Thermocapillary Migration of a Drop

Similar documents
Thermocapillary convection due to a stationary bubble

Boundary Conditions in Fluid Mechanics

Shell Balances in Fluid Mechanics

Enhancement of Heat Transfer by an Electric Field for a Drop Translating at Intermediate Reynolds Number

Simulating Interfacial Tension of a Falling. Drop in a Moving Mesh Framework

Fluid Dynamics Exercises and questions for the course

Experimental and Theoretical Study of Motion of Drops on Horizontal Solid Surfaces with a Wettability Gradient Nadjoua Moumen

The influence of axial orientation of spheroidal particles on the adsorption

Navier-Stokes Equation: Principle of Conservation of Momentum

Numerical Studies of Droplet Deformation and Break-up

On the displacement of three-dimensional fluid droplets adhering to a plane wall in viscous pressure-driven flows

Principles of Convection

Figure 11.1: A fluid jet extruded where we define the dimensionless groups

12.1 Viscous potential flow (VPF)

THERMOCAPILLARY CONVECTION IN A LIQUID BRIDGE SUBJECTED TO INTERFACIAL COOLING

3 BUBBLE OR DROPLET TRANSLATION

Spinning of a molten threadline Steady-state isothermal viscous flows

2.25 Advanced Fluid Mechanics Fall 2013

Fundamentals of Fluid Dynamics: Elementary Viscous Flow

Motion of an air bubble under the action of thermocapillary and buoyancy forces. Abstract

EFFECTS OF INTERFACE CONTAMINATION ON MASS TRANSFER INTO A SPHERICAL BUBBLE

Chapter 5. The Differential Forms of the Fundamental Laws

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

FLUID MECHANICS. Chapter 9 Flow over Immersed Bodies

Wall Effects in Convective Heat Transfer from a Sphere to Power Law Fluids in Tubes

Introduction to Fluid Mechanics

REE Internal Fluid Flow Sheet 2 - Solution Fundamentals of Fluid Mechanics

An Adsorption Desorption-Controlled Surfactant on a Deforming Droplet

Motion of a drop in a vertical temperature gradient at small Marangoni number the critical role of inertia

Lecture 9 Laminar Diffusion Flame Configurations

J. M. Lopez*,1 and A. Hirsa

Liquids and solids are essentially incompressible substances and the variation of their density with pressure is usually negligible.

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

Numerical Heat and Mass Transfer

UNIT II CONVECTION HEAT TRANSFER

2. FLUID-FLOW EQUATIONS SPRING 2019


Solution of Partial Differential Equations

CENG 501 Examination Problem: Estimation of Viscosity with a Falling - Cylinder Viscometer

COMBINED EFFECTS OF RADIATION AND JOULE HEATING WITH VISCOUS DISSIPATION ON MAGNETOHYDRODYNAMIC FREE CONVECTION FLOW AROUND A SPHERE

Interpreting Differential Equations of Transport Phenomena

Convection. forced convection when the flow is caused by external means, such as by a fan, a pump, or atmospheric winds.

- Marine Hydrodynamics. Lecture 4. Knowns Equations # Unknowns # (conservation of mass) (conservation of momentum)

Surfactant-Influenced Gas Liquid Interfaces: Nonlinear Equation of State and Finite Surface Viscosities

Research Article Innovation: International Journal of Applied Research; ISSN: (Volume-2, Issue-2) ISSN: (Volume-1, Issue-1)

Gravitational effects on the deformation of a droplet adhering to a horizontal solid surface in shear flow

FORMULA SHEET. General formulas:

Introduction to Marine Hydrodynamics

Parash Moni Thakur. Gopal Ch. Hazarika

The steady propagation of a surfactant-laden liquid plug in a two-dimensional channel

Fluid Dynamics: Theory, Computation, and Numerical Simulation Second Edition

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

FLUID MECHANICS PROF. DR. METİN GÜNER COMPILER

CHAPTER 4 BOUNDARY LAYER FLOW APPLICATION TO EXTERNAL FLOW

Combined Effect of Magnetic field and Internal Heat Generation on the Onset of Marangoni Convection

Unsteady Flow of a Newtonian Fluid in a Contracting and Expanding Pipe


Chapter 10. Solids and Fluids

Week 2 Notes, Math 865, Tanveer

10.52 Mechanics of Fluids Spring 2006 Problem Set 3

Review of fluid dynamics

The Bernoulli Equation

INTERACTION BETWEEN A PAIR OF DROPS ASCENDING IN A LINEARLY STRATIFIED FLUID

Pressure corrections for viscoelastic potential flow analysis of capillary instability

CONVECTIVE HEAT TRANSFER

6.2 Governing Equations for Natural Convection

ChE 385M Surface Phenomena University of Texas at Austin. Marangoni-Driven Finger Formation at a Two Fluid Interface. James Stiehl

ISCST shall not be responsible for statements or opinions contained in papers or printed in its publications.

Part II Fundamentals of Fluid Mechanics By Munson, Young, and Okiishi

Colloidal Particles at Liquid Interfaces: An Introduction

Lecture notes Breakup of cylindrical jets Singularities and self-similar solutions

Capillary-gravity waves: The effect of viscosity on the wave resistance

2.5 Stokes flow past a sphere

CAPA due today. Today will finish up with the hinge problem I started on Wednesday. Will start on Gravity. Universal gravitation

MAE 101A. Homework 7 - Solutions 3/12/2018

Detailed Outline, M E 521: Foundations of Fluid Mechanics I

Nonlinear Analysis: Modelling and Control, 2008, Vol. 13, No. 4,

Fluid Dynamics for Ocean and Environmental Engineering Homework #2 Viscous Flow

Fundamentals of Fluid Dynamics: Ideal Flow Theory & Basic Aerodynamics

BFC FLUID MECHANICS BFC NOOR ALIZA AHMAD

Potential flow of a second-order fluid over a sphere or an ellipse

Studies on flow through and around a porous permeable sphere: II. Heat Transfer

Absorption of gas by a falling liquid film

FORCED CONVECTION FILM CONDENSATION OF DOWNWARD-FLOWING VAPOR ON HORIZONTAL TUBE WITH WALL SUCTION EFFECT

Game Physics. Game and Media Technology Master Program - Utrecht University. Dr. Nicolas Pronost

ENGR Heat Transfer II

L = I ω = const. I = 2 3 MR2. When the balloon shrinks (because the air inside it cools down), the moment of inertia decreases, R = 1. L = I ω = I ω.

Department of Mechanical Engineering

Water is sloshing back and forth between two infinite vertical walls separated by a distance L: h(x,t) Water L

NUMERICAL INVESTIGATION OF THERMOCAPILLARY INDUCED MOTION OF A LIQUID SLUG IN A CAPILLARY TUBE

LECTURE 4: Marangoni flows

ESS314. Basics of Geophysical Fluid Dynamics by John Booker and Gerard Roe. Conservation Laws

F11AE1 1. C = ρν r r. r u z r

1. Comparison of stability analysis to previous work

Offshore Hydromechanics Module 1

ENGR Heat Transfer II

MHD Free convection flow of couple stress fluid in a vertical porous layer

Temperature fields in a liquid due to the thermocapillary motion of bubbles and drops

HEAT TRANSFER OF SIMPLIFIED PHAN-THIEN TANNER FLUIDS IN PIPES AND CHANNELS

Solution Set Two. 1 Problem #1: Projectile Motion Cartesian Coordinates Polar Coordinates... 3

Transcription:

Thermocapillary Migration of a Drop An Exact Solution with Newtonian Interfacial Rheology and Stretching/Shrinkage of Interfacial Area Elements for Small Marangoni Numbers R. BALASUBRAMANIAM a AND R. SHANKAR SUBRAMANIAN b a National Center for Microgravity Research, NASA Glenn Research Center, Cleveland, Ohio, USA b Department of Chemical Engineering, Clarkson University, Potsdam, New York, USA ABSTRACT: In this paper we analyze the effects of the following phenomena associated with the thermocapillary migration of a drop. The first is the influence of Newtonian surface rheology of the interface and the second is that of the energy changes associated with stretching and shrinkage of the interfacial area elements, when the drop is in motion. The former occurs because of dissipative processes in the interfacial region, such as when surfactant molecules are adsorbed at the interface in sufficient concentration. The interface is typically modeled in this instance by ascribing to it a surface viscosity. This is a different effect from that of interfacial tension gradients arising from surfactant concentration gradients. The stretching and shrinkage of interfacial area elements leads to changes in the internal energy of these elements that affects the transport of energy in the fluids adjoining the interface. When an element on the interface is stretched, its internal energy increases because of the increase in its area. This energy is supplied by the neighboring fluids that are cooled as a consequence. Conversely, when an element on the interface shrinks, the adjoining fluids are warmed. In the case of a moving drop, elements of interfacial area are stretched in the forward half of the drop, and are shrunk in the rear half. Consequently, the temperature variation on the surface of the drop and its migration speed are modified. The analysis of the motion of a drop including these effects was first performed by LeVan in 98, in the limit when convective transport of momentum and energy are negligible. We extend the analysis of LeVan to include the convective transport of momentum by demonstrating that an exact solution of the momentum equation is obtained for an arbitrary value of the Reynolds number. This solution is then used to calculate the slightly deformed shape of the drop from a sphere. KEYWORDS: thermocapillary migration; Newtonian interfacial rheology; small Marangoni numbers; drop; stretching and shrinking of interface Address for correspondence: R. Balasubramaniam, National Center for Microgravity Research, NASA Glenn Research Center, Mail Stop 0-3, Cleveland, OH 4435, USA. Voice: 26-433-2878; fax: 26-433-3793. bala@grc.nasa.gov Ann. N.Y. Acad. Sci. 027: 8 (2004). 2004 New York Academy of Sciences. doi: 0.96/annals.324.024

2 ANNALS NEW YORK ACADEMY OF SCIENCES INTRODUCTION The thermocapillary migration of drops was first investigated theoretically by Young, Goldstein, and Block. When a drop is present in an unbounded liquid that has a uniform temperature gradient, a thermocapillary stress is generated at the interface arising from the variation of interfacial tension with temperature. Typically, the drop is propelled toward the warm portion of the continuous phase. Young et al. obtained a result for the steady velocity of a drop, including a contribution from the effect of gravity, in the limit when the Reynolds and Marangoni numbers approach zero, so that the effect of convective transport of momentum and energy can be neglected. To verify their theory, Young et al. performed experiments on air bubbles in silicone oils, in which the effects of thermocapillarity and gravity were counterbalanced, so that the bubbles were nearly stationary. Balasubramaniam and Chai 2 showed that the solution obtained by Young et al. is an exact solution of the Navier Stokes equations for any value of the Reynolds number, when the Marangoni number is equal to zero and the effect of gravity is not present. A discussion of other pertinent literature on the motion of drops due to thermocapillarity can be found in Subramanian and Balasubramaniam. 3 Our objective in this paper is to consider the effect of Newtonian interfacial rhelogy and stretching/shrinkage of interfacial area elements on the thermocapillary motion of a drop. In the limit when the Reynolds and Marangoni numbers approach zero, this problem has been addressed by LeVan 4 and discussed by Subramanian and Balasubramaniam. 3 The rheology of the surface needs to be considered in special situations, for example, when surfactant molecules are adsorbed on the interface in sufficient quantity so that an interfacial viscosity can be ascribed to the interface. As a consequence the drop encounters additional resistance to its motion. Scriven 5 has provided the framework for a Newtonian model of the interface, in which the contribution to the surface stress depends linearly on the rate of deformation of area elements on the interface. A detailed discussion of surface rheology models is available in Edwards, Brenner, and Wasan. 6 The second effect that we consider is the consequence of the deformation of interfacial area elements on the energy balance at the interface. Even when a drop moves in a continuous phase fluid that is isothermal, temperature variations can be generated on the interface. This is because elements of surface area grow as they move from the stagnation point to the equatorial region in the forward half of the drop. The increase in area increases the interfacial internal energy, which must be provided by the fluids adjoining the interface. Thus, the neighboring fluids are cooled. In the rear portion of the drop, the situation is opposite, and the adjoining fluids are warmed. Therefore, an interfacial temperature variation can occur, and the consequent thermocapillary stress influences the motion of the drop. This effect has been considered by Harper et al., 7 Kenning, 8 LeVan, 4 and Torres and Herbolzheimer. 9 One infers from the analysis presented by Subramanian and Balasubramaniam 3 that the thermocapilliary migration velocity of the drop is unaffected by the surface shear viscosity, but is influenced by the surface dilatational viscosity and the effect of stretching and shrinkage of interfacial area elements. The velocity of the drop is reduced as a consequence. We shall revisit the analysis performed by LeVan 4 and Subramanian and Balasubramaniam 3 below and show that their results are valid even

BALASUBRAMANIAM & SUBRAMANIAN: MIGRATION OF A DROP 3 when the convective transport of momentum is included in the analysis. The effect of inertia is shown to alter the pressure field, and lead to deformation of the drop. The small change in the shape of the drop from a sphere is obtained using a technique similar to that used by Balasubramaniam and Chai. 2 FORMULATION We consider the steady migration of a spherical drop of radius R 0 immersed in a liquid that is of infinite extent. The liquid has a density ρ, viscosity µ, thermal conductivity k, and thermal diffusivity κ. The corresponding properties in the drop are denoted by γρ, αµ, βk, and λκ. The rate of change of interfacial tension between the drop and the continuous phase liquid is denoted by σ T, and is assumed to be a negative constant. All the physical properties of the fluids, with the exception of the interfacial tension, are assumed to be constant. A temperature gradient of magnitude G is imposed in the continuous phase liquid. We assume that gravitational effects are not present. The surface shear and dilatational viscosities in the Newtonian model are denoted by µ * s and λ s *, respectively. The corresponding dimensionless surface viscosities, scaled by µr 0, are denoted by µ s and λ s. The quantity e s σ*, where e s is the surface internal energy per unit area and σ* is the interfacial tension, appears in the energy balance condition at the interface to accommodate the stretching and shrinkage of interfacial area elements. We assume that e s σ* is a constant over the drop surface, and define a dimensionless parameter E s = ( e s σ* )σ T ( µk). A reference velocity for motion in the fluids is obtained by balancing the tangential stress at the interface in the continuous phase liquid with the thermocapillary stress. This velocity scale is ( σ v T )GR 0 = -------------------------. 0 () µ In addition to the property ratios α, β, γ, and λ, the dimensionless interfacial viscosities µ s and λ s, and the interfacial internal energy parameter E s, the dimensionless parameters that are important in determining the motion of the drop are the Reynolds and Marangoni numbers. These parameters are defined using the above velocity scale and the properties of the continuous phase liquid. ρv Re 0 R = -------------- 0 (2) µ v Ma 0 R = -----------. 0 (3) κ The problem is analyzed in a reference frame attached to the moving drop. We use a spherical polar coordinate system, with the origin located at the center of the drop. The radial coordinate, scaled by R 0, is r. The polar coordinate measured from the direction of the temperature gradient is θ. The azimuthal coordinate is φ; however, the problem posed below for the velocity and temperature fields is independent of φ due to axial symmetry about the direction of the imposed temperature gradient. The scaled radial and tangential velocity fields in the continuous phase liquid are denoted by u and v, respectively, and those in the drop by u and v. The scaled

4 ANNALS NEW YORK ACADEMY OF SCIENCES migration velocity of the drop is denoted by v. These are obtained by dividing the physical velocities by v 0. The scaled pressure fields in the two fluids are p and p and are obtained by dividing the physical pressure by ρv 2 0. The temperature field in the continuous phase is scaled by subtracting the temperature in the undisturbed continuous phase at the location of the drop and dividing by GR 0. T = T Gv 0 v t ---------------------------, GR 0 (4) where T is the physical temperature and t denotes time. The scaled temperature T in the drop is defined similarly. The governing equations for the velocity and temperature fields in the continuous phase liquid and the drop are given below: ---- r 2 ---- r 2 ( u) + r r ----- ( vsin = 0 sinθ θ u u v ----- - ----- u v 2 p + ---- ----- ----- 2 2u 2 u ----- r r θ r r Re r 2 ---- v 2v = + r 2 ----- ----- θ r 2 cotθ u v v ----- - ----- v uv + + ----- -- p ----- ----- 2 2 v ---- r r θ r r θ Re r 2 ----- u v = + + ------------------ θ r 2 sin 2 θ v u T v ------ - T + + ------ = ------- 2 T r r θ Ma ---- r 2 ----- r 2 ( u ) + r r ----- ( v sin = 0 sinθ θ (5) (6) (7) (8) (9) u u ------- r + v --- u ------- v ------ 2 r θ r = -- p α ------- -- ----- 2 2u + u ------- γ r γ Re r 2 2 ---- v r 2 ------- θ 2v ------- r 2 cotθ u v v ------- --- v ------- u v + + -------- ---- p α ------- -- ----- 2 2 v ---- r r θ r γr θ γ Re r 2 ------- u v = + + ------------------ θ r 2 sin 2 θ v u T v ------- --- T λ + + ------- = ------- 2 T. r r θ Ma () (2) The boundary conditions are as follows. Far from the drop, the velocity field is uniform and equal to the migration velocity of the drop (i.e., in the laboratory reference frame, the fluid far away is quiescent). The temperature field far away is the imposed linear field. Therefore, as r, u v cosθ, v v sinθ (3) T rcosθ. (4) From the analysis by LeVan, 4 it can be shown that when the Reynolds and Marangoni numbers are set equal to zero, the normal stress balance is satisfied by a spherical shape of the drop. This is no longer the case when the Reynolds number is nonzero, the case considered here. Assuming the deformation from the spherical shape is small, it can be shown by using a domain perturbation scheme that it is permissible to obtain the leading order velocity and pressure fields by applying the relevant boundary conditions at the fluid fluid interface at a spherical boundary, designated here by r =. In the next section, we show how the small perturbation to this shape arising from inertia can be calculated using these leading order fields. At the fluid (0)

BALASUBRAMANIAM & SUBRAMANIAN: MIGRATION OF A DROP 5 fluid interface, the kinematic boundary condition holds, and the velocity and temperature fields are continuous. u(, = u (, = 0 v(, = v (, T (, = T (,. (5) (6) (7) The tangential stress balance and the heat flux balance at the fluid fluid interface appear below. It is in these boundary conditions that the effects of surface viscosity and surface internal energy appear. ----- v r r - (, α ----- v --- (, r r T ------ (, 2µ (8) θ s v(, ( λ s + µ s ) = ----- ---------- ----- { v(, sinθ} θ sinθ θ T ------ (, β T E ------- (, = ---------- s ----- [ v(, sinθ]. (9) θ r sinθ θ The migration velocity of the drop is determined from a force balance at steady state the drop is not accelerating, and hence, the net force acting on it must be zero, π ----- v r r - (, sin 2 θ Re p(, 2----- u (, + r cos θ sin θ dθ = 0 0 (20) RESULTS Exact Solution for Small Marangoni Number When Ma, convective transport of energy can be neglected compared with conduction, as a first approximation. In this case, the terms on the left side of Equations (8) and (2) may be ignored. Then, it can be shown that the solution first obtained by LeVan 4 and given in Subramanian and Balasubramaniam 3 for the temperature and velocity fields in the case Re = 0 continues to apply for arbitrary values of the Reynolds number and other parameters in the problem. Only the pressure field is modified by the inclusion of inertia. This exact solution is given below: B T = r + ---- (2) r 2 cosθ T = ( + B)r cosθ (22) u = v ---- (23) r 3 cosθ v = v + ------- 2r 3 sinθ 3 u = --v 2 ( r 2 ) cosθ (24) (25)

6 ANNALS NEW YORK ACADEMY OF SCIENCES v 3v r 2 = -- sinθ 2 v2 p p ----- cos 2 θ ---- 2 2 r 3 sin θ 2 = + + ------- 2 2r 3 9 p A 0 --γ v2 8 sin 2 θ( r 4 r 2 ) 2cos 2 θ r 2 r + ---- 4 5 v = + ---------r α cosθ 2 Re β Ω B = ---------------------- 2 + β+ Ω (26) (27) (28) (29) Ω = 2E ------------------------------ s 2+ 3α + 2λ s (30) 2 Ω v = ---------------------------------------------------------------- = (3) ( 2+ 3α + 2λ s )( 2 + β+ Ω) ---------------------------------- E s ( 2 + β+ Ω). The constant A 0 in the pressure field inside the drop can be determined only when the interfacial normal stress balance is considered. The velocity field outside the drop is irrotational and the pressure field outside the drop satisfies Bernoulli s equation. The flow field within the drop is the well known Hill s spherical vortex. The temperature and the velocity fields are coupled by the effects of surface viscosity and interfacial internal energy. The surface shear viscosity µ * s has no influence on the dynamics. The surface dilatational viscosity λ* s and the effect of the interfacial internal energy reduce the migration velocity of the drop. When λ s and E s are zero, the results given above reduce to those given by Balasubramaniam and Chai. 2 Small Deformation of the Drop We now consider the normal stress balance at the fluid fluid interface and calculate the slight deformation of the drop from a sphere. Let R = R 0 [ + f(] denote the shape of the drop. We assume that f. It will be seen shortly that this requires that the Weber number We = Re Ca. Here Ca is the Capillary number, defined by Ca = µv 0 /σ 0, where σ 0 is a reference value for the interfacial tension. For a slightly deformed drop, the normal stress balance at the interface can be written as 2Hσ 2λ τ rr τ rr = ---------- + ---------- s ----- [ v(, sinθ], (32) Ca sinθ θ where τ rr denotes the dimensionless normal stress at r = evaluated from the leading order fields given above, H is the mean curvature of the interface, and σ is the interfacial tension scaled by σ 0. Note that the surface viscosity makes a direct contribution to the normal stress difference (see Scriven 5 ). When the deformation from a spherical shape is small, 2H 2 2f ---------- ----- d sinθ df -----. (33) sinθdθ dθ The normal stress balance can then be written as follows:

BALASUBRAMANIAM & SUBRAMANIAN: MIGRATION OF A DROP 7 v2 Ca Re A 0 p ----- 9 2 4 --γ 9 + + -- ( γ ) sin 2 θ 4 (34) = 2 [ Ca( + B) cosθ] 2 f + ---------- ----- d sinθ----- df. sinθdθ dθ From Equation (34), one can conclude that, for a fixed Reynolds number, f( O(Ca) as Ca 0. In this asymptotic limit, the leading order differential equation for f( is a nonhomogeneous version of the differential equation for the Legendre Polynomial P (cos. Note the absence of any terms proportional to cosθ in the inhomogeneity. If such a term is present, it can be shown that it will lead to unboundedness of the solution for f along θ = 0 and π. In fact, if a force balance on the drop had not been used to determine the speed of the drop, the speed can be obtained by requiring the term proportional to cosθ in the inhomogeneity to be zero to eliminate such singular behavior. We impose the constraints that the volume of the drop is a constant (equal to 4πR 3 0 3) and that the origin of the coordinates is the center of mass of the drop. Linearization of these constraints for small f leads to the following equations: π f ( sinθdθ = 0 (35) 0 π f ( cosθsinθdθ = 0. (36) 0 The solution can be obtained in a straightforward manner, as shown by Brignell. 0 The unknown constant A 0 is determined as part of the solution, 2 v2 A 0 -------------- p 3 = + (37) Re Ca ----- + --γ 4 2 3 f ( = -----Re Ca v 2 (38) 32 ( γ ) ( 3cos 2 θ ). The dimensionless surface dilatational viscosity λ s and the surface internal energy parameter E s affect the deformation of the drop via their influence on v (given in (3)). Because we have restricted the deformation of the drop to be small, we see that We = ReCa must be a small parameter. Drops less dense than the continuous phase liquid are predicted to contract in the direction of the temperature gradient; that is, they attain an oblate shape. Drops more dense than the continuous phase attain a prolate shape. CONCLUDING REMARKS We have shown that a solution in the limit of zero Reynolds and Marangoni numbers for the velocity and temperature fields, obtained elsewhere 4,3 for the thermocapillary migration of a drop accommodating Newtonian surface rheology and the influence of internal energy changes associated with the stretching and shrinkage of surface elements, holds for arbitrary values of the Reynolds number so long as the Marangoni number is set equal to zero. The pressure fields are modified by inertia. The leading order result for the migration speed of the drop is the same as that obtained when the Reynolds number is zero. Furthermore, whereas the drop will be

8 ANNALS NEW YORK ACADEMY OF SCIENCES spherical when the Reynolds number is zero, this is no longer the case for non-zero values of the Reynolds number. A perturbation result for the small deformation of the drop from a spherical shape is given in the case where the Capillary number is asymptotically small, for fixed Reynolds number. REFERENCES. YOUNG, N.O., J.S. GOLDSTEIN & M.J. BLOCK. 959. The motion of bubbles in a vertical temperature gradient. J. Fluid Mech. 6: 350 356. 2. BALASUBRAMANIAM, R. & A.T. CHAI. 987. Thermocapillary migration of droplets: an exact solution for small Marangoni numbers. J. Colloid Interface Sci. 9(2): 53 538. 3. SUBRAMANIAN, R.S. & R. BALASUBRAMANIAM. 200. The Motion of Bubbles and Drops in Reduced Gravity. Cambridge University Press, London/New York. 4. LEVAN, M.D. 98. Motion of a droplet with a Newtonian interface. J. Colloid Interface Sci. 83(): 7. 5. SCRIVEN, L.E. 960. Dynamics of a fluid interface. Chem. Eng. Sci. 2: 98 08. 6. EDWARDS, D.A., H. BRENNER & D.T. WASAN. 99. Interfacial Transport Processes and Rheology. Butterworth-Heinemann, Boston. 7. HARPER, J.F., D.W. MOORE & J.R.A. PEARSON. 967. The effect of the variation of surface tension with temperature on the motion of bubbles and drops. J. Fluid Mech. 27(2): 36 366. 8. KENNING, D.B.R. 969. The effect of surface energy variations on the motion of bubbles and drops. Chem. Eng. Sci. 24: 385 386. 9. TORRES, F.E. & E. HERBOLZHEIMER. 993. Temperature gradients and drag effects produced by convection of interfacial internal energy around bubbles. Phys. Fluids A 5(3): 537 549. 0. BRIGNELL, A.S. 973. The deformation of a liquid drop at small Reynolds number. Quart. J. Mech. and Applied Math. XXVI(): 99 07.