˆi nˆ )0 − (τnˆenˆ )0 = − 2Ma0 + − Ma0
Ca Ca
In this section, we assume that the interface of the droplet ∞

changes slightly about its mean position. Solving both, the flow (n(n + 1 ) − 2 )L0n (t )Sn0 (θ , φ ). (69)
field and the equation of the surface of the deformed sphere, si- n=2
multaneously is mathematically intractable. Hence, we follow a
Substituting the values of (τnˆi nˆ )0 and (τnˆenˆ )0 , we obtain for the
similar procedure as explained in Hetsroni and Haber (1970) to
leading order interface equation,
find the equation of the interface. We first solve the equations for
constant radius droplet to find the flow fields at first-order ap- 2
−μ pi 0∞ + pe 0∞ = − 2Ma0 , (70)
proximation, and then we use the normal stress balance to find Ca
the equation of the interface. We can use this interface equation to and
find the velocity fields at this deformed interface iteratively.
Ca
The area of a given surface given in spherical coordinates by a L0n (t ) =
function r = r (θ , φ , t ) is by Landau and Lifshitz (1987)
(1 − CaMa )(n(n + 1 ) − 2 )
 λ2 t  2 
  2  2 μe i λi (n + 1 )ᾱn0 + 2n(n + 1 ) (n − 1 )ᾱn0
π 2π 
∂r 1 ∂r + β̄n0 (λi fn+1 (λi ) + (n − 1 ) fn (λi ))
F = r2 + + r sin θ dθ dφ . (60)
θ =0 φ =0 ∂θ sin
2
θ ∂φ 2  
− eλe t λ2e ((n + 1 )αn0 − nαˆ n0 ) + 2n(n + 1 ) (n − 1 )αn0
Substituting r = a(1 + f (θ , φ , t )) (where max |f(θ , φ )|  1) in
− (n + 2 )αˆ n0 + βn0 (λe fn+1 (λe ) + (n − 1 ) fn (λe ))
Eq. (60), we obtain 
   2
π 2π + βˆn0 (−λe gn+1 (λe ) + (n − 1 )gn (λe )) . (71)
1 ∂f
F ≈ ( a + a f )2 + a2
θ =0 φ =0 2 ∂θ From Eq. (69), we can observe that for n = 1, we get an identity.
 2  This indicates that the assumption of a spherical drop remains cor-
1 ∂f rect up to two terms in Eq. (66), irrespective of the Capillary num-
+ a2 sin θ dθ dφ . (61)
2
sin θ ∂φ ber. This result is analogous to that of Hetsroni and Haber (1970).
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 89

This also shows that the assumption of spherical drop is true for The migration velocity is given by
creeping flows, irrespective of the Capillary number, up to first or-
U = U0 + Pes U1 + Pe2s U2 + O(Pe3s ). (77)
der.
Proceeding in a similar manner as in leading order solution In this case the zeroth order migration velocity U0 , given in
method, we obtain for the next order constant Eq. (26) reduces to
1

Enm 3 Y + μX + α P
L1n (t ) U0 =
1
Fnm 2 ρi + ρe W + μ Z + α G
−MaCa2 (2n + 1 )(n − m )!

3 Y + μX + α P
−1
= × + λ2e [v0∞ ]0 , (78)
(1 − CaMa )2 (n(n + 1 ) − 2 ) 2π (n + m )! 2 ρi + ρe W + μ Z + α G
π 2π  
  m cos mφ
where [v0∞ ]0 = eλe t iˆ. The first order migration velocity U1 , given
2
× 1 (μτnˆ nˆ )0 − (τnˆ nˆ )0 Pn (cos θ )
i e
θ =0 φ =0 sin mφ in Eq. (38) reduces to
1
⎛ ⎞  
Enm 3 2Maλ2e f2 (λi )g1 (λe )
2MaCaeλe t
2

sin θ dθ dφ + ⎝F 1 ⎠ U1 =
(1 − CaMa )(n(n + 1 ) − 2 ) nm 2 ρi + ρe (W + μZ + α G )

Y + μX + α P
−1
⎛1
⎞ ×
3
+ λ2e 1 λe t ˆ
E11
2
e i, (79)
Enm 2 ρi + ρe W + μ Z + α G
Ca
+ ⎝ 1
Fnm ⎠
(1 − CaMa )(n(n + 1 ) − 2 ) and the second order migration velocity U2 , given in Eq. (51) re-
duces to
 2    
μeλi t λ2i (n + 1 )ᾱn1 + 2n(n + 1 ) (n − 1 )ᾱn1 3 2Maλ2e f2 (λi )g1 (λe )
 U2 =
2 ρi + ρe (W + μZ + α G )
+ β̄n1 (λi fn+1 (λi ) + (n − 1 ) fn (λi ))
2  
3 Y + μX + α P
−1
− eλe t λ2e ((n + 1 )αn1 − nαˆ n1 ) + 2n(n + 1 ) (n − 1 )αn1 − (n + 2 )αˆ n1 × + λ2e 2 λe t ˆ
E11
2
e i. (80)
2 ρi + ρe W + μ Z + α G
+ βn1 (λe fn+1 (λe ) + (n − 1 ) fn (λe ))
 Here,
+ βˆn1 (−λe gn+1 (λe ) + (n − 1 )gn (λe )) . (72) 
1
1
E11 =− 3g1 (λe )(U0x − 1 )λ2e ( f2 (λi )(2αμ − 1 )
(2 + λ2e Prs )
6. Results and discussion 
− α f1 (λi )μλi ) ) g1 (λe )λ2e ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi )
We have so far derived leading order, first order and second or- + g2 (λe )λe (2 f2 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )
der terms for the migration velocity, drag force and concentration +3g1 (λe )( (2α + 1 ) f1 (λi )μλi − 2 f2 (λi )(2αμ + μ − 1 ) )),
as functions of α , μ, Ma, λe , λi . We have further considered limit-
(81)
ing cases when α → 0 (absence of interfacial slip) and λe = λi = 0
(steady case) and observed that these agree with those derived by and
Pak et al. (2014). Now, we present important observations with ref-
2E111
f2 (λi )Ma
erence to some special cases such as uniform ambient flow, Cou- 2
E11 = (3g (λ )(ρ + 2ρi )
ette flow and Poiseulle flow. ( ρe + 2 ρi + 3 ) 1 e e
 
− g2 (λe )λe )/ P rs λ2e + 2 g1 (λe )λ2e (ρe + 2ρi + 3 )
6.1. Analysis of uniform ambient flow case
( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi )
Consider a uniform flow along the x−axis, past a liquid drop of + g2 (λe )λe (ρe + 2ρi + 3 )(2 f2 (λi )(2αμ + μ − 1 )
unit radius whose center is at its origin. In this case, v∞ = v0∞ = − (2α + 1 ) f1 (λi )μλi ) − 3g1 (λe )(ρe + 2ρi )(2 f2 (λi )
iˆeλe t (non-dimensional form).
2
×(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )) ) (82)
Therefore, the corresponding scalar functions χ0∞ and η0∞ are
given by It is noted that the corresponding migration velocity vector sur-
vives only in the direction of the flow.
1 − U0x
χ0∞ θ cos φ eλ , η0∞ = 0.
2
= r sin et We have shown the surfactant distribution on the surface of the
2
drop at different times (please refer to Fig. 2). It may be noted
1−U
The above choice indicates that α10 = 2 0x , β10 = 0, γ10 = 0 in that the net surfactant distribution on the drop decreases with
Eqs. (A.19) and (A.20). Therefore the corresponding surfactant con- time. The migration velocity that is computed depends on vari-
centration distribution on the spherical drop is given by ous parameters such as Pes , μ, α , Ma. Correspondingly, we have
shown the variation of migration velocity U = U0 + Pes U1 + Pe2s U2
 = 0 + Pes 1 + Pe2s 2 + O(Pe3s ), (73)
with Pes for various slip coefficients (Fig. 3) and with Ma for vari-
where ous viscosity ratios (Fig. 4). For a fixed viscosity ratio, the slip pa-
rameter α reduces the resistance exerted by the drop. Accordingly,
0 = 1, (74)
the migration velocity increases, what can also be seen from Fig. 3.
Furthermore, we remark that the surface Péclet number measures
the importance of convection relative to diffusion. Therefore, as Pes
1 = E11 sin θ cos φ eλe t ,
2
1
(75)
increases, the migration velocity increases. The same is observed in
Fig. 3. With increasing viscosity ratio, the drop behaves like a solid
 2  2 and hence, the migration velocity decreases.
2 = E11 sin θ cos φ + E20
2 0
P2 (cos θ ) + E22
2
cos 2φ P22 (cos θ ) eλe t . The Marangoni number Ma is the ratio of surfactant concentra-
(76) tion effects to the convection forces and viscous forces. For a fixed
90 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

Fig. 2. Variation of the second order surfactant distribution with the time t corresponding to uniform flow, with λ2e = −0.04, λ2i = −0.04, μ = 5, Ma = 1 and α = 0.1.

viscosity ratio, with increasing Ma, the surfactant concentration ef-


fects dominate the convection forces. As a consequence, the mi-
gration velocity increases with Ma. The same is observed in Fig. 4.
From the solution presented in this manuscript, we note that the
first order terms of the terminal velocity are proportional to Pes Ma
and the second order terms are proportional to Pe2s Ma2 . Since we
are expanding the concentration in low Pes , this implies, respec-
tively, Pes Ma  1 and Pes  1. We may refer to Pak et al. (2014) for
a similar assumption. Correspondingly, we have followed this re-
striction and chosen parameter range while showing graphical re-
sults.

6.1.1. Comparison with the experimental results


In this section, we present comparison of our results with the
available experimental results. We pick up the migration veloc-
ity (as given in Eqs. (77)–(80)) and concentration (as given in
Eqs. (73)–(76)), in order to make this comparison. We have fixed all
the non-dimensional variables by using the data given in Table 1
(taken from Table 3 of Bratukhin et al., 2005) while showing the
Fig. 3. Variation of the migration velocity with Pes for different slip parameters
graphical results in this section. Relevant experimental data pre-
corresponding to uniform flow, α , λ2e = −0.01, λ2i = −0.01, Ma = 1 and μ = 5.
sented as figures in Bratukhin et al. (2005) is extracted by using
the GetData Graph Digitizer 2.26 software. In Fig. 5, we display the
variation of the dimensionless migration velocity (Eq. (77)) with
the Marangoni number for an air bubble, with the methanol con-
centration gradient 2.5%/cm. Here, the experimental data is taken
from the Fig. 15 of Bratukhin et al. (2005). One observes a strik-
ing coincidence of our results with the experimental results. In
Fig. 6, we show the time dependence on the dimensional migration
velocity with an air bubble of diameter 7 mm, in methanol con-
centration gradient 2.5%/cm. Here, the experimental data is taken
from the Fig. 16 of Bratukhin et al. (2005). We see a good qual-
itative agreement. In Fig. 7, we see the variation of the dimen-
sional migration velocity (normalized to bubble radius divided by
the solution dynamic viscosity) with the gradient of methanol con-
centration. Here, the experimental data is taken from Fig. 14 of
Bratukhin et al. (2005). We can see a good agreement of our results
with the experiments conducted by Bratukhin et al. (2005) (see
Figs. 5–7).
Next, in Fig. 8, we see the variation of dimensionless migra-
tion velocity with time. Here, the experimental data is taken from
Fig. 4. Variation of the migration velocity with Marangoni number (Ma) for dif- the Fig. 11 of Zuev and Kostarev (2008). We can see a good agree-
ferent viscosity ratios corresponding to uniform flow, μ, λ2e = −0.04, λ2i = −0.04, ment of our results with the experiment conducted by Zuev and
α = 0.2 and Pes = 0.01. Kostarev (2008).
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 91

Table 1
Physical data of methanol aqueous solution at 20 °C taken from Bratukhin et al. (2005).

Fluid  (%) ρ , 103 (kg/m3 ) viscosity, 10−3 (kg/(m s)) ν , 10−6 (m2 /s) D, 10−10 (m2 /s)

Methanol 80 0.8469 1.140 1.346 –


90 0.8202 0.860 1.049 –
100 0.7917 0.580 0.733 12.8

Fig. 5. Variation of the dimensionless migration velocity with different Marangoni Fig. 7. Variation of the migration velocity (normalized to bubble radius divided by
numbers (Ma) for an air bubble. The methanol concentration gradient is 2.5%/cm. the solution dynamic viscosity) with the gradient of methanol concentration. Ex-
Experimental data is taken from the Fig. 15 of Bratukhin et al. (2005). perimental data is taken from the Fig. 14 in Bratukhin et al. (2005).

Fig. 8. Variation of the dimensionless migration velocity with an air bubble of di-
Fig. 6. Time dependence of dimensional migration velocity with an air bubble of
ameter 7 mm, methanol concentration gradient 2.5%/cm. Experimental data is taken
diameter 7 mm. The methanol concentration gradient is 2.5%/cm. Experimental data
from the Fig. 11 in Zuev and Kostarev (2008).
is taken from the Fig. 16 of Bratukhin et al. (2005).

F L−U
The above choice indicates that α10 = 2
0x
, α20 = 36
F
, β10 = 0,
6.2. Couette flow
γ1 = 2λe from Eqs. (A.19) and (A.20). Therefore the corresponding
0 −3F

Consider a Couette flow past a liquid drop of unit radius whose surfactant concentration distribution on the spherical drop is given
center is at the origin (see Fig. 9). In this case, v∞ = (F (y + L )iˆ)eλe t , by
2

where L is the distance of the center of the droplet from the point  = 0 + Pes 1 + Pe2s 2 + O(Pe3s ), (83)
of zero velocity and F is the shear rate (Hetsroni and Haber, 1970).
Therefore the corresponding scalar functions χ0∞ and η0∞ are given where
by 0 = 1, (84)

FL − U
F 2  1  2
χ0∞ = 0x
r P11 (cos θ ) cos φ + r sin 2φ P22 (cos θ ) eλe t ,
2

2 36 1 = E11 sin θ cos φ + F22


1
sin 2φ P22 (cos θ ) eλe t , (85)
−3F  2
η0∞ = f1 (λe r )P1 (cos θ ). 2 = E11 sin θ cos φ + F11
2
sin θ sin φ + F22
2
sin 2φ P22 (cos θ )
2λe
92 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

Fig. 9. (a) Geometry of the problem and velocity vector corresponding to Couette ambient flow, (b) surface velocity vector field corresponding to Couette flow.

2
+ F31 sin φ P31 (cos θ ) + F332 sin 3φ P33 (cos θ ) + E20
2 0
P2 (cos θ ) concentration is constant forcing vanishing concentration gradient.
Accordingly, at leading order, we recover the case of a clean spher-
2
+ E22 cos 2φ P2 (cos θ ) + E40 P4 (cos θ )
2 2 0
 2 ical drop in an ambient flow (characterized by the velocity scale
2
+ E44 cos 4φ P44 (cos θ ) eλe t , (86) Uc ). It is well known that there can be no cross stream migra-
1 , F 1 etc. are listed in Appendix D. tion in the absence of inertia, deformation, or surfactant concen-
where the quantities E11 22 tration (Pak et al., 2014). This fact has been established now for
For an unbounded Couette flow, the migration velocity of a
the case of drop in an unbounded Couette flow. This phenomenon
force free drop is calculated as
can also be supported mathematically as follows. The dimensional
U = U0 + Pes U1 + Pe2s U2 + O(Pe3s ). (87) zeroth order migration velocity satisfies U∗0 = U0Uc ∝ Uc . Hence, re-
versal of the background flow (Uc → −Uc ) changes the direction of
In this case the zeroth order migration velocity U0 , given in
U∗0 . This prevents cross migration since cross migration should oc-
Eq. (26) reduces to
cur towards the center line. Further, in case of first order, we have

Y + μX + α P

3 U∗1 = Pes U1Uc ∝ Pes MaE11
1 U ∝ Pe MaU . The product Pe Ma is inde-
c s c s
U0 =
2 ρi + ρe W + μ Z + α G pendent of Uc which implies that, if U∗ . jˆ
= 0, the same is pro-
1

3 Y + μX + α P
−1 portional to Uc . Hence, following similar arguments as in the case
+ λ2e F Leλ iˆ.
2
× et (88)
2 ρi + ρe W + μ Z + α G of leading order, one can conclude that there cannot be any cross
stream migration in this case as well.
The first order migration velocity U1 , given in Eq. (38) reduces to However, turning to the second order migration velocity reveals
  that U∗2 .iˆ = Pe2s U2Uc .iˆ ∝ Pe2s MaE11
2 U ∝ Pe2 Ma2U ∝ U , and U∗ . jˆ =
1 ˆ
3E11 i 2Maλ2e f2 (λi )g1 (λe ) c s c c 2
U1 = Pes U2Uc . jˆ ∝ Pes MaF Uc ∝ Pes MaUc ∝ Uc2 . Therefore, the transverse
2 2 2 2
2 ρi + ρe (W + μZ + α G ) 11
migration is invariant upon reversal of the direction of the ambi-

3 Y + μX + α P
−1
ent Couette flow.
+ λ2e eλe t ,
2
× (89)
2 ρi + ρe W + μ Z + α G We also note that the transverse migration depends linearly on
1 ∝ L) which respects the symmetry requirement that the
L (as E11
and the second order migration velocity U2 , given in Eq. (51) re-
direction of the transverse migration should reverse when the drop
duces to
  is placed at the same distance but on the opposite side with re-
3(E11
2 ˆ
j ) 2Maλ2e f2 (λi )g1 (λe )
2 ˆ
i + F11 spect to the center of the Couette flow (see Fig. 9). The same is
U2 =
2 ρi + ρe (W + μZ + α G ) noted by Pak et al. (2014) for the case of Poiseuille flow.

3 Y + μX + α P
−1 We can see the drop deformation with the increasing Capillary
+ λ2e eλe t .
2
× (90) number in Fig. 12. The first and second order surfactant distribu-
2 ρi + ρe W + μ Z + α G tions for the case of ambient Couette flow are plotted with specific
By symmetry, it is expected that there will be no velocity com- values of parameters for visualization in Figs. 10 and 11. Here, we
have seen that the concentration increases with L (since E11 1 ∝ L).
ponent in z-direction. In case if there is a cross migration (i.e.,
motion transverse to the flow direction), the same occurs towards This may be due to enhanced convection far from the origin. We
the center line and should be in the y-direction (see Fig. 9). In have observed the variation of axial migration velocity and cross-
Ref. Pak et al. (2014), a detailed explanation on cross migration stream migration velocity for different parameters. We emphasize
of a surfactant coated viscous drop in Poiseuille flow is presented. that the cross migration is due to the linear term, and the axial mi-
Similar arguments are deduced in Mandal et al. (2015) while dis- gration is due to the constant term present in the ambient Couette
cussing the migration of deformed drop with interfacial slip in an flow. As a consequence, the axial migration velocity in the case of
unbounded Poiseuille flow. We also follow similar arguments to Couette flow behaves in the same manner as the migration veloc-
show that cross migration occurs only at second order with re- ity in the case of uniform ambient flow.
spect to the expansion of the migration velocity in terms of the We have observed the variation of cross migration velocity with
surface Péclet number. We have shown that at leading order the the amplification factor λe (see Fig. 13). The cross migration ve-
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 93

Fig. 10. Variation of the first order surfactant distribution for different values of L corresponding to Couette flow, with λ2e = −0.04, λ2i = −0.04, μ = 5, Ma = 1, F = 1, t = 1,
Pes = 0.01 and α = 0.1.

Fig. 11. Variation of the second order surfactant distribution for different values of L corresponding to Couette flow, with λ2e = −0.04, λ2i = −0.04, μ = 5, Ma = 1, F = 1,
t = 1, Pes = 0.01 and α = 0.1.

Fig. 12. Variation of the drop interface for different values of Ca, Capillary number, corresponding to Couette flow, with λ2e = −0.04, λ2i = −0.04, μ = 0.5, Ma = 1, F = 1,
t = π /2, Pes = 0.01 and α = 0.1.
94 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

Fig. 13. Variation of the cross migration velocity with iλe for different viscosity Fig. 15. Variation of the cross migration velocity with Pes for different slip parame-
ratios corresponding to Couette flow, μ, with λ2i = −0.04, α = 0.2, Ma = 1, F = 1, ters corresponding to Couette flow, α , with λ2e = −0.04, λ2i = −0.04, Ma = 1, F = 1,
L = 2 and Pes = 0.01. L = 2 and μ = 5.

b2 + 2br sin θ cos φ , and the velocity is non-dimensionalized with


the characteristic velocity Ub , which is at a dimensionless distance
b from the drop, and R0 is the dimensionless distance to the point
of zero velocity of the flow, λe is the amplification factor. We ex-
panded v∞ in the form of a series for small λe to get χ0∞ and η0∞ .
These are given by
 


χ = β12 f1 (λe r )S1 (θ , φ ) +

αn2 r Sn (θ , φ ) eλe t , η0∞ = 0,
2
n
0
n=1

η0∞ = γ10 f1 (λe r )T10 (θ , φ )eλe t ,


2

where,
1

1
−1  b2 λ2e

α10 = − 1− 1− − U0z /2, (91)
2 J0 (iλe R0 ) 4J0 (iλe R0 )

1
−1  bλ2e

α =− 1−
0
, (92)
2
J0 (iλe R0 ) 36J0 (iλe R0 )
Fig. 14. Variation of the cross migration velocity with viscosity ratio, μ, for dif-
ferent α corresponding to Couette flow with λ2e = −0.04, λ2i = −0.04, α = 0.1,
 
−1
λ2e 1
Ma = 1, F = 1, L = 4 and Pes = 0.01. α =
0
1− , (93)
3
120J0 (iλe R0 ) J0 (iλe R0 )

−3

1
−1
locity show an oscillatory behavior with this amplification factor.
β10 = 1− , (94)
From Fig. 14, we see that, with the increasing viscosity ratio, the 2λe J0 (iλe R0 ) J0 (iλe R0 )
magnitude of the cross migration decreases as expected (as the  
−1
viscosity ratio increases, the drop behaves like a solid). It may be 3λ2e b 1
γ =
0
1− . (95)
noted that the surface Péclet number measures the role of convec- 1
4J0 (iλe R0 ) J0 (iλe R0 )
tion relative to diffusion. Therefore, as Pes increases, the magnitude
of migration velocity increases. Also, for a fixed viscosity ratio, the Therefore the corresponding surfactant concentration distribution
slip parameter reduces the resistance exerted by the drop. Accord- on the spherical drop is given by
ingly, the migration velocity increases which can be observed in  = 0 + Pes 1 + Pe2s 2 + O(Pe3s ), (96)
Fig. 15.
where

6.3. Poiseuille flow 0 = 1, (97)

Consider a Poiseuille flow past a liquid drop of unit ra-  1  2


1 = E10 cos θ + E21
1
cos φ P21 (cos θ ) + E30
1 0
P3 (cos θ ) eλe t , (98)
dius whose center is at origin (Refs. Hetsroni and Haber, 1970;
Pak et al., 2014 to see the geometrical setup of the problem).  2
In this case, we have calculated the ambient velocity as v∞ =


−1 2 = E10 cos θ + E21
2
cos φ P21 (cos θ ) + E30
2 0
P3 (cos θ )
J0 (iλe R )
v0∞ = kˆ eλe t 1 − . Here R2 = r 2 sin2 θ +
2 1
J0 (iλe R0 )
1− J0 (iλe R0 )
2 1
+ E11 P1 (cos θ ) cos φ + E20
2 0
P2 (cos θ ) + E22
2
cos 2φ P22 (cos θ )
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 95

Fig. 16. Variation of the drop interface for different values of Ca, Capillary number, corresponding to Poisuelle flow, with λ2e = −0.04, λ2i = −0.04, μ = 0.5, Ma = 1, b = 2,
R0 = 5, t = π /2, Pes = 0.1 and α = 0.1.

2
+ E31 cos φ P31 (cos θ ) + E42
2
cos 2φ P42 (cos θ ) where
2
+ E51 cos φ P51 (cos θ ) 1  2
1
E10 =− 3λe ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi )
+ E62 cos 2φ P62 (cos θ ) + E71
2 2
cos φ P71 (cos θ ) (2 + λ2e Prs )
 2
+ E82 cos 2φ P8 (cos θ )
2 2 × b g1 (λe )λ2e +4(g1 (λe )(U0z −1 )J0 (iR0 λe ) + f1 (λe )g2 (λe )
 2
2
+ E52 cos 2φ P52 (cos θ ) + E40
2 0
P4 (cos θ ) + E60
2 0
P6 (cos θ ) eλe t , + f2 (λe )g1 (λe ) − g1 (λe )U0z )) )/
 
(99) × 4(−1 + J0 (iR0 λe ) ) g1 (λe )λ2e ( f2 (λi )(2αμ − 1 )
1 , E 1 , etc. can be computed using Eqs. (35), − α f1 (λi )μλi ) + g2 (λe )λe (2 f2 (λi )(2αμ + μ − 1 )
where the constants E10 21
(36), (48) and (49). −(2α + 1 ) f1 (λi )μλi ) + 3g1 (λe )( (2α + 1 ) f1 (λi )μλi
For an unbounded Poiseuille flow, the migration velocity of a − 2 f2 (λi )(2αμ + μ − 1 ) )) ), (106)
force free drop is calculated as
Similar to the arguments made in Section 6.2, by symmetry, it
U = U0 + Pes U1 + Pe2s U2 + O(Pe3s ). (100) is expected that there will be no velocity component in y direction.
In this case the zeroth order migration velocity U0 , given in In case if there is a cross migration, the same occurs towards the
Eq. (26) reduces to center line and should be in x direction (Pak et al., 2014). Similar

3

Y + μX + α P V + μU + α H
arguments show that there is no cross stream migration in leading
U0 = [v0∞ ]0 + [∇ 2v0∞ ]0 and first order. This implies that the symmetry condition at leading
2 ρi + ρe W + μ Z + α G W + μZ + α G order is satisfied (Pak et al., 2014).

3 Y + μX + α P
−1 If we consider the limiting case of no oscillations in the hydro-
+ λ2e , (101) dynamic flow field, i.e., λi = λe = 0, and zero slip coefficient, i.e.,
2 ρi + ρe W + μ Z + α G
α = 0, then the first order terminal velocity reduces to
where
  4Ma
I0 (bλe ) − I0 (λe R0 ) U1 = − kˆ . (107)
[v0∞ ]0 = eλe t kˆ .
2
(102) 3(3μ + 2 )2 R20
(−1 + J0 (iλe R0 ))I0 (λe R0 )
The above result matches with the one obtained by
and Pak et al. (2014).
λ2e I0 (bλe ) Further, the second order migration velocity U2 , given in
[∇ 2v0∞ ]0 = eλe t kˆ .
2
(103) Eq. (51) reduces to
−I0 (λe R0 ) + I0 (λe R0 )2
 
If we consider the limiting case of no oscillations in the hydro- 3(E11
2 ˆ 2 ˆ
i + E10 k ) 2Maλ2e f2 (λi )g1 (λe )
U2 =
dynamic flow field, i.e., λi = λe = 0, and zero slip coefficient, i.e., 2 ρi + ρe (W + μZ + α G )
α = 0, then the zeroth order terminal velocity reduces to
−1
  3 Y + μX + α P
+ λ2e eλe t .
2
2
μ 4 × (108)
U0 = 1−
b
− kˆ , (104) 2 ρi + ρe W + μ Z + α G
R20 4 + 6μ R20
If we consider the limiting case of no oscillations in the hydro-
which is exactly matching with the one that is obtained by dynamic flow field, i.e., λi = λe = 0, and zero slip coefficient, i.e.,
Pak et al. (2014). α = 0, then the second order terminal velocity reduces to
The first order migration velocity U1 , given in Eq. (38) reduces  
2Ma b 70μ2 + 109μ + 40 Ma
to U2 = ˆ−
k iˆ. (109)
  (3μ + 2 )2 R20 105(μ + 1 )2 (3μ + 2 )2 R40
1 ˆ
3E10 k 2Maλ2e f2 (λi )g1 (λe )
U1 =
2 ρi + ρe (W + μZ + α G ) If we look at the second order migration velocity, we see U∗2 .kˆ =

3 Y + μX + α P
−1 2 U ∝ Pe2 Ma2U ∝ U , and U∗ .iˆ = Pe2 U U .iˆ ∝
Pes U2Uc .kˆ ∝ Pe2s MaE10
2
c s c c 2 s 2 c
+ λ2e eλe t ,
2
× (105) Pes MaE11Uc ∝ Pes MaUc ∝ Uc2 . Therefore, the transverse migration is
2 2 2
2 ρi + ρe W + μ Z + α G
96 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

invariant upon reversal of the direction of the ambient Poiseuille viscosity ratios up to second order in the surface Péclet number,
flow. i.e., up to O(Pe2s ). We have analyzed the variation of surfactants
We have presented analytical solution for the problem of sur- for different viscosity ratios and Marangoni number. We have ob-
factant coated spherical drop in an unsteady arbitrary viscous flow, served that the impurities residing on the surface do not show
which is not explored earlier. Further, it may be noted that up much effect on the behavior of the drop for increasing viscosity ra-
to our knowledge the other cases like Uniform flow, Couette flow, tios. We considered various special cases and computed drag and
Poisuelle flow have been computed here for the first time (Fig. 16). migration velocity up to second order in the surface Péclet number
in each case. We have also compared the results with the existing
7. Conclusions literature for some limiting cases including analytical and experi-
mental data.
In this paper, we have considered an arbitrary transient Stokes
flow with a given ambient flow past a spherical drop. We have an- Acknowledgments
alyzed the effects of surface-active agents on the motion of the
drop. We have solved the unsteady convection-diffusion equation One of the authors (VS) would like to acknowledge the finan-
to find the surfactant transport on the surface of the drop for low cial support by CSIR-UGC (F.No. 17-06/2012 (i) EU-V dated, 05-10-
surface Péclet number. We have also considered the effects of in- 2012), India. The other author (GPRS) acknowledges the support of
terfacial slip. We found a closed form expression for drag and mi- Alexander von Humboldt Foundation for the fellowship, and Pro-
gration velocity in terms of Marangoni number, slip parameter and fessor Timothy J. Pedley for useful discussions.
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 97

Appendix A. The leading order hydrodynamic problem

The Leading order hydrodynamic problem is given by for r < 1


∂ u i0 dU0
+ = −∇
 pi + ∇ 2 u
 i0 ; ∇
 .u
 i0 = 0, (A.1)
∂t dt 0

and for r > 1


∂ u e0 dU0
+ = −∇
 pe + ∇ 2 u
 e0 ; ∇
 .u
 e0 = 0. (A.2)
∂t dt 0

with the boundary conditions


 e0 → u
u  e0∞ = ve0∞ − U0 as r → ∞, (A.3)

 e0 .nˆ = 0; u
u  i0 .nˆ = 0, (A.4)

u  i0 .tˆ = α (τnˆetˆ )0 ,
 e0 .tˆ − u (A.5)

(τnˆetˆ )0 − μ(τnˆi tˆ )0 = Ma∇


 s 0 .tˆ, (A.6)
By eliminating the pressure from the unsteady Stokes Eqs. (A.1) or (A.2), one can verify that the velocity fields inside and outside the
droplet satisfy
 

∇2 ∇2 − u i = 0, (A.7)
∂t 0
 

∇2 ∇2 − u e = 0 = 0. (A.8)
∂t 0
By using the general solution for the unsteady Stokes equation together with the equation of continuity (Venkatalaxmi et al., 2004), we
can have the following representation for the velocity and pressure fields
 e0 = ∇
u  ×∇
 × ( rχ e ) + ∇
0
 × ( rη e ) ,
0 (A.9)

 
∂ ∂χ0e
p0 e = pe∞ + ρe r ∇ 2 χ0e − , (A.10)
∂r ∂t
where the scalars χ0e and η0e are solutions of
 

∇2 ∇2 − χ e = 0, (A.11)
∂t 0
 

∇2 − η e = 0. (A.12)
∂t 0
j
Here r is the position vector and p∞ is a constant. The above representation is the so-called poloidal-toroidal decomposition of a solenoidal
vector field as given by Backus (1986). In fact, very recently, Suman Kumar and Amaranath (2018) have presented a complete general
solution of an unsteady Brinkman equations. A similar representation can be written for the velocity and pressure fields inside the drop
in terms of the two scalar functions χ0i and η0i . Hence, the problem can now be handled in terms of the scalars χ0e , η0e , χ0i and η0i .
Accordingly, the boundary conditions in terms of χ0e , η0e , χ0i and η0i are given by
Vanishing normal component of the velocity
χ0e = 0; χ0i = 0 on r = 1 . (A.13)
Slip in the tangential component of velocity
 
∂χ0e ∂χ0i ∂ 2 χ0e ∂ η0e
− =α , η e
− η i
= α on r = 1 . (A.14)
∂r ∂r ∂ r2 0 0
∂r r
Tangential stress balance
 
∂ ∂ 2 χ0e ∂ 2 χ0i ∂ 0
−μ = Ma on r = 1 , (A.15)
∂θ ∂r 2 ∂ r2 ∂θ
 
∂ ∂ 2 χ0e ∂ 2 χ0i ∂ 0
− μ = Ma on r = 1 , (A.16)
∂φ ∂ r2 ∂ r2 ∂φ
98 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

   
∂ η0e ∂ η0i
=μ on r = 1. (A.17)
∂r r ∂r r
Finite velocity and pressure fields inside the drop require that
χ0i < ∞, η0i < ∞. (A.18)
We represent the far-field ambient flow in terms of χ0∞ and η0∞ , given by
∞ 
 
χ0∞ = αn0 rn + βn0 fn (λe r ) Sn0 (θ , φ )eλe t ,
2
(A.19)
n=1

∞ 
 
η0∞ = γn0 fn (λe r ) Tn0 (θ , φ )eλe t ,
2
(A.20)
n=1

where

n  
Sn0 (θ , φ ) = Pnm (η ) A0nm cos mφ + B0nm sin mφ , (A.21)
m=0


n  
Tn0 (θ , φ ) = Pnm (η ) Cnm
0
cos mφ + D0nm sin mφ (A.22)
m=0

are spherical harmonics, and αn0 , βn0 , γn0 , A0nm , B0nm , Cnm
0 and D0 are the known coefficients. These coefficients are controlled by the choice
nm
of the ambient flow. For example, in the case of uniform ambient flow, χ0∞ = 12 r cos θ eλe t , η0∞ = 0; and hence α10 = 12 , αn0 = 0 for n
= 1,
2

βn = 0, γn = 0, A10 = 1, Anm = 0 for n


= 1 or m
= 0, Bnm = 0, Cnm = 0 and Dnm = 0. Here, fn (λj r) and gn (λj r) ( j = i, e) are modified spherical
0 0 0 0 0 0 0

Bessel functions of first and second kind, respectively. Note that, for the bounded solution as t → ∞, we require λ2j < 0. In the presence
of the spherical drop, the resultant flow due to the disturbance can be represented as general solution of (A.11) and (A.12) as follows, for
r<1
∞ 
 
χ0i = ᾱn0 rn + β̄n0 fn (λi r ) Sn0 (θ , φ )eλi t ,
2
(A.23)
n=1

∞ 
 
η0i = γ̄n0 fn (λi r ) Tn0 (θ , φ )eλi t ,
2
(A.24)
n=1

and for r > 1



 
 αˆ n0
χ = α +β (λe r ) + βˆn0 gn (λe r ) Sn0 (θ , φ )eλe t ,
2
e 0 n 0
0 nr + n fn (A.25)
r n+1
n=1

∞ 
 
η0e = γn0 fn (λe r ) + γˆn0 gn (λe r ) Tn0 (θ , φ )eλe t ,
2
(A.26)
n=1

where ᾱn0 , β̄n0 , γ̄n0 , αˆ n0 , βˆn0 , γˆn0 are the unknown coefficients which are to be determined from the boundary conditions (A.13)–(A.18), and
λi , λe are the amplification factors corresponding to the flow inside and outside of the drop which can be found if the initial conditions
are provided (Venkatalaxmi et al., 2004). Moreover, the far field condition turns out to be χ0e → χ0∞ and η0e → η0∞ as r → ∞. The unknown
coefficients can be expressed in terms of the known ambient flow variables using the boundary conditions.
The unknown coefficients in Eqs. (A.23)–(A.26) can be found using the boundary conditions given in Eqs. (A.13)–(A.18) which are given
as follows:

αˆ n0 = −2 fn+1 (λi )gn+1 (λe )αn0 λe + 2 fn+1 (λi )g2 (λe )μαn0 λe − 2 fn+1 (λe ) fn+1 (λi )gn (λe )βn0 λe
− 2 fn (λe ) fn+1 (λi )gn+1 (λe )βn0 λe + 2 fn+1 (λe ) fn+1 (λi )gn (λe )μβn0 λe
+ 2 fn (λe ) fn+1 (λi )gn+1 (λe )μβn0 λe
− fn+1 (λi )gn (λe )αn0 λ2e − fn (λi )gn+1 (λe )μαn0 λe λi − fn (λi ) fn+1 (λe )gn (λe )μβn0 λe λi
− fn (λe ) fn (λi )gn+1 (λe )μβn0 λe λi + 4 fn+1 (λi )gn+1 (λe )αμαn0 λe
+ 4 fn+1 (λe ) fn+1 (λi )gn (λe )αμβn0 λe
+ 4 fn (λe ) fn+1 (λi )gn+1 (λe )αμβn0 λe + 2 fn+1 (λi )gn (λe )αμαn0 λ2e − 2 fn (λi )gn+1 (λe )αμαn0 λe λi
− 2 fn (λi ) fn+1 (λe )gn (λe )αμβn0 λe λi

− 2 fn (λe ) fn (λi )gn+1 (λe )αμβn0 λe λi − fn (λi )gn (λe )αμαn0 λ2e λi
/(−2 fn+1 (λi )gn (λe ) − 4 fn+1 (λi )gn (λe )n + 2 fn+1 (λi )gn (λe )μ + 4 fn+1 (λi )gn (λe )nμ
+ 2 fn+1 (λi )gn+1 (λe )λe − 2 fn+1 (λi )gn+1 (λe )μλe + fn+1 (λi )gn (λe )λ2e − fn (λi )gn (λe )μλi
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 99

− 2 fn (λi )gn (λe )nμλi + fn (λi )gn+1 (λe )μλe λi + 4 fn+1 (λi )gn (λe )αμ + 8 fn+1 (λi )gn (λe )nαμ
− 4 fn+1 (λi )gn+1 (λe )αμλe − 2 fn+1 (λi )gn (λe )αμλ2e − 2 fn (λi )gn (λe )αμλi

− 4 fn (λi )gn (λe )nαμλi + 2 fn (λi )gn+1 (λe )αμλe λi + fn (λi )gn (λe )αμλ2e λi , (A.27)


βˆn0 = 2 fn+1 (λi )αn0 + 4 fn+1 (λi )nαn0 − 2 fn+1 (λi )μαn0 − 4 fn+1 (λi )nμαn0 + 2 fn (λe ) fn+1 (λi )βn0
+ 4 fn (λe ) fn+1 (λi )nβn0 − 2 fn (λe ) fn+1 (λi )μβn0 − 4 fn (λe ) fn+1 (λi )nμβn0
+ 2 fn+1 (λe ) fn+1 (λi )βn0 λe − 2 fn+1 (λe ) fn+1 (λi )μβn0 λe − fn (λe ) fn+1 (λi )βn0 λ2e
+ fn (λi )μαn0 λi + 2 fn (λi )nμαn0 λi + fn (λe ) fn (λi )μβn0 λi
+ 2 fn (λe ) fn (λi )nμβn0 λi + fn (λi ) fn+1 (λe )μβn0 λe λi − 4 fn+1 (λi )αμαn0
− 8 fn+1 (λi )nαμαn0 − 4 fn (λe ) fn+1 (λi )αμβn0 − 8 fn (λe ) fn+1 (λi )nαμβn0
− 4 fn+1 (λe ) fn+1 (λi )αμβn0 λe + 2 fn (λe ) fn+1 (λi )αμβn0 λ2e + 2 fn (λi )αμαn0 λi
+ 4 fn (λi )nαμαn0 λi + 2 fn (λe ) fn (λi )αμβn0 λi + 4 fn (λe ) fn (λi )nαμβn0 λi

+ 2 fn (λi ) fn+1 (λe )αμβn0 λe λi − fn (λe ) fn (λi )αμβn0 λ2e λi /
(−2 fn+1 (λi )gn (λe ) − 4 fn+1 (λi )gn (λe )n + 2 fn+1 (λi )gn (λe )μ + 4 fn+1 (λi )gn (λe )nμ
+ 2 fn+1 (λi )gn+1 (λe )λe − 2 fn+1 (λi )gn+1 (λe )μλe + fn+1 (λi )gn (λe )λ2e
− fn (λi )gn (λe )μλi − 2 fn (λi )gn (λe )nμλi + fn (λi )gn+1 (λe )μλe λi
+ 4 fn+1 (λi )gn (λe )αμ + 8 fn+1 (λi )gn (λe )nαμ − 4 f2 (λi )g2 (λe )αμλe
− 2 fn+1 (λi )gn (λe )αμλ2e − 2 fn (λi )gn (λe )αμλi − 4 fn (λi )gn (λe )nαμλi

+ 2 fn (λi )gn+1 (λe )αμλe λi + fn (λi )gn (λe )αμλ2e λi , (A.28)

 
ᾱn0 = −et (λe −λi ) fn (λi )λ2e (gn (λe ) + 2gn (λe )n )αn0 + ( fn+1 (λe )gn (λe ) + fn (λe )gn+1 (λe ))βn0 λe
2 2

 
/ λi gn (λe )λ2e ( fn+1 (λi ) + αμ(−2 fn+1 (λi ) + fn (λi )λi ))
+ gn+1 (λe )λe ( fn (λi )μλi (1 + 2α ) − 2 fn+1 (λi )(−1 + μ + 2αμ ))
+ gn (λe )(1 + 2n )(− fn (λi )μλi (1 + 2α ) + 2 fn+1 (λi )(−1 + μ + 2αμ )))), (A.29)

 
βˆn0 = et (λe −λi ) λ2e (gn (λe ) + 2gn (λe )n )αn0 + ( fn+1 (λe )gn (λe ) + fn (λe )gn+1 (λe ))βn0 λe
2 2

 
/ λi gn (λe )λ2e ( fn+1 (λi ) + αμ(−2 fn+1 (λi ) + fn (λi )λi ))
+ gn+1 (λe )λe ( fn (λi )μλi (1 + 2α ) − 2 fn+1 (λi )(−1 + μ + 2αμ ))
+ gn (λe )(1 + 2n )(− fn (λi )μλi (1 + 2α ) + 2 fn+1 (λi )(−1 + μ + 2αμ )))), (A.30)

  
γˆn0 = γn0 fn+1 (λe )λe fn (λi ) + et λe αμ( fn (λi )(−1 + n ) + fn+1 (λi )λi )μe
2

   2  
+ fn (λe ) fn+1 (λi )μλi −1 + −1 + et λe n α
  2  
+ fn (λi )(−1 + n ) 1 − μ + −1 + et λe n αμ /
  
gn+1 (λe )λe fn (λi ) + et λe αμ( fn (λi )(−1 + n ) + fn+1 (λi )λi )μe
2

   2  
− gn (λe ) fn+1 (λi )μλi −1 + −1 + et λe n α
  2  
+ fn (λi )(−1 + n ) 1 − μ + −1 + et λe n αμ , (A.31)


2 2  
2 
γˆn0 = et (λe −λi ) γn0 ( fn+1 (λe )gn (λe ) + fn (λe )gn+1 (λe ))λe −1 + −1 + et λe α
  
/ −gn+1 (λe )λe fn (λi ) + et λe αμ( fn (λi )(−1 + n ) + fn+1 (λi )λi )μe
2

   2  
+ gn (λe ) fn+1 (λi )μλi −1 + −1 + et λe n α
  2  
+ fn (λi )(−1 + n ) 1 − μ + −1 + et λe n αμ . (A.32)

Appendix B. Symbols

The constants given in Eq. (21) are given as follows:


X = λ3e {λi f1 (λi ) − 2 f2 (λi )}g2 (λe ),

Y = λ3e {λi g1 (λe ) + 2g2 (λe )} f2 (λi ),


100 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

P = −μλ3e {−λi f1 (λi ) + 2 f2 (λi )}{2g2 (λe ) + g1 (λe )},

G = μ{2g2 (λe )λe − 6g1 (λe ) + λ2e g1 (λe )}{λi f1 (λi ) − 2 f2 (λi )},

Z = {λi f1 (λi ) − 2 f2 (λi )}{λe g2 (λe ) − 3g1 (λe )},

W = {2λe g2 (λe ) − (6 − λ2e )g1 (λe )} f2 (λi ),

S = 3{ f2 (λe )g1 (λe )}{λi f1 (λi ) − 2 f2 (λi )},

T = 6 f2 (λi ){ f2 (λe )g1 (λe ) + f1 (λe )g2 (λe )},

Q = −3μ{ f2 (λe )g1 (λe ) + f1 (λe )g2 (λe )}{4 f2 (λi ) − 2λi f2 (λi )},

X Y P
U =S− ;V = T − ;H = Q − .
λ2e λ2e λ2e
Appendix C. The first order problem

The first-order pressure and velocity fields satisfy the unsteady Stokes and continuity equations. Correspondingly, we express χ1i , η1i ,
χ1e and η1e as follows
∞ 
 
χ1i = ᾱn1 rn + β̄n1 fn (λi r ) Sn1 (θ , φ )eλi t ,
2
(C.1)
n=1

∞ 
 
η1i = γ̄n1 fn (λi r ) Tn1 (θ , φ )eλi t ,
2
(C.2)
n=1


 
 αˆ n1
χ = α +β (λe r ) + βˆn1 gn (λe r ) Sn1 (θ , φ )eλe t ,
2
e 1 n 1
1 nr + n fn (C.3)
r n+1
n=1

∞ 
 
η1e = γn1 fn (λe r ) + γˆn1 gn (λe r ) Tn1 (θ , φ )eλe t ,
2
(C.4)
n=1

where Sn1 (θ , φ ) and Tn1 (θ , φ ) are spherical harmonics of order n. The interfacial surfactant that is coupled via the boundary conditions
(A.15) and (A.16) together with the form of  1 given in Eq. (30) enforces Sn1 (θ , φ ) = R1n (θ , φ ). However, we have
n 
 
Tn1 (θ , φ ) =  cos mφ + F  sin mφ P m (cos
Enm θ ). (C.5)
nm n
m=0

The unknown coefficients given in Eqs. (C.1)–(C.4) are given by


 2 2
ᾱ = et λe −t λi fn (λi )Ma(gn (λe ) + 2gn (λe )n − gn+1 (λe )λe + 2gn (λe )α
1
n

+ 4gn (λe )nα − 2gn+1 (λe )αλe − gn (λe )αλ2e
/(λi (2 fn+1 (λi )gn (λe ) + 4 fn+1 (λi )gn (λe )n − 2 fn+1 (λi )gn (λe )μ − 4 fn+1 (λi )gn (λe )nμ
− 2 fn+1 (λi )gn+1 (λe )λe + 2 fn+1 (λi )gn+1 (λe )μλe − fn+1 (λi )gn (λe )λ2e + fn (λi )gn (λe )μλi
+ 2 fn (λi )gn (λe )nμλi − fn (λi )gn+1 (λe )μλe λi − 4 fn+1 (λi )gn (λe )αμ
− 8 fn+1 (λi )gn (λe )nαμ + 4 fn+1 (λi )gn+1 (λe )αμλe + 2 fn+1 (λi )gn (λe )αμλ2e
+ 2 fn (λi )gn (λe )αμλi + 4 fn (λi )gn (λe )nαμλi

− 2 fn (λi )gn+1 (λe )αμλe λi − fn (λi )gn (λe )αμλ2e λi , (C.6)

 2 2
β̄n1 = − et λe −t λi Ma(gn (λe ) + 2gn (λe )n − gn+1 (λe )λe + 2gn (λe )α

+ 4gn (λe )nα − 2gn+1 (λe )αλe − gn (λe )αλ2e
/(λi (2 fn+1 (λi )gn (λe ) + 4 fn+1 (λi )gn (λe )n − 2 fn+1 (λi )gn (λe )μ − 4 fn+1 (λi )gn (λe )nμ
− 2 fn+1 (λi )gn+1 (λe )λe + 2 fn+1 (λi )gn+1 (λe )μλe − fn+1 (λi )gn (λe )λ2e + fn (λi )gn (λe )μλi
+ 2 fn (λi )gn (λe )nμλi − fn (λi )gn+1 (λe )μλe λi − 4 fn+1 (λi )gn (λe )αμ
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 101

− 8 fn+1 (λi )gn (λe )nαμ + 4 fn+1 (λi )gn+1 (λe )αμλe + 2 fn+1 (λi )gn (λe )αμλ2e
+ 2 fn (λi )gn (λe )αμλi + 4 fn (λi )gn (λe )nαμλi

− 2 fn (λi )gn+1 (λe )αμλe λi − fn (λi )gn (λe )αμλ2e λi , (C.7)

γ̄n1 = 0, (C.8)

αn1 = 0, (C.9)

αˆ n1 = ( fn+1 (λi )gn (λe )Ma )


/(2 fn+1 (λi )gn (λe ) + 4 fn+1 (λi )gn (λe )n − 2 fn+1 (λi )gn (λe )μ − 4 fn+1 (λi )gn (λe )nμ
− 2 fn+1 (λi )gn+1 (λe )λe + 2 fn+1 (λi )gn+1 (λe )μλe − fn+1 (λi )gn (λe )λ2e + fn (λi )gn (λe )μλi
+ 2 fn (λi )gn (λe )nμλi − fn (λi )gn+1 (λe )μλe λi − 4 fn+1 (λi )gn (λe )αμ
− 8 fn+1 (λi )gn (λe )nαμ + 4 fn+1 (λi )gn+1 (λe )αμλe + 2 fn+1 (λi )gn (λe )αμλ2e
+ 2 fn (λi )gn (λe )αμλi + 4 fn (λi )gn (λe )nαμλi

− 2 fn (λi )gn+1 (λe )αμλe λi − fn (λi )gn (λe )αμλ2e λi , (C.10)

βn1 = 0, (C.11)

βˆn1 = −( fn+1 (λi )Ma )


/(2 fn+1 (λi )gn (λe ) + 4 fn+1 (λi )gn (λe )n − 2 fn+1 (λi )gn (λe )μ − 4 fn+1 (λi )gn (λe )nμ
− 2 fn+1 (λi )gn+1 (λe )λe + 2 fn+1 (λi )gn+1 (λe )μλe − fn+1 (λi )gn (λe )λ2e + fn (λi )gn (λe )μλi
+ 2 fn (λi )gn (λe )nμλi − fn (λi )gn+1 (λe )μλe λi − 4 fn+1 (λi )gn (λe )αμ
− 8 fn+1 (λi )gn (λe )nαμ + 4 fn+1 (λi )gn+1 (λe )αμλe + 2 fn+1 (λi )gn (λe )αμλ2e
+ 2 fn (λi )gn (λe )αμλi + 4 fn (λi )gn (λe )nαμλi

− 2 fn (λi )gn+1 (λe )αμλe λi − fn (λi )gn (λe )αμλ2e λi , (C.12)

γn1 = 0, (C.13)

γˆn1 = 0. (C.14)

Appendix D. Coefficients Couette flow

(F L − U0x )  
1
E11 = 3g1 (λe )λ2e ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi ) /
(2 + λe Prs )
2

g1 (λe )λ2e ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi )
+ g2 (λe )λe (2 f2 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )
+3g1 (λe )( (2α + 1 ) f1 (λi )μλi − 2 f2 (λi )(2αμ + μ − 1 ) )) ) (D.1)

5F g2 (λe )λ2e ( f3 (λi )(2αμ − 1 ) − α f2 (λi )μλi )


1
F22 =   (D.2)
6 P rs λ2e + 6 δ2
where
δ2 = g2 (λe )λ2e ( f3 (λi )(2αμ − 1 ) − α f2 (λi )μλi )
+ g3 (λe )λe (2 f3 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f2 (λi )μλi )
+ 5g2 (λe )( (2α + 1 ) f2 (λi )μλi − 2 f3 (λi )(2αμ + μ − 1 ) ), (D.3)
and
1
2E11 f2 (λi )Ma
2
E11 =   (3g1 (λe )(ρe + 2ρi ) − g2 (λe )λe (ρe + 2ρi + 3 ) )/
P rs λ2e + 2

g1 (λe )λ2e (ρe + 2ρi + 3 )( f2 (λi )(2αμ − 1 )
102 V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103

− α f1 (λi )μλi ) + g2 (λe )λe (ρe + 2ρi + 3 )


(2 f2 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )
− 3g1 (λe )(ρe + 2ρi )(2 f2 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )) (D.4)

1
6F22 f3 (λi )Ma(−5g2 (λe ) + g3 (λe )λe )
2
F22 =   , (D.5)
Pr λ2e + 6 δ2

e−λe t ((−6AF22 ))
2
1
+ 5CE11
1
+ 18BE11
1
2
F11 =−   , (D.6)
5 P rs λ2e + 2

4(96(AF22 ))e−λe t
2
1
+ 2BE11
1
2
F31 =   , (D.7)
(5 )35 Prs λ2e + 12

(2(AF22 ))e−λe t
2
1
+ 2BE111
2
F33 =−   , (D.8)
5 P rs λe + 12
2

(−2(7AE11 ))e−λe t
2
1
+ 72BF22 1
2
E20 =   , (D.9)
7 P rs λ2e + 6

(AE11 )e−λe t
2
1
− 4CF221
2
E22 =   , (D.10)
2 P rs λe + 6
2

(144BF22 )e−λe t
2
1
2
E40 =  , (D.11)
7 P rs λe + 20
2

(−3BF22 )e−λe t
2
1
2
E44 =  , (D.12)
7 P rs λ2e + 20
where
 
A = −(F L − U0x ) 3g1 (λe )λ2e ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi ) /
 
2 g1 (λe )λ2e ( f2 (λi )(2αμ − 1 ) − α f1 (λi )μλi )
+ g2 (λe )λe (2 f2 (λi )(2αμ + μ − 1 ) − (2α + 1 ) f1 (λi )μλi )
+3g1 (λe )( (2α + 1 ) f1 (λi )μλi − 2 f2 (λi )(2αμ + μ − 1 ) )) )eλe t
2
(D.13)

5F g2 (λe )λ2e ( f3 (λi )(2αμ − 1 ) − α f2 (λi )μλi ) λ2e t


B= e (D.14)
36δ2
 
3F et λe ( f1 (λe )g2 (λe ) + f2 (λe )g1 (λe ))
α f2 (λi )μλi et λe + f1 (λi )
2 2

C=   2     (D.15)
f2 (λi )g1 (λe )μλi α eT λe − 1 − 1 − g2 (λe )λe α f2 (λi )μλi eT λe + f1 (λi )
2

Supplementary material Chan, P.H., Leal, L., 1979. The motion of a deformable drop in a second-order fluid.
J. Fluid Mech. 92 (01), 131–170.
Chisnell, R., 1987. The unsteady motion of a drop moving vertically under gravity. J.
Supplementary material associated with this article can be Fluid Mech. 176, 443–464.
found, in the online version, at doi:10.1016/j.ijmultiphaseflow.2018. Choudhuri, D., Padmavati, B.S., 2014. A study of an arbitrary unsteady stokes flow
05.008. in and around a liquid sphere. Appl. Math. Comput. 243, 644–656.
Choudhuri, D., Raja Sekhar, G.P., 2013. Thermocapillary drift on a spherical drop in
a viscous fluid. Phys. Fluids 25 (4), 043104.
References Clift, R., Grace, J.R., Weber, M.E., 2005. Bubbles, Drops, and Particles. Courier Corpo-
ration.
Alke, A., Bothe, D., 2007. VOF-simulation of fluid particles influenced by soluble sur- Cox, R., Brenner, H., 1968. The lateral migration of solid particles in Poiseuille flow-I
factant. In: 6th International Conference on Multiphase Flow, ICMF2007, Leipzig theory. Chem. Eng. Sci. 23 (2), 147–173.
(Germany). Das, S., Mandal, S., Chakraborty, S., 2017. Migration of a surfactant-laden droplet in
Backus, G., 1986. Poloidal and toroidal fields in geomagnetic field modeling. Rev. non-isothermal poiseuille flow. Phys. Fluids 29 (1), 012002.
Geophys. 24 (1), 75–109. Das, S., Mandal, S., Chakraborty, S., 2018. Effect of temperature gradient on the
Bratukhin, Y.K., Kostarev, K., Viviani, A., Zuev, A., 2005. Experimental study of cross-stream migration of a surfactant-laden droplet in Poiseuille flow. J. Fluid
marangoni bubble migration in normal gravity. Exp Fluids 38 (5), 594–605. Mech. 835, 170–216.
Carslaw, H.S., Jaeger, J.C., 1959. Conduction of Heat in Solids, 1959, 2nd ed. Oxford: Dill, L.H., Balasubramaniam, R., 1992. Unsteady thermocapillary migration of iso-
Clarendon Press. lated drops in creeping flow. Int. J. Heat Fluid Flow 13 (1), 78–85.
V. Sharanya et al. / International Journal of Multiphase Flow 107 (2018) 82–103 103

Fleckenstein, S., Bothe, D., 2013. Simplified modeling of the influence of surfactants Polyanin, A.D., Kutepov, A., Kazenin, D., Vyazmin, A., 2001. Hydrodynamics, Mass
on the rise of bubbles in VOF-simulations. Chem. Eng. Sci. 102, 514–523. and Heat Transfer in Chemical Engineering, vol. 14. CRC Press.
Goldsmith, H., Mason, S., 1962. The flow of suspensions through tubes. I. Single Ramachandran, A., Gary Leal, L., 2012. The effect of interfacial slip on the rheol-
spheres, rods, and discs. J Colloid Sci 17 (5), 448–476. ogy of a dilute emulsion of drops for small capillary numbers. J. Rheol. 56 (6),
Hanna, J.A., Vlahovska, P.M., 2010. Surfactant-induced migration of a spherical drop 1555–1587.
in stokes flow. Phys. Fluids 22 (1), 013102. Ramachandran, A., Tsigklifis, K., Roy, A., Leal, G., 2012. The effect of interfacial slip
Happel, J., Brenner, H., 2012. Low Reynolds Number Hydrodynamics: with Special on the dynamics of a drop in flow: Part I. Stretching, relaxation, and breakup. J.
Applications to Particulate Media, vol. 1. Springer Science & Business Media. Rheol. 56 (1), 45–97.
Hetsroni, G., Haber, S., 1970. The flow in and around a droplet or bubble submerged Sadhal, S., Johnson, R.E., 1983. Stokes flow past bubbles and drops partially coated
in an unbound arbitrary velocity field. Rheol. Acta 9 (4), 488–496. with thin films. Part 1. Stagnant cap of surfactant film–exact solution. J. Fluid
Holbrook, J.A., Levan, M.D., 1983. Retardation of droplet motion by surfactant. Part Mech. 126, 237–250.
1. Theoretical development and asymptotic solutions. Chem. Eng. Commun. 20 Sharanya, V., Raja Sekhar, G.P., 2015. Thermocapillary migration of a spherical drop
(3–4), 191–207. in an arbitrary transient Stokes flow. Phys. Fluids (1994-present) 27 (6), 063104.
Holbrook, J.A., Levan, M.D., 1983. Retardation of droplet motion by surfactant. Part Stan, C.A., Ellerbee, A.K., Guglielmini, L., Stone, H.A., Whitesides, G.M., 2013. The
2. Numerical solutions for exterior diffusion, surface diffusion, and adsorption magnitude of lift forces acting on drops and bubbles in liquids flowing inside
kinetics. Chem. Eng. Commun. 20 (5–6), 273–290. microchannels. Lab Chip 13 (3), 365–376.
Ho, B., Leal, L., 1974. Inertial migration of rigid spheres in two-dimensional unidi- Stone, H., Leal, L., 1990. The effects of surfactants on drop deformation and breakup.
rectional flows. J. Fluid Mech. 65 (2), 365–400. J. Fluid Mech. 220, 161–186.
Karnis, A., Mason, S., 1967. Particle motions in sheared suspensions: XXIII. Wall mi- Subramanian, R.S., 1983. Thermocapillary migration of bubbles and droplets. Adv.
gration of fluid drops. J. Colloid Interface Sci. 24 (2), 164–169. Space Res. 3 (5), 145–153.
Landau, L.D., Lifshitz, E. M., Fluid Mechanics, 1987, Pergamon Press. Subramanian, R.S., Balasubramaniam, R., 2001. The Motion of Bubbles and Drops in
Leal, L.G., 2007. Advanced Transport Phenomena: Fluid Mechanics and Convective Reduced Gravity. Cambridge University Press.
Transport Processes. Cambridge University Press. Suman Kumar, T., Amaranath, T., 2018. A complete general solution of the unsteady
Leven, M.D., Newman, J., 1976. The effect of surfactant on the terminal and interfa- Brinkman equations. J. Math. Anal. Appl. 461 (2), 1365–1373.
cial velocities of a bubble or drop. AlChE J. 22 (4), 695–701. Venkatalaxmi, A., Padmavathi, B., Amaranath, T., 2004. A general solution of un-
Mandal, S., Bandopadhyay, A., Chakraborty, S., 2015. Effect of interfacial slip on the steady stokes equations. Fluid Dyn. Res. 35 (3), 229–236.
cross-stream migration of a drop in an unbounded poiseuille flow. Phys. Rev. E Wong, H., Rumschitzki, D., Maldarelli, C., 1996. On the surfactant mass balance at a
92 (2), 023002. deforming fluid interface. Phys. Fluids (1994-present) 8 (11), 3203–3204.
Oguz, H., Sadhal, S., 1988. Effects of soluble and insoluble surfactants on the motion Young, N., Goldstein, J.S., Block, M.J., 1959. The motion of bubbles in a vertical tem-
of drops. J. Fluid Mech. 194, 563–579. perature gradient. J. Fluid Mech. 6 (3), 350–356.
Pak, O.S., Feng, J., Stone, H.A., 2014. Viscous Marangoni migration of a drop in a Zuev, A.L., Kostarev, K.G., 2008. Certain peculiarities of the solutocapillary convec-
poiseuille flow at low surface Péclet numbers. J. Fluid Mech. 753, 535–552. tion. Phys. Usp. 51 (10), 1027.

You might also like