Interfaces for st the 21 Century New Research Directions in
Fluid Mechanics and Materials Science
Marc K. Smith, Michael J. Miksis, Geoffrey B. McFadden, G. Paul Neitzel & David R. Canright editors
Imperial College Press
Interfaces for st the 21 Century New Research Directions in Fluid Mechanics and Materials Science
This page is intentionally left blank
Interfaces for st the 21 Century New Research Directions in Fluid Mechanics and Materials Science
A collection of research papers dedicated to Steven H. Davis in commemoration of his 60f birthday
editors
Marc K. Smith Georgia Institute of Technology, USA
Michael J. Miksis Northwestern University, USA
Geoffrey B. McFadden National Institute of Standards and Technology, USA
G. Paul Neitzel Georgia Institute of Technology, USA
David R. Canright Naval Postgraduate School, USA
ICP
Imperial College Press
Published by Imperial College Press 57 Shelton Street Covent Garden London WC2H 9HE Distributed by World Scientific Publishing Co. Pte. Ltd. P O Box 128, Farrer Road, Singapore 912805 USA office: Suite IB, 1060 Main Street, River Edge, NJ 07661 UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library.
INTERFACES FOR THE TWENTY-FIRST CENTURY: NEW RESEARCH DIRECTIONS IN FLUID MECHANICS AND MATERIALS SCIENCE Copyright © 2002 by Imperial College Press All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the Publisher.
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from the publisher.
ISBN 1-86094-319-5
Printed in Singapore by Uto-Print
PREFACE In the science of mechanics, the term interface is generally used to refer to a material boundary between two phases, i.e., liquid-gas, solid-liquid or solid-gas. Specific examples in fluid mechanics are the spreading of a liquid droplet and the dynamics of a thin liquid film; specific examples in materials science are the solidification of a solid in a melt and an elastically stressed solid interface which evolves by surface diffusion. The primary difficulty in the modeling of interfaces is the fact that there locations are, in general unknown a priori. This difficulty has resulted in the development of many analytical and numerical solution techniques. Here a number of new and exciting solution methods as well as the physical problems which have initiated them will be discussed. Our hope is that the collection of work presented here will encourage others to work in and initiate new research into this important field. The material contained in this volume is the bulk of that presented in oral form at a conference, Interfaces for the Twenty-First Century, held August 16-18, 1999, in Monterey California. These oral presentations were invited from experts renowned for their research in the field of interfacial mechanics. In addition, abstracts of contributed posters which were also presented are included. The focus of the conference was on interfacial topics in fluid mechanics and materials science. The presentations covered experimental as well as theoretical approaches with an overlying philosophy of the conference being to investigate new techniques for modeling, both theoretically and numerically, the dynamics of interfaces. The conference was supported by the National Science Foundation and the National Aeronautics and Space Administration, for which the organizing committee and conference participants are extremely grateful. The genesis of the conference which produced the present volume was to hold a meeting to pay tribute to one of the world's foremost researchers in the field of interfacial mechanics, Professor Stephen H. Davis, on the occasion of his sixtieth birthday (September 7, 1999). Steve, through his outstanding research and mentoring of a generation of graduate and post-doctoral students, has contributed profoundly to our present knowledge in the field. It is with gratitude and pleasure that we present the contributions of some of the major investigators in this field as a tribute to the work of Steve Davis. Marc K. Smith Michael J. Miksis Geoffrey B. McFadden G. Paul Neitzel David R. Canright January 2002
This page is intentionally left blank
CONTENTS
Preface
v
PART 1: INVITED PAPERS
1
The effect of a stabilising gradient on interface morphology T. Maxworthy
3
Spreading of a liquid drop with mass loss L. M. Hocking
21
Viscous gravity currents with solidification M. Bunk, P. Ehrhard, and U. Mutter
35
Coarsening dynamics of roll waves H.-C. Chang and E. A. Demekhin
51
Thermo capillary control with feedback of large wavelength interfacial instabilities R. E. Kelly
61
Pattern formation in thin liquid films D. Gallez and E. R. de Souza
73
Molecular aspects of contact-line dynamics J. Koplik and J. R. Banavar
89
Computational methods for advancing interfaces J. A. Sethian
105
Direct numerical simulations of multiphase flows G. Tryggvason and B. Bunner
121
A phase-field model with convection: numerical simulations D. M. Anderson, G. B. McFadden, and A. A. Wheeler
131
Phase field model of multicomponent alloy solidification with hydrodynamics R. F. Sekerka and Z. Bi
147
The effects of a stress-dependent mobility on interfacial stability P. W. Voorhees and M. J. Aziz
167
Non-constant growth characteristics of pivalic acid dendrites in microgravity J. C. LaCombe, M. B. Koss, A. O. Lupulescu, J. E. Frei, and M. E. Glicksman
177
viii
Contents
Interfaces on all scales during solidification and melting M. G. Worster
187
Phase and microstructure selection in peritectics W. Kurz and S. Dobler
203
Model phase diagrams for an FCC alloy R. J. Braun, J. Zhang, J. W. Cahn, G. B. McFadden, and A. A. Wheeler
213
PART 2: CONTRIBUTED ABSTRACTS
231
Influence of contact-angle conditions on evolution of solidification fronts V. S. Ajaev and S. H. Davis
233
Creeping steady thin film on an inclined plane with an edge N. Aksel
234
Phase-field simulation of convective effects on dendritic growth G. Amberg and R. Tonhardt
235
A model for a spreading and melting droplet on a heated substrate D. Manderson, M. G. Forest, and R. Superfine
236
Instabilities of a three-dimensional liquid droplet on a heated solid surface S. W. Benintendi and M. K. Smith
237
A laboratory model for the solidification of the Earth's inner core and the inner core's seismic anisotropy M. I. Bergman
239
Suppression of channel convection in solidifying Pb-Sn alloys via an applied magnetic field M. I. Bergman, D. R. Fearn, and J. Bloxham
240
Dynamics and stability of Van-der-Waals-driven thin film rupture A. J. Bernoff and T. P. Witelski
241
New approaches to front-tracking and front-capturing methods J. Brackbill, D. Jamet, 0. Lebaigue, and D. Torres
242
Manipulation of intravascular gas embolism dynamics with exogenous surfactants A. B. Branger and D. M. Eckmann
244
Adiabatic hypercooling of binary melts K. Brattkus
245
An insoluble surfactant model for a draining vertical liquid film R. J. Braun
246
Contents
ix
The effect of time-periodic airway wall stretch on surfactant and liquid transport in the lung J. L. Bull, D. Halpern, and J. B. Grotberg
247
The dynamic effects of surfactants on stationary gas bubbles in liquid flows D. P. Cavanagh and D. M. Eckmann
248
Buckling instabilities in thin viscous sheets S. Chaieb, R. da Siliveira, L. Mahadevan, and G. H. McKinley
249
Fluid-fluid interface experiments at the University of Chicago /. Cohen, S. R. Nagel, M. P. Brenner, J. Eggers, R. O. Grigoriev, and T. F. Dupont
250
Asymptotic estimates for 2-D sloshing modes: theory and experiment A. M. J. Davis and P. D. Weidman
251
Buoyancy-driven interactions of viscous drops with deforming interfaces R. H. Davis, J. Kushner, and M. A. Rother
252
Anomaly and uncertainty when liquid films flow over solid surfaces W. Debler
253
Flow behavior of Langmuir monolayers M. Dennin and R. S. Ghaskadvi
254
Soluble surfactants and contact-angle dynamics D. M. Eckmann
255
Experimental studies of the hydrodynamics near moving contact lines S. Garoff and E. Rame
256
Axisymmetry-breaking instabilities in axisymmetric freezing of ice A. Yu. Gelfgat, P. Z. Bar-Yoseph, A. Solan, and T. A. Kowalewski
257
Separation mechanics of thin interfacial liquid layers: the role of viscous fingering A. Gopinath
258
Large finite-element modeling of axially symmetric free-surface flows R. Grigoriev and T. Dupont
259
Molecular simulations of interface phenomena: an alternative approach N. Hadjiconstantinou
260
Experimental investigation of environment-oxygen content in solder-jet technology E. Howell, S.-Y. Lee, C. M. Megaridis, M. McNallan, and D. Wallace ...261 Droplet spreading with surfactant: modeling and simulation J. Hunter, Z. Li, and H. Zhao
263
Air entrainment at low viscosities A. Indeikina, I. Veretennikov, and H.-C. Chang
264
x
Contents
Numerical simulations of vibration-induced droplet ejection A. James, M. K. Smith, and A. Glezer
265
Interfacial dynamics associated with evaporation of LNG in a storage tank S. W. Joo, C. Park, and S. Hong
266
Oscillatory thermocapillary convection generated by a bubble M. Kassemi and N. Rashidnia
267
About computations of thin-film flows L. Kondic, J. Diez, and A. Bertozzi
268
Visualization of convection in liquid metals J. N. Koster
269
Time-evolving interfaces in viscous flows M. C. A. Kropinski
270
Influence of a nonlinear equation of state on contamination fronts at air/water interfaces J. M. Lopez and A. H. Hirsa
271
Stabilization of an electrically conducting capillary bridge far beyond the Rayleigh-Plateau limit using feedback control of electrostatic stresses M. J. Marr-Lyon, D. B. Thiessen, F. J. Blonigen, and P. L. Marston
272
Stabilization of capillary bridges in air far beyond the Rayleigh-Plateau limit in low gravity using acoustic radiation pressure M. J. Marr-Lyon, D. B. Thiessen, and P. L. Marston
273
Interactions between Hele-Shaw flows and directional solidification: numerical simulations E. Meiburg
274
Modelling the contact region of an evaporating meniscus with a view to applications S. J. S. Morris
275
Aspects of vortex dynamics at a free surface B. Peck, L. Sigurdson, P. Koumoutsakos, and J. Walther
277
Instabilities at the "interface" between miscible fluids — emergence of an effective surface tension P. Petitjeans, P. Kurowski, and J. Fernandez
279
Secondary instabilities of falling films using models C. Ruyer-Quil and P. Manneville
280
Interfacial phenomena in suspensions U. Schaflinger and G. Machu
282
Sessile drop solidification W. W. Schultz, M. G. Worster, and D. M. Anderson
283
A homogenized Monte-Carlo model for film growth T. Schulze
284
Contents
xi
Influence of the surrounding conditions near the interface on the stability of liquid bridges V. M. Shevtsova, M. Mojahed, and J. C. Legros
285
Convective-diffusive lattice models of interfacial and wetting dynamics Y. Shnidman
286
A level-set approach to domain growth in multicomponent fluids K. A. Smith, F. J. Solis, and M. 0. de la Cruz
287
Morphological instability in strained alloy films B. J. Spencer, P. W. Voorhees, and J. Tersoff
288
Stability issues in spin-casting molten metals P. H. Steen
289
The stability of thermocapillary convection in half zones with deformed free-surface profiles L. B. S. Sumner and G. P. Neitzel
291
Direct write of passive circuitry using ink-jet technology J. Szczech, C. M. Megaridis, D. Gamota, and J. Zhang
292
Stability of the meniscoid particle band at advancing interfaces in Hele-Shaw suspension flows H. Tang, W. D. Grivas, T. J. Singler, J. F. Geer, and D. Homentcovschi
293
Instabilities in thin fluid sheets B. S. Tilley, D. T. Papageorgiou, and R. V. Samulyak
295
Thermal effects of internal interfaces: equilibrium microstructure and kinetics A. Umantsev
296
The role of long-range forces in the stability of nematic films M. P. Valignat, F. Vandenbrouck, and A. M. Cazabat
297
Vibration-induced drop atomization B. Vukasinovic, M. K. Smith, and A. Glezer
298
Capillarity-driven instabilities and the evolution of solid thin films H. Wong, M. J. Miksis, P. W. Voorhees, and S. H. Davis
300
Interfacial wave theory for dendrite growth J.-J. Xu
301
Uniformly valid asymptotic solutions for dendrite growth with convection J.-J. Xu and D.-S. Yu
302
Comparison of asymptotic solutions of phase-field models to a sharp-interface model G. W. Young and S. I. Hariharan
303
xii
Contents
A new approach to measure the contact angle and the evaporation rate with flow visualization in a sessile drop N. Zhang and D. F. Chao
PART 3: PANEL DISCUSSION SESSION New research directions in interfacial science M. K. Smith
304
305 307
PART 1
INVITED PAPERS
This page is intentionally left blank
THE EFFECT OF A STABILISING G R A D I E N T O N INTERFACE MORPHOLOGY
T. MAXWORTHYt Department
of Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, CA 90089-1191, U.S.A.
We have been studying the motion of sedimenting, surface gravity currents and the resultant motion of particles through the interface between a heavy, ambient fluid and a lighter surface layer (Maxworthy [1,2]). As noted by Green [3] and Chen [4] this latter motion has many features in common with double-diffusive interfaces, but further study has revealed a similarity to a wide range of problems involving the stability and morphology of interfaces in general. Some of these similarities have been discussed before (Michalland [5] and others) in another context and we discuss these and related applications.
1
Experiments to Study the Dynamics and Stability of a Sedimenting Interface
We have undertaken a series of experiments on the dynamics of sedimenting, surface gravity currents in a rectangular tank 220 cm long, 57 cm deep, and 15.7 cm wide (Maxworthy [1]). The physical situation concerns a current, with total density pc, evolving at the surface of a fluid of greater density, p^. In turn, pc is made up of interstitial fluid of density pi and heavy particles with a concentration, by weight, c and a density pp. The relevant dimensionless parameter is denoted as R = iPA—pc)/(pc—pi), which has a close relationship [R = (Rp — l)} to the densitydifference parameter, Rp, used in the study of double diffusion (Turner [6], p. 274). In summary it was found that the sedimentation of the particles, plus some of the interstitial fluid, through the interface between the two fluids has a profound effect upon the motion of the current. In order to study this trans-interfacial particle/fluid motion in detail a subsidiary experiment was set-up by running a mixture of particles and pure water on top of a deep layer of salt water in a smaller, narrower tank (10 x 7 x 0.4 cm) (Maxworthy [2]). In order to observe the resultant flow only relatively large values of R were used, since, for small values, the instability evolved so quickly that interpretation was difficult. In Fig. la, a series of photographs of the resultant particle motion is shown for R = 3.1. In the first of these photographs one can see a number of the larger particles falling out of the upper region, the instability then appears on the interface after a sufficient number of the smaller particles in the distribution had entered the lower layer and its effective density, and thickness, was large enough for the layer to undergo a Rayleigh-Taylor type of instability. Of especial interest is the observation that the instability is locked in its axial location by the underlying stable fluid distribution. In Fig. lb we compare the shape of a typical cell with the Saffman T
This contribution is dedicated to my good friend Professor Stephen H. Davis on the occasion of his sixtieth birthday. He has been and continues to be a source of inspiration to us all.
3
4
T.
Maxworthy
Figure 1. a) Sequence of photographs of an unstable, sedimenting interface between heavier-lower and lighter-upper fluid layers. Instability first appears between times ii) and iii). The majority of the heavy particles drain from the upper layer through thin fingers that are about 0.5-0.7 mm wide. During this motion they drag a substantial amount of upper-layer fluid downwards with them, b) Comparison between the experimental interface shape and that due to Saffman and Taylor [7] for an immiscible interface in a Hele-Shaw cell. For a value of A = 0.95. c) Idealised representation of the interface before sedimentation starts, d) After a time At the monodispersed particles have formed a heavier layer in the lower layer that is v$ At thick. This is then susceptible to a gravitational instability of the Rayleigh-Taylor type.
and Taylor [7] solution for an air finger penetrating into a viscous fluid with the same width ratio (A), i.e., the width of the particle-free region divided by the total
The Effect of a Stabilising
Gradient on Interface Morphology
5
width of the cell. Here, we have chosen a value for A — 0.95 that appears to be about the correct value for the sedimenting finger. The solution given is for a theory that does not take surface tension into account and gives the flattened shape shown in Fig. lb. The finger shape including this latter effect is more rounded (see e.g., Pitts [8], McLean and Saffman [9]) and such finger shapes are similar to the shapes found in flows with free surfaces at which surface tension is an important dynamical effect, as discussed below. In summary, the sedimenting interface is then one example of a class of problems in which the axial gradient of a stabilising quantity sets the morphology of the resultant instability and often this morphology is similar to that found in Saffman-Taylor fingers. In the photographs, note also the lateral interaction of the cells as the finite-amplitude instability establishes itself. This behaviour is also characteristic of systems stabilised by an axial gradient, as discussed in what follows. Based on these observations, we have explored possible theoretical explanations for the instability of the particle-laden layer (Maxworthy [2]). For our present purposes, we consider only an idealized situation in which the initial density distribution is two-layered, with an interface of zero thickness and with a monodispersed particle distribution, as shown in Fig. lc. After a short time, At, the particle-front enters the lower layer to produce an unstable, heavy layer of thickness, vsAt, and a density close to [PA + (pc — Pi)], sandwiched within the stable fluid density distribution (Fig. Id). Here, vs is the settling velocity of the particles. After some time this heavy layer becomes unstable and produces a sequence of fingers that drain particles and fluid from the upper layer. For this simple model, the use of the theoretical results of Chandler and Redekopp [10] gives a time to instability that is somewhat smaller than that measured experimentally. In a more complete discussion of this problem (Maxworthy [2]), we have obtained better agreement between theory and experiment by considering a number of extra effects: e.g., the unsteady growth of the intermediate layer, the finite thickness of the density interface, the initial particle distribution within that interface, the polydispersion of that particle distribution, and the fact that the collective effect of the particles is to drag some of the light, upper-layer fluid downwards with them, reducing their effective buoyancy and hence velocity, as they traversed the stable layer. We wish to extract two points from this introductory section: firstly, to show the basic physical mechanisms that lead to instability in this previously unexplored example and, secondly, to show the morphology of the finite-amplitude interface that is formed under the conditions of these experiments. It is this second point that we wish to explore further in what follows. That is, we present diverse physical situations that produce similar morphologies, at finite amplitude, to the ones shown in Fig. la. The requirement to do so seems to be that there is some physical parameter representing a gradient of force that acts normally to the interface to prevent axial competition between the growing instability waves and, secondarily, reduces the growth rate. In the case discussed above this gradient is created by the difference in density between the upper and lower fluid layers. In what follows, we look at different mechanisms that create similar patterns. In order to inter-compare mechanisms that inhibit axial competition, it is probably appropriate to first show examples of this competition between growing insta-
6
T.
Maxworthy
Figure 2. Growth of a gravitationally unstable, immiscible interface in a Hele-Shaw cell, showing the axial competition between fingers that results in one finger dominating the others. Reprinted with permission from The Physics of Fluids.
bility waves so that their characteristics are evident and it is clear what we mean by the term. There are numerous examples, so we show just one that is the result of non-linear growth of an immiscible interface in a Hele-Shaw cell of constant gap width (Fig. 2). In this case, when one wave grows a little faster than its neighbours it produces a flow field that prevents the growth of these neighbours so that it can outrun them more readily the larger it becomes. In the literature on the subject this is sometimes called a "shielding" effect. If the growth discussed above takes place in a gap with a width that is not constant, the instability is regularised if the front is moving perpendicular to, and in the direction of, the negative gradient in gap width. That is, the attempt by any one wave to outrun neighbouring waves, that are slightly retarded, is inhibited by the weaker backfiow it generates at their location where the gap width is larger. The result is a growth that is uniform along the length of the interface. Contrariwise, we assume that competition in the direction of the positive gradient in gap width is enhanced since the flow field produced at the retarded perturbation is increased in magnitude by the decrease in width at that location. The parameter that seems to organise this behaviour, in the case of immiscible Hele-Shaw flows, is the ratio T of the wavelength of the instability to a length scale characteristic of the gap-width variation, ho/(dh/dx), where x is the distance perpendicular to the interface, h is the local gap height, and the O-subscript indicates the gap width at the x location of the interface. Since, from linear theory at low values of the capillary number Ca, the most unstable wavelength scales approximately
The Effect of a Stabilising Gradient on Interface Morphology
7
as ho/Ca1/2 (Maxworthy [11]), then Y = {dh/dx)/Ca1/2, where Ca = Ufi/a, U is a characteristic interface velocity, /i the absolute viscosity of the fluid, and a the surface tension at the interface. On the other hand, at values of Ca of order one and greater the most unstable wavelength scales as the width of the gap ho (Maxworthy [11]) so that r = dh/dx. In the region around Y = 0, classical SafFman-Taylor growth takes place and there is weak axial competition. For Y relatively large and negative, i.e., the unstable interface movement is in the direction of the negative gradient of height, axial competition is inhibited and the cells can only interact laterally. It is this case that interests us most in what follows. For Y relatively large and positive one would expect that axial competition would be enhanced, although, as far as we know, no experiments have been undertaken to look at this case. 2
Interface Stability in a Tapered Gap, as in a Journal Bearing, a.k.a.: Printer's Instability, Coating Instability, and Directional Viscous Fingering
In what follows, we discuss a number of examples of the effect of a stabilising gradient in layer thickness starting with the one that appears to have chronological priority: "Journal-Bearing Instability," sometimes called "Printer's Instability" (Michalland [5]), "Coating Instability," or "Directional Viscous Fingering," the latter name to emphasis the analogy with the problem of directional solidification, to be discussed later. This type of instability occurs under many different circumstances, with the section title revealing several examples. The generation of streakiness in fluid films in many industrial processes has been observed for many years, but it was with Pearson [12], who first analysed the streaks generated by a simplified version of a paint brush, that attempts were made to place these on a scientific footing. Recent work on this type of instability in journal bearings appears to have started with the papers of Banks and Mill [13] and Cole and Hughes [14], followed by Pitts and Greiller [15], Floberg [16], and Taylor [17]. Of special interest in the latter case are his Plates 1 and 2, with the latter being reproduced in Fig. 3. These are photographs of the fingers found in a tapered gap of divergence angle 2.8°. Comparison with the finger shape found in the study of SafFman and Taylor [7] is not satisfactory. However, when this theory is modified to take the effect of surface tension into account (Pitts [8] and McLean and Saffman [9]) the agreement is much better, as shown in Fig. 3. This is confirmed by the experiments of Michalland [5], in which the shapes were compared with a theory by Ben Amar (unpublished, Fig. 4) that is apparently similar to that of McLean and Saffman [9]. Michalland's experiments are among the most comprehensive known to us and will form the basis for much of the discussion that follows. Parts of the thesis have been extracted as published papers and these are the most readily available [18-23]. In Fig. 5a, we show a simple schematic of the double-roller apparatus used by Michalland, and in Fig. 5b a sequence of the shapes taken by the trailing interface (location xm\ of Fig. 5a) as the speed of the inner cylinder is increased, with a non-rotating outer cylinder. The progression from a sinusoidal shape through
8
T.
Maxworthy
Figure 3. The form of the fingers found in a journal bearing with the shape compared with the solutions of McLean and Saffman [9] and Pitts [8].
shapes that are more and more non-linear is apparent, with the final ones being characteristic of a wide range of problems with a stabilising axial gradient (in this case of gap thickness), as will be shown in what follows. This very basic system displays a wide range of interesting behaviour. In particular, various types of lateral interaction is possible between cells. In Fig. 6, we show an example of space-time traces that exhibit cell splitting, wave propagation, chaos, etc. Michalland also presents the formal similarity between these types of instability and those found in directional solidification, to be discussed in more detail in Sec. 3. 2.1
Peeling Instability
The Peeling Instability is intimately related to the cases discussed above and was explored in detail by McEwan and Taylor [24]. The theory, the apparatus devised for the experiments, and the experiments themselves, show the imagination and skill inherent in the body of work from both authors. A horizontal roller supports the free end of a sheet of plastic that is, at the same time, loaded with a weight of known magnitude. The other end of the sheet is "glued" to a flat surface using a Newtonian fluid of known properties that was rolled to a constant, known thickness before an experiment began. An experiment was started by placing weights in the
The Effect of a Stabilising
Gradient on Interface Morphology
9
Figure 4. Comparison between the interface shapes found by Michalland [5] and the theory of Ben Amar (unpublished) for the double-roller apparatus of Fig. 5a.
pan and then raising the roller at a known rate. The speed of peeling and the angle of the sheet (5 were then set by these external parameters and the properties of the meniscus at the line of detachment. For our present purposes, we are interested in the shape of the instability cells formed at the line of detachment. Since the photographs taken from the original paper are hard to reproduce informatively, a photograph, taken using the apparatus described in Sec. 2.2, which is dynamically identical to the McEwan-Taylor arrangement, is shown in Fig. 7. The correspondence between these shapes and those found in the printer's instability is quite clear, especially when one inter-compares them with the Saffman-Taylor solutions with surface tension (McLean and Saffman [9] and Pitts [8]), for the shape of a finger with the same width ratio. 2.2
A Simple Home/'Lecture
Demonstration
The examples given above are easily demonstrated by an experiment that can be performed with a very simple apparatus that is easy to transport. All that is needed is two pieces of thin plastic sheet approximately 30 x 15 x 0.15 cm. To each of the two short ends of one is attached a commercial spring (paper) clip. The sheet is then bent and a piece of wire, tied to the clips, is used to maintain the curvature. A small
10
T.
Maxworthy
Figure 5. a) A schematic of the double-roller, journal-bearing apparatus used by Michalland [5] to study stability of the interface at the leading edge ( x m i ) . b) Photographs of the interface shapes found by Michalland [5] for various values of the velocity of the inner cylinder. In this case, the outer cylinder was stationary. Velocity increases from top to bottom. At the highest velocity the cell shapes are well represented by the Pitts [8], McLean and Saffman [9], and Ben Amar (unpublished ) solutions as shown in Fig. 4.
amount of a readily available, but transparent, viscous fluid (honey, syrup, or olive, canola, motor oils, etc., shampooing liquid, oil painting oil, etc.) is poured onto the flat sheet and the curved one is pressed into it (see Fig. 8). By rocking the latter back and forth or dragging it over the fluid layer one can generate the instability and cells of interest. The onset of instability, the spacing between cells, the lateral interaction between cells, and other behaviour can be seen to depend on the speed of rocking or towing, the viscosity and surface tension of the oil, its thickness far from the gap and the minimum value under the curved sheet. By working out the relative velocities between the two sheets one can determine that the system involving rocking the sheet is an exact analoque of the peeling instability as well as of some variations of the printer's instability. Photograph of typical "experiments" using this apparatus are shown in Figs. 7 and 8b. It is especially useful as a lecture demonstration since it can be placed on an overhead projector and viewed, at large scale, on a lecture-hall screen. 2.3
Experiments in a Tapered, Hele-Shaw Cell
A useful review of the effects of various perturbing mechanisms on unsteady interface dynamics in a Hele-Shaw cell has been presented by McCloud and Meyer [25]. In particular, they look at the effects of a uniform gradient in thickness on the
The Effect of a Stabilising
Gradient on Interface Morphology
11
Figure 6. A spatio-temporal representation of the regime of intermittant behaviour in the experiments of Michalland [5]. Locations A correspond to regions of regular behaviour and B to regions of chaotic behaviour. Regions of wave propagation and cell splitting and coalescence can also be seen clearly. From Michalland [5].
stability of an immiscible interface. The basic conclusion from the theoretical work is that the effect is small unless the gap slope is very large. They report on some experimental work in a relatively narrow cell, using a very small slope in thickness, in which only a few waves could be produced. This made the accurate determination of the instability wavelength very difficult as is apparent in their large data scatter. Recently, we have extended the work reported by Maxworthy [11] to include the effect of gap-width variation. Here, we have manufactured a gap that varies radially in a cylindrical geometry at an angle of 2.50° and used the same techniques as Maxworthy [11]. A photograph of the interface evolution in such a case is shown in Fig. 9. At values of Ca below about 0.2 it was impossible to generate a clean, unstable interface. As a result, a different method was used that involved tilting a flat plate in a thin pool of viscous fluid. Here, the angle was varied as the plate was tilted, but the gap width at the location of the interface remained approximately constant at 6.0 mm. Typically, the instability took place when the angle was approximately 1.50°, while the values of the most unstable wavelength were a seamless continuation of the values from the radial experiment (see Fig. 10). The most unstable wavelength, L, divided by the gap width, 6, as a function of Ca is shown in Fig. 10 (Maxworthy [26]), where they are compared with the results
12
T. Maxworthy
Figure 7. Three examples of the interface shapes found in the peeling instability. The apparatus used here is described in Sec. 2.2 and is formally identical to that used in the original experiments of McEwan and Taylor [24].
using a gap of constant width (Maxworthy [11]). It appears that with a gap gradient of between 1.5 and 2.50 the region of L/b approximately equal to 5-6 is extended to smaller values of Ca than in the case of constant gap width, with no indication of a tendency towards the linear instability results that give values of L/b greater than approximately 20-25. This would suggest, in line with previous results, that three- dimensional, viscous effects (Paterson [27]) dominate the dynamics at these small, but finite angles and at all values of Ca that result in an instability. As found by Michalland [5], in the case of the Printer's Instability, it was not possible to create an unstable interface below values of Ca of about 0.1, probably due to the stabilising effect of gravity on the non-uniform layer in the gap (Maxworthy [26]). 3
Effect of a Stabilising Temperature Gradient on a Solidifying Interface: Directional Solidification
The stability of a solidifying interface is of great practical importance as demonstrated by the enormous literature it has generated. Here, we concentrate on a small corner of this research and look at the effects of a stabilising temperature gradient on the stability of such a front. A typical apparatus consists of a HeleShaw cell filled with a transparent, organic, liquid alloy, e.g., succinonitrile/acetone, that is traversed through the temperature gradient formed between a cold and a hot reservoir. The relevent stability parameter is the morphological number M: M =
mVC0{k~l) kDLG
(1)
where V is the towing speed, G the temperature gradient, m the liquidus slope, k the segregation coefficient, CQ the solute concentration at infinity, and DL the
The Effect of a Stabilising
Gradient on Interface Morphology
Rocking Motion \
Spring Clip
13
.Cord Spring Clip
TT^'X^s
/ / s /\s / s /1}
Fluid
Spacer
r
~=~s b). Figure 8. a) A sketch of the apparatus used to demonstrate the Peeling Instability, b) A threequarter view of an experiment performed in the apparatus of Fig. 8a.
solute's diffusivity. For a given alloy when V and G are such that M is slightly above unity the interface becomes unstable to a sinusoidal instability that evolves over time to the non-linear shapes shown in Fig. lla,b; c.f., compare to Fig. 5b for the case of the Printer's Instability. This asymptotic shape is in every way similar to those discussed before, and a direct comparison with the Saffman-Taylor finger shape, with surface tension, is favourable (Fig. lie). If M is substantially above unity then the cells eventually develope perturbations at their tips that evolve into dendrites (Fig. lid) that also suppress the growth of their neighbours. In this case, the temperature gradient is not large enough initially to prevent axial competition between cells, but as the cell spacing becomes larger this is no longer true and axial competi-
14
T.
Maxworthy
m®&
Figure 9. Images from a video recording of the evolution of an immiscible interface in a circular Hele-Shaw cell with a gap thickness that varies radially (Maxworthy [26]). In this case L/b as 5.2.
20 L/b (402&230cP;Circular Plate) L/b(137cP) L/b(63cP)
15
10
L/b (63cp;Tilting Plate)
-
0.01
0.1
Ca
10
Figure 10. The variation of L/b with Ca over a range of the latter of almost two orders of magnitude. The results for a constant gap width are shown by the curve marked Maxworthy [11]. The error bars are ± one standard deviation in both cases. Note that L/b has a value close to that found at large Ca for constant gap width at all values of Ca for which the interface is unstable.
tion is inhibited. In the fluid dynamical systems we have considered up to now the dendritic form of instability seems to only occur if some mechanical anisotropy is placed at the tip of the growing cell in the Hele-Shaw apparatus. Examples of such disturbances include, a bubble, a longitudinal wire, a groove in one of the plates of a Hele-Shaw cell or in one of the cylinders in Michalland's experiments, etc. An
The Effect of a Stabilising
=>•
Gradient on Interface Morphology
15
d).1§^il/U,
Figure 11. a, b) Examples of interface shapes in directional solidification, (Billia and Trevedi [28]). Note the similarity to previous examples, c) Comparison of interface shape with the McLean and Saffman [9] solution for an interface with surface tension, d) Axial competition between cells in directional solidification when the axial temperature gradient is not large enough to prevent it. Eventually, when the largest dendrites have emerged, the temperature gradient is large enough to prevent competition and the cells evolve uniformly, e) "Dendritic" growth of a gravity-driven liquid/air interface in a Hele-Shaw cell with an anisotropic tip condition, in this case a small tip-bubble.
example is shown in Fig. l i e for the case of a large bubble with a small tip-bubble in a Hele Shaw cell. The latter subject is covered extensively in Couder et al. [29]. Also, McCloud and Meyer [25] explore such matters in considerable detail and give numerous examples of these effects, while Billia and Trivedi [28] have discussed the qualitative similarity between Saffman-Taylor fingers and solidifying interfaces and Michalland [5] the mathematical analogy.
16
T.
Maxworthy
4
Combustion-Front Instability
The stability of pre-mixed, combustion fronts has been treated in the aerodynamic, thin-flame limit by Markstein [30]. The effect that gives rise to the appearance of quasi-steady, cellular flames is the difference in diffusivity between the fuel and oxidiser in a curved flame. In the theory, this is parameterised by allowing the local flame speed to be a function of flame curvature: U = U0(1+L/R),
(2)
where U is the local flame speed, UQ the speed of a flat flame, R the flame radius of curvature, and L a phenomenological parameter that describes the effect of the preferential diffusion on flame speed. One can, then, derive an effective surface tension a for a flame (Maxworthy [31]): o = 2PU*L(\-l),
(3)
where p is the unburned gas density and A is the density ratio across the flame. The presence of this effect acts to stabilise the flame, generates a cut-off wavelength and a wavelength at which the growth-rate is maximum. At finite amplitude, the shape saturates at a certain amplitude, a measure of which has been calculated by Markstein [?] for example. Experimentally it is possible to generate finite-amplitude, unstable flames on a slot-burner of the type shown in Fig. 12a. A photograph of a finite-amplitude flame is shown in Fig. 12b, where the similarity to the Printer's Instability front of Fig. 5b should be noted. Agreement with the calculated shapes is excellent. In some regions of the flow-rate/equivalence-ratio parameter space the flame becomes unsteady with cells appearing and disappearing as shown in the location vs. time photographs of Figs. 12c,d. The similarity to cell motion in the Printer's Instability, Fig. 6 for example, is again striking (Michalland [5]). 5
Viscous Gravity Current Instability
The instability of interest here is well known to anyone who has attempted to paint a vertical surface and found, on applying an excess of paint, that it forms rivulets that destroy the smooth appearance of the surface. The first attempt at a scientific explanation of the phenomenon was due to Huppert [32], as far as we are aware. A two-dimensional, viscous gravity current was formed on a sloping surface by releasing a known volume of viscous fluid from behind a dam. Initially, the current was two-dimensional, but it finally became unstable to a fingering instability. Huppert presented an expression that scaled the wavelength of the disturbance, L*, as follows:
gp sin a where A is the cross-sectional area of the current, a is the surface tension of the fluid, g is the acceleration of gravity, p is the density of the fluid, and a is the slope of the plate over which the current is flowing. Johnson et al. [33] have considered the case with a constant Row rate per unit width, Q (units of cm 2 /s). Their
The Effect of a Stabilising
Gradient on Interface Morphology
17
Figure 12. a) A slot burner used to study the stability of two-dimensional flames, b) T h e flamefront shape for a rich propane-air mixture (Maxworthy [31]). Note the similarity in shape to examples given in prior sections, c and d) Spatio-temporal photograph of a part of a cellular flame front (Markstein [30]). Note examples of cell coalescence, cell propagation, and oscillatory behaviour.
experiments gave L* = 13.9d/(3Ca)1^3, if Huppert's scaling arguments were used, where Ca — Qn/da, and d is the layer thickness equal to (3QfJ,/pg sin a ) 1 / 3 . The best fit to the data is actually given by L* = 9.2d/(3Ca)0A5, which suggests that a scaling on Ca~1^2 is probably closer to the true behaviour. A photograph of the instability by Johnson et al. [33], is shown in Fig. 13 using a fluorescence technique to determine the fluid depth. This method had been used previously by Goodwin [34] in the case of release of a constant volume of fluid. Troian, Herbolzheimer, Safran, and Joanny [35] have solved the linear-stability problem for this flow noting that the instability is due to the development of a raised nose at the front of the current which then becomes unstable under the ac-
18
T.
Maxworthy
•*|ll|l*«:
'^' " fs 'M «l Si S it *•
I
i s <;j Is »l 'i» S3 s i- •
Figure 13. Evolution of an initially two-dimensional liquid front for the case of steady, constant flow down an inclined glass plate (Johnson et al. [33]). Here, at initiation, the plate was wiped clean of fluid using a squeegee. The instability progressed from this state with a constant flow both into the fingers and into the fluid beyond the borders of the scraped region.
tion of gravity. Spaid and Homsy [36] showed that the growth is due to a kinematic instability in which, under a buoyancy/viscous balance, any small bulges that appear are thicker and thus heavier and less affected by viscous forces. These are
The Effect of a Stabilising Gradient on Interface Morphology 19
able to out-run regions that are thinner. As can be seen in Fig. 13 the shapes of the regions between the fingers are more pointed than the corresponding SaffmanTaylor fingers, even when surface tension forces are taken into account. This is almost certainly due to the rather different boundary conditions that are at play at the advancing contact-line, where the fluid/air/solid intersection takes place. In particular, the troughs appear, under most circumstances, to be locked in place by the contact-line dynamics and this accounts for the similarity in appearance between this flow and the others considered before. There is no evidence of lateral interaction between fingers in this application. A similar instability can be generated by placing an interface in a temperature gradient (Kataoka and Troian [37]) in which case upward motion is created by the differential of surface tension caused by the gradient (the Marangoni effect). The initially two-dimensional front becomes unstable to form fingers as in the gravitydriven case. Here, however, gravity is a stabilising effect and can maintain a twodimensional front under the appropriate circumstances (Kataoka and Troian [38]). 6
Conclusions and Discussion
We have presented a number of fluid systems that exhibit cellular structures having gross similarities. In all cases this is due to the stabilising effect of a gradient, in some physical property of the system, in the direction of cell propagation. Although we have tended, in some cases, to emphasise the steady-state shape of the cells formed in terms of Saffman-Taylor fingers with or without surface tension, they also have many interesting dynamical, spacio-temporally varying properties that have been studied using complex equations of various types. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16.
T. Maxworthy, J. Fluid Mech. 392, 27 (1999). T. Maxworthy, in preparation for J. Fluid Mech. (1999). T. Green, Sedimentology 34, 319 (1987). C. F. Chen, Deep Sea Res. 44, 1645 (1997). S. Michalland, Etudes des Differents Regimes Dynamique de I'Instabilite de I'Imprimeur, Ph.D. Thesis, Paris VI, France (1992). J. S. Turner, Buoyancy Effects in Fluids, Cambridge University Press, London, New York (1976). P. G. Saffman and G. I. Taylor, Proc. Roy. Soc. A 245, 312 (1958). E. Pitts, J. Fluid Mech. 97, 53 (1980). J. W. McLean and P. G. Saffman, J. Fluid Mech. 102, 455 (1981). F. O. Chandler and L. G. Redekopp, sub judice Euro. J. Mech. B/Fluids (1999). T. Maxworthy, Phys. Rev. A 39, 5863 (1989). J. R. A. Pearson, J. Fluid Mech. 7, 481 (1960). W. H. Banks and C. C. Mills, Proc. Roy. Soc. A 223, 414 (1954). J. A. Cole and C. J. Hughes, Proc. Inst. Mech. Eng. 170, 499 (1956). E. Pitts and J. Greiler, J. Fluid Mech. 11, 33 (1961). L. Floberg, Dissertation, Chalmers Tech. Univ., Gothenberg (1961).
20
T.
Maxworthy
17. G. I. Taylor, J. Fluid Mech. 16, 595 (1963). 18. M. Rabaud, S. Michalland, and Y. Couder, Phys. Rev. Lett. 64, 184 (1990). 19. M. Rabaud, Y. Couder, and S. Michalland, Euro. J. Mech. B/Fluids 10, 253 (1991). 20. H. Z. Cummins, L. Fourtune, and M. Rabaud, Phys. Rev. E 47, 1727 (1993). 21. S. Michalland and M. Rabaud, Physica D 61, 197 (1992). 22. S. Michalland, M. Rabaud, and Y. Couder, Europhys. Lett. 22, 17 (1993). 23. V. Hakim, M. Rabaud, H. Thome, and Y. Couder, Proc. NATO Adv. Workshop in New Trends in Nonlinear Dynamics and Pattern Forming Phenomena: the Geometry of Nonequilibrium, Cargese, France (1988). 24. A. M. McEwan and G. I. Taylor, J. Fluid Mech. 26, 1 (1966). 25. K. V. McCloud and J. V. Meyer, Physics Reports 260, 139 (1995). 26. T. Maxworthy, sub judice, Phys. Fluids (1999). 27. L. Paterson, Phys. Fluids 28, 26 (1985). 28. B. Billia and R. Trevedi, Handbook of Crystal Growth, Vol. l b , Fundamentals, Transport and Stability, D. T. Hurle, Ed., North Holland, Amsterdam, pp. 8991065 (1989). 29. Y. Couder, O. Cardoso, D. Dupuy, P. Tavernier, and W. Thorn, Europhys. Lett. 2, 437 (1986). 30. G. Markstein, Unsteady Flame Propagation, NATO Agardograph, Pergamon (1965). 31. T. Maxworthy, Flame Propagation in Tubes, Ph.D. Thesis, Harvard University, Cambridge, MA (1960). 32. H. H. Huppert, Nature 300, 427 (1982). 33. M. F. T. Johnson, R. A. Schluter, M. J. Miksis, and S. G. Bankoff, J. Fluid Mech. 394, 339 (1999). 34. R. Goodwin, Ph.D. Thesis, Department of Chemical Engineering, Stanford University, CA (1991). 35. S. M. Troian, E. Herbolzheimer, S. A. Safran, and J. F. Joanny, Europhys. Lett. 10, 25 (1989). 36. M. A. Spaid and G. M. Homsy, Phys. Fluids 8, 460 (1996). 37. D. E. Kataoka and S. M. Troian, J. Coll. Int. Sci. 192, 350 (1997). 38. D. E. Kataoka and S. M. Troian, J. Coll. Int. Sci. 203, 335 (1998). 39. M. Zhang, and T. Maxworthy, Meeting on Microgravity Materials Science, Huntsville, AL (1998).
S P R E A D I N G OF A LIQUID D R O P W I T H M A S S LOSS L. M. H O C K I N G Department of Mathematics, University College London, Gower Street, London WC1E 6BT, U.K. A thin liquid drop on a flat surface will spread under the action of gravity and capillarity. In this paper, the effect of mass loss on the size and lifetime of the drop is determined. The simple case of a uniform and constant rate of depletion is analysed in detail. An example of a physical mechanism that results in mass loss is when the surface supporting the drop is heated and the drop loses mass by evaporation. The dynamic behaviour of the contact angle is found, and it is shown that there is no simple relationship between contact angle and speed when mass loss is present.
1
Spreading of Thin Drops
When a liquid drop is placed on a solid surface, it will spread under the action of gravity and capillarity until it reaches an equilibrium size. The motion of the liquid forming the drop is resisted by viscosity, so the relevant parameters are the density and viscosity of the liquid, the surface tension between the liquid and the surrounding gas, gravity, the volume of liquid in the drop, and the static contact angle at the contact line between liquid, gas, and solid substrate. A further parameter, the slip length, has to be introduced if a slip model is postulated to remove the force singularity at the contact line. Most studies of this problem have been limited to thin drops, in which the height of the drop is much smaller than its radius, which requires the contact angle to be small also. In this limit, the fluid dynamics if simplified by the validity of lubrication theory, and an evolution equation can be derived for the height of the drop as a function of the horizontal co-ordinate and the time, with boundary conditions that serve to predict the radius of the drop as a function of time. In this model, the slip length is supposed to be small compared with the height of the drop. In the vicinity of the contact line rapid changes of slope are possible, and it is important to distinguish between the apparent contact angle, measured or deduced from the shape of the drop away from the vicinity of the contact line and the real contact angle measured at the contact line. It is assumed here that the real contact angle is always equal to the static angle, and any dynamic variation is confined to the apparent angle, and can be determined as a consequence of the spreading of the drop. The extension to this standard and much studied problem that is presented here is the inclusion of a process that diminishes the mass of liquid in the drop. Such a mass loss is produced by, for example, porosity in the supporting base of the drop, which allows liquid to be sucked into the permeable base. Another possibility is that the drop loses mass by evaporation into the surrounding gas. The main aim of the analysis in these cases is to predict the lifetime of the drop, that is, the time at which the whole mass of liquid in the drop is either imbibed by the base or is vaporized. Fig. 1 is a simple sketch of this model problem.
21
22
L. M. Hocking
\y
\k
\y
M/
Figure 1. Sketch of drop with mass lost into the base or the atmosphere.
After formulating the problem in non-dimensional terms, the solution for a constant rate of mass loss is found, which elucidates some general aspects of the motion. Then, the specific example of mass loss by evaporation is discussed. 2
Governing Equations
The height h'(r',t') of a thin axisymmetric drop, at time t' and at a distance r' from the centre of the drop, satisfies an evolution equation, that represents the conservation of mass in the drop. Suppose that the rate of mass loss per unit area is denoted by pq1 , where p is the density of the liquid. Then the equation for h' can be written in the form
dh/_ J ~dF
+
d_ u2,u
OM^ / 9 f
/ d2ti
dh' \
,,1
3/Li r'dr'
0, (1)
where p is the viscosity of the liquid, a is the surface tension, A' is a slip coefficient, and g is gravity. The precise way by which the edge singularity is relaxed by slip is not significant, but the form chosen here is appropriate for one of the particular cases examined in Sec. 4. If the drop radius at time t' is denoted by a'(t'), the boundary conditions for Eq. (1) are that dh'
d3 h' <9r'3 = 0, 9,
* ' - •
at r' = 0,
(2a)
at r' = a',
(2b)
%
where 8 is the contact angle. The volume of liquid in the drop is given by V " ( 0 = 2TT / ° r'ti(r',t')dr'. (3) Jo The evolution equation is only valid if the horizontal length scale is much larger than the vertical one. The effect of the capillary and hydrostatic terms is to redistribute the material forming the drop, while the mass loss controls the lifetime of the drop. Thus, the time scale can be chosen so that the first and third terms in Eq. (1) are of comparable size. Non-dimensional variables are defined by h! = hoh,
A' = hoX/3,
r' = a0r,
a1 = aoa,
Spreading of a Liquid Drop with Mass Loss 23
q' = Qq,
V = 2TT a20h0V.
t' = hot/Q,
The non-dimensional forms of Eqs. (1) and (3) can then be written as 2 dh 1 d , , d (d2
Wr{wt
V{t)=
.... dh
Bh
\
7Fr- )
+ 9 = 0,
r
// rh(r,t)dr. r h(r, t) dr. Jo Jo
(4a)
(4b)
where the Bond number B and the capillary number C are defined by B =
3
Pqap
c
=
fap^4 h0
M
The contact angle 9 must be small if the lubrication approximation is to be valid, and the height scale ho can be chosen so that ho = Oa^. The non-dimensional forms of the boundary conditions (2) are then given by d3h
dh 0
h = 0,
, atr
a ^ = 0'
a^ = '
dh — = —1 or
N
(5a)
= 0:
at r = a.
(5b)
The global mass balance can be found by integrating Eq. (4a) over the drop, and gives the condition that (dh
\
,
[a
dV
or[Tt+q)dr=-+Jorqdr
= 0.
(6)
The mass-depletion rate q will have different forms depending on the processes that are removing liquid from the drop. In general, it may be a function of r, t, and the local value of h. It may also depend on the global value of h and, in nonisothermal situations, on the temperature. In this case, the quantity q may provide a link between the mass and energy conservation equations. Before examining particular physical situations, it is instructive to consider first a very simple case, in which the local depletion rate is assumed to be a given constant. 3
C o n s t a n t Depletion R a t e
Suppose that the dimensional depletion rate q' = Q, & constant, so that q = 1. Suppose further that, initially, the drop is in static equilibrium, subject to capillarity and gravity, and that the initial radius of the drop is chosen as the length scale ao. Then, at t — 0, and for 0 < r < 1, h
~
fo(6)-fo(fcr) bh(b) •
h=^{l-r2),
I2(b) -2bh(by
v
V = i,
forB = 0,
(7a)
(7b)
24
L. M. Hocking
where the In are modified Bessel functions, B = 62, and the boundary conditions (5) are satisfied with a = 1. The main goals of the analysis are to determine the time t* at which the volume of the drop becomes equal to zero, and to establish any significant changes in the shape of the drop. The capillary term in Eq. (4) contains a factor 1/C, and the limiting cases when C is very large and very small can be examined. Suppose first that C » 1. Then, to leading order, the capillary term is negligible, and Eq. (4) becomes simply
(8,
£ + 1-0. so that, from the given initial conditions,
The radius a of the drop at time t, the apparent contact angle 9app, and the lifetime of the drop t* are determined by Io(ba) =
I0(b)-btI1{b),
8h] dr\r=a
J
app
=
h{ba) h(b) '
=
J0(ft) - 1 bh(b) • (10)
In the limiting case 6 = 0, h=-(l-r2)-t,
a = Vl~
2t,
eapp = a,
t* = i .
(11)
Thus, the apparent contact angle and the inward speed U — —da/dt of the contact line are related by the equation U = l/0app for all B. This solution is also valid for C of order one, since the capillary term in Eq. (4) vanishes when h is given by Eqs. (9) or (11). It must be modified near the contact line, where capillarity is active in a narrow region in which the slope adjusts from its outer value 8app to 1 at the edge. The most significant features of the drop as its mass decreases, however, are determined by the outer region. The analysis of the contact-line region is given later for intermediate values of C. At the opposite extreme, suppose that C « 1. Now capillarity acts on a short time scale C, whereas the change in drop size takes place on a time scale of order one. On the short time scale, the drop adjusts to a quasi-steady equilibrium shape appropriate for the current value of the drop volume. This quasi-steady shape and the drop volume, when the radius of the drop is a, are given by Iojba) - I0(br) _ a2I2(ba) l ] bh(ba) ' 26/! (6a)" To determine the time-dependence of a and V, it is not necessary to consider the perturbations to these expressions; instead, the global mass balance provides the required information. With q = 1, Eq. (6) gives the condition %
+
2
n
\*-0.
(13)
and, with V given in Eq. (12), dadV
1
[l 2 dV ,
Spreading of a Liquid Drop with Mass Loss
20
15
10
25
Figure 2. The lifetime of a drop as a function of the Bond number B = b2 for small and 0 ( 1 ) capillary numbers C.
After some manipulation of the Bessel functions, the lifetime of the drop can be expressed as
2 r = Mr + bh(b) bj0
h{x) Xh(x)
dx.
(15)
If 6 = 0, 9
2
h = az — r 2a
V=\a
a=
l--t,
t*
(16)
'•
In the limiting case of small C, the shape of the drop on the slow time scale is similar to its static shape, and there is no difference between the apparent and real contact angles. The lifetime of the drop as a function of b is shown in Fig. 3 for C small and for C of 0(1), calculated from the expressions for t* given in Eqs. (10) and (15). The asymptotic values of t* for large b are ^ ln6 +0.4325 t* ~ 2In 6 - 0.783
L5 b2
. _ 1 for C < 1, for C = 0(1).
.,_ , (17a) (17b)
In the spreading of a drop when there is no mass loss, three regions can be identified: the outer region, covering the bulk of the drop, a slip region near its edge,
26
L. M. Hocking
and an intermediate region linking these two regions together (see Hocking [1]). In that case, the time scale for spreading is proportional to ln(l/A), so that the time scales appropriate for spreading and for mass loss are comparable when Cln(l/A) is of 0(1). To demonstrate the significant features of the solution of this model problem, it is sufficient to treat the special case of zero Bond number. The initial conditions are given in Eq. (7). In the outer region, the height profile of the drop can be written as h = hc + Ch1}
hc = — (a2 - r2) ,
0app = —,
y
= - / ,
(18)
where the last equation represents the global mass balance (6). After one integration, the evolution equation (4) gives dr \ dr2
rdr J
a3 dt J
\
2a2
The solution of this equation subject to the boundary conditions can easily be determined. The quantity of interest is the limiting value of the slope of the drop as the edge is approached from the outer region, and this is given by dh 8V „ ~ 3 \-C dr a
3
a 3 V (,
SVda\
., .
.
, „
, as r
(20) The constant A can be determined from the global mass balance, since this implies that hi makes a zero contribution to the volume of the drop. In the inner region, the scaled variables H and x are defined by r = a — Xx and h = XH(x,t), and to leading order in A the evolution equation (4) gives, after one integration,
ff2(H + 1 ) 0 + c ( ^ +
x)=O.
(H)
The solution of this equation subject to the boundary conditions is easily found, and the slope of the drop away from the vicinity of the edge is given by 1 + Cf — + 1 1 (lnz + l)
asx^oo.
(22)
Since the slopes given by Eqs. (20) and (22) do not match, an intermediate layer is required. In terms of a new small parameter e, defined by e = l/ln(l/A), new variables X and s are defined by X = e \nx, H = xs(X, t), and to leading order in e, Eq. (21) becomes ds s3——: = c(-us uX
+ 1),
where c = C/e and da/dt = —u,
s ~ 1 + c(-u + 1)X
asX->0,
(23a) (23b)
which matches with Eq. (22) to leading order in e. Hence, s3
1 — us
ds = cX.
(24)
Spreading of a Liquid Drop with Mass Loss
27
0.8
07 0.6 -• 0.5 --
t* 0.4 -• 0.3 0.2 0.1
0 0
0.5
0.25
0.75
Figure 3. The lifetime of a drop at zero Bond number as a function of c = C l n ( l / A ) .
In terms of the new variables, the limiting slope (20) can be written to leading order as l — uS.„ 3~S+C—jp-(X-l),
_.
8V
S
(25)
Hence the equations that determine u — —da/dt and V are given by dV_ dt
1 - us
(26)
-2a
If c •—> 0, S —> 1, and u —* 4/3, which agrees with the small-C limit (16). If c —> oo, S —> 1/u, and the solution in this limit is given by V — a 4 /8 and a2 = 1 — 2i, which agrees with the C = 0(1) limit (11). In the general case, the integral in Eq. (26) can be evaluated explicitly. If the indefinite integral is denoted by F(s, u), and if a is chosen as the independent variable, the equations to be solved can be written in the form F(S,u) - F{l,u) = c,
dV _ a2 da 2u'
dt da
(27)
with the initial conditions t h a t V — 1/8, S = 1, t = 0, a = 1, and
F(S,u) = ~A i ( l - uS)3 - hi
-uS)2+
3(1 - uS) - ln(l
-uS)
(28)
The apparent contact angle 6apP = S = 8V/a3. A numerical solution for this problem was found and the lifetime t* of the drop as a function of c is shown in Fig. 3.
28 L. M. Hocking
4
Evaporation
An example of a mechanism by which a drop may lose mass is that of evaporation of the liquid forming the drop into the surrounding vapour. This is a physically complicated situation, and the starting point for the present discussion is the model formulated by Burelbach, Bankoff, and Davis [2], and the consequences for the spreading of a plane volatile drop elucidated by Anderson and Davis [3]. The model includes Marangoni effects and vapour recoil, but the results given in Ref. [3] showed that these two processes had only a relatively small effect on the spreading of the drop, the major effect coming from the mass loss by evaporation at the free surface of the drop. With this term only included, the evolution equation examined in Ref. [3], amended to its axisymmetric form and in the notation used here, has the form dh dt
1 d d_ /(Ph, {hs + Xh2) C rdr dr \ dr2
dh_ r dr
K
Bh
+K
+h
0,
(29)
where Q
C =
ph0K'
Oo
pahoK
(30)
\ho
In the terminology of Ref. [3], E is the evaporation parameter, and K the nonequilibrium parameter. The depletion rate is equal to one at the contact line and decreases towards the centre of the drop. When ] ( " » 1, it is approximately constant across the drop, but when K
0 only. In addition, the hydrostatic effect will be ignored, that is, B — 0. The solution for large and small C can be found as in Sec. 3. For C » 1, the capillary term is absent to leading order, and the solution satisfying the initial conditions is 1,~
^2
1
(K + hy
1 + 2K-
2^K2
K+\(l
r2)
+ 2Kt,
(31a)
-Kt,
t* = ]- +
^-.
(31b)
In Sec. 3, the large-C solution was valid for C — 0(1) since the capillary term vanished identically in the leading-order solution, and this extension of the range of validity for the solution cannot now be made. For C C 1, negligible evaporation takes place on the fast time scale during which the drop adjusts to a quasiequilibrium shape. The global mass balance then gives dV_
+
Kr dr = 0 K +h
with V •
1 2a
(32)
Spreading of a Liquid Drop with Mass Loss 29
Hence, 3 2 da 8a ~dl
-Kaln
:*+£).
da. ( 3 3 ) ]n(l + a/2K)~" The integral can be expressed in terms of exponential integrals, but must be evaluated numerically. The asymptotic values for large and small K are given by t
^ l
+
iK^K»1'
,0 8 Wo
t
*~16KHl/2K)^K«1-
M
The large K asymptotic value for t* is accurate to within less than one per cent for K > 1. Some calculated values of t* are shown in Table 1. It should be noted Table 1. The lifetime of a drop for C
t* 33.0 5.6 1.5 0.8
that the scaling of t' depends on K, and that the physical value of the lifetime of the drop is proportional to Kt*. In these limiting solutions for C either large or small, the contact-line regions are of secondary importance. For large C, the radius of the drop and its lifetime can be determined to leading order without reference to the contact angle. If desired, the solution near the contact line can be determined as in Sec. 3. For small C, the dynamic behaviour near the contact line and the spreading of the drop take place on a fast time scale, and on the slower scale, the drop has a quasi-static shape, with the contact angle taking its static value. For intermediate values of C, and if the initial drop radius is not close to its equilibrium value in the absence of evaporation, as in the results presented in Ref. [3], the distinction between the two time scales loses significance. It is then necessary to include the capillary terms in the analysis, as was done in the model problem of Sec. 3, but with the added restriction that C is small. The initial drop profile and radius are, as before, given by h = (1 — r) 2 /2, a = 1, V = 1/8, and the evolution equation is Eq. (29). The analysis proceeds on similar lines to that presented in Sec. 3 for intermediate values of C , with c = Cln(l/A) of order one. The only difference is that the constant mass depletion rate in Sec. 3 is now dependent on h through the evaporation term K/(K + h). When K S>> 1, the results of Sec. 3 still hold. It is also assumed that K 3> A. In the central region, the drop profile can be written as [compare to Eq. (18)], h = ho + Chi, dV dt
h0 =
^(a*-r%
fa Kr KaA , A AV dr = — ^ - l n 1 + Ka2 o K + h0 8V
SV
(35a) (35b)
30
L. M. Hocking
The solution of the equation for hi can be found as in Sec. 3, and it follows that dh
„{a3\3/
8V
8Vda\,
,
This is identical to the corresponding result for a constant depletion rate. The dependence of the mass loss on h enters the solution through the global mass balance equation in Eq. (35), but in Eq. (36) it is only the magnitude of the depletion rate at the edge of the drop that is required. For the edge region, the analysis is the same as in Sec. 3, since K 3> h near the edge. With r = a — Ax, and h = XH(x), the equation corresponding to Eq. (22) is dH , / da\, -r— ~ 1 + C 1 + — lnx ox \ at J •
as a; -»oo.
(37)
An intermediate region is required, and the analysis is identical with that presented in Sec. 3. Matching across the intermediate layer and the global mass balance provides the equations to determine a and V as functions of time. These equations are, from Eqs. (26) and (35), fS
s3
_,
dV
K a
4
/
W \
8V
n
With a as the independent variable, the equations to be solved can be written in the form 8V 3
\
_
,
dV
• « ) - ^ . « ) = «.
Ka4,
(,
4V \
Ht-^V+Ka*
dt
>
1
da = -u> <39>
with the initial conditions V = 1/8, t = 0, a = 1, and F(S, u) as given by Eq. (28). The apparent contact angle 6app = S — 8V/a3. In the limit as c —• 0, the value of u determined by the first equation in Eq. (39) is positive if 0 < S < 1 and negative if S > 1. Thus, the apparent contact angle is less than or greater than the static contact angle if the drop radius is decreasing or increasing. The values of u determined by Eq. (39) for c positive and a given value of S lie in the ranges defined by u > 1/S, 0 < u < 1/S, u < 0, x 4
when 0 < S < 1, when 1 < S < cc, when S > cc,
(40a) (40b) (40c)
where c0 = (l + 4c) / . The capillary spreading and the mass loss are in competition when the apparent contact angle exceeds the static value, and the drop radius increases only when c > cc. For smaller values of c the apparent contact angle may be greater than the static value when the drop radius is decreasing. These limitations on the range of possible values aid the numerical search for the value of u for a given S and c. Some calculated values of the lifetime of the drop are shown in Fig. 4.
Spreading of a Liquid Drop with Mass Loss
31
3-r Ar={).\
2.5--
1 v..
0
05
1
2
3
4
5
c Figure 4. The lifetime of a drop as a function of c for a range of values of K.
When C = 0(1), the full Eq. (29) must be solved in the central region with A = 0. There does not seem to be any analytical simplification in this case, and a numerical solution is required. The solution for this range of values for C is not attempted here. 5
Dynamic Contact Angle
A key assumption in this analysis is that the contact angle remains constant throughout the evaporation of the drop. In the slip and intermediate regions near the contact line the slope changes perceptibly, and an apparent contact angle can be denned for the central region. For an isothermal drop, this apparent contact angle 8 is related to the appropriate static value 9S by an equation of the form 03-63s=
U,
(41)
where U is the velocity of the contact line, so that the apparent contact angle is greater (less) than the static angle for expanding (contracting) drops. This equation and the constant of proportionality can be determined analytically for a specific slip model (see Hocking [1]), and is consistent with experimental observations. This relationship between contact angle and speed is sometimes referred to as Tanner's law, but that title is more correctly given to its consequence, that the radius of a spreading drop is proportional to the one-tenth power of the time. With the real contact angle assumed to be constant, the analysis in this paper allows the apparent contact angle to be deduced. It is then possible to investigate whether or not there is a simple extension of Eq. (41) to spreading with mass loss. The first equation in Eq. (39) relates the values of S, u, and c, so the apparent contact angle is given by the implicit equation F(S,u) - F(l,u)
= c,
(42)
with F(S, u) defined by Eq. (28), and some limitations on the possible roots of this equation are given by Eq. (40). Hence, the dynamic contact angle for a contact line
32
L. M. Hocking
Figure 5. The dynamic contact angle for contracting (u positive) and advancing (u negative) contact lines, for different scaled mass depletion rates. The asymptotic value Eq. (46) for c = 0.2 is also shown.
moving inwards with speed u satisfies the equation \u3(l
- S3) + \u2{\
- S2) + u(l - 5) + In
1-u 1-uS
cu
(43)
where 5 and c are positive. When c —> oo, which corresponds to large evaporation, u > 0, and the solution is given by S = l/u.
(44)
For large c, 5 reaches a maximum as u decreases, with 5 m a x = 0.73c 1/4
when u = 1.24c"1/4.
(45)
When c < < 1, the solution is given by 1 - S3 = 3cu,
(46)
which is the dynamic contact angle for spreading without mass loss Eq. (41). Note that the angle has been scaled relative to its static value. In the small-C solution of the evaporating drop given in Sec. 4, Eq. (46) holds on the short time scale during which the angle adjusts to the static value, and on the slow time scale, the apparent contact angle 5 = 1 . For intermediate values of C a numerical solution of Eq. (43) is required and the results are shown in Fig. 5, together with the small-c asymptotic value from Eq. (46) calculated for c = 0.2. For negative u (advancing contact line) the value of 5 for no mass loss holds when — cu S> 1, as expected. For positive u (retracting contact line), Eq. (46) is not applicable. It predicts that the slope will be zero for 3eu = 1, whereas when there is mass loss, 5 > 0 for all c. Also, when mass loss is present, there is a range of speeds for which the contact angle exceeds its static value and the contact line
Spreading of a Liquid Drop with Mass Loss
33
is retracting. With mass loss, all the solution curves pass through the point where u=l, 5 = 1 . For the constant depletion rate of Sec. 3, S < 1 throughout the lifetime of the drop and S —+ 0 as a —• 0. The same result holds for the evaporating drop of Sec. 5 when K is not too small. However, if K is sufficiently small, the mass loss occurs only near the edge of the drop, and it is then expected that the slope will increase, at least initially. Calculated values of the slope show that, when K exceeds about 0.6, the slope decreases monotonically. For smaller values of K the slope initially increases, reaching a maximum value of 1.98 when K = 0.2 and 2.27 when K = 0.1, before decreasing. The results given in Sec. 4 for the large-C limit show that, when K > 1, the slope decreases monotonically, but for smaller values of K, the slope reaches a maximum value given by Smax = -g ( — — J 6
when a = ( — - — J
.
(47)
Conclusion
The size and lifetime of a thin liquid drop that is losing mass has been discussed for a constant depletion rate and when the drop is evaporating. The dynamic behaviour of the apparent contact angle has been determined. Another example of mass loss occurs when the base supporting the drop is porous. This problem is discussed in Davis and Hocking [4,5] for planar drops, and their analysis has been extended to axisymmetric drops, but there is no space to discuss this example here. References 1. 2. 3. 4. 5.
L. M. Hocking, Quart. J. Mech. Appl. Math. 36, 55 (1983). J. P. Burelbach, S. G. Bankoff, and S. H. Davis, JFM 195, 463 (1988). D. M. Anderson and S. H. Davis, PF 7, 248 (1995). S. H. Davis and L. M. Hocking, PF 11, 48 (1999). S. H. Davis and L. M. Hocking, PF 12, 1646 (2000).
This page is intentionally left blank
VISCOUS GRAVITY CURRENTS WITH
SOLIDIFICATION
M. B U N K , P. E H R H A R D f , A N D U. M U L L E R Forschungszentrum Karlsruhe, Institut fur Kern- und Energietechnik P.O. Box 3640, D-76021 Karlsruhe, Germany E-mail: [email protected] The plane spreading of a viscous melt on a horizontal plate, driven by gravitational forces, for isothermal conditions represents already a problem with a free liquid/gas interface. This problem has been treated in the literature both theoretically and experimentally (cf., Huppert [1], Didden and Maxworthy [2]). If heat is removed at the plate or at the liquid/gas interface, we may have solidification of the melt in the proximity of these locations and, thus, further free interfaces (solid/liquid) will be present in the problem. This has relevance for several applied problems in engineering and geophysics. Model experiments with metallic (Pr < 1) and oxide (Pr 3> 1) melts have been conducted in a plane spreading geometry for constant pouring of the melt. The experiments feature (i) isothermal conditions and (ii) basal cooling. Thus, the experiments are suitable to examine the influence of a basal crust on the spreading flow. Further, an asymptotic model is presented, which for P r > 1 captures the spreading of an oxide melt in the presence of a thin crust. Hereby the liquid/gas interface and the solidified zone (solid/liquid interface) are predicted. A comparison of experimental and theoretical results completes the picture.
1
Introduction
Consider the flow of lava down a volcano dome or the flow of core melt in the basement of a nuclear reactor after a severe accident. Such flows are, in general, three-dimensional and are driven by gravitational forces across a solid substrate of given geometry (cf., Fig. 1). The interface between liquid melt and ambient air is free to adjust according to the pressure field. Thus, a time-dependent liquid/gas interface is present, which in turn is responsible for a time-dependent (moving) contact line (melt front). For a purely liquid melt, viscous forces resulting from shear within the spreading layer will balance the gravitational driving forces. Capillary and wetting effects are of minor importance as long as the spreading layer is of sufficiently large thickness. In both lava and core-melt cases, the temperature of the melt is above the solidification temperature and much higher than the ambient temperature. Therefore, heat losses are present on both the bottom and top sides of the spreading layer. On the bottom side, heat is removed by (unsteady) heat conduction into the solid substrate; on the top, heat is dominantly radiated to the ambient environment. The heat losses on both sides of the spreading layer eventually lead to temperatures below the solidification temperature of the melt. Thus, the buildup of solid crusts on both sides of the spreading layer is the consequence. This eventually adds further moving solid/liquid interfaces to the problem. The decrease of the melt temperature at both boundaries of the spreading layer, firstly, will lead to an increase of viscosity and consequently to large viscous forces. Secondly, the basal t Corresponding author. 35
36
M. Bunk, P. Ehrhard & U. Miiller
Figure 1. General problem of lava or core melt flow.
crust will be attached to the solid substrate and will, therefore, modify the shape of the solid substrate upon which the spreading proceeds. Thirdly, the interfacial crust will modify the boundary conditions at the top free interface. All these effects resulting from heat losses may substantially alter the progression of the spreading and may even lead to a complete solidification of the melt. A better understanding of these phenomena appears to be a worthwhile goal, since the prediction of the front progression or of the melt height is relevant from a safety point of view. In one case, the safety of villages, streets, or train tracks may be judged based on these predictions. In the other case, the distribution of the core melt in the reactor basement after solidification provides a foundation for assessing the removal of decay heat. The prediction of viscous gravity currents in the presence of solidification thus remains highly desirable. 2
Problem
We shall concentrate here on the aspect of basal solidification within a viscous gravity current. We further restrict our attention to a two-dimensional problem in (x,z). A sketch of the addressed problem is given in Fig. 2. A denned volumetric
Viscous Gravity Currents with Solidification
37
u0,To
s(x,t) a(t) x
Figure 2. Problem of spreading with basal solidification.
flux of melt is fed from the left into the spreading region at temperature To with characteristic velocity UQ and height ho- The spreading occurs on a horizontal plate of given temperature Tw < T0. The liquid/gas interface is denoted by h(x,t) and the moving contact line is at a(t). The upper surface of the crust formed due to the presence of the cold plate is given by s(x,t). If we denote the characteristic horizontal extent of the spreading layer by ^o > we shall expect after some time that e = HO/IQ
Experiments Experimental Setup
The experimental setup is sketched in Fig. 3. The plane spreading is arranged on a horizontal plate of 100 cm length and 20 cm width. A well-controlled volumetric flux of melt V enters the thermally-insulated reservoir on the left side of the spreading area. The volumetric flux V allows, via V oc UOHQ, directly for a determination of the Reynolds number. Once the reservoir is filled, spreading occurs across the plate to the right. The temperatures of both the melt To at various locations and the plate Tw at various positions are monitored using thermocouples. The spreading of the melt, particularly the progression of the moving contact line, is captured by a CCD camera positioned directly overhead. This allows, via an image-processing system, a fairly accurate measurement of the wetted area, which can be transformed into a y—averaged position a(t) of the contact line. Alternatively, a triangulation sensor can be positioned at any position xo above the spreading area. This allows the measurement of the liquid/gas interface h(xo,t) at the respective position, as it develops in time. Invoking the frozen-shape approximation, this information can be transformed into the liquid/gas interface h(x, to) at some averaged time to. This method is used to obtain profiles of the liquid/gas interface, particularly around
38 M. Bunk, P. Ehrhard & U. Miiller CCD c a m e r a : a(t)
M
/
/
\
\
triangiilation: h(t)
I thermocouples:! 1 ,,
* Twl
^w2
'w3
Figure 3. Experimental setup and measuring technique.
the moving contact line. The triangulation technique is restricted to diffusively reflecting interfaces. This means that only for Carnauba wax are these measurements possible. For the Woods-metal melt the interface has a mirror-like character, which inhibits the use of this technique. To obtain at least some information on the basal crust within the experiments, we have further arranged an array of needles with spacing Ax = 20 mm, which can be positioned along the x-axis above the spreading plate (cf., Fig. 4). At a chosen time to the needles are released and fall down through the melt. If a basal crust is present, the needles shall stop at the solid/liquid interface and from their final position above the plate an estimate of s(x,to) can be made. This method is surely not of high accuracy, however, a reasonable estimate is obtained in all cases. As a consequence of the poor accuracy, this method is applicable only for crusts thicker than 0.1 mm. Such crusts are present only for metallic melts (Woods-metal); therefore, measurements for the oxide melt (Carnauba-wax) cannot be obtained in this fashion.
3.2
Properties of the Model Melts
We have chosen two different model melts, representing both low and high Prandtl numbers. Moreover, the solidification temperatures of both melts lie in a convenient regime. The most important properties of both melts are summarized in Table 1.
Viscous Gravity Currents with Solidification
39
Figure 4. Measuring technique for the basal crust.
Table 1. Properties of the Model Melts.
3.3
property
Woods-metal MCP58 (75° C)
Carnauba-wax (90° C)
density p dynamic viscosity /j, specific heat c p heat conductivity A solidification temperature Ts Prandtl number Pr
9.2 • W kg m - a 37.3 • IQ~6 N s m~2 2.0- W2 W s k g " 1 K " 1 10.0 W m " 1 K " 1 58.°C 0.74
8.7 • 10^ kg m ~ a 48.0- 1 0 - a N s m.~'z 1.8- 10 3 W s k g " 1 K - 1 0.14 W m " 1 K " 1 80. - 84.°C 594.
Results for a Metallic Melt
It is obvious that, for the metallic melt with Pr = 0.74, the penetration of the lower-plate temperature into the melt is much deeper than occurs for the largePrandtl-number (oxide) melt. This can be inferred from results for a fiat plate under forced flow conditions as explained by Schlichting [3], for example. Therefore, we expect a relatively thick thermal boundary layer and correspondingly thick crust developing at the cooled plate. This will have a strong influence on the spreading. In Fig. 5, an actual photograph is provided, showing the spreading of the lowPrandtl-number melt (Woods-metal) on a cool plate of temperature Tw — 40°C and at a Reynolds number of Re ~ 40. We observe a fairly smooth contact line and a terraced structure of the liquid/gas interface. This structure results from a repeated flow of melt over several solidified basal-crust layers. The progression of the contact line can be discussed more systematically based on our complete results for Re ~ 40 and various plate temperatures, shown in Fig. 6. Here, the averaged position of the contact line is plotted as a function of time. A time offset between the curves is artificially introduced to clarify the picture. The two curves marked by circles correspond to the case with To = Tw = 75°C. As both initial temperature To and plate temperature Tw are above the solidification temperature of the melt (Ts = 58° C), we observe isothermal and purely liquid
40
M. Bunk, P. Ehrhard & U. Midler
Figure 5. Photograph of a metallic melt spreading on a cool plate.
spreading. Both curves are parallel and approximate straight lines, indicating a contact line moving at constant speed across the plate. A regression analysis of the experimental curves gives power laws, which are fairly close to the theoretically
Viscous Gravity Currents with Solidification
41
B u
o O (X
o
Figure 6. Position of the front as a function of time.
expected law a(t) oc i 4 / 5 . A comparison of both curves further demonstrates that the experiments are reproducible to a reasonable degree. Our results for a cool plate with Tw = 50° C and identical initial temperature T0 = 75° C and Reynolds number Re ~ 40 are given by the curves marked by squares. Here, a distinct difference to the isothermal results is obvious. Initially, the contact line moves at roughly the same speed as observed in the isothermal experiments. Then a pronounced retardation of the contact line, resulting from the buildup of a basal crust, is evident. A second wave of melt then spreads over the basal crust and re-accelerates the contact line to roughly the initial speed. The second wave of melt, in turn, likewise develops a basal crust and a complete stop of the contact line at x ~ 43 cm occurs, in conjunction with complete solidification of the melt. For an even lower plate temperature of Tw = 40°C our experimental curves are marked with triangles. In principle, the same phenomena observed for Tw = 50°C are seen in this case, although the lower plate temperature leads to an earlier stop of the progression of the contact line at a position X ~ 36 cm. We have studied the development of the basal crust using an array of needles. A characteristic set of measurements for Re = 4.4, T0 = 75°C, and Tw — 55°C is given in Fig. 7. For two instants of time the solid/liquid interface s(x, U) and some data of the solidified liquid/gas interface are shown. We recognize at t = 63 sec a basal crust of thickness up to smax ~ 0.4 mm. At a later stage (t = 90 sec) the thickness has grown to smax — 0.9 mm. Even though the solidified liquid/gas interface is not equivalent to the actual liquid/gas interface position h(x,U), these data confirm that the front of the basal crust runs behind the moving contact line in all cases. The profile of the basal crust is of wedge type, i.e., it rises slowly with increasing x and falls rapidly to the contact line. From the solidified liquid/gas
42
M. Bunk, P. Ehrhard & U. Midler
length x [mm] Figure 7. Measurements of basal crust thickness.
interface at t = 90 sec we can, moreover, confirm the terraced structure of this interface. 3.4
Results for an Oxide Melt
As discussed in Sec. 3.3, for large-Prandtl-number (oxide) melts we expect only a relatively thin thermal boundary layer to develop and, therefore, only a thin basal crust can form. The effect of this thin crust on the spreading will remain weak. Similarly, changes in viscosity due to changes in temperature within this thin thermal boundary layer will have only minor effects. In Fig. 8, we show an actual photograph of a (Carnauba wax) high-Prandtl-number melt for the parameters Re ~ 5, T0 = 90°C,7V = 30°C, and Pr = 594. We recognize a smooth contact line as well as a smooth liquid/gas interface. Both the overall picture and the contact-line progression are virtually indistinguishable from the isothermal case. This confirms our expectations with respect to the weak influence of solidification in this case. A more systematic analysis of the contact-line progression is possible from the graph in Fig. 9. The contact-line position for isothermal conditions with
Viscous Gravity Currents with Solidification
43
Figure 8. Photograph of an oxide melt spreading on a cool plate.
a plate temperature of Tw = 90°C is given by the curve marked with circles. The experiments with Tw = 60, 50, and 30°C plate temperature are marked by triangles, squares, and diamonds, respectively, and again an artificial offset on the time axis is introduced. The speed of the contact line is slightly slower when a basal crust is present. Again the weak effect of the basal crust on the contact-line speed is consistent with the physical reasoning for large-Prandtl-number melts. We finally give some examples from the measurements of the liquid/gas interface. In Fig. 10, profiles are shown for different Reynolds numbers Re and plate temperatures Tw- The Reynolds number characterizes (due to Re oc V) the intensity of the volumetric flux V into the spreading region. The liquid/gas interface profiles obtained for Tw = To = 90°C and Re = 0.56,1.2, and 5.2, in the upper graph in Fig. 10 reflect, through an increasing thickness of the melt layer, the increasing volumetric flux of melt entering the spreading region. In addition to the thickness, the contact angle experiences a distinct increase with increasing Reynolds number. This is linked to an increasing speed of the contact line with increasing Reynolds number. The large speed of the contact line, in turn, causes a larger contact angle. The lower graph in Fig. 10 shows the influence of the plate temperature Tw on the liquid/gas interface. We recognize that all experiments with the plate temperature Tw below the solidification temperature Ts give roughly identical profiles. In contrast, the isothermal experiment with Tw = TQ = 90°C shows a melt layer of half the thickness. Moreover, the contact angle in the experiments with Tw = 60,40, and 20°C appears to be larger — around 7r/2. In fact, supplementary
44
M. Bunk, P. Ehrhard & U. Muller
e o
a o TO
o a
o
time [sec] Figure 9. Position of the front as a function of time.
measurements by regular photography have shown that the contact angle is larger than 7r/2, SO that, qualitatively, the dashed line is the true profile near the contact line. The triangulation technique, being directed vertically downward, is, of course, not able to resolve this type of profile. The dramatic change of the contact angle occurs at roughly identical speeds of the contact line and is purely dependent on the plate temperature. In fact, the wetting characteristics in the cases Tw > Ts and Tw < Ts are different. For Tw > Ts the liquid melt proceeds on a thin liquid layer of melt, always present on the plate from former experiments. In contrast, for Tw < Ts the liquid melt proceeds on a solid thin melt layer. The second situation leads to a poorly wetting system with a larger contact angle, independent of the precise temperature of the plate. Thus, the change of the wetting characteristic (depending on whether the pre-wetted melt layer is solid or liquid) is responsible for the largely different thickness of the melt layer observed for these cases. Of course, the above phenomena are most pronounced for small Reynolds numbers (Re = 0.56 in Fig. 10). Increasing Reynolds numbers lead to a weaker influence of the wetting characteristics on the spreading. 4 4-1
Theoretical Model for Pr » 1 Formulation and Assumptions
We do not intend to describe the mathematical model developed for Pr 3> 1 in full detail; a complete presentation may be found in Bunk [4]. We summarize here the basic ideas, methods, and assumptions entering the model and apply the model to a set of parameters that corresponds to the experiments conducted with an oxide melt described in Sec. 3.4. Thus, a comparison of experimental observations and theoretical results is possible, validating the major part of the theoretical assumptions
Viscous Gravity Currents with Solidification 45
XI
'55
-40.0
-30.0
-20.0
-10.0
0.0
length x [ m m ]
XI
'55 XI
-30.0
-20.0
-10.0
length x [mm] Figure 10. Profiles of the liquid/gas interface.
and results. The model takes advantage of the separate length scales present in the problem. Based on the scaling
t0
liQ
(1)
46
M. Bunk, P. Ehrhard & U. Muller
W
[/ = _ ,
W
w 7 T - „ (u0ho/lo)
= -r^
UQ
(2)
P (fil0uo/hl)'
(3)
t (lo/u0)'
(4)
T-Tw
e = f—-^, J-o -
(5)
J-w
which involves the small parameter e = ho/lo, we obtain to leading order in e a nonisothermal, thin-layer approximation to the problem. Within this approximation the conservation equations are Ux+Wz 0 = -Px
= 0,
(6)
+ UZZ
(7)
eRe
0=-Pz-
(8)
~¥r
ePrRe{QT + U&x + W@z} - ®zz-
(9)
Physically, this means that gravitational forces cause a vertical pressure gradient that drives the flow. Dissipation is by viscous shear within the melt layer. Heat is transported passively by the flow field, i.e., there is no primary influence of the temperature field on the flow. The dimensionless groups within the conservation equations (6-9) are the Reynolds, Froude, and Prandtl numbers, defined by Re=^°,
(10)
Fr=^,
(11) hog
Pr = -.
(12)
The obvious boundary conditions, which correspond to the problem in Fig. 2 are, to leading order,
z = o:
e = o,
Z = S(X,T):
U = 0,
Z = H(X,T):UZ
X = 0:
= 0,P
W = 0,Q = 0,
=
Qs,
W = HxU
+ HT, GZ
= 0,
9 = 1.
Further, we have a constraint on the volume of melt V(r), namely,
[
'
Viscous Gravity Currents with Solidification
47
A(T)
V{T)=
J H(X,r)dX
= CVTa.
(14)
0
The problem defined by Eqs. (6-14) can be further simplified, provided we concentrate on the limit Pr —• oo. In this limit, the thermal boundary layer at the plate is very thin and a matched-asymptotic, quasi-steady solution for the thermal field can be constructed. Thus, the assumptions for the parameters are e 1, Re ~ 1, Fr ~ 1. Within this approximation the liberation of latent heat at the solid/liquid interface S(X,T) is negligible and capillary forces have a likewise negligible contribution for sufficiently-thick melt layers. We further ignore any changes of density upon solidification. We use an iterative scheme to obtain a solution to the problem through the following steps: • We solve for the liquid/gas interface H(X,T) and the flow field {U,W) of an isothermal spreading on the solid contour S(X,T). In the first iteration, we have SQ(X,T) = 0, i.e., we have isothermal spreading on a horizontal plate, which has been solved by Huppert [1]. For a more general solid contour S(X, r ) , we have to solve a modified evolution equation (with respect to Huppert [1]), namely
HT-\e-^{(H-S?Hx}x
= Q,
(15)
to obtain the liquid/gas interface H(X,T). For general parameters, the evolution equation (15) is solved numerically by the method of lines. Based on H(X,T), the flow field is readily obtained from analytical expressions (cf.,
Bunk [4]). • Given the flow field, we seek a solution to the temperature field. We employ here a quasi-steady solution and use matched-asymptotic solutions for the thermal boundary layer and the isothermal core of the flow. • The temperature field now allows for the identification of the isotherm G = Qs, which is assumed to coincide with the position of the solid/liquid interface Si(X,r). Of course, Si(X,r) is derived for both convective and conductive heat transport present in the entire melt layer. In truth, below Si(X,r) only conductive heat transport is present. Therefore, Si(X,r) will only be an approximation for the true solid/liquid interface at S(X,T). • In order to obtain a reasonably good approximation to S(X, r ) , we have to repeat the above steps until Si(X,r) converges to S(X,T). This typically occurs after 3-5 iteration loops.
48
M. Bunk, P. Ehrhard & U. Muller
AM
Figure 11. Propagation of the contact line and interfaces.
4-2
Results
In Fig. 11, we show typical results for the contact-line position A(T), the liquid/gas interface H(X,T) and the solid/liquid interface S(X,r) for varied 0 S , the scaled solidification temperature T s , i.e., Qs — (Ts - 7 V ) / ( T 0 — I V ) . 6 S enters the problem via the boundary conditions (13) and can be viewed as the fourth dimensionless parameter in the problem. The melt enters the spreading region at temperature 0o = 1 and, due to the thermal boundary condition (13) at the plate, the temperature of the melt lies in the range 0 < 0 < 1. Thus, if 0^ = 0, we have pure melt above the solidification temperature. Consequently, the curves obtained in Fig. 11 for 0 S = 0 relate to isothermal spreading without solidification. As 0 S increases, an increasing portion of the melt is below the solidification temperature and, thus, immobilized. Therefore, the curves for 0 S = 0.25,0.5, and 0.75 show increasing thickness of the basal crust below S(X, r) and increasing thickness of the total melt layer below H(X, r ) . This has an influence on the progression of the contact line A(T), yielding a reduced speed of the contact line for increasing Qs. Physically, the reduced speed of the contact line is consistent with an increasing basal-crust thickness, which reduces of the driving gravitational pressure head.
Viscous Gravity Currents with Solidification
49
Figure 12. Experimental and theoretical data on the contact-line history.
4-3
Comparison with Experiments
The results from the theoretical model can be validated against the experimental observations from the high-Prandtl-number melt, given in Sec. 3.4. In Fig. 12, we give an example for such a comparison. For Carnauba wax, we use an initial temperature of To = 90° C in all experiments and have a solidification temperature of Ts ~ 81°C. Based on the respective material properties, the dimensionless groups in the given experiment are e = 0.2, Re = 0.56, Pr = 594, Fr = 0.09, and Qs = 0.55. Our theoretical results for the contact-line position a{t) are given in Fig. 12 by two solid lines, valid for plate temperatures of Tw = 70 and 90°C. Our experimental data are given by symbols for identical plate temperatures. We note that the experimental isothermal data (TV = 90° C) lie somewhat below the theoretical curve. This slight discrepancy is likely caused by the finite width of the experimental spreading channel, which introduces additional shear at the lateral walls. This additional shear leads to a slower speed of the contact line. In other words, the experiment does not give a perfect realization of plane spreading. This effect should be present in all experiments. The experimental data for spreading with basal solidification (Tw = 70°C) are, in contrast, clearly above the theoretical curve. This is caused by the quasi-steady approximation within the model. Bunk [4] demonstrates that the quasi-steady temperature field overpredicts the basal crust thickness by up to 20%. Thus, the theoretical model tends to slightly overpredict the effect of solidification.
50
M. Bunk, P. Ehrhard & U. Miiller
5
Summary
We have conducted a set of experiments with both a metallic (Pr = 0.74) and an oxide (Pr = 594) melt spreading in a plane fashion on a horizontal plate of controlled temperature. With the plate temperature equal to the initial melt temperature, we obtain isothermal, pure-liquid spreading. For plate temperatures below the solidification temperature of the melt, non-isothermal spreading occurs in conjunction with the buildup of a basal crust. In all cases we deliver a volumetric flux of melt V into the spreading region which is constant in time. Aside from the plate temperature Tw and the Prandtl number Pr we vary the Reynolds number Re in the spreading layer via Re oc V in a range 0.56 < Re < 40. We further present a theoretical model for Pr S> 1 that allows the prediction of histories of both the interface shape and the contact-line position for oxide melts and arbitrary V(T). The model is validated against results obtained for (oxide) Carnauba-wax with constant volumetric feeding (V oc t). For the metallic melt, we find from the experiments a strong influence of solidification on spreading. We observe a pulsating progression of the contact line caused by a periodic buildup of a basal crust with subsequent overflow. This leads to a terraced structure of the liquid/gas interface that persists until complete solidification. Moreover, the contact line tends to develop fingering in cases of low plate temperatures. For the oxide melt, both the experiments and the theoretical model consistently demonstrate a weak effect of solidification on spreading. The progression of the contact line is slightly slower than in the case of isothermal spreading and the contact line remains quite straight. The theoretical predictions of contact-line history compare reasonably well with the experimental data. This validates major parts of the theoretical assumptions. An interesting feature is observed in the experiments with respect to the wetting characteristics: as the plate temperature falls below the solidification temperature of the melt a strong increase of the contact angle is observed for all contact-line speeds. The above experiments and theoretical model both concentrate on the aspect of a basal crust. Of course, an interfacial crust developing due to interfacial cooling is of equal importance. In fact, Bunk [4] has developed a theoretical model that allows prediction of the spreading of an oxide melt under similar conditions, but in the presence of an interfacial crust. Moreover, a set of experiments with identical model melts and interfacial cooling has already been completed at the Research Centre in Karlsruhe and will be published shortly. References 1. H. Huppert, J. Fluid Mech. 121, 43 (1982). 2. N. Didden and T. Maxworthy, J. Fluid Mech. 121, 27 (1982). 3. H. Schlichting and K. Gersten, Grenzschicht-Theorie, Springer-Verlag, Berlin (1997). 4. M. Bunk, Modellierung zur Ausbreitung von Schmelzen bei gleichzeitiger Erstarrung, Ph.D. Thesis, Univ. of Karlsruhe, FZKA-Report No. 6365 (1999).
COARSENING D Y N A M I C S OF ROLL WAVES H S U E H - C H I A C H A N G A N D E V G E N Y A. D E M E K H I N Department
of Chemical Engineering, University of Notre Notre Dame, IN 46556, U.S.A.
Dame,
Channel roll waves are observed to amplify, accelerate, coalesce, and coarsen downstream. It is shown here that such strongly nonlinear dynamics are driven by localized interfacial coherent structures in the form of equilibrium hydraulic jumps. Their localized structure, due to a balance among inertia, kinematics, and dissipation, allows us to normalize the amplitude of every jump along a channel into a single value via a power-law scaling. The same scaling applies for the differential speed between an equilibrium jump and a non-equilibrium one that results from coalescence. We are then able to produce a texture-invariant coarsening rate that is favorably compared to experimental and numerical results.
1
Introduction
Our community will be forever grateful to Steve Davis for exposing us to the myriad of fascinating interfacial dynamics that exist on thin films. One of us (HCC) entered the field after reading Steve's clear exposition of the Benney equation for falling films. Due precisely to the film's thinness, which Benney exploited in his longwave lubrication simplification, thin-film or shallow-water waves are typically dissipative and strongly nonlinear. As such, they are distinctly different from deepwater waves and their rich dynamics are beyond the classical inverse scattering and inviscid wave theories. Fortunately, their strongly nonlinear and dissipative nature allows a completely different coherent-structure approach that exploits the prevalence of robust and localized solitary-wave structures. Here, we demonstrate this new approach on roll waves that appear on inclined shallow-water channel flow. Along straight stretches of canals with constant inclination angles and water depths, roll waves often appear. The best examples are seen on cement urban canals like the Santa Anita wash in Los Angeles, but they also appear on natural rivers [1]. Such roll waves often intensify in amplitude and speed downstream and cause river bank erosion when they reach sufficiently large scales. Brock [1, 2] carried out the first channel-flow experiments to study this wave intensification dynamics and attributed them to irreversible coalescence events between roll waves. Such coalescence events reduce the number of roll waves and hence increase their separation. To produce the same flow rate as before, these liquid-carrying roll waves must then increase in amplitude and speed. Curiously, the coarsening dynamics that yield larger roll waves are scale invariant. Brock observed that the time-averaged wave period < t > at every station increases linearly downstream as < t > increases by an order of magnitude. The rate of change in < t > seems to be independent of < t >. Such scale-invariant coarsening was also observed in our simulation of thin-film capillary waves on a vertical plane (Chang et al. [3,4]) and seems to be a universal phenomenon for waves on falling liquid films. As elucidated in our earlier study [4], a fundamental reason for the coalescencedriven coarsening dynamics is the robustness and solitary structure of the waves.
51
52
H.-C. Chang & E. A.
Demekhin
Unlike solitons of integrable systems like KdV and nonlinear Schrodinger equations, such solitary waves do not pass through each other, but rather coalesce and form a larger solitary wave. This larger solitary wave has a higher speed than an average one in front and quickly captures it to perpetuate a coalescence cascade. That such a solitary wave is stable, even though the flat substrate around it is strictly unstable in this active, dispersive, and dissipative system, has been explained with a convective stability theory of localized wave structures [5]. Such stability and the lack of momentum and energy conservation as in integrable systems roughly explain why the coalesced waves remain as one. It is energetically more efficient for the coalesced waves to approach a stable equilibrium wave by slowly draining liquid out than splitting rapidly into two equilibrium waves. What has not been clarified, however, is why the coalescence cascade produces a constant rate of change in the wave texture. This we explain here for the roll waves. 2
Hydraulic Equations and Roll Waves
The reason the coarsening dynamics are easier to quantify for roll waves is that they are accurately described by a simple set of dissipative/active hyperbolic hydraulic equations derived by Dressier [6], du
du
W+^
fih
,
u2
+^ ^ - T '
g + J ^ ) = o,
,„ , (la)
(ib)
where a bulk friction factor formulation has been used to capture viscous dissipation. The lone parameter G = cosO/Fr2 is a modified Froude number with Fr = ghw cos 9/u2N = Cf/ tan#, where the subscript N denotes the Nusselt flat film, Cf the friction coefficient, 9 the inclination angle from the horizontal, and g is the acceleration of gravity. Roll waves are travelling wave solutions h(x — ct) and u(x — ct) to Eq. (1) which approach constant values at the two infinites, h —» \ a n d u —» y/x where x is the substrate thickness at infinity. This roll-wave family is hence a two-parameter family parameterized by G and %• Due to the symmetry of the hydraulic equations to an affine stretching in x, one can actually normalize x to unity by the transformation h = xH, x = x£, u = Uy/x, and c = Cy/x or £ = Tyfx- The parameter G remains invariant under this transformation. For a given channel, there is hence only one normalized "equilibrium" roll wave with unit substrate thickness. This family is depicted in Fig. 1 and its speed C and amplitude A have been derived explicitly by Dressier [6]. We have simulated Eq. (1) over a long channel with wide-banded disturbances at the inlet. To simulate how such small-amplitude noise evolves into roll waves, we added an artificial viscous dissipation term in Eq. (1) to allow length-scale selection during roll-wave inception. We find, however, that, once the roll waves are formed, their coarsening rate is independent of the dynamics in the inception region and hence independent of the artificial dissipation term. A snapshot of one simulation is shown in Fig. 2. The simulation is carried out for a sufficient duration
Coarsening Dynamics t '-
' • "I
T
— . — i - ™
•
£.2.5
1.7S
of Roll Waves
53
! -
G=0.1
/
-
0.12
; /
0.14
^^X
-
——"""^
j
0.16
/
^i^A . \
JM
^
i £ b
Figure 1. Wave profiles of the roll-wave family as a function of the modified Proude number G. Each profile has been shifted vertically as each should approach a unit substrate thickness.
to allow meaningful statistics at every station. The time-averaged wave period < t > is plotted as a function of x' = (^r-)x (Brock's coordinate) in Fig. 3. The inception region occurs between x' — 0 and x' = 2000 where wave compression actually occurs. Beyond x' = 2000, however, roll waves begin to form and coalesce. Their linear coarsening dynamics persist until increases by a factor of 4. The inception length x' = 2000 and the < t > at that position are achieved by adjusting the artificial dissipation term and the noise amplitude. It is the only fitting done and the downstream coarsening beyond inception is insensitive to either quantity. As seen in Fig. 3, excellent agreement with Brock's data is obtained. The world-
54
H.-C. Chang & E. A.
Demekhin
Figure 2. A snapshot of the simulated wave profile for G = 0.04.
Figure 3. The time-averaged wave period < t > as a function of position from simulations (curve) and Brock's experiments (circles and stars from two different experiments) for G = 0.04.
lines tracking the solitary roll waves beyond the inception region are shown in Fig. 4. Cascade coalescence events are clearly evident as one large excited roll wave seems to capture its front neighbors successively. 3
Similarity and Scale-Invariant Dynamics
As evident in Fig. 3, the downstream roll waves are clearly solitary although they are not identical: they tend to increase in amplitude, separation, and speed downstream. Nevertheless, they are locally stationary travelling waves as seen from the world lines of Fig. 4. Since G remains constant over the channel, this family of roll
Coarsening Dynamics
10
12
14
Ifl
of Roll Waves
55
IB
X Figure 4. World lines of the simulation for G = 0.04 in the x — t plane.
waves down the channel must be parameterized by \- Moreover, if we normalize by the local substrate thickness, all roll waves on a given channel must collapse into the corresponding member of the normalized equilibrium family in Fig. 1. It is difficult to estimate \ accurately since it does not change much as seen in Fig. 2. However, the time-averaged local speed c increases rapidly downstream and we use it to estimate \ = (c/C1)2 where C is obtained from Dressier [6]. When the timeaveraged amplitude at every location is scaled by the estimated x, a single value is obtained as seen in Fig. 5. This self-similarity downstream is valid for five channels (G values). Moreover, when the same normalization is carried out with Brock's measured time-averaged speed and amplitude, values at up to six very different locations along any given channel nearly collapse into one point, as seen in Fig. 5. All collapsed values are also close to Dressler's amplitude for the normalized roll wave. This self-similarity can also be exploited, in conjunction with the invariance of time-averaged flow rate to downstream position or Xi to obtain how the wave separation < / > and wave period < t > are parameterized by x downstream. However, the coarsening rate of interest is independent of the local wave texture and these two quantities are irrelevant in this respect. We examine the coalescence dynamics in the normalized coordinate with a unit substrate. For the coarsening to be scale invariant, the "excited" wave created by the previous coalescence event must have the same normalized amplitude at every station. This amplitude is not that of the equilibrium roll wave in Fig. 1, but it is nevertheless a universal constant after normalization. These excited waves are larger and faster than the equilibrium ones in Fig. 1, and are, in fact, the driving
56
H.-C. Chang & E. A. 10
-
Demekhin
1
,
A - Fr»3.45, sine - 0.05011 x\ 0 - Fr.4.63. Sin9 » 0.08429 8•
o\
- Fr.4.96, sine • 0.08429
a -Fr=5.06, sine-0.08429 $ -Fi*5.60, sine. 0.1192 •
*
*
!
\
\
x - calculation
y
D
y4«
.
-
•
i
0.05
0.1
L_
0.15
1—^~T~"-T-.
0.2
.
0.25
G
Figure 5. Collapse of all simulated and measured roll-wave amplitudes along five different channels.
wave in the coalescence cascades in Fig. 4. However, our statistics show that the distribution in wave separation, t or I, does not distinguish between excited and equilibrium waves at every station. We can hence use a dilating periodic lattice in a mean-field theory for both excited and equilibrium waves. The dilation occurs downstream because of the coarsening. However, the ratio of the average separation of equilibrium waves < t > to that of the excited waves < r > must remain the same, even though both increase downstream. Consider then, a train of waves spaced by < t > with every < r > / < t > being an excited one at every position. The excited one travels faster and will capture its front neighbor after the train has travelled a distance of lc. This increases the average wave period by 4 ^ - < t > and hence d_ dx
^ 4 -
(2)
< T > lc
Since there are very few excited waves as seen in Fig. 4, < t > / < T > « 1 and the average wave speed is that of the equilibrium wave c = C^/x- Consequently, < t > ~ c and d_ dx lc ^ ' We have argued that < £ > / < r > i s independent of position, it hence only remains to show < I > /lc is also constant along the channel to obtain a scale-invariant coarsening rate. This requires an estimate of lc, or equivalently, the differential speed between the excited and equilibrium waves.
Coarsening Dynamics
of Roll Waves
57
G=0.04
0.4
0.35
0.3
0.25
0.2
M„=1.1 A 0.1
0.05
1.5 AA
0.5
2.5
Figure 6. The decay of excited waves at G — 0.04 whose initial reduced amplitudes AA0 are indicated. The dashed line is the theoretical C(G)/A(G) for the correlation constant between A C and AA.
Simulations of excited waves have been carried out in the normalized unit substrate coordinate. We find that their instantaneous speed A C normalized with respect to the equilibrium value is always linear with respect to the corresponding normalized incremental amplitude A A This is shown in Fig. 6 as the decay of several waves with different degrees of excitation is depicted. The proportionality between the two can be estimated by modelling the excited wave as a transient shock with a unit substrate in front and a thicker substrate of thickness x behind. The wave profile is quasi-steady as it decays and is assumed to follow the same % scalings of the equilibrium solitary wave. As such, A C = (y/x — 1)C and AA = (x~ l)A and 7
=
AC
AA
=
C
A(y/x + lY
U
Since x varies from unity to 2, for an excited wave that is twice as tall as the equilibrium one, 7 is relatively constant. To be more precise, we choose its value as x — 2 to obtain 7 = 2 U4A(G) a n c ^ a ^ e r m s e r t m g Dressler's values of C(G) and A(G), we find it to be in quantitative agreement with the linear correlation in the simulations, as seen in Fig. 6. Neglecting the decay of an excited wave, the time required for the next coalescence event in the unit-substrate normalized coordinate is Tc =< L > /AC where < L > = < I > jx is the wave separation in the normalized coordinate. Hence,
58
H.-C. Chang & E. A.
Demekhin
Figure 7. The invariance of lc/ < I > and < t> / < T > along the channel for G = 0.04 from our simulations.
Lc = lc/X = CTC and lc
Lr
^ = U/x-i),
(5)
where x represents the ratio between the amplitude of the excited wave immediately after the previous coalescence event to the equilibrium one in the normalized coordinate. This number must be identical from station to station. A reasonable number is 2, corresponding to an excited wave that is twice as tall as the equilibrium one and < I > /lc = \/2 - 1 ~ 0.414 or lc/ < I > ~ 2.4. We have verified the invariance of lcj < I > and < t > / < T > numerically as seen in Fig. 7. The latter quantity is almost constant at every position while the former seems to vary from 3.0 to 3.5 along the channel, i.e., slightly higher than the predicted 2.4 and with a small variation. The simulations suggest that, while < t > / < T > is independent of \ or position, it is a weak function of G as seen in Fig. 8. When these results for < t > / < r > are inserted into Eq. (4) and if < I > /lc is taken to be 0.414 from the model or 1/3.5 from the simulation, the linear coarsening rates at different G are satisfactorily approximated in Fig. 8. 4
Summary
The scale-invariant coarsening dynamics of roll waves are simulated and modelled. Linear coarsening results because the excited wave has a differential speed AC that
Coarsening Dynamics O.OB
1
1
1
1
1
of Roll Waves
1
59
1
o 0
d < 1 > J>^_ ck \ 0.04
0 ^ - ^ < :
^*\.\
-
/ >/lc=
0.414 -
o
-
^ ^ — ^ ^ — _
< I >/lc
—
0
= 1/3.5 = 0.286
t
-
(> <>
-
<
0.1Q
_
( >/<
T >
_
)
i
-
(
i
< i
1
1
0.03
0.04
1 0.0S
0.06
1 0.0?
1
-
1
0.08
0.09
G Figure 8. The quantity < t > / < r > and the linear coarsening rate as functions of G from our simulations. The two theoretical curves correspond to Eq. (4) with < Z > /lc = 0.414 and 1/3.5.
scales linearly with respect to C in the normalized coordinates and because the density ratio of excited and equilibrium waves < t > / < r > remains constant down the channel. The existence of a normalized coordinate, on the other hand, arises from the self-similarity of the hydraulic equations and greatly simplifies the modelling effort. The only remaining question is how < £ > / < r > i s determined within the roll-wave inception region. The excited roll waves must be created differently from the equilibrium ones near the inlet. We suspect the modulation
60
H.-C. Chang & E. A.
Demekhin
instabilities of periodic wave trains give rise to these two distinct classes of roll waves. Acknowledgment This work is supported by an NSF grant and a NASA grant. References 1. 2. 3. 4.
R. R. Brock, J. of Hydraulics Div. 4, 1401 (1969). R. R. Brock, J. of Hydraulics Div. 12, 2565 (1970). H.-C. Chang, E. A. Demekhin, and E. N. Kalaidin, AIChE J. 42, 1553 (1996). H.-C. Chang, E. A. Demekhin, E. N. Kalaidin, and Y. Ye, Phys. Rev. E 54, 1467 (1996). 5. H.-C. Chang, E. A. Demekhin, and E. N. Kalaidin, SIAM J. App. Math. 58, 1246 (1998). 6. P. S. Dressier, Pure App. Math. 2, 149 (1949).
THERMOCAPILLARY CONTROL W I T H F E E D B A C K OF LARGE WAVELENGTH INTERFACIAL INSTABILITIES R. E. KELLY Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, Los Angeles, CA 90095-1597, U.S.A. E-mail: [email protected] Various methods, both passive and active, of controlling interfacial instabilities are first discussed briefly in order to provide a framework. Attention is then focused on active control with feedback when thermocapillarity is used to suppress instability. For the case of Benard-Marangoni convection, the results of Or et al. [1] for largewavelength disturbances involving surface deformation are compared to those of Bau [2] who used a different control law. Extension of the results is then made to the control of a viscous film flowing down a heated, inclined wall.
1
Introduction
It is with great pleasure that I make this contribution to a volume honoring Professor Stephen H. Davis on the occasion of his sixtieth birthday. I have enjoyed knowing Steve both as a friend and a professional colleague for approximately 36 of those 60 years, dating back to when we were both graduate students. Over that time, we have shared a common interest in the topic of fluid instabilities, especially in regard to interfacial instabilities and the instability of unsteady fluid states [3]. Steve is a good listener (at least if he finds the topic to be of interest), and he has made many helpful remarks concerning my own research. During the spring of 1985, I enjoyed visiting with Steve and his colleagues at Northwestern University while on sabbatical leave. We had each been working separately with students on the instability of a heated film flowing down an inclined plane and decided to combine our efforts in regard to thermocapillary effects. The resulting research was presented at the Eighth International Heat Transfer Conference and published in the proceedings [4]. I will later refer back to this paper and extend the results so as to include the effects of feedback control. The topic of flow control is an old one in fluid mechanics but, as discussed by Gad-el-Hak et al. [5], the use of feedback control to affect transitional and turbulent shear flows is of fairly recent origin and is currently a very active area of research. By feedback control, we mean that a sensor is used to detect and relay information about a disturbance to an actuator which then proceeds to influence the flow so as to affect the disturbance. Feedback control has also been used during the 1990's to influence the onset and development of thermal convection [6-11]. However, it has hardly yet been applied to the control of interfacial instabilities. For such instabilities, various methods of passive control have been proposed, e.g., use of steady magnetic [12,13] or electric [14] fields, imposing a mean flow [15], thermocapillary action achieved by steady cooling [16], and viscosity stratification, either in a continuous manner [17,18] or in discrete layers [19]. Active control achieved by imposing oscillatory shear in a prescribed manner has also been suggested [20-23].
61
62
R. E. Kelly
Active control of the Rayleigh-Plateau instability of a nominally isothermal liquid jet issuing into air has been demonstrated by Nahas and Panton [24] who used a laser beam as a controller. The beam was directed normal to the direction of the jet so that its energy was absorbed in part by the surface of the jet, thereby thermally influencing the surface tension. The beam was tuned to the same frequency as a controlled acoustic input disturbance, and so feedback was not involved. Without the laser beam, the acoustic disturbance diminished the breakup length of the jet. With the beam, the effect of the acoustic disturbance was cancelled, and breakup occurred as in natural conditions. A demonstration using feedforward control to damp out pre-existing hydrothermal waves in thermocapillary convection (occurring when a mean temperature gradient exists parallel to the surface) was reported in [25]. In this case, infrared sensing of a wave arising from instability occurred at one instant and was used to actuate a laser beam at a later instant so as to change the surface tension and diminish the disturbance. Whereas Nahas and Panton [24] concluded that they achieved their objective by controlling the magnitude of the surface tension, Benz et al. [25] realized their goal by inducing a thermocapillary flow that counteracted the flow associated with the hydrothermal wave. Both of these demonstrations [24,25] are examples of wave cancellation for which the basic stability characteristics of the fluid system are left unchanged, in contrast to the cases to be discussed later for which feedback control is used. The distinction becomes important when the governing stability parameter is well beyond the critical value without control because very large growth rates can then occur if feedback is not used. Nonetheless, the above two investigations can certainly be said to concern one type of thermocapillary control. Another approach is to use feedback control to alter the linear-stability characteristics of the system, assuming that the linear dynamics are understood. MarrLyon et al. [26] have been able to suppress experimentally the Rayleigh-Plateau instability of an isothermal liquid bridge by controlling the acoustic radiation pressure (time-averaged pressure) of an ultrasonic standing wave generated in the fluid surrounding the liquid bridge. The two fluids were chosen so that the cylindrical liquid bridge was neutrally buoyant. An optical sensor was used to detect deformation of the bridge, and its output was used to control on a linear basis the transducer driving the standing wave. The authors concluded that the effect of feedback was to increase the effective stiffness of the system, i.e., the linear-stability characteristics were changed. This general approach has also been investigated theoretically for Benard-Marangoni instability when the temperature gradient is normal to the interface by Bau [2] and by Or et al. [1]. Bau considered the effect of control on arbitrary-wavelength disturbances on a linear basis whereas Or et al. used a weakly nonlinear analysis to focus on the long-wavelength mode for which interfacial deflection is essential. This mode is the most critical one in a microgravity environment. The instability occurs on a subcritical basis which means that, without control, the value of the Marangoni number at the onset of convection depends on the level of the background disturbances. Hence, it is desirable in this problem to control also the finite-amplitude, subcritical instability in order to ensure that the linear critical value is meaningful. Or et al. [1] proposed the use of a cubic-order thermal control that would: (i) shift the critical value of the Marangoni number (Mac)
Thermocapillary
Control of Large Wavelength Interfacial
Instabilities
63
to higher values; (ii) eliminate the quadratic nonlinearity that gives rise to subcritical hexagonal convection; and (iii) construct a cubic-order nonlinearity such that a supercritical bifurcation occurs. Apart from the subcritical nature of the instability, control of the long-wavelength mode is desirable because, without control, the instability leads to rupture of the liquid layer [27]. Thus, the instability is catastrophic in nature because film rupture is to be avoided in most technical applications. The goal of the control then is to stabilize the long-wavelength mode associated with rupture while allowing, or even promoting, the hexagonal, finitewavelength mode so as to increase, say, heat transfer without rupturing the film. The reader is referred to [1] for details of the nonlinear analysis, but the results of the corresponding linear analysis are given in the next section. Demonstration of the control of unstable states exhibiting large-amplitude oscillations has been given by Petrov et al. [28], who controlled experimentally the intrinsic frequency of an unstable (but initially cylindrical) liquid bridge extending from a heated to a cooled boundary. Without control, steady convection occurred initially as the Marangoni number was increased, and then oscillatory convection set in, first with one frequency and then with two distinct frequencies. A single thermistor was used as a sensor and, on the opposite side of the bridge, a single thermoelectric device was used to heat the surface of the bridge. A model independent, nonlinear control algorithm was used to determine the behavior of the actuator. The goal of reducing the two-frequency state to a single-frequency state without changing the Marangoni number was achieved. In a later paper, Petrov et al. [29] used two sensors and actuators to stabilize the primary oscillatory mode, a result that is not possible with only one sensor and actuator. With the actuators used, the critical temperature difference for the onset of oscillations could be increased by about 8.5%. 2
Thermocapillary Control W i t h Feedback of Benard-Marangoni Convection
The basic concept of thermocapillary convection at large wavelengths is illustrated in Fig. 1(a). A fluid interface with surface tension that is assumed to decrease with temperature is located between a lower, heated wall and an upper, cooled wall. The horizontal dashed line indicates the undisturbed level of the interface from which the vertical coordinate (z) is measured. The overall depth of the layer or the value of gravity is assumed to be small enough that buoyancy can be ignored. The basic state is one of thermal conduction. Consider the interface to be displaced as shown in Fig. 1(a) with a wavelength much greater than the depth. For this situation, conduction also governs approximately the temperature in the disturbed state. Hence, the crest of the disturbance shown becomes cooler and has greater surface tension relative to the trough which is warmer. Say that the upper layer has negligible dynamical effect upon the lower layer. The net effect then, as is well-known [30], is to convect the lower-layer fluid from the trough to the crest, thereby augmenting the original disturbance and leading to instability if diffusive effects are relatively weak. Although interfacial deflection is not essential in general for thermocapillary convection to occur, it is essential for instability to occur at large wavelengths.
64
R. E. Kelly
cold wall
0(x,y,-l,t)=0 hot wall (a)
cooler
cooler
9(x,y,-l,0 = ec(x,y,t) cooler
cooler (b)
Figure 1. Thermocapillary effect on surface deflection: (a) without control; and (b) with control strong enough to reverse the sign of the disturbed temperature at the interface. The arrows show the direction on the induced flow.
Now assume that we have a means of controlling, say, the wall temperature so that its value can be increased or decreased locally. For the experimental control of Rayleigh-Benard convection, Tang and Bau [7] used individual heaters to establish both the mean and the spatially-varying control temperature. Increased heating below the crest by the control temperature and decreased heating below the trough will tend to weaken the driving mechanism for instability. If the flow induced by the thermal control is sufficiently strong, as shown in Fig. 1(b), the disturbed flow can be reversed, resulting in stabilization of the interface. This simple concept is referred to here as thermocapillary control, which we assume is applied with the aid of a sensor and feedback. It is important to realize that the control temperature only needs to be of the same order as the disturbance temperature for stabilization to occur. Thus, the control is energy efficient. The concept can be applied to many other situations, a simple one being to increase the damping rate of small-amplitude interfacial disturbances that can occur without the presence of mean heating.
Thermocapillary
Control of Large Wavelength Interfacial
Instabilities
65
In order to complete the control system, we need a sensor and a control law. Or et al. [1] assumed that the interface displacement (77*) can be measured locally; various methods of performing this measurement are given in [31]. On the other hand, Bau [2] assumed that the temperature of the interface Tj(z* = 77*) can be measured optically, as in the experiment by Benz et al. [25] who used an infrared camera for this purpose. Other choices could be made, such as the disturbance shear stress at the lower wall or even the velocity at the interface. The information is then fed to a thermal controller (which we assume to be located at the lower wall) via a control law. In order to control subcritical instability, Or et al. [1] argued that a nonlinear control law is necessary. However, we restrict the discussion here to shifting Mac and so a linear, proportional control is assumed,
T^x\y*,-d*,t*)-rv,
= Klr,%x*,y*,n-K;{TJ{z*=r,*)-ri{0))t
(l)
where an asterisk denotes a dimensional quantity, Tw is the constant reference temperature of the lower wall, TI(Q) is the interfacial temperature in the undisturbed state, and K* and K'2 are dimensional gains. If we define the nondimensional gains Kx = K^d*/AT* and K2 = K%AT*, where AT* = T*W- T,(0), we can write the control law as T*(r* 11* —d* t*\ —T*w w[ ,V
' "' AT
'
—
=K1r,~K2{TI(z
= r1)-TI(0)}.
(2)
Taking K2 = 0 corresponds to the case (I) studied by Or et al. [1] where K\ > 0 means heating under a crest (77 > 0); taking K1 = 0 corresponds to the case (II) investigated by Bau [2] where K2 > 0 again corresponds to heating under a crest where (Tj — Tj(0)) < 0. On a linear basis the term in curly brackets in Eq. (2) can be approximated by — dT T^z = 77) - T/(0) w 77—(0) + 0{x,y,O,t),
(3)
where the first term on the right of Eq. (3) represents the effect of interfacial displacement. Besides being used in Eq. (2), Eq. (3) is also used in the linearizedshear-stress interfacial condition. The nondimensional surface tension is taken as a(TI) = l-1(TI-TI(0)).
(4)
After using a diffusive scaling for the disturbance velocity, the interfacial shear stress condition becomes
where Ma is the Marangoni number (y*d*AT //io«o)Let Majv(fc) denote the neutral curve, where k is a nondimensional wavenumber. Without control, we assume that conditions are such that long-wavelength disturbances are the most critical, so MaN(fi) — Mac. For case I, Or et al. [1] found that with control, MaN(0) _ Mac{Kx = 0 )
1 \-Kx-
(g)
66
R. E. Kelly
which is greater than unity for 0 < K\ < 1, i.e., stabilization occurs in the limit k —* 0. As K\ —• 1, the ratio tends to infinity. This result occurs because, when K\ exceeds unity, the disturbance temperature at the interface Eq. (3) changes sign (see next section) and so, in accordance with the discussion of Fig. 1, stability should be expected as k —* 0. Of course, this means that Mac occurs at some finite wavelength kc. It remains to be seen whether or not kc can be increased enough so that rupture does not occur. For K\ > 1, Eq. (6) predicts that the layer will be stable on a linear basis if Ma > 0. For case II, Bau [2] found that Majv(0) < Mac(K2 = 0) for K2 > 0, implying that destabilization occurs. However, scrutiny of his analysis indicates that, although the full expression given in Eq. (3) was used in the shearstress condition, the part representing surface deflection was omitted in the control law given by Eq. (2). When the full expression is used in Eq. (2), the result for case II turns out to be, Majy(O) _ 1 + Bi + K2 Mac(K2 = 0)~ 1 + Bi '
. > '
[
where Bi is the Biot number, equal to a heat-transfer coefficient at the free surface multiplied by the depth of the liquid layer and divided by the thermal conductivity of the liquid; see, for instance, [30]. Thus, Majv(O) increases gradually as K2 increases from zero. For this case, the sign of the disturbance temperature on the interface cannot be reversed. If nothing else, this example indicates that it is worthwhile to consider on a physical basis various sensing and control strategies in order to design the control system. It might also be mentioned that, for large K2, Bau [2] found that oscillatory instability is possible for finite wavenumbers, just as Tang and Bau [6] found for controlled Rayleigh-Benard convection. For Rayleigh-Benard convection, this phenomenon seems to be due to destabilization of a higher mode by control, resulting in mode coalescence, as discussed by Or and Kelly [32]. As k —> 0, the higher modes here move off to infinity and presumably cannot contribute to oscillatory instability. Before going on, it is worth noting some of the assumptions and limitations of the above model. Information about interfacial quantities is assumed to be known continuously, and control is also assumed to be exerted continuously at the wall. In reality, thermal control at least can only be exerted over discrete regions on the wall. However, large-wavelength instabilities tend to vary gradually at first, and probably only a few actuators per wavelength would be necessary. VanHook et al. [27] reported that the "dry spot" resulting from the large wavelength instability has a scale of about lOOd*, which we take to be representative of a wavelength. In the experiment, d* « 0.02 cm, and so a wavelength is approximately equal to 2 cm. The heaters used in the experiment of Tang and Bau [7] each had a typical length scale of 6 mm, and so, in principle, several actuators could be used per wavelength. However, wall heating is probably not the most effective way of exerting thermal control at the interface anyway. It has already been mentioned that the use of a laser can allow for energy absorption in a thin surface layer, which seems ideal as a way of controlling surface tension. The other point worth mentioning is that Or et al. [1] proposed implementing control spatially so that a localized disturbance can be controlled, whereas Bau [2] assumed control of a given, though arbitrary,
Thermocapillary
air:
T
Control of Large Wavelength Interfacial
Instabilities
67
£nb
Figure 2. The basic configuration for a heated film flow.
wavenumber. Although Bau's approach is certainly suitable for small-aspect-ratio containers, more and more wavenumbers have to be controlled as the aspect ratio increases in size. By direct spatial control, one hopes that the dominant unstable wavenumbers characterizing the disturbance would be controlled automatically and that inaccuracies in the control would be associated with stable wavenumbers. 3
Film Flow Down a Heated, Inclined Plane
The flow is shown in Fig. 2. As in [4], only thermocapillarity and shear are considered to contribute to instability. Other mechanisms relevant to heated films are discussed by Oron et al. [33]. For the flow shown in Fig. 2, the surrounding air is assumed to have negligible dynamical effect. However, it does serve as a heat sink, and heat transfer at the surface is characterized by an empirical Biot number (Bi). Without control, the analysis and results are presented in [4], where a Fourier representation for the disturbance is used. In a small-wavenumber expansion (k « 1), the complex wave velocity is expanded as c = c0 + ka+O{k2).
(8)
Both inertial and thermocapillary effects enter at O(k), and so CQ = 2, as for an isothermal film if the surface velocity is used as a reference velocity in the convective scaling. The stability of the flow is determined at 0(k) by setting cu = 0, and the critical Reynolds number is determined as, (Rec)
Ma =0
= - cot /?,
(9)
where (3 is the angle of inclination. With Ma ^ 0, the following expression for the growth rate of two-dimensional disturbances was given in [4]: k cu
8fe2 15i?e
„
2
5
„„
15 /
Ma
Re2 - - cotpRe + — 4 16 v (1 + Bi)Pr
(10)
68
R. E. Kelly
This expression can be factored as 8k2
fc2ci
* = i l k {(Re - R e ^ ) (Re ~ Re<+))}'
(11)
where ife(±) = ^ C O t / ? i l ±
YIMa 5(1 + Bi)Pr cot 2 (3_
1/2^1
(12)
The root Re[ ' clearly tends to the isothermal result as Ma —• 0 and so is associated with the shear mode of interfacial instability for Re > Rec . For Ma > 0, Rec is reduced and so the shear mode is destabilized by thermocapillary effects for the case of a heated wall. If the wall is cooled, however, Rec increases, and so stabilization occurs. The other root Rec is positive only if Ma > 0 and corresponds to thermocapillary instability as influenced by shear. For Re < Rec , instability occurs only if Re < Re^ , i.e., shear tends to stabilize the thermocapillary mode. Goussis and Kelly [34] give similar results for arbitrary wavenumbers and three-dimensional disturbances. Now consider the two types of control used for thermocapillary convection. For case I control, the disturbed surface temperature can be shown to be proportional to (1 — Ki)/(1 + Bi) which changes sign as K\ exceeds unity. For case II control, the same quantity is proportional to (1 + Bi + i ^ ) - 1 - These results are the same as for thermocapillary convection without shear because the mean-flow effect upon the surface temperature is of higher order for disturbances with k —> 0. Hence, case I control offers an advantage here as well as for the control of pure thermocapillary convection. Because we have already discussed the effect of control on thermocapillary convection, we will concentrate on the effect of control on the shear mode of instability. Analysis shows that Rec has the same form as Eq. (12) with the (+) sign, except that Ma is replaced by Ma where Ma={l-Kl)Ma,
(13)
for case I control, whereas
Ma = (—L±J*L-) Ma,
(14)
for case II control. In order to stabilize the shear mode, we want Ma < Ma since Ma > 0 is destabilizing, and this is accomplished for both cases if K\ and Ki are positive. However, the consequences are quite different. For case I, Ma vanishes as K\ —y 1 and then becomes negative, meaning that the shear mode is stabilized even though Ma > 0. For case II, however, Ma vanishes only as K2 —> oo, i.e., one cannot improve upon the isothermal result with case II control. Large values of gain tend to amplify background noise which is undesirable, and so only K2 ~ 0(1) should be considered anyway. Finally, we note that control of nominally isothermal film flow can be evaluated by setting AT = 0 and instead using (K*d*) as a characteristic temperature for case I control. Thus, if we define a new Marangoni number as Ma = (K*d*/AT*)Ma,
(15)
Thermocapillary
Control of Large Wavelength Interfacial
Instabilities
69
one can show that the critical Reynolds number with case I control is 1/2'
YIMa Rec = - cot j3 { 1 + 1 + 5Pr(l + Bi) cot 2 j3
(16a)
or Rec = -ReCi0
15Ma {1 + 1 + 4 P r ( l + Bijite^ Q
-,1/2'
(16b)
where i?eC]o is the critical value without control, i.e., K\ — 0. Although (Rec/Recfl) > 1 for Ma > 0, appreciable stabilization is realizable only for reasonably low values of i?eCjo, i.e., for situations when the plane is close to being vertical. We note that Pr > 1 for most liquids, the exception being liquid metals. If Ma is very large, then other non-Boussinesq effects, e.g., variable viscosity, might also have to be included for an accurate prediction. Case II control is not relevant to an isothermal base flow. Although there are applications in materials processing when film flow instability is undesirable, it is desirable as far as heat transfer is concerned. However, film rupture is still undesirable. For heated film flow, rupture seems to be associated with the onset of three-dimensional waves [35,36]. Hence, control should be used to stabilize the finite-amplitude, two-dimensional waves that arise initially against three-dimensional disturbances. Although this is clearly a more ambitious goal than the ones discussed so far, it indicates that thermocapillary control using feedback can be applied to a wide variety of phenomena.
4
Conclusions
It has been shown that thermocapillary control with feedback consisting of a sensor and an actuator that can cause a spatial variation in surface tension can, in principle, increase the stability of fluid systems with an interface. Although the emphasis has been placed on the control of large-wavelength disturbances due to practical considerations associated with thin films, the concept is also relevant to finite-wavelength disturbances. However, there are many flows for which largewavelength disturbances are the most critical [37]. Besides the case of the falling film discussed here, analysis of the effect of thermocapillary control upon the instability discovered by Yih [38] for two-layer Couette and Poiseuille flows when different viscosities exist in each layer is being investigated; the results will be presented later. For such flows, the stabilization of long waves presumably stabilizes the flow as a whole, although a separate analysis is required to determine the degree of stabilization. It should, however, be emphasized that in order to implement this method of control, other questions need to be answered first. For instance, the number of individual actuators required to simulate effectively the spatially-continuous control assumed here needs to be estimated.
70
R. E. Kelly
Acknowledgments This paper has been prepared with support from the NASA Microgravity Fluid Physics Program (Grant No. NAG3-1819). The author is indebted to Dr. A. C. Or for his contributions to the earlier thermocapillary results as well as for helping in the preparation of this paper. References 1. A. C. Or, R. E. Kelly, L. Cortelezzi, and J. L. Speyer, J. Fluid Mech. 387, 321 (1999). 2. H. H. Bau, Int. J. Heat Mass Transfer 42, 1327 (1999). 3. S. H. Davis, Ann. Rev. Fluid Mech. 8, 57 (1976). 4. R. E. Kelly, S. H. Davis, and D. A. Goussis, Heat Transfer 1986, (Proc. Eighth Int. Heat Transfer Conference) 4, 1937 (1986). 5. M. Gad-el-Hak, A. Pollard, and J. P. Bonnet (eds.), Flow Control, SpringerVerlag, Berlin (1998). 6. J. Tang and H. H. Bau, Proc. Roy. Soc. London A 447, 587 (1994). 7. J. Tang and H. H. Bau, J. Fluid Mech. 363, 153 (1998). 8. J. Tang and H. H. Bau, Phys. Fluids 10, 1597 (1998). 9. L. E. Howie, Int. J. Heat Mass Transfer 40, 817 (1997). 10. L. E. Howie, Phys. Fluids 9, 1861 (1997). 11. L. E. Howie, Phys. Fluids 9, 3111 (1997). 12. J. A. Nicolas, Phys. Fluids A 4, 2573 (1992). 13. T. E. Morthland and J. S. Walker, J. Fluid Mech. 382, 87 (1999). 14. S. Sankaran and D. A. Seville, Phys. Fluids A 5, 1081 (1993). 15. B. J. Lowry and P. H. Steen, J. Fluid Mech. 330, 189 (1997). 16. S. P. Lin, Lett. Heat Mass Transfer 2, 361 (1975). 17. D. A. Goussis and R. E. Kelly, Phys. Fluids 28, 3207 (1985). 18. D. A. Goussis and R. E. Kelly, Phys. Fluids 30, 974 (1987). 19. D. D. Joseph and Y. Y. Renardy, Fundamentals of Two-Fluid Dynamics, Springer-Verlag, New York (1993). 20. S. P. Lin, L. N. Chen, and D. R. Woods, Phys. Fluids 8, 3247 (1996) 21. S. P. Lin and J. N. Chen, Phys. Fluids 10, 1787 (1998). 22. A. C. Or and R. E. Kelly, Int. J. Heat Mass Transfer 38, 2269 (1995). 23. A. C. Or and R. E. Kelly, J. Fluid Mech. 360, 21 (1998). 24. N. M. Nahas and R. L. Panton, J. Fluid Engr. (Trans. ASME) 112, 296 (1990). 25. S. Benz, P. Hintz, R. J. Riley, and G. P. Neitzel, J. Fluid Mech. 359, 165 (1998). 26. M. J. Marr-Lyon, D. V. Thiessen, and P. L. Marston, J. Fluid Mech. 351, 345 (1997). 27. S. J. VanHook, M. F. Schatz, J. B. Swift, W. D. McCormick, and H. L. Swinney, J. Fluid Mech. 345, 45 (1997). 28. V. Petrov, M. F. Schatz, K. A. Muehlner, S. J. VanHook, W. D. McCormick, J. B. Swift, and H. L. Swinney, Phys. Rev. Lett. 77, 3779 (1996). 29. V. Petrov, A. Haaning, K. A. Muehlner, S. J. VanHook, and H. L. Swinney,
Thermocapillary Control of Large Wavelength Interfacial Instabilities
71
Phys. Rev. E 58, 427 (1998). 30. S. H. Davis, Ann. Rev. Fluid Mech. 19, 403 (1987). 31. S. V. Alekseenko, V. E. Nakoryako, and B. G. Pokusaev, Wave Flow of Liquid Films, Begell House, New York (1994). 32. A. C. Or and R. E. Kelly, J. Fluid Mech. 440, 27 (2001). 33. A. Oron, S. H. Davis, and S. G. Bankoff, Rev. Mod. Phys. 69, 931 (1997). 34. D. A. Goussis and R. E. Kelly, J. Fluid Mech. 223, 25 (1991). 35. S. W. Joo, S. H. Davis, and S. G. Bankoff, J. Fluid Mech. 230, 146 (1996). 36. B. Ramaswamy, S. Krishnamoorty, and S. W. Joo, J. Comp. Physics 131, 70 (1997). 37. B. S. Tilley, S. H. Davis, and S. G. Bankoff, Phys. Fluids 1, 3906 (1994). 38. C. S. Yih, J. Fluid Mech. 27, 337 (1967).
This page is intentionally left blank
P A T T E R N FORMATION IN THIN LIQUID FILMS
D. GALLEZ Service de Chimie Physique and Centre for Nonlinear Phenomena and Complex CP 231, Universite Libre de Bruxelles, Blvd. du 1050 Brussels, Belgium
Systems, Triomphe,
E. R A M O S D E SOUZA Centro Federal de Educacao Tecnologica da Rua Emidio Santos, s/n, Barbalho, 40300-010 Salvador, Bahia, Brazil
Bahia,
The problem of pattern formation in thin liquid films with neutral and with electrically charged, insoluble surfactants is addressed. A thin fluid film bounded by a wall is modeled by a set of N + 1 nonlinear evolution equations for the film thickness and for the concentrations of N species of surfactants on the free interface. A weakly nonlinear analysis for the case of a film with neutral surfactants predicts the appearance of morphological patterns with a homogeneous distribution of surfactants that is confirmed by numerical simulations. For the case of electrically charged, surface-active molecules, the system displays patterns with varying film thickness and an inhomogeneous distribution of surfactants. In both cases, morphological patterns are explained in terms of the competition between attractive and repulsive forces. The variable surfactant concentration for electrically charged, surface-active molecules is due to the competition between diffusion and ion migration. In this case, surface-tension gradients lead to the formation of roll cells driven by Marangoni convection.
1
Introduction
Thin liquid (macroscopic) films are ubiquitous entities in a variety of applications from chemical engineering to biology [1]. In chemical engineering, they occur in foam dynamics, wetting phenomena, and in the coalescence of emulsions [2,3]. In biology, they appear as membranes [4], as the lining of mammalian lungs, as tear films in the eye [5], and as separation contact regions in cell-cell or cell-substrate adhesion [6,7]. These films exhibit a variety of spatio-temporal instabilities leading either to rupture or to the formation of a steady-state pattern. Rupture means that instabilities evolve until the thickness of the film approaches zero at some point in space. In the formation of patterns, the film surface reaches a new stationary state with varying film thickness (morphological pattern) where the surfactant concentration can be nonhomogeneous. Until a few years ago most attention was concentrated in analyzing instabilities leading to film rupture. The idea that rupture is the most important phenomenon comes from the belief that the van der Waals attraction is the dominant mechanism for ultrathin layers (100 — 1000 A). However, steady patterns have been observed in dewetting experiments (nanodrops) [8], in cell-cell or cell-glass interactions (periodic contacts of the order of 1 /zm) [9-11], and in lipid vesicle-supported membrane interactions (blistering) [7]. This suggest that other forces than attractive ones may be acting to stabilize the film. In the last five years some efforts have
73
74
D. Gallez & E. R. de Souza
been made to model physical situations in which instabilities in thin liquid films lead to pattern formation [2,4,12-15]. These studies show that the formation of morphological steady patterns depends on the competition between attractive and repulsive forces. Whereas to attain steady states with nonhomogeneous concentration of surfactants other physicochemical effects such as the diffusion and the ionic migration of surfactant molecules may act. The mathematical description of the full problem of thin-film dynamics leads generally to strongly nonlinear equations that are difficult to solve analytically. Mathematical tools have been employed to render the problem tractable and at the same time still complex enough to retain the main features of the problem [1]. As instabilities generally occur as long-scale phenomena it is usual to employ a lubrication approach to obtain an asymptotic version of the problem. As a result the film dynamics is governed by a set of nonlinear partial differential evolution equations for the film thickness and the surfactant concentration, which are tractable both by analytical approaches [16,17] and by numerical techniques [12,15,17]. An analytical proof of the possibility of supercritical bifurcation leading to pattern formation in a thin liquid film with constant surface tension was given [16]. Because in thin liquid films the attractive interactions are dominated by intermolecular surface forces that rapidly increase with decreasing film thickness, the formation of morphological patterns depends on the competition between attractive and repulsive body forces. We also investigated the dynamics of a thin film with neutral insoluble surfactants [17]. The set of two governing equations for the film thickness and the surfactant concentration was analyzed both analytically and numerically. The film also attains morphological patterns but, despite transient inhomogeneities, at steady state the surfactant concentration is always homogeneous. The presence of neutral surfactants only alters the time in which the patterns are attained. Lastly the dynamics of a thin aqueous film with charged surfactant is modeled. In this case the film reached nontrivial steady states displaying both morphological patterns and nonhomogeneous concentration of surfactants. The model considers an aqueous film sandwiched by a free surface with charged diffusing surfactants and a substrate bearing an electric non-diffusing charge of opposite sign. It follows that the morphological patterns present higher surface-charge concentrations in the narrower regions of the film. The nonhomogeneous distribution of surfactants comes from the competition between diffusion and migration of the ionic species. As a consequence of surface-tension gradients, Marangoni convection leads to the emergence of roll cells in the steady-state solutions [25,30]. In this paper, we present the evolution equations for a thin liquid film on a fixed substrate with N species of surfactant molecules [18]. Then we will discuss, for some applications [16-18], the conditions in which steady patterns are formed in the case of thin films with one specie of surface-active molecules. The model accounts for a thin film sandwiched by a fixed substrate and a bounding phase. The later could be either an infinite gaseous phase (Sec. 3) or a thin membranous phase in contact with a infinite adjacent phase (Sec. 4). The relationship between the viscosities of the film and of the bounding phase gives rise to some interesting questions related to the role of Marangoni convection and the problem of roll-cell formation. These questions will also be addressed.
Pattern Formation
in Thin Liquid Films
75
z adjacent phase
h 0
substrate
X Figure 1. The model. An asymmetric thin liquid film of mean thickness ho is sandwiched between two interfaces: a fixed film-substrate interface, Sf, located at z = 0, and a mobile film-bounding phase interface, 5 m , located at z = h(x,t).
2
The Model and the Governing Equations
In many applications, such as wetting phenomena, foam dynamics, coalescence of emulsions, and the spreading of surfactants in the lungs, the liquid films can be modeled as a thin aqueous film bounded by a gas phase on a substrate. However, in others applications, such as cell-substrate and vesicle-substrate adhesion, a thin electrolyte layer on a substrate is bounded by a thin dielectric (membrane-like) layer, which is in contact with a far more extended liquid phase. Here, we will adopt, for these two classes of phenomena, a general model (Fig. 1) of an asymmetric thin liquid film between a fixed substrate and a thin bounding phase which is in contact with an infinite adjacent phase. In the first class of phenomena the bounding phase and the adjacent phase form together an infinite gaseous phase. In the second class, the bounding phase is a membraneous phase and the adjacent phase is an infinite liquid phase. We consider a thin aqueous layer of mean thickness ho, density p, and viscosity ~p, bounded by a fixed substrate and a bounding phase of density ~p' and viscosity Ji' (Fig. 1). The thin aqueous layer will be modeled as a thin liquid film sandwiched between the film-substrate and the film-bounding phase interfaces. Both interfaces are under bulk forces and the film asymmetry relies on the requirement that the velocity field on the substrate vanishes. Each interface has a thickness far smaller than those of the film and of the bounding phase and is considered as a twodimensional singular phase with intrinsic rheological properties [19], like surface tension and surface coverage. We neglect inhomogeneities in the y direction and the film thickness h(x,t) is a function of the lateral coordinate x and time t. The free surface separates the film from the bounding phase and is located at ~z = h(x,t), where I is the normal coordinate. It represents a film-bounding phase interface with N species of surfactant molecules. Each species moves laterally in the interface with a diffusion coefficient D™>, j = 1 , . . . , N. The fixed surface represents a filmsubstrate interface located at J = 0.
76
D. Gallez & E. R. de Souza
The two-dimensional motion inside the film is governed by the Navier-Stokes equations and the film is under body-forcing stresses. The external stresses generate a conservative body force with potential $ on the fluid as well as normal II and tangential r stresses on the interfaces [1]. The disjoining pressure in the film will be (x,t), the surface tension £ ( f (•?)) will vary, according to the local surfactant concentration, and produce a solutal Marangoni effect [20]. In addition, a conservation equation is adopted for each surfactant concentration, where T^(x,t) varies in space and in time due to convection and diffusion effects [21]. For neutral surfactants, purely chemical diffusion will be considered, whereas for electrically charged surfactants we must consider both chemical diffusion and the migration of ionic components. The surface radius of curvature will be considered to be so large that the rate of change of T^\x,t) due to dilatation of the surface is neglected. The equations of motion as well as the corresponding boundary conditions at the free surface are considerably simplified whenever the film deformation occurs at a length scale that far exceeds the film thickness [22]. The long-wavelength approximation [1], when applied to a thin film on a solid substrate with N species of insoluble surfactants, leads to a system of N + 1 coupled nonlinear evolution equations [18] that (in dimensionless form) read: h3
N
2
ir £ M « l f > - T) - y (p0hxx -
hT^ u J2 Mij)rij) - T ) - YTU) ( S o / l - - *)x
(la)
, (lb)
where subscripts t and x denote differentiation. The dimensionless quantities h = h/ho, x = x/ho, z = z/ho, t = Vt/hQ, and T^ = li^t^ are respectively, the dimensionless film thickness, lateral coordinate, normal coordinate, time, and surfactant concentration. Here, V = ~p/~p is the kinematic viscosity. The first term on the right hand side (RHS) of Eq. (lb), - V • J ( i ) = -(/ig/P)V • J ( i ) , accounts for the dimensionless diffusion flux at the free surface due to the j t,l -species. J^ is given by minus the gradient of the respective chemical potential and V- is the surface divergence operator. If ideal solutions are considered, the j t/l -diffusion term reads -V • jW = D&TW,
(2)
for the case of neutral surfactants, and
V - J W = b « ( r « + r(j)
Zlpx RT
(3)
Pattern Formation in Thin Liquid Films
77
for the case of charged surfactants. In Eqs. (2) and (3), D^ = D^/u is the dimensionless diffusion coefficient. In Eq. (3), the dimensionless quantities Z = (pu2hl)~1/2Z, i> = (pp2)-1'2^, and RT = (pv2h0)-lRT are, respectively, the electrolytic molar charge, the electric potential at the free surface, and the thermal energy (the gas constant R times the absolute temperature T). Although nonlinear terms might be involved in the variation of the surface tension with the local surfactant concentration, we will assume here that the dimensionless surface tension £ = (ho/p~V2)H satisfies a linear equation of state N
E = So-X] M ( J ' ) r 0 ) '
(4)
where M^> = —d'S/dT^' is the Marangoni number at equilibrium. The viscosity of the bounding phase, ~p!, is taken into account in the film evolution through pr = (1 — JL'jji)~l. If the bounding phase plus the adjacent phase constitute together a gas phase, then JL' JL, as it occurs, for example, when the film is a thin electrolyte phase and the bounding phase is a biological membrane, then pr < 0 and the term involving the Marangoni number in the evolution equations (1) has the opposite sign with respect to the case of a water-gas interface. In Eqs. (1), I>2)(<& — II) = (h0lJ>V2) is the dimensionless disjoining pressure. The jump of shear stress through the film-bounding phase interface due to surface curvature is accounted for by r = (h2)/pP2)f. The long-wavelength approximation is carried out by supposing that the wavelength of the film deformation A is far greater than the film thickness h. A small parameter rj = h/X
78
D. Gallez & E. R. de Souza
3
Thin Film W i t h Neutral Surfactants: Morphological Patterns
In this section, we consider a thin aqueous film bounded by a fixed substrate and a gas phase containing one species of neutral, insoluble, surface-active molecules in the water-gas interface [17]. Indeed the model is an extension of the previously considered model of a thin film on a substrate with constant surface tension [16]. Because the surfactants are neutral, in accordance with Eq. (2) the diffusive flux will only contain a term describing the diffusion of the surfactant molecules in a concentration gradient. It follows that the governing equations (1) now read [21,17]: h? ih = M—Tx
i f > = DTX
h3 r-(£ 0 h X x
0
(5a)
+ Mhvvx- yr(s 0 h r a -<^
(5b)
The film is under external apolar van der Waals forces and polar interactions due to hydration, relevant to describe dewetting experiments [12]. The potential function >, the dimensionless disjoining pressure, is given by: A
(
h—
(6)
where the first RHS term accounts for the apolar van der Waals forces and the second RHS term describes the polar interactions. The dimensionless parameter P = h0P/~pV2 measures the amplitude of the polar forces P and the Born repulsion is controlled by the equilibrium distance d = d/ho and the correlation length L = L/ho. The signs of A and P depend on the apolar and polar components of the spreading coefficients. The apolar van der Waals force will be attractive (for A > 0) or repulsive (for A < 0), whereas the polar interactions will be repulsive (for P > 0) or attractive (for P < 0). A weakly nonlinear analysis of the system of coupled, nonlinear equations (5) in the vicinity of the first critical point was performed in order to analyze the evolution of small disturbances from the reference state (h, T) = (1,1); see [17]. The Hamaker constant A was taken as the bifurcation parameter allowing the discussion of the stability of the nonuniform steady states in terms of the repulsive parameters. The analysis showed that nontrival steady solutions may exist and are given by (C(A - Ac))1'2
sin(fcia;) + 0{\A-
Ac\
(7)
where Ac is the value of the control parameter at the critical point, k\ is the wavenumber of the perturbation, and C with p =
2
Z0k
PCSoh)'1
76 3
% exp ( - i ^ j x (1 - 20L2 + f L(l - 4L)) + £Texp(-!^)
(8)
Pattern Formation
lOCh
79
stability diagram c>0
\ \
p
\ 0o
1
/ /
C<0
J
o ^---T"--i»-^-==irrT
o.i / 1
-100-
in Thin Liquid Films
I
/
/ C<0 /
1
C>0
if. 5 ^
/
/
/
^
c<0
Figure 2. Limits of the stable bifurcations. The solid lines represent the solution p = P/Tk\ from the equation C _ 1 = 0 as a function of the correlation length L, where C is defined by Eq. (8), with d = 0.2. The regions where C > 0 correspond to stable patterns while t h e regions where C < 0 correspond to unstable structures (rupture). The dashed line represents the points in which Ac = 0. Points (p, L) above (below) the dashed line correspond to Ac > 0 (Ac < 0).
Bifurcation in a film with neutral insoluble surfactants obeys the same criteria as those obtained by Erneux and Gallez [16] for a film with constant surface tension. A supercritical bifurcation (stable solution) for h does exist for A > Ac if C > 0 and a subcritical bifurcation (unstable solution) for h exists for A < Ac if C < 0. However, for the model of a thin liquid film with insoluble surfactants that we are considering here, Eq. (7) shows that while h bifurcates to a stable nonhomogeneous solution, the long-time steady-state solution for T is homogeneous. Thus, in this case, the system will only display morphological patterns. The bifurcation diagram is represented in Fig. 2 where nontrivial steady solutions are possible for values of film parameters p and L inside the two regions in which C > 0. Accordingly, morphological patterns can be formed in two cases: a) A > 0 (attractive apolar force) and P > 0 (repulsive polar force); and b) A < 0 (repulsive apolar force) and P < 0 (attractive polar force). The linear analysis [17] shows that for A < 0 (repulsive) and P > 0 (repulsive) the homogeneous stationary state (h, T) = (1,1) is always stable and neither pattern formation nor film rupture is possible. Taken together, the results of the linear and of the nonlinear analysis [17] show that for A > 0 (attractive) and P < 0 (attractive) only film rupture is possible. Numerical simulations for parameter values in the upper C > 0 region of the bifurcation diagram (Fig. 2), using periodic boundary conditions (PBC) and random initial conditions (RIC), confirm the analytical predictions [17]. Fig. 3 shows that, in accordance with Eq. (7), a small-amplitude disturbance on the reference state (h, T) = (1,1) leads to the formation of a morphological steady pattern. However, at the steady state, the concentration of neutral surfactants is homogeneous.
80
D. Gallez & E. R. de Souza
Figure 3. Formation of a stable pattern for (p « 76, L = 0.15), and P = 280, T = 30, M = 0.1, S c = 30, d = 0.2, £ = 18, A = 5.0 (A c « 4.56), PBC, and RIC. (a) min(/i) (solid line) and min(r) (dashed line), through (0,1), as function of time. A slow decrease of min(/i) is followed by a fast-slow evolution towards a new stationary state at t = tp. After a transient (downward) peak, min(r) evolves back to T(x,tp) = 1. (b) Successive plots of h(x,t) separated in time by an interval of At = 80. The initial disturbance (curve to) is amplified towards a new stationary (nonhomogeneous) state (curve tp). (c) Successive plots for T(x,t) at the same time interval show that the initial state T(x, 0) = 1 (curve to) is at first amplified when surfactants are more concentrated on the crests than on the film depressions. Then r ( x , t) evolves back to the homogeneous state at t = tp.
In addition, the numerical simulations show that slow transient patterns for the concentration of neutral, surface-active molecules are attained as the concentration approaches uniformity. The reader is addressed to Ref. [17] for many other interesting results of the nonlinear analysis and of the numerical simulations of the evolution equations (5). Amongst them, a wavelength-selection phenomenon is observed numerically. Starting from a random initial condition the film develops a transient two-wave pattern followed by the selection of a single-wave pattern whose amplitude increases up to a stationary state. The mechanism of pattern formation where transient stable modes with multiple humps coalesce into a one-hump pattern is analyzed elsewhere [15].
Pattern Formation
in Thin Liquid Films
81
These multiple-hump patterns have been named kinetically stable structures because the system dwells on them for an exceedingly long time. This mechanism is similar to the selection of wavelengths shown in this paper, especially when the value of the control parameter is above the second or the third critical point. During the time where the system is dwelling on a kinetically stable structure, other instability mechanisms can become operative, like hole growth or a fingering instability [2]. For the case of a film bounded by a gas phase considered in this section, the Marangoni term does not depend on the viscosity of the bounding phase and is always positive; see Eqs. (5). The numerical results also show that the smaller the Marangoni number M the greater the time required for pattern stabilization, that is, Marangoni convection retards fluid movement from the narrower to the broader regions of the film until pattern formation. Let us note that at steady state the surfactant concentration is homogeneous and, as a consequence, the morphological patterns have constant surface tension. 4
Thin Film W i t h Charged Surfactants: Morphological Patterns and Charge Separation
Now we consider a thin aqueous film between a bounding phase with insoluble charged surfactants on the water-bounding phase interface and a fixed substrate bearing an electrical charge of the opposite sign. Blistering of a giant lipid vesicle bearing negatively charged components on a supported membrane with unlike charges was observed by reflection interference contrast microscopy [7]. A thermodynamic analysis of this model suggests that an adhesion-induced demixing of the charged components comes from the failure of the Young-Dupre law [23]. Here, an electrohydrodynamical approach is adopted instead. In this case, the electric disjoining pressure has attractive and repulsive contributions [24]. Consequently, if the diffusing surfactants have surface concentration T(x,i) = a(x,t)/Z, where Z is the molar charge and a(x,t) is the surface-charge concentration, the surface tension E(r) will vary according to the local surfactant concentration due to a solutal Marangoni effect [20]. Because the diffusing surfactant molecules on the water-bounding phase interface are electrically charged there is an electro-chemical diffusion term in the evolution equation for the surfactant concentration [17] rather than the purely chemical one considered in Sec. 3. The diffusion term in Eq. (lb) is given by Eq. (3) and the film dynamics is governed by the following set of two evolution equations for the film thickness h(x,t) and for the surface-charge concentration at the mobile interface a(x,t): ht
flrM 2
g
n ox
0,3,
1,3
E
<*t
(9a)
r-fi nxxx + -n c,
D(ax + Zatpx) + firMhaax
0 ,
,
- ~—h2zahxxx Z
,
1,2,
+ -h2o4>x Z
,
(9b)
J x
where M = —dHjda is the (constant) reduced Marangoni number and Z = I / O
(pF 2 ) Z/(RT), with R being the gas constant and T the absolute temperature. The contribution of uncharged surfactant molecules to the surface tension are
82
D. Gallez & E. R. de Souza
already included in Eo since we are assuming that variations of S are essentially due to charged surfactants only. Note that here the bounding phase viscosity Jl' is taken into account on the film evolution through /u.r = (1 —Ji'/~p)~l. If the viscosity of the bounding phase is larger than that of the aqueous medium (JL1 > 71), the term involving the Marangoni number in the evolution equations (9) has the opposite sign with respect to the case of a water-gas interface considered in Sec. 3 in which the gas viscosity is neglected (JL' is given by
where e is the film-phase dielectric constant, K = hoK is the (dimensionless) inverse /—2
\1/2
of the Debye length, and as = (h0/jw2) as is the (dimensionless) charge density on the substrate. In the long-wavelength approximation, the order of magnitude of the tangential electric stress r is smaller than the orders of magnitude of the remaining terms in the tangential-stress balance, so r = 0. I / O
The electric potential is set to zero at infinity and ip = (jw2) the electric surface potential at the free surface, which reads [24] i(h
\
/ M
ip accounts for
2as + c(e«h + e~Kh)
Here, and ip were obtained at the low surface-potential approximation and hold for ijj < 25 mV at T = 25°C. Eqs. (9) were solved numerically as an initial value problem with spatially periodic boundary conditions (PBC). Random initial values for the thickness of the film were considered. A uniform initial value was assumed for the vesicle charge density, but random initial values lead essentially to the same results. The shape of the film surface as well as the vesicle charge-density profile obtained from the numerical integration as a function of x at various times t are represented in Figs. 4a and 4b. Fig. 4c shows the temporal behaviour of the minimal and maximal values of h(x, t) and a(x, t) as the film approaches a steady profile. Notice that for the charged, surface-active molecules considered here, the surfactant concentration reaches a steady nonhomogeneous pattern due to the presence of electrochemical diffusion. There is a clustering of the charged surfactants at the closest contact points, that is, at steady states the film thickness and concentration of charged surfactants are 180° out-of-phase. Numerical integration also shows that Eqs. (9) admit several steady solutions leading to the formation of kinetically stable structures. Unlike the case of a liquid film bounded by a gas phase considered in Sec. 3, the Marangoni terms in Eqs. (9) depend on the viscosity of the bounding phase Jl' through \xr. For the case of an aqueous film bounded by a membrane, fj,r is negative. In this case, the linear analysis shows that the greater the Marangoni number the faster the growth rate, that is,
Pattern Formation
in Thin Liquid Films
h(x,t)
83
(a)
0
20
40
80
100
0.025
a(x,t)
(b)
0.020 - - ~ ^ J ± ^ ^
0.015 100
h max (t)
/ " " 1.5 -
1.0
:
o m a x (t)/c 0
i=^r~x-czr o m i n (t)/a 0
0.5 -
0.0 -
hmin(t)
V^^^ I
1
,
,
1
I
1
I
1
1
1
.
I
1
2
1
,
I
i
•
•
•
•
3
tx10"8 Figure 4. (a) Shape of the film surface, h{x,t), and (b) the shape of the vesicle charge-density profile, a(x,t), obtained from numerical integration as a function of x at various times t, starting from a random initial condition for the film thickness (e = 0.15) and crrj = 0.02 at t = 0 (light gray), at intervals of At = 4 X 10 7 up to t — 4.8 x 10 8 (black). The bounded domain with P B C is taken as 0 < x < L, where L = 100. Integration was performed over a grid of 100 cells, (c) The temporal behaviour of the minimal and maximal values of h(x,t) and
84
D. Gallez & E. R. de Souza
100
Figure 5. Spatial profiles of the film thickness, surface-charge concentration, components of the fluid velocity at the interface, and of the diffusion flux at the interface. The profiles are plotted for the same film parameters as those of Fig. 4, but for a longer time t = 4.2 x 10 9 . (A) Profile of film thickness; (B) morphological pattern with more concentrated charged surfactants where the film is thinner; (C) positive (negative) values of the lateral component of the velocity at the interface u correspond to fluid movement to the right (left); (D)the fluid moves up (down) if the value of the vertical component of the velocity at the interface v is positive (negative); and (E) positive (negative) values of the surfactant diffusion flux correspond to movement of surfactant molecules to the right (left).
Marangoni convection accelerates the instabilities [18]. However, the competition between the diffusion and the ionic migration of surfactant molecules, not present in the model of Sec. 3, can reverse the profile of the surfactant concentration (compare Figs. 3 and 4). ft follows that the sense of the Marangoni convection also changes leading to the formation of roll cells, as deduced from Fig. 5. A detailed study of such roll cells has been performed [25,30]. 5
Discussion
The nonlinear dynamics of a thin liquid film on a substrate with TV species of insoluble surfactants is described by the set of N + 1 governing equations (1), derived from lubrication theory [1], It was shown that for free symmetric films, that is, a thin film with two mobile surfaces, the evolution equations for the film thickness and for the surfactant concentration are also coupled to an equation for
Pattern Formation
in Thin Liquid Films
85
the lateral component of the velocity at the interface [21]. The difference between the two cases of asymmetric and symmetric films relies on the symmetry conditions that must be imposed for the model of free symmetric films, which require a higherorder analysis in the derivation of the governing equations in the long-wavelength approximation. Thus, the dynamics of a free film with N species of insoluble surfactants is described by a set of N + 2 governing equations [18]. In many applications, such as wetting phenomena, foam dynamics, coalescence of emulsions, and the spreading of surfactants in the lungs, the liquid films are well modeled as a thin aqueous film bounded by a gas phase. However, in others applications of biological interest, such as cell-substrate and vesicle-substrate adhesion, our model accounts for a thin film bounded by a thin membraneous layer that is in contact with an infinite liquid phase. The general model (Fig. 1) of an asymmetric thin liquid film between a fixed substrate and a thin bounding phase that is in contact with an infinite adjacent phase is described by the set of governing equations (1). The equations for a film bounded by a gas phase are recovered for Jl ;§> "pi, so that nr = 1. Let us note that the derivation of the disjoining pressure must consider suitable boundary conditions in each case, that is, in the case of a film bounded by a gas phase or that of a film bounded by a thin membraneous phase. The corresponding general model for a free symmetric film with N species of insoluble surfactants is given elsewhere [18]. Pattern formation in thin liquid films with one species of surface-active molecules has been found both for neutral surfactants (Sec. 3) and for electrically charged surfactants (Sec. 4). Analytical proofs that the formation of morphological patterns requires a competition between attractive and repulsive intermolecular surface forces were already given [16,17]. However, in the case of Rayleigh-Taylor instabilities, pattern formation in thicker liquid films just subjected to attractive forces has also been found [26]. We point out that in the case of ultrathin liquid films the attractive intermolecular forces that are operative drastically increase as the film thickness h decreases [27]. The /i-dependence of the disjoining pressures for the following attractive intermolecular interactions are illustrative: h~3 for van der Waals forces [27], exp(-kh), where k is the respective decay length, for hydration [12,13], depletion molecular [11,28,29], and electrostatic forces [24]. In this case, because of the rapid decrease of the negative disjoining pressure, a positive disjoining pressure is needed for the system to attain a new steady state. On the other hand, for the case of thicker films that exhibit the Rayleigh-Taylor instability [26], the dominant attractive interaction is a gravitational force that decreases as the film thickness decreases [26], and the film surface tension suffices to balance the negative disjoining pressure. As concerns the concentration of surfactants, inhomogeneous concentrations were attained for one species of electrically charged surfactant molecules (Sec. 4), but it was not the case for one species of neutral surfactant molecules (Sec. 3). In the former case, patterns appeared from the competition between diffusion in a concentration gradient and ionic migration in an electric field. In the case of neutral surfactants, despite the transient inhomogeneous distribution of surfactants, the homogeneity of the reference state is recovered in the non-trivial, steady solutions. In both situations reported here just one species of surface-active molecules was
86
D. Gallez & E. R. de Souza
considered. The conditions in which patterns of surfactant concentration are formed for more than one species of surface-active material is an open question. Surface-tension-gradients lead to Marangoni convective flows [20] inside the film whose direction depends on the relationship between the viscosities of the film, 72, and of the bounding phase, 72'. If 71 > 72', Marangoni convection opposes the fluid motion (Sec. 4) and retards the appearance of the patterns. However, if 71 < 72' (Sec. 4) Marangoni convection acts in the same direction as the fluid motion and accelerates the pattern formation. More interestingly, in the case where patterns of surfactant concentration were formed Marangoni convection began in the same direction as the fluid motion and then reversed its direction leading to the formation of roll cells [18]. A detailed study of the velocity field inside the film is in progress [25] in which roll cells are derived from the numerical solution of the governing equations (9). In the investigation of pattern formation in thin liquid films, periodic structures have appeared throughout as kinetically stable structures (KSS) [15]. In the numerical simulations, a KSS first appears as an n-humped metastable structure that dwells for a given time and then decomposes into an, also metastable, (n — 1)humped structure, and so on, until a 1-humped structure is attained. Yet, the last 1-humped structure has appeared in both the numerical simulations and the nonlinear analysis as a stable structure. Compare, for example, the patterns for the film thickness and the surfactant concentration in Figs. 4a,b and 5a,b, which correspond to the same film parameters but different times. The patterns at t = 4.8 x 108 (Fig. 4a,b) are 2-humped, while at t = 4.2 x 109 (Fig. 4a,b) they are 1-humped. It is not clear whether the KSS reflect limitations of the model equations [15] or if they could correspond to experimental situations. For instance, in the blistering of a giant vesicle on a supported membrane [7], the fusion of smaller blisters into larger ones was observed. This experimental observation could be explained by the phenomenon of wavelength selection computed from a model based on thin liquid film theory [18].
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11.
A. Oron, S. H. Davis, and S. G. Bankoff, Rev. Mod. Phys. 69, 931 (1997). A. Sharma and G. Reiter, J. Coll. Int. Sci. 178, 243 (1996). G. Elender and E. Sackmann, J. Phys. II France 4, 455 (1994). E. Ramos de Souza, C. Anteneodo, N. M. Costa Pinto, and P. M. Bisch, J. Coll. Int. Sci. 187, 313 (1997). A. Sharma and E. Ruckenstein, J. Coll. Int. Sci. 106, 12 (1985). D. Gallez and W. T. Coakley, Heterog. Chem. Rev. 3, 443 (1996). J. Nardi, T. Feder, R. Bruinsma, and E. Sackmann, Europhys. Lett. 37, 371 (1997). G. Reiter, Phys. Rev. Lett. 68, 75 (1992). D. Gallez and T. Coakley, Progr. Biophys. Molec. Biol. 48, 155 (1986). N. E. Thomas, W. T. Coakley, and C. Winters, Colloids Surface B: Biointerf. 6, 139 (1995). W. T. Coakley, D. Gallez, E. Ramos de Souza, and H. Gauci, Biophys. J. 77,
Pattern Formation
12. 13. 14. 15. 16. 17. 18.
19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30.
in Thin Liquid Films
87
817 (1999). A. Sharma and A. Jammel, J. Coll. Int. Sci. 161, 190 (1993). A. T. Jameel and A. Sharma, J. Coll. Int. Sci. 164, 416 (1994). V. S. Mitlin, J. Coll. Int. Sci. 156, 491 (1993). V. S. Mitlin and N. V. Petviashvili, Phys. Lett. A 192, 323 (1994). T. Erneux and D. Gallez, Phys. Fluids 9, 1194 (1997). E. Ramos de Souza and D. Gallez, Phys. Fluids 10, 1804 (1998). E. Ramos de Souza, aEstudo da Adesao de Celulas e Vesiculas atraves da Teoria Eletro-Hidrodinamica" Ph.D. Thesis, Instituto de Biofisica Carlos Chagas Filho, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil (1999). R. Aris, Vectors, Tensors, and the Basic Equations of Fluid Mechanics," Dover Publications, Inc., New York (1962). C. V. Sternling and L. E. Scriven, AIChE J. 5, 514 (1959). A. De Wit, D. Gallez, and I. Christov, Phys. Fluids 6, 3256 (1994). M. B. Williams and S. H. Davis, J. Coll. Int. Sci. 90, 220 (1982). J. Nardi, T. Feder, R. Bruinsma, and E. Sackmann, Phys. Rev. E 58, 6340 (1998). V. A. Parsegian and D. Gingell, Biophys. J. 12, 1192 (1972). C. Anteneodo, E. Ramos de Souza, D. Gallez, and P. M. Bisch, Physica A 283, 243 (2000). S. G. Yiantsios and B. G. Higgins, Phys. Fluids A 1, 1484 (1989). J. N. Israelachvili, Intermolecular and Surface Forces, Academic Press (1989). E. Evans and D. Needham, Macromolecules 21, 1822 (1988). E. Evans, Macromolecules 22, 2227 (1989). E. Ramos de Souza, C. Anteneodo, D. Gallez, and P. M. Bisch, J. Coll. Int. Sci., in press.
This page is intentionally left blank
MOLECULAR ASPECTS OF CONTACT-LINE D Y N A M I C S JOEL KOPLIK Benjamin Levich Institute and Department of Physics, City College of the City University of New York, New York, NY 10031, U.S.A. E-mail: koplik@sci. ccny. cuny. edu JAYANTH R. BANAVAR Department of Physics, The Pennsylvania State University, University Park, PA 16802 E-mail: [email protected]. edu We summarize and elaborate upon molecular dynamics results bearing on the microscopic properties of the flow near a moving contact line, such as slip, rheology, fluctuations and hysteresis, and present some new work on contact-line motion in dewetting.
1
Introduction
The moving contact line appears when a liquid displaces vapor or a second immiscible liquid from a solid surface, and represents the continuum idealization of the finite but microscopic region where the three phases coexist. Although it would be computationally convenient and perhaps conceptually desirable to replace this region by a simple mathematical line where a contact-angle boundary condition is to be applied, experiment and theory both indicate that much more is involved [1]. It is observed that contact angles are not simply fixed by the equilibrium surface tensions involved, as Young's equation would predict in the static case, and not simply a function of the velocity of the interface and the material properties of the phases, but depend on the geometry and history of the entire flow [2]. Theoretically, a straightforward hydrodynamic analysis [3] leads to a singular solution of the Navier-Stokes equation, and while slip models provide a fix that is mathematically acceptable [4], they do not realistically incorporate the small-scale physics involved and do not account for the flow dependence of the contact angle. Since the continuum description runs into problems associated with a very small subdomain of the flow, it is natural to refine one's spatial resolution and pursue these questions at a level where the results are not biased by theoretical prejudice, and where the microscopic input is both well characterized and subject to explicit control. Systems of thousands or more interacting atoms can be studied numerically in this way using the method of molecular dynamics (MD) simulations [5]. It has been found that, in many circumstances, systems of nanometer size observed over nanosecond intervals display flow dynamics governed in quantitative detail by the Navier-Stokes equations [6,7]. It is therefore possible to perform a numerical experiment which resolves atomic scale motion while having time- and space-averaged behavior consistent with the usual macroscopic equations. There is of course a price associated with this trick: substantial statistical fluctuations are present in small systems and often it is impossible to obtain fully-resolved flow
89
90
J. Koplik & J. R. Banavar
fields, and stress tensors in particular. Nonetheless, until experimental techniques catch up, MD simulations provide the only detailed atomic-scale information on the contact-line region, and give quantitative results in favorable cases and invaluable insight in general. 2
Molecular Dynamics Simulations
The principles of MD simulations are discussed in detail in monographs such as [5], and the specific choices of geometries, interaction parameters, etc., useful in simulating a miniature experimental fluid cell are given in reviews such as [7] and in the papers cited below, but it is useful to summarize the essential features here. The principle of any MD computation is to place (on a computer) a certain number of particles in a volume with random velocities, giving on the one hand the flow geometry and density and temperature operating conditions, and on the other initial values and derivatives of the coordinates, which then evolve according to the classical equations of motion
a
b^a
In this expression, a, b refer to the constituent atoms or molecules, m is the mass and x is the position of these atoms or molecules, and the potential energy has been taken to be a sum of two-body potentials V. If one is concerned with Newtonian fluids, it often suffices to employ the simplest form that includes the generic properties of the interaction between two molecules — the Lennard- Jones potential: VLj(r)=4e
(;)"'-*©'
(2)
where a and e are length and energy scales, respectively, and i,j refer to different species of material. The two terms provide a sharp repulsion at short distances to avoid atomic overlap, and an attractive tail at larger separations to control the chemistry. In particular, the miscibility of various fluid and solid species can be adjusted through the coefficients c^. To have a computation time proportional to the number of atoms, the potential is cut off, typically at r = 2.5
= —^loS 1
(3) r'o2 between adjacent atoms along a linear chain [8,9]. If solid walls are present, rigidity can be maintained by attaching the wall atoms to fixed lattice sites. Furthermore, if the flow near the wall is of interest, it is appropriate to allow the atoms of the solid VFENEM
Molecular Aspects of Contact-Line
Dynamics
91
to have thermal vibrations, which can be achieved by "tethering" them to the fixed lattice sites with linear springs [10]. An additional advantage of such a wall is that it allows heat to enter and leave the fluid system — a relatively realistic form of thermostat. Indeed, in viscous flow the fluid heats up, quite noticeably in a small system, and some thermalization is needed to maintain a constant temperature. At high shear rates, or in the absence of a thermalizing wall, the temperature is fixed by artificial means such as rescaling the velocities by hand, or coupling the molecules individually to a heat bath. The atomic positions and velocities are updated by numerical integration of the system of Eq. (1), for example using predictor-corrector methods. Most of the computation time is spent computing the forces; if the sum of the two-body potentials is evaluated by brute force, the time scales quadratically in the number of molecules. However, one may take advantage of the short-range nature of the interaction to reduce the time to O(N), for example, by using a layered linked cell list. To interpret the results of the calculation, the continuum fields are required. For an inhomogeneous flow, one divides the region occupied by the molecules into finite-sized sampling bins, so that the local density, say, is defined as the number of molecules occupying a bin about the point divided by the bin volume, averaged over some time interval. Likewise, the mean molecular velocity provides the Eulerian velocity field, and the stress tensor is given by the Kirkwood-Irving [5] expression involving the atomic velocities, separations, and forces for atoms in a sampling bin. In the results below, we will mostly discuss flows that vary in two directions, and the sampling bins are parallelepipeds of roughly one-atom cross section whose length extends across the passive direction. Note that the combined spatial and temporal averaging in MD precisely reflects the finite resolution of laboratory measurements. In most cases (but see the discussion of matching below) one constructs a nanometer-scale version of an entire three-dimensional experiment, including solid or other boundaries, temperature control, and external forcing, if needed. The typical configurations studied here are the Wilhelmy-plate system, Couette or Poiseuille channel flows with a spanwise interface, and the free spreading of a drop or the withdrawal of a film placed on a planar substrate. It should be emphasized that once the interactions, solid boundaries, and operating conditions are specified one has no further control over the contact-line dynamics. The system then exhibits the physically-correct behavior of the microscopic "laboratory apparatus" in question. 3
A Summary of Advancing Contact-Line Results
A number of independent MD simulations of contact-line motion have appeared in the last decade, and here we wish to summarize their common results. Although this research has only answered a subset of the relevant questions so far, there are some firm conclusions available.
92
J. Koplik & J. R.
Banavar
Figure 1. Snapshot of a Wilhelmy-plate simulation: the plate is translating upwards at the velocity 0.03CT/T.
3.1
The contact angle is well-defined down to atomic scales, with microscopic fluctuations, and varies systematically with flow velocity [10,11],
We illustrate this point with results from a Wilhelmy-plate simulation [12] illustrated in Fig. 1. The figure shows a side view of a three-dimensional system consisting of a solid plate straddling the interface between two immiscible monatomic liquids, with force fields at top and bottom to confine the fluids, and periodic boundary conditions in both horizontal directions. If the plate is translated up or down, the liquids are dragged along and the interface shape is distorted by the resulting shear stress. The symbols are the positions of the atomic centers, with atoms of the two fluids indicated by dots and circles, respectively, and those of the solid by asterisks. The system is fully three-dimensional but observed from a large distance, so that apparently overlapping atoms are actually separated in the third direction. The lattice spacing in the solid is about 0.8cr, corresponding to a few Angstroms in physical units. In this simulation, the two liquids are chosen to have the same attraction to the solid, so the static contact angle is 90°. The fact that the liquid adjacent to the walls forms layers is characteristic of compact molecules near a plane [13], and layers need not appear for some other systems, such as tethered polymer brushes. At these scales atomic fluctuations are clearly discernible, so nothing is truly "static." For example, as seen in Fig. 2, when the plate is at rest the microscopic contact angle fluctuates very noticeably. The snapshots in the latter figure are taken at intervals of lOr, physically in the nanosecond range.
Molecular Aspects of Contact-Line
Dynamics
93
Figure 2. Snapshots of a "90°" contact line demonstrating statistical fluctuations in the microscopic contact angle for liquids at rest. The format of the figure is the same as in Fig. 1, with only the atomic positions in the lower liquid shown while those of the second liquid and the solid are not indicated.
Figure 3. The variation of the microscopic contact angle with velocity. Snapshots of configurations with (downward) plate velocities: (a) U = 0.1, (b) U = 0.15, and (c) U = 0.2a/r.
T h e microcopic contact angle does vary systematically with plate velocity, beyond the fluctuations just noted, as shown in Fig. 3. We should note, however, t h a t in these simulations the capillary number, defined as Ca = /j,U/~f, where U is the plate velocity and /j, and 7 are the liquid's viscosity and surface tension, respectively (which have been determined by separate MD simulations), is not so small. Here, Ca — 0 ( 0 . 1 - 1 . 0 ) , and the effect could be unobservable in many laboratory situations where Ca <SL 1.
94
J. Koplik & J. R.
Banavar
. i !
i I I , ! I
.• ! . 1 1 . I I
. ! I , , I
11
Figure 4. (a) The velocity, and (b) the shear stress near a moving contact line in the left half of a simulation similar to Fig. 1. The plate moves upwards at U = 0.05tr/r, which is the magnitude of the largest liquid velocity vector shown, and the maximum stress value is 0.48 m/crr 2 .
3.2
The no-slip boundary condition breaks down within a few atomic diameters of the contact line, accompanied by locally non-Newtonian behavior [10,11,14,15].
The original studies involved Poiseuille and Couette channel flows, but we again give a more recent example involving a Wilhelmy-plate simulation. The velocity field near the contact line is shown in Fig. 4a, where each arrow is the average of the atomic velocities with a resolution of about an atomic diameter. One sees that the no-slip condition holds along the solid plate except within a few atoms of the contact line. This breakdown of the no-slip condition is accompanied by a prominent but finite peak in the shear stress, shown in Fig. 4b in a contour plot. Physically, the liquids would "like to" satisfy no-slip, but could only do so at the expense of a divergent shear stress. Because the interactions arise from smooth potentials, the liquid is compliant, and effects a compromise with a small amount of slip and a large but finite shear stress. (Note that the term slip here refers to the average Eulerian velocity field only; the liquid atoms are not bound to the solid here, and in fact execute a biased random walk [11].) Furthermore, the peak in the shear stress is not matched by a corresponding enhancement in the strain-rate tensor and the liquid behaves locally as a non-Newtonian fluid, an observation first made for a contact line in a Couette geometry [16]. In fact, this slip or no-slip behavior is not purely a property of contact lines, and
Molecular Aspects of Contact-Line
Velocity (a h)
Dynamics
95
Vcloclly (a 1%)
Figure 5. Force vs. velocity for a Wilhelmy-plate simulation: (a) without surface roughness, and (b) with surface roughness.
analogous behavior occurs in other "corner" flows. For example, if one solid moves across another as in the sliding-plate problem in CFD [17] or in Taylor's scraper problem [18], there is a velocity discontinuity at the corner, which in combination with an assumed no-slip boundary condition leads to a shear-stress divergence of the same form (~ r - 1 / 2 ) as in the contact-line problem. An MD simulation of such a corner flow [15] gives localized slip and shear-stress peaks similar to those shown above. Because of the fixed geometry, one can improve the statistics and analyze the flow fields in more detail. In the corners, slip occurs for any boundary velocity, and the microscopic and Newtonian expressions for the shear stress always disagree. Away from the corners, there is a transition in behavior as the global Reynolds number of the flow increases: no-slip and Newtonian, followed by no-slip and non-Newtonian, and eventually slip and non-Newtonian. The non-Newtonian behavior occurs once the global Deborah number (defined as De = jr, where 7 is the strain rate) is 0(1), but the presence of slip depends on other aspects of the flow as well. In particular, slip is not well-described by a Navier boundary condition in this situation. 3.3
Surface heterogeneity leads to contact-angle hysteresis [12].
Hysteresis is often loosely defined as a non-constant contact angle, but here we mean history dependence rather than just variation with velocity. The canonical example is the range of static contact angles often observed in the lab for given materials, and it has been known experimentally for some time that this phenomenon arises from chemical and structural heterogeneity of the solid (assuming the liquid is uncontaminated). MD simulations have provided an unambiguous verification of this statement, as well as some new information. First, we show in Fig. 5a results from the simulation discussed above on the force on the Wilhelmy plate as a function of its velocity. In the lab, in a very low capillary number flow the force is just the sum of surface tension and gravity contributions, and is linearly proportional to the cosine of the contact angle, while having the advantage of less subjective measurement. In the simulations Ca = O(0.1), so
96
J. Koplik & J. R.
Banavar
Figure 6. The contact-line region near a rough plate that moves with Velocity: (a) U = 0.1, and (b) U = -0.05cr/r.
there is a substantial viscous drag contribution as well, making the angle harder to disentangle, but the force measurement is still statistically more precise than direct observation of the angle. To obtain the figure, the plate began at rest, was ramped up to the first velocity value and run in steady state to obtain the force, then ramped up again to find the force at the next velocity value, etc. Eventually, the contact line approaches the top of the box, and then the plate velocity was ramped down to a minimum value, and then back up to zero velocity, forming a velocity loop. We see that while the force is velocity-dependent, it is not history-dependent, meaning that its value is a function of the current velocity only and is not sensitive to the past history of the motion. However, if we modify the structure of the solid plate, by displacing the solid's atomic positions randomly in the horizontal direction from perfect lattice sites, the force-velocity curve in Fig. 5b shows an open hysteresis loop. This curve is analogous to that of magnetization vs. applied magnetic field in ferromagnets [19]. In the magnetic cases the hysteresis originates in the metastable magnetic domains produced at high field that do not readily relax back to their previous state when the field is lowered. In our case, the rough solid produces a rough interface, as seen in Fig. 6, whose irregularity increases with velocity and does not relax back during the time interval covered by the simulations. 3-4
Layered spreading films appear for some interactions and molecule shapes [20-22].
In Fig. 7, we show three side views of monatomic three-dimensional liquid drops spreading on a solid substrate for different values of the interaction strength appearing in the Lennard-Jones potential. At CLS = 1-0, we have partial wetting and a structureless liquid drop, which fluctuates in shape but does not spread. At CLS = 1-2, there is complete wetting but the liquid remains disordered, while at CLS — 1-4, the liquid forms well-defined monomolecular layers as it spreads completely. The layering arises because the strong attraction pulls the liquid atoms to
Molecular Aspects of Contact-Line
,
.
.
.•>
la)
•
'
»
•-.
•<
. . t I n e = 500 .
-; ->V
(b)
W
•
•
Dynamics
97
'-
>
.... ' . ' » W $ ,
W
t i m e = 500
. i';sW;s >-^Tf^r^mrn^^^^WTr,v^;'v-r.>;
(c)
t i m e = 340
Figure 7. A three-dimensional drop of Lennard-Jones liquid on a solid substrate with different values of c L g : (a) cLS = 1.0, (b) cLS = 1.2, and (c) c L S = 1.4.
the substrate, and because they are essentially spherical they can minimize their energy by packing into a hexagonal layer. If the liquid is made of chains, all of whose monomers are strongly attracted to the solid [21], the entire molecule is pulled to the surface and one again sees obvious layering, but now the disordered shape of the chain prevents the formation of in-plane hexagonal structure. A more complicated liquid, only some of whose atoms are strongly attracted to the solid, does not form obvious layers at all [23]. These examples illustrate the (obvious) point that the possibility of layering depends on both interaction and steric effects. An interesting but as yet unresolved question at the molecular level is that of precursor films. It is tempting to regard the bottom-most spreading liquid layers in Figs. 7b,c as precursor films, since they appears to advance ahead of the "bulk" drop, but the systems studied are too small to identify distinct spreading rates. Furthermore, at least in the latter case, the entire spreading process occurs in the form of layers, and corresponds experimentally to spreading in very specific
98
J. Koplik & J. R.
Banavar
systems [24], and may represent exceptional behavior. 3.5
Deviations from standard continuum behavior are confined to the immediate vicinity of the contact line.
In some theories of contact-line dynamics [25], one assumes that there are regions large enough to be resolved at the continuum level that have anomalous elastic or other properties. There is no evidence for any such effects in MD simulations. The region where slip occurs is only a few molecules thick, any layering effects are induced by the presence of a nearby solid and are not specific to the contact-line region, and molecules are observed to diffuse in and out of the contact-line region without any anomalous behavior. 4
Dewetting Processes
An interesting variant of the contact-line problem occurs in dewetting, the result of covering a solid surface with a liquid in an out-of-equilibrium configuration with a contact angle smaller that that given by Young's equation. For example, if a partially-wetting liquid is spread out mechanically to cover a solid, it may break up and eventually withdraw into discrete drops, or perhaps more intricate shapes if the liquid-solid interactions are suitably complicated [26]. Whether or not such an instability occurs depends on a competition between the solid-liquid interaction forces, of van der Waals type in the simplest cases, and restraining forces such as surface tension [29] or gravity [30]. Here, we focus on the hydrodynamic aspects of the flow of a withdrawing liquid, in the case where the initial coating film is thin enough for van der Waals forces to play a dominant role and for gravity to be negligible. Suppose such an unstable film has just opened a hole, that is to say a dry patch has just appeared on the solid. The liquid will try to lower its energy by pulling back to replace liquid-solid interface by vapor-solid, and the contact line at the edge of the hole must move. The continuum description of this problem has a number of difficulties: aside from the usual singularity and total ignorance of the appropriate dynamic contact angle, one must have a sufficiently fine resolution to resolve the sub-micron region where van der Waals forces enter. To date, there is only one serious theoretical attempt to consider the fluid mechanics of the process [27], based on the strong simplifying assumptions that the liquid forms a growing rim as it dewets, and that the rim is an arc of a circle and that particular angles appear at the edges of the rim. It is then possible to derive power laws for the speed and shape evolution of the rim. We have conducted MD simulations of dewetting [23], and find partial agreement with these predictions. The MD calculation begins with a solid substrate initially covered by a thin uniform liquid film. The liquid is made of short chains, four atoms per molecule, to reduce the vapor pressure compared to that of a pure Lennard- Jones system and to display a sharper interface. The molecular binding is accomplished via the FENE potential given in Eq. (3), although the stress is generally not large enough for non-Newtonian effects to enter. The geometry is a long narrow strip, to discourage
Molecular Aspects of Contact-Line
Dynamics
99
^mi^dx^M'M^-ivm^&sid ->J ^ * ^ ' - - . ^ -i*> ,'vfsi^yfc,;,
. ^i&*?J^*jt3t&M*.
m^%^i£^$&iT
, -'s-i^^j^jafrofiu:- - ^ ^ M S i f i
Figure 8. The evolution of a dewetting film a t times: (a) 100, (b) 1250, (c) 2500, (d) 3750, and (e) 5 0 0 0 T .
transverse instabilities of the interface as it withdraws, and periodic boundary conditions are applied in the plane of the substrate. Note that van der Waals forces, treated in continuum calculations as an external disjoining potential proportional to (film thickness) , are simply the summation of the Lennard-Jones interactions with solid atoms [28]. The liquid film is first equilibrated with parameters corresponding to complete wetting and then the solid-liquid interaction coefficient CLS is reduced at the edge of the solid while some liquid is removed there "by hand" to kick off the dewetting motion. (The liquid will spontaneously rupture if the film is thin enough, but this takes time, and in any case represents a separate calculation [23].) A side view of the film is shown in Fig. 8, just after this step, and subsequently as the liquid withdraws. We see that the dewetted liquid indeed collects in a rim that grows in size as it moves, but only the outer edge is circular, while the inner boundary is more ramp-like. The solid-liquid-vapor contact angle is seen to fluctuate with time, as usual at this scale, but its mean value of roughly 90° is distinctly less than the static equilibrium value of 105° ± 5°, obtained in a separate simulation with identical physical parameters where a cylindrical drop was allowed to freely come
100
J. Koplik & J. R.
200f
Banavar
'
^ ^ .
" ^ ^ ^ ^
150^ ^ ^ ^
100-
50-
\ \
^ ^
°0
1000
2000
3000 t
4000
5000
0
1000
2000
3000
4000
""5000
t Figure 9. The time dependence of: (a) the location of the edge, and (b) the height of the rim, for the dewetting film shown in Fig. 8.
to equilibrium on a solid. The position of the edge of the film as a function of time is shown in Fig. 9a, and as predicted [27] its speed is roughly constant. (The glitch at time 1000T is related to the first appearance of an obvious rim.) The height and width of the rim were predicted to vary as %/i, but we observe a much slower growth: the rim is subject to avalanching and grows irregularly at a rate closer to i 1 / 4 , while the width of the rim stabilizes. Although these results refer to a single, rather large-scale simulation with 72,000 fluid atoms on a 547
Molecular Aspects of Contact-Line
»
^
^
4,
«
4
, . , • : >
(
,
i
r
« -i
i
*• ,
f
» t . V , - , « .
<
i
> J
< *
t
,
V
<
, « . . » * • * - * -
T
,
\
» *
*» \
*
*
*
A
i
A
v *
*
^
*• T- »
^
T >
.
.
.
.
,
,
,
,
<
.
»
,
.
,
»
t
x
i
t
h 1
>
>
i
1 r
<
<
<
T T
*
T
i
»
»
r
-
»
>
% < »
<
*
> '
H
> *
r
<
« *
*.
•
r
1
t-
h
>
> t
T
:•
< t
,
:•
r
•
>
I
•
<
f
I
i
f
V >
' 1
'
T
»
t
•
i
j
* ,
* .
» »
•
•
,
,
v
•
.
,
*
, ! * > > .
,
»
.
'•
*
1
>
^
'
j
> •
i
>
X ' »
»
< * J
>. >
/
T
I- r
r
,
h
,
101
V--.
'*- >- ' • * ' ,
*
X ",
>• . , , ' - »
^ - * ^ ^ l l > l •
t
V >
Dynamics
"xj
V .. , .
,,
• i v > • ^ T •
1
•
>
I
'
'
1
I
1
>
V •
»
r
<
>
'
k
•
< T r
1
!
< <
I
'
1 i
J- i
>
T
1
,
/
» •< 1 4
>
I
Figure 10. The velocity field: (a) in the rim, and (b) in the center of the film shown in Fig. 8.
the contact line, but is smooth elsewhere. Note that the rim attaches smoothly to the static flat part of the film, and there is no sign of the predicted [27] liquidliquid-vapor contact-angle region at the interior edge of the rim. The continuum modeling of this problem remains a challenge. 5
Conclusions
We have discussed the use of molecular-level simulations in contact-line problems, where they provide unique information which complements the results of experiments and continuum calculations. MD simulations have quite unambiguously shown that the contact-line continuum singularity is in fact a tiny region that displays slip, high but finite shear stress, and non-Newtonian behavior. Clear evidence for the existence of velocity dependence and hysteresis in the contact angle is found, and in a manner that is manifestly independent of contamination effects and questions of resolution. Simulations of dewetting provide a theoretically unbiased view
102
J. Koplik & J. R.
Banavar
of the flow fields and liquid film shapes. A major issue in the further development of these techniques is computational speed. While it is true that increases in pure computing power have greatly increased the size of systems and the complexity of phenomena that are accessible to MD studies, there is the fundamental difficulty that larger systems require longer physical time scales for their dynamics, and unless one has unlimited CPU time the net gain is modest. Contact-line problems typically involve low Reynolds number, and their physical time evolution is controlled by diffusion of vorticity, increasing as (volume) 2 / 3 . This observation implies not that prospects are grim, but rather that better thinking is required. Simulating an entire experiments with molecules is an inefficient procedure because it provides globally high resolution whereas coarse is adequate outside of the small sub-regions where the continuum approximation may be inadequate. Recent work has begun to develop algorithms that match the two types of calculation in the appropriate parts of a domain [14,31] and methods of this type are likely to push molecular simulation of fluid flows into the future. Acknowledgments This research was supported by the NASA Microgravity Science and Applications Division, with computer resources provided by the NASA Center for Computational Sciences and the San Diego Supercomputer Center. References 1. E. B. Dussan V., Ann. Rev. Fluid Mech. 11, 371 (1979); P. G. de Gennes, Rev. Mod. Phys. 57, 827 (1985); J. F. Joanny and L. Leger, Rept. Prog. Phys. 55, 431 (1992). 2. C. G. Ngan and E. B. Dussan V., J. Fluid Mech. 118, 27 (1982); T. D. Blake, M. Bracke, and Y. D. Shikhmurzaev, Phys. Fluids 11, 1995 (1999). 3. C. Huh and L. E. Scriven, J. Coll. Int. Sci. 60, 11 (1977). 4. E. B. Dussan V., J. Fluid Mech. 230, 97 (1991). 5. M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids, Oxford (1987). 6. M. Mareschal and B. L. Holian, eds. Microscopic Simulations of Complex Hydrodynamical Phenomena, Plenum, New York (1994). 7. J. Koplik and J. R. Banavar, Ann. Rev. Fluid Mech. 27, 257 (1995). 8. M. Kroger, W. Loose, and S. Hess, J. Rheol. 37, 1057 (1993). 9. R. B. Bird, C. F. Curtiss, R. C. Armstrong, and O. Hassager, Dynamics of Polymeric Liquids, 2nd ed., 2 vols., Wiley, New York (1987). 10. P. A. Thompson and M. O. Robbins, Phys. Rev. Lett. 63, 766 (1989). 11. J. Koplik, J. R. Banavar, and J. F. Willemsen, Phys. Rev. Lett. 60, 1282 (1988); Phys. Fluids A 1, 789 (1989). 12. W. Jin, J. Koplik, and J. R. Banavar, Phys. Rev. Lett. 78, 1520 (1997). 13. W. J. Ma, J. R. Banavar, and J. Koplik, J. Chem. Phys. 97, 485 (1992). 14. N. G. Hadjiconstantinou and A. T. Patera, Int. J. Mod. Phys. C 8, 967 (1997); N. G. Hadjiconstantinou, J. Comput. Phys. 154, 245 (1999).
Molecular Aspects of Contact-Line Dynamics
103
15. J. R. Kossoff and R. L. Street, J. Fluids Engr. 106, 390 (1984). 16. P. A. Thompson, W. B. Brinckerhoff, and M. O. Robbins, J. Adhes. Sci. Tech. 7, 535 (1993). 17. J. Koplik and J. R. Banavar, Phys. Fluids A 7, 3118 (1995). 18. G. K. Batchelor, An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge, pp. 224-27 (1967). 19. C. Kittel, Introduction to Solid State Physics, 7th ed., Wiley, NY (1996). 20. J.-X. Yang, J. Koplik, and J. R. Banavar, Phys. Rev. Lett. 67, 3539 (1991); Phys. Rev. A 46, 7738 (1992). 21. J. De Coninck, U. D'Ortona, J. Koplik, and J. R. Banavar, Phys. Rev. Lett. 74, 928 (1995); Phys. Rev. E 53, 562 (1996). 22. J. Niemenen, D. Abraham, M. Karttienen, and K. Kaski, Phys. Rev. Lett. 69, 124 (1992); J. Niemenen and T. Ala-Nissila, Phys. Rev. E 49, 4228 (1994). 23. J. Koplik and J. R. Banavar, Phys. Rev. Lett. 84, 4401 (2000). 24. F. Heslot, N. Fraysse, and A. M. Cazabat, Nature 338, 640 (1989). 25. Y. D. Shikhmurzaev, J. Fluid Mech. 334, 211 (1997). 26. G. Reiter, Science 282, 888 (1999), and earlier references therein. 27. F. Brochard-Wyart and J. Daillant, Can. J. Phys. 68, 1084 (1989); F. Brochard-Wyart, J.-M. di Meglio, and D. Quere, C. R. Acad. Sci. Paris 304, 553 (1987). 28. J. Israelachvili, Intermolecular and Surface Forces, 2nd ed., Academic, New York, (1992). 29. R. K. Jain and E. Ruckenstein, J. Coll. Int. Sci. 54, 108 (1976). 30. G. I. Taylor and D. H. Michael, J. Fluid Mech. 58, 625 (1973). 31. S. T. O'Connell and P. A. Thompson, Phys. Rev. E 52, 5792 (1995).
This page is intentionally left blank
COMPUTATIONAL M E T H O D S FOR A D V A N C I N G INTERFACES J. A. S E T H I A N * Department
of Mathematics, University of California, Berkeley, CA 94720, U.S.A. E-mail: sethian@math. berkeley.edu
Berkeley,
A large number of computational problems and physical phenomena involve the motion of interfaces separating two or more regions. These can include problems in such areas as fluid mechanics, combustion, materials science, meteorology, and computer vision. In these problems, challenging issues often involve interfaces that change topology, form sharp corners and singularities, depend on delicate geometric quantities such as curvature and normal direction, and involve subtle feedback between the physics and chemistry of the interface and the position/motion of the front itself. In this paper, we will explain some of the issues involved in tracking interfaces, focus on a particular set of numerical techniques that arise from an implicit representation of the interface, and provide an overview of some of the applications that are possible with this view.
1
Characterizations of Moving Interfaces
Suppose we are given an interface separating two regions, and speed F in a direction normal to the interface (see Fig. 1). Typically, this speed F can depend on the position of the front, the local geometry, and the solution to associated partial differential equations on either side of the interface, in addition to given jump conditions across the boundaries. Nonetheless, let us assume for now that the speed F is given.
B
Figure 1. An evolving interface.
'Dedicated to Professor Steve Davis in thanks for having so generously welcomed me into the warm community of those celebrating his birthday and his work.
105
106
1.1
J. A.
Sethian
Mathematical Formulations
There are at least three ways to characterize a moving interface, and none of them are new. Interestingly, each comes from its own branch of mathematics. For simplicity, we discuss the issues in two space dimensions, that is, a one-dimensional interface that is a simple closed curve T(t) moving in two dimensions. Assume that a given velocity field u = (u, v) transports the interface. All three constructions carry over to three dimensions. • The Geometric View: Suppose one parameterizes the interface, that is, T(t) = x(s,t),y(s,t). Then one can write (see [23]) the equations of motion in terms of individual components (see Fig. 2) x = (x, y) as
xt= u
*{w^)>yt = -v*(w^)-
(1)
Figure 2. A parameterized view.
This is a differential geometry view; the underlying fixed coordinate system has been abandoned, and the motion is characterized by differentiating with respect to the parameterization variable s.
The Set-Theoretic View: Consider the characteristic function \{x,y,t), where x IS o n e inside the interface T and zero otherwise (see Fig. 3). Then one can write the motion of the characteristic function as Xt = u-VX-
Figure 3. A set-theoretic view.
(2)
Computational
Methods for Advancing
Interfaces
107
In this view, all the points inside the set (that is, where the characteristic function is unity) are transported under the velocity field. • The Analysis View: Consider the implicit function <j> : R2 x [0, oo) —> R, denned so that the zero level set <> / = 0 corresponds to the evolving front T(t) (see Fig. 4).
Figure 4. The implicit view.
Then the equation for the evolution of this implicit function corresponding to the motion of the interface is given by & + u - V ^ = 0. 1.2
(3)
Discretizations
Each of these views is perfectly reasonable, and each has spawned its own numerical methodology to discretize the equations of motion: • Marker particle methods, also known as string methods and nodal methods, discretize the geometric view, and take a finite number of points to divide up the parameterization space S (see Fig. 5).
Figure 5. Discretized parameterization into markers.
• Volume-of-fiuid methods, also known as cell methods and volume-fraction methods, use a fixed underlying grid and discretize the characteristic function, filling each cell with a number that reflects the amount of characteristic function contained in that cell (see Fig. 6).
108
J. A.
Sethian
Figure 6. Discretized characteristic function into cell fractions.
• Level set methods, introduced by Osher and Sethian [14] approximate the partial differential equation for the time-dependent implicit function through a discretization of the evolution operators on a fixed grid (see Fig. 7).
Figure 7. Discretized implicit function onto grid.
These discretizations contain keys to both the virtues and the drawbacks of the various approaches. • The geometric/marker-particle view keeps the definition of a front sharp. It requires special attention when marker particles collide; these can create corners and cusps, as well as changes in topology. These techniques often go by names such as contour surgery, reconnection algorithms, etc.; at their core, they reflect user-based decisions about the level of resolution. In addition, this discrete parameterized characterization of the interface can be intricate for two-dimensional surfaces moving in three dimensions. • The characteristic/volume-of-fiuid approach straightforwardly applies in multiple dimensions, and handles topological merger easily, since this results from Boolean operations on sets. It requires some method of differentiating the characteristic function x; since by definition this object is discontinuous, one must devise an approximation to Vx in order to perform the evolution update. This is typically done through algorithms which locally reconstruct the front from the volume or cell fractions, and then use this reconstruction to build the appropriate transport terms. • The implicit/level set approach extends to multiple dimensions and handles topological changes easily. In addition, because the function <j> is defined everywhere and smooth in many places, calculation of gradients in the transport
Computational
Methods for Advancing
Interfaces
109
term, as well as geometric quantities such as normal derivatives and curvature are straightforward. However, a numerically consistent and accurate way to produce the correct viscosity solution to the equations of interface propagation is required, as discussed below. 1.3
Implicit Formulations of Interface Motion
Our goal here is to discuss the implications and implementations of this implicit approach. We begin writing the implicit form for the equation of motion of a front evolving with speed F in its normal direction (see [14]), namely t + F\V<(>\ = 0. In order to approximate this equation and solve for the evolving level-set function <j), there are three central issues. • First, an appropriate theory and strategy must be chosen in order to select the correct weak solution once the underlying smoothness is lost; this is provided by the work on the evolution of curves and surfaces and the link between hyperbolic conservation laws and propagation equations, see Sethian [17-19]; leading up to the introduction of level set methods by Osher and Sethian in [14]. • Second, the Osher-Sethian level set technique that discretizes the above equation requires an additional space dimension to carry the embedding, and hence is computationally inefficient for many problems. This is rectified through the adaptive Narrow Band Method given by Adalsteinsson and Sethian in [1]. • Third, since both the level-set function and the velocity are now defined away from the original interface, appropriate extensions of these values must be constructed. These extension velocities have been explicitly constructed for a variety of specific problems. One general technique for doing so for arbitrary physics and chemistry problems is given by Adalsteinsson and Sethian in [2] through the use of Fast Marching Methods to solve an associated equation that constructs these extensions. 1.4
Interrelations Between Techniques
It is important to state that each of the above techniques has evolved to the point where they provide practical, efficient, and accurate methodologies for computing a host of computational problems involving moving interfaces. Marker-particles methods have been around for a very long time, and have been used in a collection of settings, including, for example, Bunner and Tryggvason [4]. Volume-of-fluid techniques, starting with the initial work of Noh and Woodward [13], have been used to handle shock interactions and fluid interfaces (see, for example, Puckett [15]). Level set techniques have been applied to a large collection of problems; general reviews may be found in [22,23]; and an introductory web page may be found at www.math.berkeley.edu/~sethian/level_set.html.
(4)
110 J. A. Sethian
Finally, we note that the strict delineations between various approaches is not meant to imply that the various techniques have not influenced each other. Modern level set methods often use a temporary marker representation of the front to help build the extension velocities; volume-of-fluid methods use differentiation ideas from level set methods to help construct normal vectors and curvature values; and marker models often use an underlying fixed grid to help with topological changes. Good numerics is ultimately about getting things to work; the slavish and blind devotion to one approach above all others is usually a sign of unfamiliarity with the range of troubles and challenges presented by real applications. 2
2.1
Algorithms for Front Propagation: Level Set Methods and Fast Marching Methods The Link With Hyperbolic Conservation Laws
In order to approximate the level-set equation given by & + F | V 0 | = O,
(5)
we must construct suitable approximations to the gradient term. As discussed in [17,18], an evolving interface can develop corners as it evolves, and, once this happens, the correct weak solution must be constructed beyond the development of singularities in the curvature. In order to construct numerical schemes that build this correct solution, we briefly follow the argument given in [18] that links moving fronts to hyperbolic conservation laws. Consider the initial front given by the graph of f(x), with / and / ' periodic on [0,1], and suppose that the propagating front remains a graph for all time. Let ip be the height of the propagating function at time t, and thus ip(x, 0) = fix). The tangent at (x, ip) is (1, tjjx). The change in height V in a unit time is related to the speed F in the normal direction by v_
=
[ + VJ
f
(6)
and thus the equation of motion becomes 4,t = F(l + 1>2x)1'2.
(7)
Use of the speed function F(K) = 1 - en and the formula K = —ipXx/(^ + ~>Pl)3^2 yields
fc - (l + £)^=e-^.
(8)
This is a partial differential equation with a first-order time and space derivative on the left side, and a second-order term on the right. Differentiation of both sides of this equation yields an evolution equation for the slope u = dtp/dx of the propagating front, namely, ut
[-(l+ U 2 ) 1 / 2 ]* = e
1 + u2
(9)
Computational
Methods for Advancing
Interfaces
111
Thus, as shown in [19], the derivative of the curvature-modified equation for the changing height ip looks like some form of a viscous hyperbolic conservation law, with G(u) = — (1+ii 2 ) 1 / 2 for the propagating slope u. Hyperbolic conservation laws of this form have been studied in considerable detail and the strategy for picking the correct weak solution beyond the occurence of a singularity is equivalent to the one for propagating shocks in hyperbolic conservation laws (see [18,19]). 2.2
Link to Numerical Schemes for Hyperbolic Conservation Laws
Given this connection, the next step in development of PDE-based interface advancement techniques was to in fact exploit the considerable numerical technology for hyperbolic conservation laws to tackle front propagation itself. In such problems, schemes are specifically designed to construct entropy-satisfying limiting solutions and maintain sharp discontinuities wherever possible; the goal is to keep fluid variables such as pressure from oscillating, and to make sure that discontinuities are not smeared out. This is equally important in the tracking of interfaces, in which one wants corners to remain sharp and to accurately track intricate development. Thus, the strategy laid out in [19] was to transfer this technology to front-propagation problems, and led up to the level set method introduced in [14]. 2.3
Numerical Algorithms for Solving the Level Set Equation
The above discussion focused on curves that remain graphs. The numerical OsherSethian "level set method" recasts the front in one higher dimension, and uses the implicit analytic framework given above to tackle problems that do not remain graphs; in addition, that work developed multi-dimensional upwind schemes to approximate the relevant gradients. For the sake of completeness, and using the usual notation, a straightforward first-order explicit advancement scheme for the level-set equation is given by hjk
)k - At[ma,x{Fijk, 0)V+^> + min{Fijk, 0)V"
(10)
where m a x p ^ O ) 2 + min(D+^,0)2+-' max(Dr.^,0)2+min(D+^,0)2+ max(D^,0)2+min(i)+^,0)2
V"4>=
1/2
m a x ( ^ ^ , 0 ) 2 + m i n ( D - . ^ , 0 ) 2 + -| 1 / 2 max(D+^,0)2 + m m ( Z ^ 0 ) 2 + . max(D+^,0)2 + min(D-.^,0)2 .
For information about these and higher-order variations, see [14]. 2.4
Adaptivity and Efficiency
Considerable computational speedup in the level set method comes from the use of the "Narrow Band Level Set Method", introduced by Adalsteinsson and Sethian in
112
J. A. Sethian
[1]. It is clear that performing calculations over the entire computational domain is wasteful. Instead, an efficient modification is to perform work only in a neighborhood (or "narrow band") of the zero level set. This drops the operation count in three dimensions to 0(kN3), where k is the number of cells in the narrow band. This is a significant cost reduction; it also means that extension velocities need only be constructed at points lying in the narrow band, as opposed to all points in the computational domain. 2.5
Fast Marching Methods
There is a different view of propagating interfaces given by Fast Marching Methods [20], which exchanges the initial value perspective for a boundary value approach. Fast Marching Methods are finite-difference techniques, more recently extended to unstructured meshes, for solving the Eikonal equation of the form \VT\F{x,y,z) = l,
T= 0onT.
This can be thought of as a front propagation problem for a front initially located at T and propagating with speed F(x, y, z) > 0. We note that this is a boundary value partial differential equation as opposed to an initial value problem given by level set methods, even though it describes a moving interface. This Eikonal equation describes a large number of physical phenomena, including those from optics, wave transport, seismology, photolithography, and optimal path planning, and Fast Marching Methods have been used to solve these and a host of other problems. We refer the reader to [23] and [24] for a large collection of applications based on this technique. Fast Marching Methods are very fast (0(N log N)) methods for solving the Eikonal equation, and rely on a marriage of upwind finite-difference schemes, heapsort techniques, and a Dijkstra-like update ordering that reduces the problem to a single-pass algorithm. We note only briefly two critical facts that lie at the core of Fast Marching Methods: • An upwind difference scheme can be used to approximate the Eikonal equation in a viscosity-satisfying framework, namely 1+
>,0)22 max(D:>,-D+^,0) + ijk ' ijk 2 U , -D+Au, 0) m a x ( 'ijk Z& i ~^ijk
1/2
1 (ii) •Tijk
(see Rouy and Tourin [16]) • The order in which the grid values produced through these finite-difference approximations are obtained is intimately connected to Dijkstra's method [7], which is a depth-search technique for computing shortest paths on a network. In that technique, the algorithm keeps track of the speed of propagation along the network links, and fans out along the network links to touch all the grid points. The Fast Marching Method exploits a similar idea in the context of a continuous finite-difference approximation to the underlying partial differential equation, rather than discrete network links. The resulting technique is 0(NlogN), and hence is extraordinarily fast.
Computational
Methods for Advancing
Interfaces
113
The Fast Marching Method evolved in part from examining the limit of the Narrow Band level set method as the band was reduced to one grid cell. Fast Marching Methods, by taking the perspective of the large body of work on higherorder, upwind, finite-difference approximants from hyperbolic conservation laws, allow for higher-order versions on both structured and unstructured meshes. The Fast Marching Method has been extended to higher-order finite-difference approximations by Sethian in [24], first-order unstructured meshes by Kimmel and Sethian [10], and higher-order unstructured meshes by Sethian and Vladimirsky [26]. Some early applications include photolithography in [21], a comparison of this approach with volume-of-fluid techniques in [9], a fast algorithm for image segmentation in [12], and computation of seismic travel times by Sethian and Popovici [25]; see also [27] for a different Dijkstra-like algorithm that obtains the viscosity solution through a control-theoretic discretization that hinges on a causality relationship based on the optimality criterion. 3
Applications
The range of applications of level set and Fast Marching Methods is vast. Fig. 8 gives a perspective on how some of these topics are related. There are many other aspects in the evolution of these ideas; the chart is meant to give a perspective on how the theory, algorithms, and applications have evolved. The text and bibliography of [23] gives a somewhat more complete sense of the literature and the range of work underway. In this paper, we discuss only one application, namely sintering and flow under surface diffusion. Sintering (see [11, 28]) is the process under which a compact consisting of many particles is heated to such a high temperature that the particles become a viscous creeping fluid, and the particles begin to coalesce together. One of the oldest technological examples involves the production of bricks; other examples include formation of rock strata from sandy sediments and the motion of thin films of metals in the microfabrication of electronic components. At issue is the solution of the equations for creeping flow, in which the body forces on the boundary of the materials depend on the tangential stress derivative on the boundary. In one component of this model, the interface speed F in its normal direction depends on the second derivative of the curvature, where the derivatives are taken with respect to arc length a. Thus, in our level set framework, one wants (in two dimensions) to follow a curve propagating with speed F = —tKaa. Thus, + e/cQQ|V0| = 4>t + e
-eV
V
|V0|
(y,~x
|V<^|J
|V0| = 0,
(y,-
|V#
(12)
This means that the speed depends on the fourth derivative of the level set function. We immediately note that a circle is a stable object, since the curvature is constant. A little examination leads one to think that an ellipse undergoes a restoring force that brings it back into a circle.
114
J. A.
Sethian
T h e o r y of C u r v e a n d Surface Evolution: Shocks, Singularities a n d E n t r o p y Conditions (Ref. [17,181)
Tracking Interface M o t i o n w i t h Schemes from Hyperbolic Conservation Laws {Ref. [19])
Level Set M e t h o d \ = 0 Initial Value Problem
Stationary Perspective
(Ref. [U])
(Ref- [8]) T adaptivity
| VT|F = 1 Boundary Value Problem
T
adaptivity
1
1
NARROW BAND LEVEL SET M E T H O D S
FAST M A R C H I N G METHODS
{Ref. [1])
(Ref. [20])
ADDITIONAL FORMULATIONS Unstructured Mesh Level Set M e t h o d s (Ref- [3])
Coupling t o Physics: Extension Velocities (Ref- [2])
Unstructured Mesh Fast M a r c h i n g M e t h o d s (Ref [10])
I
APPLICATIONS Geometry
Grid G e n e r a t i o n
Seismic Analysis
C o m p u t a t i o n a l G e o m e t r y C o m p u t e r Vision O p t i m a l i t y a n d C o n t r o l Fluid Mechanics
Combustion
M a t e r i a l s Sciences
Semiconductor M a n u f a c t u r i n g Figure 8. Algorithms and applications for interface propagation.
Iter=l
Computational
Methods for Advancing
Iter=100
Iter=200
Figure 9. Motion of an ellipse under the speed F =
Interfaces
115
Kaa.
What about more complex shapes? The problem is quite subtle. Numerical experiments are notoriously unstable when they involve computing fourth derivatives", and are eloquently described by Van de Vorst [28]; he uses marker-particle schemes together with elaborate remeshing strategies to keep the calculation alive. A level set approach to this problem was developed by Chopp and Sethian [5]. In that work, the individual derivatives in the above expression were approximated by central-difference approximations and were used to study the motion of a sequence of closed curves to analyze flow under the second derivative of curvature given by Eq. (12). Here, we summarize some of the results in [5]. First, in Fig. 9, we show the evolution of a simple ellipse under this motion. The transformation shows the elliptical initial state on the left, followed by the evolution into a circle, which then remains fixed after a large number of calculations. This might seem to indicate that a convex shape remains convex as it flows under this evolution equation. This, in fact, is not true, as seen by examining the motion of a slightly more elongated ellipse in Fig. 10. At points of high curvature, the interface moves inward, leaving a bulge that propagates around to the flatter sides until the interface balances itself out. In the case of sharp corners, the effect is more pronounced, as seen in Fig. 11. Fig. 11 shows the evolution of several nonconvex initial shapes, all of which approach the stable state of a circle. The curves are shown at uneven times, and the flows are not completed. Second-derivative flow becomes even more murky in the face of topological change. Imagine two ellipses, each with a large ratio between the major and minor axes. If they are put side by side (rather than end to end), the flatter sides will cross over each other, and one expects (at least in many physical situations) some sort of merger, as in Fig. 12. This example underlies the difference between curvature flow (F = —K) and flow by the second derivative of curvature. In the former, a maximum principle ensures that two separate closed curves will always remain separate under this flow. This allows the sort of natural embedding prescribed by a level set interpretation. In contrast, flow by the second derivative of curvature has no such property, as demonstrated in the previous example. Hence, the notion of embedding the motion of the two ellipses in a single "level set function" requires thought. In order to execute true merger of two regions moving under the second derivative of curvature, care must be taken. In Fig. 13, taken from [5], we show the merger a
For example, computing the solution to the biharmonic equation is delicate.
116
J. A.
Sethian
Iter=l
Iter=100
Iter=200
Figure 10. Motion of an elongated ellipse under the speed F =
a.
ii.
Kaa.
^
H C # a.
$
$
Mr • C • Figure 11. Motion of non-convex curves under the speed F =
Kaa.
of two rectangular regions. The results show how the regions come together. Finally, in Fig. 14, also taken from [5], we show an example of three-dimensional flow under the Laplacian of curvature, revealing the smoothing effects of this flow. For details, see [5].
Computational
f\f\
Methods for Advancing
Interfaces
117
m
w U
Figure 12. Expected motion of two ellipses under the second derivative of curvature.
Finally, we point out that the above algorithm is slow, because it is restricted by the time step due to the Courant condition. In [6], an alternative technique is presented that uses an approximation to the equations of motion that allows for a much larger time step.
Figure 13. Merger of two separate regions under F ••
Acknowledgments All calculations were performed at the University of California at Berkeley and the Lawrence Berkeley Laboratory. All of the work on sintering is joint with Professor David Chopp of Northwestern University. A detailed web page on level set methods and Fast Marching Methods may be found at www.math.berkeley.edu/~sethian/level_set.html. References 1. 2. 3. 4.
D. D. T. B.
Adalsteinsson and J. A. Sethian, J. Corn-put. Phys. 118, 269 (1995). Adalsteinsson and J. A. Sethian, J. Comput. Phys. 148, 2 (1999). J. Barth and J. A. Sethian, J. Comput. Phys. 145, 1 (1998). Bunner and G. Tryggvason, Phys. Fluids 11, 1967 (1999).
118
J. A.
Sethian
Figure 14. Flow under the Laplacian of curvature.
5. D. L. Chopp and J. A. Sethian, Motion by Intrinsic Laplacian of Curvature, CPAM Report PAM-746, Dept. of Mathematics, University of California, Interfaces and Free Boundaries, p. 1 (1999). 6. D. L. Chopp, A. Tongen, and J. A. Sethian, submitted for publication, J. Comput. Phys. (Jan. 2000). 7. E. W. Dijkstra, Numerische Mathematic 1, 269 (1959). 8. P. Garabedian, Partial Differential Equations, Wiley, New York (1964). 9. J. Helmsen, E. G. Puckett, P. Colella, and M. Dorr, Two New Methods for Simulating Photolithography Development, in Proceedings of the SPIE 1996 International Symposium on Microlithography, Santa Clara, CA (June 1996). 10. R. Kimmel and J. A. Sethian, Proc. Nat. Acad. Sci. 95, 8341 (1998). 11. H. K. Kuiken, J. Fluid Mech. 214, 503 (1990). 12. R. Malladi and J. A. Sethian, Proc. Nat. Acad. Sci. 93, 9389 (1996). 13. W. Noh and P. Woodward, in Proceedings of the Fifth International Conference on Fluid Dynamics, eds. A. I. Van de Vooran and P. J. Zandberger, SpringerVerlag (1976). 14. S. Osher and J. A. Sethian, J. Comput. Phys. 79, 12 (1988). 15. E. G. Puckett, in Proceedings of the 4th International Symposium on Computational Computational Fluid Dynamics, Davis, CA (1991). 16. E. Rouy and A. Tourin, SIAM J. Num. Anal. 29, 867 (1992). 17. J. A. Sethian, An Analysis of Flame Propagation, Ph.D. Thesis, Department of Mathematics, University of California, Berkeley, CA (1982). 18. J. A. Sethian, Comm. in Math. Phys. 101, 487 (1985). 19. J. A. Sethian, in Variational Methods for Free Surface Interfaces, eds. P. Concus and R. Finn, Springer-Verlag, NY (1987). 20. J. A. Sethian, Proc. Nat. Acad. Sci. 93, 1591 (1996). 21. J. A. Sethian, Fast Marching Level Set Methods for Three-Dimensional Photolithography Development, in Proceedings of the SPIE 1996 International Symposium on Microlithography, Santa Clara, CA (June 1996). 22. J. A. Sethian, in Acta Numerica, Cambridge University Press (1996).
Computational
Methods for Advancing
Interfaces
119
23. J. A. Sethian, Level Set Methods and Fast Marching Methods, Cambridge University Press, Cambridge, U.K. (1999). 24. J. A. Sethian, SIAM Review 4 1 , 199 (1999). 25. J. A. Sethian and M. Popovici, Geophysics 64, 2 (1999). 26. J. A. Sethian and A. Vladimirsky, Proc. Nat. Acad. Sci., submitted (1999). 27. J. N. Tsitsiklis, IEEE Transactions on Automatic Control 40, 1528 (1995). 28. G. A. L. Van de Vorst, Modeling and Numerical Simulation of Viscous Sintering, Ph.D. Thesis, Eindhoven University of Technology, Febodruk-Enschede, The Netherlands (1994).
This page is intentionally left blank
D I R E C T N U M E R I C A L SIMULATIONS OF M U L T I P H A S E FLOWS G. T R Y G G V A S O N Mechanical Engineering Department, Worcester Polytechnic Institute, Worcester, MA 01609-2280, U.S.A. E-mail: gretar@wpi. edu B. B U N N E R Coventor, Inc., 625 Mount Auburn Street, Cambridge, MA 02138, U.S.A. Direct numerical simulations of multiphase flows are discussed. The Navier-Stokes equations are solved by a finite-difference/front-tracking technique that allows the inclusion of fully deformable interfaces and surface tension, in addition to inertial and viscous effects. Studies of the motion of a few two- and three-dimensional, finite-Reynolds-number, buoyant bubbles in a periodic domain are reviewed. A parallel version of the method makes it possible to use large grids and resolve flows containing 0(100) bubbles. Applications of the numerical method to other multiphase flows are also discussed.
1
Introduction
Multiphase and multifluid flows are common in many natural and technologically important processes. Rain, spray combustion, spray painting, and boiling heat transfer are just a few examples. While it is the overall, integral characteristics of such flows that are of the most interest, these processes are determined to a large degree by the evolution of the smallest scales in the flow. The combustion of sprays, for example, depends on the size and the number density of the drops. Generally, these small-scale processes take place on a short spatial scale and a fast temporal scale, and in most cases visual access to the interior of the flow is limited. Experimentally, it is therefore very difficult to determine the exact nature of the small-scale processes. Direct numerical simulations, where the governing equations are solved exactly, offer the potential to gain a detailed understanding of these flows. Such direct simulations, where it is necessary to account for inertial, viscous, and surface forces in addition to a deformable interface between the different phases, still remains one of the most difficult problems in computational fluid dynamics. Here, a numerical method that has been found to be particularly suitable for direct simulations of flows containing moving and deforming phase boundaries is briefly described. Applications of the method to the study of bubbly flows are reviewed in some detail and other applications are mentioned briefly.
2
Formulation and Numerical Method
A single vector Navier-Stokes equation can be written for a flow field containing many immiscible phases by adding a singular force term to account for interfacial forces at the phase boundaries, or the "front." In conservative form, the equation
121
122
G. Tryggvason & B.
Bunner
is:
dpu
+ V • puu = - V P + V • p(Vu + X?uT) +a
nf nf 6(x - xf) dAf.
(1)
Here, u is the velocity, P is the pressure, and p and p are the discontinuous density and viscosity fields, respectively. 6 is a three-dimensional delta function constructed by repeated multiplication of one-dimensional delta functions, K is twice the mean curvature, and n is a unit-vector normal to the front. Formally, the integral is over the entire front, thereby adding the delta functions together to create a force that is concentrated at the interface, but smooth along the front. The point x is the point at which the equation is evaluated and xj is the position of the front. These equations implicitly enforce the proper stress conditions at the fluid interface. In most applications of the method, all the phases simulated have been taken to be incompressible, so the velocity field is divergence free: V • u = 0.
(2)
This equation, when combined with the momentum equations leads to a nonseparable elliptic equation for the pressure. We also have equations of state for the density and the viscosity:
Here, D/Dt is the material derivative and these last two equations simply state that the density and the viscosity of each fluid remain constant. These equations are solved by a finite-difference/front-tracking method originally developed by Unverdi and Tryggvason [1]. The Navier-Stokes equations are solved by a second-order accurate projection method, using centered-differences on a fixed, staggered grid. In order to keep the boundary between the phases sharp and to accurately compute the surface tension, the phase boundary is tracked by connected marker points (the "front"). The front points are advected by the flow velocity, interpolated from the fixed grid (see Fig. 1). As the front deforms, surface markers are dynamically added and deleted. The surface tension is represented by a distribution of singularities (delta functions) located at the front. The gradient of the density and viscosity becomes a delta function when the change is abrupt across the boundary. To transfer the front singularities to the fixed grid, the delta functions are approximated by smoother functions with a compact support on the fixed grid. At each time step, after the front has been advected, the density and the viscosity fields are reconstructed by integration of the smooth grid-delta function. The surface tension is then added to the nodal values of the discrete Navier-Stokes equations. Finally, an elliptic pressure equation is solved by a multigrid method to impose a divergence-free velocity field. This one-field approach, where the governing equations are solved simultaneously for both fluids, and the interface singularities are approximated on a stationary grid, is also used in the well-known volume-of-fluid (VOF) method and in the more recent level set method. However, in those methods, the interface is located by a marker function that is advected on the fixed grid instead of by explicit marker
Direct Numerical Simulations
of Multiphase
Flows
123
A*'
A
• •
*. *
• •
••
.•
t . fv-
'*n
' ' . ' • ' • ' '
»
. I
• A' f y •. "
*
c
^O-o-
Figure 1. The representation of a phase boundary by connected marker points (front) overlaid on a fixed grid that is used to solve the Navier-Stokes equations.
particles as done here. For a detailed description of the method and various validation tests, see Unverdi and Tryggvason [1], Tryggvason et al. [2], and Esmaeeli and Tryggvason [3]. Accurate and fast simulations of large, well-resolved, three-dimensional flows can only be obtained with parallel computers. The method was therefore reimplemented for distributed-memory parallel computers supporting the Message Passing Interface (MPI) standard. Different strategies are employed for the grid and the front. The Navier-Stokes solver, including the multigrid pressure solver, is parallelized by simple domain decomposition. The flow domain is partitioned into equisized subdomains, each subdomain is supported by a different processor, and boundary data must be exchanged between adjacent subdomains. The method was parallelized in order to conduct simulations of a large number of bubbles. The front computations were parallelized by a master-slave technique, which took advantage of the nature of the physical problem to limit programming complexity and provide good performance. When a bubble is spread over more than one subdomain, one processor is designated as the "master" for that bubble. It handles and centralizes the corresponding data and sends it to the other processors, or "slaves," which then distribute the front data onto the fixed grid. While most of the code parallelizes very efficiently due to the natural load balancing of the physical problem, the parallelization efficiency is somewhat degraded by the multigrid solver. Coarse-grain parallelism is therefore employed in most of our computations. 3 3.1
Results Bubbly Flows
Multiphase bubbly flows occur in both industrial and natural processes. Boiling heat transfer, cloud cavitation, aeration and stirring of reactors in water purification and waste-water treatment plants, bubble columns and centrifuges in the petrochemical industry, the cooling circuits of nuclear reactors, propagation of sound in
124
G. Tryggvason & B,
Bunner
the ocean, the exchange of gases and heat between the oceans and the atmosphere, and explosive volcanic eruptions are just a few examples. Understanding the evolution and properties of bubbly flows is therefore of major technological as well as scientific interest. For engineering applications with a large number of bubbles, computational modeling of bubbly flows relies on equations that describe the average flow field. The two-fluid model, where separate equations are solved for the gas and the liquid phases, is the most common approach. Since no attempt is made to resolve the unsteady motion of individual bubbles, closure relations are necessary for the unresolved motion and the forces between the bubbles and the continuous phase. Closure relations are usually determined through a combination of dimensional arguments and correlation of experimental data. The situation is analogous to computations of turbulent flows using the Reynolds averaged Navier-Stokes equations where momentum transfer due to unsteady small-scale motion must be modeled. For the turbulent motion of single phase flows, direct numerical simulations, where the unsteady Navier-Stokes equations are solved on fine enough grids to fully resolve all flow scales, have had a major impact on closure modeling. Although the need for a similar approach is widely recognized by multiphase flow researchers, the computational difficulties in dealing with unsteady phase boundaries have limited such direct simulations to very simple systems. In the limit of high and low Reynolds numbers, it is sometimes possible to simplify the flow description considerably by either ignoring inertia completely (Stokes flow) or by ignoring viscous effects completely (potential flow). Most success has been achieved when the bubbles are undeformable spheres because, in both of these limits, it is possible to reduce the governing equations to a system of coupled ordinary differential equations for the bubble positions. For intermediate Reynolds numbers it is necessary to solve the full unsteady Navier-Stokes equations. Such simulations for the unsteady motion of many bubbles or particles are relatively recent. Feng, Hu, and Joseph [4, 5] simulated the two-dimensional, unsteady motion of one and two rigid particles, and Tomiyama et al. [6] showed a computation of four bubbles. Fully three-dimensional simulations of a hundred particles have been presented by Johnson and Tezduyar [7]. Unverdi and Tryggvason [1,8] computed the interactions of two- and three-dimensional bubbles and Esmaeeli and Tryggvason [9,10] simulated the unsteady motion of several three-dimensional bubbles. Esmaeeli and Tryggvason [3] followed the evolution of a few hundred two-dimensional bubbles and Hu [11] computed the motion of a few hundred two-dimensional solid particles. The rise of a single buoyant bubble is governed by four nondimensional numbers. Two are the ratios of the bubble density and viscosity to the ones of the outer fluid: Pb/po a n d fJ-b/Ho- Here, the subscript o denotes the ambient fluid and b stands for the fluid inside the bubble. The ratios of the material properties are usually small and have little influence on the motion. The remaining two numbers can be selected in a number of ways. If we pick the density of the outer fluid, p0, the effective diameter of the bubble, de, and the gravitational acceleration, g, to make
Direct Numerical Simulations
of Multiphase
Flows
125
the other variables dimensionless, we obtain: N=P^-
and
Eo=P-^.
(4)
The first number is usually called the Gallileo or the Archimedes number (see Clift, Grace, and Weber [12]), and the second one is the Eotvos number. For a flow with many bubbles, the void fraction, a, must also be specified. To examine the behavior of complex multiphase flows, we have done a large number of simulations of the motion of several bubbles in periodic domains. In Esmaeeli and Tryggvason [9], we examined a case where the average rise Reynolds number of the bubbles remained relatively small, 1-2, and in Esmaeeli and Tryggvason [10] we looked at another case where the Reynolds number is 20-30. In both cases, the deformation of the bubbles was small. The results showed that while freely-evolving bubbles at low Reynolds numbers rise faster than a regular array (in agreement with Stokes-flow results), at higher Reynolds numbers the trend is reversed and the freely-moving bubbles rise slower. Preliminary results for even higher Reynolds numbers indicate that once the bubbles start to wobble, the rise velocity is reduced even further compared to the steady rise of a regular array of bubbles at the same parameters. We also observed that there is an increased tendency for the bubbles to line up side-by-side as the rise Reynolds number increases, suggesting a monotonic trend from the nearly no-mode preference found by Ladd [13] for Stokes flow, to the strong layer formation seen in the potential-flow simulations of Sangani and Didwania [14], for example. In addition to the stronger interactions between the bubbles, simulations with a few hundred two-dimensional bubbles at 0(1) Reynolds number by Esmaeeli and Tryggvason [3] showed that the bubble motion leads to an inverse energy cascade where the flow structures increase continuously in size. This is similar to the evolution of stirred two-dimensional turbulence, and although the same interaction is not expected in three dimensions, the simulations demonstrated the importance of examining large systems with many bubbles. To examine the usefulness of simplified models, the results were compared with analytical expressions for simple cell models in the Stokes-flow and the potential-flow limits. The simulations were also compared to a two-dimensional Stokes-flow simulation. The results show that the rise velocity at low Reynolds number is reasonably well predicted by Stokes-flow-based models. The bubble interaction mechanism, however, is quite different. At both Reynolds numbers, two-bubble interactions take place by the "drafting, kissing, and tumbling" mechanism of Fortes, Joseph, and Lundgren [15]. This is, of course, very different from either a Stokes flow, where two bubbles do not change their relative orientation unless acted on by a third bubble, or the predictions of potential-flow models, where a bubble is repelled from the wake of another one, not drawn into it. For moderate Reynolds numbers (about 20), we find that the Reynolds stresses for a freely-evolving, two-dimensional bubble array are comparable to predictions from Stokes-flow models, while in three-dimensional flow the results are comparable to predictions from potential-flow cell models. Most of these computations were limited to relatively small systems, and while Esmaeeli and Tryggvason [3] presented simulations of a few hundred two-dimensional bubbles at a low Reynolds number, the three-dimensional simulations in Esmaeeli and
126
G. Tryggvason & B.
Bunner
Tryggvason [9,10] were limited to only eight bubbles. Using a fully parallelized version of the method we have recently simulated several three-dimensional systems with up to 216 three-dimensional buoyant bubbles in periodic domains (see Bunner and Tryggvason [16,17]). The governing parameters were selected such that the average rise Reynolds number was about 20-30, depending on the void fraction, and deformations of the bubbles were small. Although the motion of the individual bubbles was unsteady, the simulations were carried out for a long enough time so that the average behavior of the system was well defined. Simulations with different numbers of bubbles have been used to explore the dependency of various average quantities on the size of the system. The average rise Reynolds number and the Reynolds stresses are essentially fully converged for systems with 27 bubbles, but convergence of the average fluctuation of the bubble velocities will require larger systems. Examination of the pair-distribution function for the bubbles shows a preference for horizontal alignment of bubble pairs, independent of system size, but the distribution of bubbles remains nearly uniform. The energy spectrum for the largest simulation quickly reaches a steady state, showing no growth of modes much longer than the bubble dimensions. To examine the effect of bubble deformation, we have done two sets of simulations using 27 bubbles per periodic domain. In one, the bubbles are spherical, in the other the bubbles deform into ellipsoids with an aspect ratio of approximately 0.8. The nearly spherical bubbles quickly reach a well-defined, average rise velocity and remain nearly uniformly distributed across the computational domain. The deformable bubbles generally exhibit considerably larger fluctuations than the spherical bubbles and bubble/bubble collisions are more common. Figs. 2 and 3 show the bubble distribution along with the streamlines and vorticity for one time from a simulation of 27 bubbles in a periodic domain. Here, N = 900, the void fraction is 12%, and Eo = 1 in Fig. 2 and Eo = 5 in Fig. 3. The streamlines in a plane through the domain and the vorticity in the same plane are also shown. In Fig. 4, the average rise (or drift) Reynolds number of the bubbles is plotted versus time for both the nearly spherical bubbles as well as the deformable ones. In a few cases, usually for small void fractions, and after the bubbles have risen for a considerable distance, the bubbles transition to a completely different state where they accumulate in vertical streams, rising much faster than when they are uniformly distributed. This behavior can be explained by the dependency of the lift force that the bubbles experience on the deformation of the bubbles. For nearly spherical bubbles, the lift force will push bubbles out of a stream, but the lift force on deformable bubbles will draw the bubbles into the stream. Although we have not seen streaming in all of the simulations of deformable bubbles that we have done, we believe that the potential for streaming is there. Since the system requires fairly large perturbations to reach the streaming state, it may take a long time for streaming to appear. Simulations starting with the bubbles in a streaming state show that deformable bubbles stay in the stream, but that spherical bubbles disperse. The large velocity spike in Fig. 4 may be an indication of a transition that almost lead to streaming.
Direct Numerical Simulations
of Multiphase
Flows
127
Figure 2. One frame from a simulation of 27 nearly spherical bubbles at 12% void fraction, rising with an average rise Reynolds number of about 20. The streamlines and gray-scale contours of the enstrophy are shown.
3.2
Other Applications
The finite-difference/front-tracking method described in Sec. 2 has also been used to study a number of other problems. We have, for example, examined the motion of a single bubble in a shear flow (Ervin and Tryggvason [18]) and a single drop in a pressure-driven channel flow (Mortazavi and Tryggvason [19]). The results show that a relatively small deformation of a bubble changes not only the value of the lift coefficient, but also its sign. The results for a single drop have shown that for modest Reynolds numbers the drop generally moves to a position about halfway between the center of the channel and its wall. For drops, we have also looked at the collision of two equal-sized drops, both by axisymmetric computations for head-on collisions (Nobari, Jan, and Tryggvason [20]) as well as by fully three-dimensional simulations for off-axis collisions (Nobari and Tryggvason [21]). The primary focus of these studies was when the drops broke up again after initial coalescence. In all cases, we have found good agreement with experimental observations. Recently, we have examined the breakup of accelerated drops (Han and Tryggvason [22]). The computations show both "bag" and "shear" breakup and have helped clarify the breakup mechanism when the density difference between the drops and the surrounding fluid is small. Two-dimensional simulations have been used to examine the dissipation of free surface waves (Yang and Tryggvason [23]) and the nonlinear evolution of the Kelvin-Helmholtz instability of stratified flows (Tryggvason and Unverdi [24]). In addition to problems where two isothermal fluids move together, we have extended the methodology to handle situations where other physical effects must
128
G. Tryggvason & B.
Bunner
Figure 3. One frame from a simulation of 27 deformable bubbles at 12% void fraction, rising with an average rise Reynolds number of about 20. The streamlines and gray-scale contours of the enstrophy are shown.
Eo=5 - - Eoal to o I
0
1
i
1
1
;
1
1
1
1
20
40
60
80
100 time
120
140
160
180
200
Figure 4. The average rise Reynolds number for 27 nearly spherical bubbles (dash-dot line) and for deformable bubbles (solid line) as a function of the nondimensional time.
be taken into account. Real bubbles, particularly in water, are rarely clean, and we have investigated the effect of contaminants on deformable bubbles at finite Reynolds numbers (Jan [25]). While the presence of contaminants usually slows the bubbles down, contamination generally reduces deformations and, for very de-
Direct Numerical Simulations
of Multiphase
Flows
129
formable bubbles at high Reynolds numbers, the reduction in pressure drag can offset the increased frictional drag. We have also done several computations of the migration of drops and bubbles due to a temperature-dependent surface tension. These have shown, for example, that for a fairly wide range of parameters the drops line up in rows perpendicular to the temperature gradient (Nas [26]; Nas and Tryggvason [27]). In Yu, Ceccio, and Tryggvason [28] we used a simple model of cavitation, where all thermal effects are ignored and the pressure inside the bubble is set to a constant, to examine the effect of fluid shear on the growth and collapse of vapor bubbles. A more realistic model of phase change was presented in Juric and Tryggvason [29], where solidification of a pure material is simulated by writing one energy equation for both phases. Heat release due to phase change is included as a source term at the phase boundary. For full multiphase flow problems, the energy equation must be coupled with the Navier-Stokes equations. Juric and Tryggvason [30] presented a method to do this and showed simulations for the growth of a two-dimensional vapor layer near a hot wall. 4
Conclusions
The results presented here show the feasibility of using direct numerical simulations to examine the dynamics of finite-Reynolds-number, multiphase flows. Large-scale simulations of systems of many bubbles have been used to gain insight into the dynamics of such flows and to obtain quantitative data that is useful for engineering modeling. The methodology has also been extended to systems with more complex physics, such as free-surface effects and phase change. Acknowledgments This work was supported by the National Science Foundation under grant CTS9503208 and by a Rackham fellowship from the University of Michigan. The computations were done on the IBM SP2 parallel computers at the Maui High Performance Center and at the Center for Parallel Computing at the University of Michigan. References 1. S. 0 . Unverdi and G. Tryggvason, J. Comput. Phys. 100, 25 (1992). 2. G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han, S. Nas, and Y.-J. Jan, J. Comput. Phys. 169, 708 (2001). 3. A. Esmaeeli and G. Tryggvason, J. Fluid Mech. 314, 315 (1996). 4. J. Feng, H. H. Hu, and D. D. Joseph, J. Fluid Mech. 261, 95 (1994). 5. J. Feng, H. H. Hu, and D. D. Joseph, J. Fluid Mech. 277, 271 (1995). 6. A. Tomiyama, A. Sou, I. Zun, and T. Sakaguchi, Proc. German-Japanese Symp. on Multiphase Flow KfK 5389, 487 (1994). 7. A. A. Johnson and T. E. Tezduyar, Comput. Methods Appl. Mech. Engng. 145, 301 (1997). 8. S. O. Unverdi and G. Tryggvason, Physica D 60, 70 (1992). 9. A. Esmaeeli and G. Tryggvason, J. Fluid Mech. 377, 313 (1998).
130
G. Tryggvason & B.
Bunner
10. A. Esmaeeli and G. Tryggvason, J. Fluid Mech. 385, 325 (1999). 11. H. H. Hu, Int. J. Multiphase Flow 22, 335 (1996). 12. R. Clift, J. R. Grace, and M. E. Weber, Bubbles, Drops, and Particles, Academic Press (1978). 13. A. J. C. Ladd, Phys. Fluids A 5, 299 (1993). 14. A. S. Sangani and A. K. Didwania, J. Fluid Mech. 250, 307 (1993). 15. A. Fortes, D. D. Joseph, and T. Lundgren, J. Fluid Mech. 177, 467 (1987). 16. B. Bunner and G. Tryggvason, Phys. Fluids 11, 1967 (1999). 17. B. Bunner and G. Tryggvason, J. Visualization 2, 153 (1999). 18. E. A. Ervin and G. Tryggvason, J. Fluid Engng. 119, 443 (1997). 19. S. Mortazavi and G. Tryggvason, J. Fluid Mech. 411, 325 (2000). 20. M. R. H. Nobari, Y.-J. Jan, and G. Tryggvason, Phys. Fluids 8, 29 (1996). 21. M. R. H. Nobari and G. Tryggvason, AIAA Journal 34, 750 (1996). 22. J. Han and G. Tryggvason, Phys. Fluids 11, 3650 (1999). 23. Y. Yang and G. Tryggvason, Computers and Fluids 27, 829 (1998). 24. G. Tryggvason and S. 0 . Unverdi, in Fluid Dynamics at Interfaces, eds. W. Shyy and R. Narayanan, Cambridge University Press, Cambridge, U.K. (1999). 25. Y.-J. Jan, Computational Studies of Bubble Dynamics, Ph.D. Thesis, University of Michigan, Ann Arbor, MI (1994). 26. S. Nas, Computational Investigation of Thermocapillary Migration of Bubbles and Drops in Zero Gravity, Ph.D. Thesis, University of Michigan, Ann Arbor, MI (1995). 27. S. Nas and G. Tryggvason, in AMD 174/FED 175 Fluid Mechanics Phenomena in Microgravity, eds. Siginer, Thompson, and Trefethen, ASME, p. 71 (1993). 28. P.-W. Yu, S. L. Ceccio, and G. Tryggvason, Phys. Fluids 7, 2608 (1995). 29. D. Juric and G. Tryggvason, J. Comput. Phys. 123, 127 (1996). 30. D. Juric and G. Tryggvason, J. Multiphase Flow 24, 387 (1998).
A PHASE-FIELD MODEL W I T H C O N V E C T I O N : N U M E R I C A L SIMULATIONS D. M. ANDERSON Department of Mathematical Sciences, George Mason Fairfax, VA 22030, U.S.A.
University,
G. B. MCFADDEN Mathematical and Computational Sciences Division, National Institute of Standards and Technology, Gaithersburg, MD 20899-8910, U.S.A. A. A. WHEELER Faculty of Mathematical Studies, University of Southampton, Highfield, Southampton, S017 1BJ, U.K. In a previously-developed phase-field model of solidification that includes convection in the melt [1], the two phases are represented as viscous liquids, where the putative solid phase has a viscosity much larger than the liquid phase. In this paper, we report numerical computations on a simplified form of this model that represents the growth of a two-dimensional dendrite in a thin gap between two parallel thermally-insulting plates. In these computations, flow in the liquid arises because of the differing densities of the solid and liquid phases.
1
Introduction
The notion of representing the boundary between two bulk thermodynamic phases as a diffuse interface dates back to the work of Poisson [2], Gibbs [3], Maxwell [4], Rayleigh [5], van der Waals [6], and Korteweg [7] in the 19th Century. The central assumption is that there is an interfacial region of small but nonzero thickness separating the two bulk phases. In such models, quantities such as surface tension, that in a sharp-interface description are regarded as localized on the interfacial surface, are instead recognized as being distributed through the interfacial region. Diffuse interface models may be based on an extended thermodynamics that incorporates effects involving gradients of the thermodynamic variables ("nonclassical terms") to account for nonlocal effects. That a model incorporating a diffuse interface in this way is referred to as "nonclassical" is perhaps ironic in light of the above history, and speaks volumes for the success of the "classical" sharp-interface description of interfacial free boundary problems. Despite the overall success of the classical approach, there are still special situations in which a diffuse-interface description of an interface between two bulk phases is a viable and even necessary approach. At least three types of situations may be identified, (i): The thickness of the interface becomes comparable to or larger than other mesoscopic length scales of interest in the problem. An example of such a situation is in the case of a fluid near its critical point, where the thickness of the interface diverges. Early diffuse interface models were developed to investigate this problem (see, e.g., van der Waals [6] and Rowlinson and Widom [8]). (ii): The length scales of interest in the problem under consideration are so small that they
131
132
D. M. Anderson,
G. B. McFadden & A. A.
Wheeler
are comparable to the thickness of the interface. Contact line problems in fluid mechanics (e.g., Davis [9]) are potentially in this category, as the diffuse nature of a fluid-fluid interface may become important at the small scales of interest near a contact line. In fact, recent calculations using diffuse-interface descriptions suggest that the force singularity associated with the classical free boundary description of a moving contact line (Dussan V. and Davis [10]) can be relieved when a nonzero interface thickness is taken into account (Jacqmin [11,12], Seppecher [13]). (hi): A diffuse-interface formulation becomes a viable computational alternative to the classical free boundary problem when the morphology of the interface becomes very complicated or changes its topology. An important example is the solidification of dendrites, where sidearms branch from the main stem of the dendrite in a complicated dynamical process that involves both growth and subsequent coarsening behavior. Many successful computations of dendritic growth have now been performed [14-21]. Diffuse-interface theories have been developed and applied successfully in a wide range of other physical situations as well, such as superconductivity [22], liquid crystals [23], spinodal decomposition [24,25], ordering transitions in alloys [26-28], and a variety of hydrodynamic phenomena [29]. Our interest here concerns a phase-field model that accounts for both solidification and fluid motion. This work extends the phase-field model of the solidification of a pure material that was first proposed by Langer [30,31] and subsequently developed by a number of researchers [32-37]. Phase-field models provide an example of a diffuse-interface model in which an order parameter, , is postulated whose value indicates the phase of the system at a particular point in space and time (in this paper = 1 and 4> = 0 denote the solid and liquid phases, respectively). Langer represented the free energy of a single-component system by a gradient energy functional of the form
T = Jv{\e2\V4>? + f{,T)}dV,
(1)
where e is the gradient-energy coefficient and T is the temperature. The free energy density, f(
where 1/M is a positive constant termed the mobility, p is the density, c is the heat capacity per unit mass, k is the thermal conductivity, and L is the latent heat per unit volume of the material. This phase-field formulation replaces the free boundary problem associated with the sharp-interface model of an interface by a coupled pair of nonlinear reaction diffusion equations. The location of the interface is represented by the level set = 1/2.
A Phase-Field
Model with Convection:
Numerical Simulations
133
An early attempt to include fluid motion within a phase-field model of solidification is due to Caginalp and Jones [38,39]. They appended the inviscid momentum equation and the continuity equation to the phase-field model, but did not address the issues of momentum balance in the solid and capillary contributions to the stress tensor. Diepers et al. [40,41] have employed the methodology of two-phase fluid flow, where , in line with our aim of directly extending conventional phase-field models of solidification to account for convection. This has the advantage that we may treat quasi-incompressible systems [45], in which the density field is taken to be a prescribed function of . We sketch how the model may be derived in the setting of irreversible thermodynamics. The quasi-incompressibility assumption restricts the form of the thermodynamic potentials that may be employed [45]. The model comprises the compressible Navier-Stokes equations with a modified stress tensor that includes additional terms related to gradients of , an energy equation, and a phase-field equation involving a material time derivative of (p. We go on to describe computations based on a simplified form of this phase-field model. In particular, we study a configuration in which a dendrite grows into an undercooled liquid between two thermally insulating plates. This allows us to avoid directly solving the generalized compressible Navier-Stokes equations by adopting a Hele-Shaw approximation. The densities of the solid and liquid phases are allowed to differ, and we study numerically the effect of the density-induced flow on the growth of the dendrite. 2
The Model
We consider a non-isothermal system consisting of a pure material that may exist in two distinct phases. We follow the standard phase-field methodology and introduce a phase-field variable, <j>(x,t), whose value indicates the thermodynamic phase of
134
D. M. Anderson,
G. B. McFadden & A. A. Wheeler
the system as a function of position, x, and time, t. A solid-liquid interface is represented by a thin layer in which the phase field varies rapidly between zero (liquid) and unity (solid). The governing equations are derived by following the formalism of irreversible thermodynamics [37,46-48] as described below. 2.1
Governing Equations
We assume that the total entropy, S, in a material volume, Q{t), of the system is given by Jn< /fi(t)
1 PS--4T\V4>)
dV,
(4)
where p is the density and s is the entropy per unit mass. The first term in the integrand, ps, is the classical entropy density per unit volume and the second is a nonclassical term associated with spatial gradients of the phase field. Here, the gradient-entropy coefficient es is assumed to be a constant for simplicity, and F is a homogeneous function of degree unity. The function T allows for a general anisotropic surface energy of the solid-liquid interface and allows the Cahn-Hoffman ^-vector formalism for sharp interfaces [49, 50] to be generalized and extended to diffuse-interface models [44,51]. An isotropic surface energy results from the choice
i W ) = |V0|. The total mass, A4, linear momentum, V, and internal energy, £, associated with the material volume are assumed to have the forms: M = / V=
/
pdV,
(5)
pudV,
(6)
JQ(t)
£=
pe+\P\u\^l-elT\V4>)
dV,
(7)
/n(t)
respectively. Here u is the velocity, e is the internal energy density per unit mass, and €E is the gradient energy coefficient, which is assumed to be constant. The thermodynamic relations
de = T ds + 4 dp + 4l # . (r
(8)
ocp
e = Ts-p/p
+ ti,
(9)
are assumed to apply locally, where p is the thermodynamic pressure and p, is the chemical potential (or Gibbs free energy per unit mass). The physical balance laws for mass, linear momentum, and internal energy are given by
dp
f
— = I "*
J6U(t)
n- mdA,
(11)
A Phase-Field
~y~ + / «*
Model with Convection:
qE-ndA=
J6Q(t)
I
Numerical Simulations
n-m-udA,
135
(12)
JgQ(t)
respectively, where h is the outward unit normal to SQ(t), m is the stress tensor, and qE is the internal energy flux. The momentum balance (11) requires that the rate of change of the total momentum of the material volume results from forces acting on its boundary 6£l(t) (for simplicity we neglect body forces such as gravity; their inclusion is straightforward). The energy balance (12) equates the rate of change of the total internal energy of Q(t) plus the energy flux through its boundary to the rate of work of the forces at its boundary. In addition, the entropy balance takes the form —-+
/ qs-ndA= Jsn(t)
at
sproddV,
(13)
Jn{t)
where qs is the entropy flux and sprod is the local rate of entropy production. The second law of thermodynamics is then expressed by the requirement that sprod is non- negative. To proceed, we recast the conservation laws (10-13) as differential equations. These are used to express the local entropy production in terms of the fluxes m, (JE, and qs, as well as Dfi/Dt. We then identify forms for these quantities that ensure that the local entropy production is non-negative. The fluxes that result from this procedure involve both classical contributions and non-classical contributions that depend on V>. In addition, we obtain an evolution equation for the phase field. The details of this procedure are given in Ref. [1] and result in the following governing equations: Dp = - p V • u, Dt p
~bl
= v m
l)t
= e2AT)V
M
p
(14) (15)
' > ' [ r ( V ^ "Pc|'
De — = V • [fcVT] + eEV • I W ) £ Dt
(16)
D4> -£ + Dt
ms:Vu,
(17)
where 1/M is a mobility, m is the stress tensor [see Eq. (24)], ms is a modified stress tensor [see Eq. (25)], ep is the Helmholtz gradient energy coefficient given by eF(T) = e\ + Tes, g(T,p,d>) = e — Ts + p/p is the Gibbs free energy per unit mass, and £ is the generalized ^-vector [44] whose components are defined by £j = dT(p)/dpj, where we have written p = V(f>. The density of the two bulk phases may be different, and we will assume that p depends on 4> alone, p((t>) = pSr (>)+ PL [ 1 - r (»)],
(18)
where r() is a monotonic increasing function with r(0) = 0 and r(l) = 1; suitable choices include r((f>) = <j> or r(<j>) = <^>2(3-2>). This assumption, in which the density does not depend on temperature or pressure, is known as quasi-incompressibility, as it still allows a nonzero divergence of the velocity vector. This assumption places
136
D. M. Anderson,
G. B. McFadden & A. A.
Wheeler
a constraint on the form of the underlying thermodynamic potentials [45] that requires the underlying Gibbs free energy (per unit mass) to have the form (P - Po) g(T,p,) =go(T, >) + •
P() where po is a reference pressure. Here, we assume that the function go(T,t the form e0 -
cTM
r{4>)L
cT In
T TMJ
' 4a"
<7o(T,<
1
1 4as
H{4>)
1
T
(19) has (20) (21)
H(),
in which case the corresponding expressions for the internal energy and entropy densities are 1 PO (22) e = e0 + c(T - TM) r^L+^-H^) 1
e0 - r{4>)L -
TM
Aas
H{4>) + cln
(23)
TM r
where l/as — 1/a—l/as- Here, 1/a is the height of the double well of the Gibbs free energy density at T = TM, and 1/CLE and l/as are the heights of the double wells in the internal energy and entropy densities, respectively. The double-well potential H{4>) is a prescribed function of <j> (see [56]). The quantity eo is a constant reference energy and both the heat capacity per unit mass c and the latent heat per unit mass L are assumed to be constant. The temperature TM is the melting-point temperature at the reference pressure po. The tensors m and ms are given by m =
ms
AT)
-p +
r 2 (v»)
4 (T)r(V<£)£ ® V + r ,
-P + ^ r 2 ( V 0 ) I - T e | r ( V ^ ) ^ V^ + r ,
(24) (25)
where r is the viscous stress tensor, T =H
Vu+(Vu)T
-(V-u)I
(26)
and fi is the viscosity, which is a function of <j>, (27)
K) = M S K 0 ) + ML[1 • r ( 0 ) ] .
By examining the isothermal one-dimensional solution of the governing equations at the melting temperature TM with ps — PL, it may be shown that the surface tension 7, interface thickness I, and interface attachment coefficient p,Q are related to the phase-field parameters by 7(n) = ^ r ( n ) 7 f ,
l(n)
eF{TM)T{n)J—, V PL
/x0(n)
6pLLlT(n) TMM (28)
A Phase-Field
Model with Convection:
Numerical Simulations
137
We will henceforth confine our attention to the case of isotropic surface energies and set r(V>) = |V^|; in this case we note that T(n) = 1. It is also convenient to define an associated capillary length by lc — TMJ/(PL[L2/C]). 2.2
Dimensionless Governing Equations
We non-dimensionalize the governing equations by introducing the following dimensionless variables, which we denote with a prime: x = lox",
t = —t',
u
V
m
L
T = TM +
n>
m ,
I'02
fj, = pLp',
PLP
p = po +
P
~f P,
(29)
(30)
k = kLk'.
Here, the reference length scale IQ is a typical length scale associated with the interface shape, such as a dendrite tip radius, the reference time scale is 12/KL, and the reference velocity scale is U = KL/IO, where KL is the thermal diffusivity in the bulk liquid phase. The dimensionless governing equations are dp + V • (pu) = 0, dt Du
(31)
„
p - ^ = V-m, 2 e
M ^ = e 2 ( l + a0)V2<; Dt
(32)
1
£)
-(i + Wfl'W + W\4>) + Ip-zz D<j>_
De
P^=V-[W)W] + eV^V 2
,(33)
(34)
•K
where, for simplicity, we have omitted the primes on the dimensionless variables. The dimensionless internal energy density is given by e
= Q-r^)
+
iH{<j>)-^--, 2 A7 p
(35)
and the dimensionless stress tensors are m = crp + (1 + aO)^
+ Prr,
(36)
ms =
(37) UL
is the kinematic (38) (39)
r = p{4>) Vu + {Vu)T - - ( V • u) I
(40)
138 D. M. Anderson,
G. B. McFadden & A . A . Wheeler
The source term in the energy equation is m
H = Y
• Vw = j j - [- (p + p*) I + a (9 + S'1) a4, + Prr] : W ,
s
(41)
and the dimensionless parameters are given by Le
aL
s
Q =
0 = P
2
CC F(TMY
e=
8=— 2aEV
ascTM'
f.
A
=^> P* = - ^ . 5 = 4 - ,
to „
=
6l0lcL2 2
K LCTM
u=-^~ = ^ PLpL 2aLe2F(TMy
6ZC =
6lo2_
pLKl jtf =
(43)
cTM
*LMTMC 2
PLI^V
(42) ^Z>
£>PLIICL
=
[KL/IC]
(U)
[L/C]H0'
and Q(
- 1) r # ) ,
M>) = l + ^ M _ i j r(>),
p{4>) = 1 + ^ - 1 ^ r(0).
(45)
We note that the parameter 7 is related to the capillary (or crispation) number Ca by 7 = 6Pr/Ca, where Ca = PL^L/(IOJ)In the absence of flow these equations reduce to the generalized phase-field equations recently studied in Refs. [52] and [53], The leading-order, free boundary problem that emerges from a sharp-interface limit of these equations depends on the distinguished limit that is taken [52,54,55]. The limit in which A = 0(1) as e —> 0 corresponds to the so-called 'thin interface' limit studied by Karma and Rappell [55]. In this analysis when the thermal conductivities of the solid and liquid phases are unequal, the leading-order temperature is discontinuous across the interface and the leading-order modified Gibbs-Thomson equation contains terms dependent on the interfacial temperature gradients. However, if A = 0(e) as e —> 0, the socalled 'classical' limit, the temperature is continuous across the interface at leading order and a nonlinear form of the modified Gibbs-Thomson equation is obtained at leading order. However, if we formally set the coefficients a, (3, S, and v to zero, thereby omitting the nonstandard terms in the generalized phase-field equations, then the classical sharp-interface analysis recovers, at leading order, a standard free boundary problem in which the interfacial temperature is continuous and the conventional modified Gibbs-Thomson equation is obtained. Here, we study a simplified form of our model by setting the constants a, j3, 6, and u to zero, and we neglect the source term H in the energy equation. The dimensionless governing equations of the simplified model are d
JL + V • (pu) = 0,
(46)
Du p
Wt=v'm'
(47)
A Phase-Field
2 £
M^=e2V20-p
Model with Convection:
I^)
+wW +
Numerical Simulations
ipAQ)
139
(48)
/>^ = V-[W)V0],
(49)
where the dimensionless stress tensor is m = crp + a't' + PrT.
(50)
We have recently examined the full system of governing equations, including flow, in the sharp-interface limit [56]. Our investigations reveal that, in the classical sharp-interface limit, the boundary conditions at a sharp interface in equilibrium comprise the normal stress balance including surface tension and the ClausiusClayperon equation all under isothermal conditions. In the nonequilibrium case we find hydrodynamic conditions on the normal and tangential velocities representing the conservation of mass and the no-slip condition. Jumps conditions on the normal and tangential stresses are also obtained. The temperature is found to be continuous across the interface while the jump in heat flux across the interface is modified by nonequilibrium effects. The temperature of the interface is found to obey a nonequilibrium version of the Clausius-Clapeyron relation. 3
Model Computations
We now describe computations, based upon the phase-field model given by Eqs. (46), (47), (48), (49), and (50) that represent the density change flow associated with the growth of a dendrite from an undercooled melt. To proceed we make a number of additional approximations in order to develop a simplified phase-field model that captures the qualitative features of this situation. First, we consider the dendrite to be two-dimensional and growing in a uniform thin gap of width d between two thermally insulated flat plates. This allows us to ignore the effects of inertia and to model the flow using a Hele-Shaw approximation. The momentum equation may then be written as + ^ V 0 | 2 ) - eV • [V> ® V>] + ^ V •
V (-±p
{LI(
= 0.
(51)
In the absence of flow it is known that it is essential to include surface energy anisotropy in order to compute dendritic structures using a phase-field model [14]. We will accordingly retain anisotropic surface energy terms in the phase-field equation alone. Specifically, an isotropic surface energy term is used in the momentum equation (51), while the phase-field equation (48) is modified to allow for anisotropic surface energy by using the Calm-Hoffman ^-vector formalism e
'
M
^
= e V
'
' [ r ( V ) ^ ~~ P \\H'{(t,)
+
X9r {
'
(52)
so that the direct effect of anisotropy is upon the interfacial surface energy rather than the flow. We note we have also omitted the pressure dependence in the freeenergy term in the phase-field equation. This is a reasonable approximation for
140
D. M. Anderson,
G. B. McFadden & A. A.
Wheeler
density-driven flows, as evidenced by the insignificant variations of the melting temperature due to pressure fluctuations in the Clausius-Clayperon relation under these conditions. In order to simplify the system further we make the approximation V • (V(/> (8> V>) w V(|V>|2). This approximation is exact in one space dimension, but not higher dimensions. We justify it by observing that the phase-field variable only changes in the thin interfacial regions where it depends primarily on the perpendicular distance through the layer and hence is approximately one dimensional. Using this approximation the momentum equation (51) becomes - V p + y V . W ) T ] = 0,
(53)
P=4p+||V0|2.
(54)
where 7 * We now integrate these equations across the narrow gap of width d « 1 [58] to obtain -3d2 /z\ (d Ca fi((f>)
© ffl *•
<«>
where here and below the operator V acts in the plane parallel to the thin gap. We now apply the continuity equation (46) to find that p satisfies 2_
2Ca
f.y(4>)d4>
p'(>) V0 • Vp. _ p-{4>) P{4>) . P'(4>)
(56)
In our numerical computations, we solved the energy equation (49), phasefield equation (52), and pressure equation (56) using VLUGR [57]. This freelydistributed package is designed to solve systems of parabolic partial differential equations in which the solution exhibits regions in space with large gradients. It employs a finite-difference discretization allied to local grid refinement and a variable time step integration of the underlying discretized equations. Computations were conducted on the rectangular domain [0,X] x [0, Y]. Neumann boundary conditions were employed on and 6 on all four sides. However, for the pressure, Neumann boundary conditions were only invoked on the sides x = 0 and y = 0, with Dirichlet boundary conditions on the other two sides. The initial condition represented a small circular solid region centered on the origin in a uniformly undercooled melt with dimensionless temperature T = C(TM — TQ)/L, where To is the initial dimensional temperature. The governing equations were solved with p(<j>) given by Eq. (45), r'{(f)) — 30^ 2 (1 — (f>)2, and H(<j>) = (p2(l — 4>)2. The surface energy had a four-fold anisotropy with T(n) = 1 + 0.015 cos(49), where n = V<£/|V<^>| is a unit vector in the (x,y) plane and 6 is the angle between n and the x-axis. The values of the dimensionless parameters used in the computations are given by Ca = 30, ps/l^L = 1, ^sl^L — 1, A = 7.5, and M = 10. In Fig. 1, we display the results of a typical computation in which PS/PL = 0.9, X = 1, and Y = 3. This figure shows the pressure field, the velocity field and
A Phase-Field
Model with Convection:
Numerical Simulations
141
Pressure
0.8
0.7
0.6
0.4
0.3
0.2
0.1
Figure 1. The phase field, pressure field, and velocity field for a computation at t = 0.3 with PS I PL = 0-9 on a 1 x 3 domain with four-fold anisotropy. The grayscale indicates the liquid ((/> = 0) and solid {<j> = 1) regions, the solid curves are the isobars, and the small arrows represent the velocity field.
the phase field at time t = 0.3. The solid curves are isobars, the arrows represent the local velocity, and the shading indicates the phase field. The x- and y-axes represent planes of symmetry in the calculation, but since X ^ Y the resulting shape has two-fold, not four-fold symmetry due to the presence of the sidewalls. The dendrite growing in the x-direction has a blunter tip than the dendrite growing in the y-direction due to its closer proximity to the sidewall, which has a significant effect on the growth dynamics at this stage. Since the density of the solid is less than that of the liquid, a given amount of material will expand upon solidification, which drives a flow away from the interface into the melt. For a sharp-interface model, the conservation of mass boundary condition takes the form Un
^n
^ - 1 PL
(57)
142
D. M. Anderson,
G. B. McFadden & A. A.
Wheeler
t
| —
^^-
11
Pressure | " ; . : > 0.9
• o.8
•0.7
- 0.6
-0.5
- 0.4
0.3
-0.2
-0.1
Jo 0
0.5
1
1.5
2
Figure 2. The phase field, pressure field, and velocity field at t = 0.75 for a computation with Ps I PL = 1.1 on a 2 x 4 domain with four-fold anisotropy. The grayscale indicates the liquid ( = 0) and solid (
where un = ft • u is the normal component of the fluid velocity at the interface, and vn is the normal velocity of the interface. Thus, for solidification with vn > 0, the flow is away from the interface (un > 0) for ps < PL, and is toward the interface (un < 0) for ps > PL- The computation shows that the flow is greatest in the vicinity of the tip of the dendrite. For this calculation with equal viscosities in the solid and liquid phases, there is also a significant flow in the solid region. This artifact is reduced for computations with ^SIP-L 3> 1; here we are illustrating an extreme example of this effect. Fig. 2 shows a similar situation, but with the density in the solid greater than that of the liquid, PS/PL = 1-1, with X — 2 and Y = 4 at a time t = 0.75. In this case, the advection is toward the interface, as expected. Between the two dendrite tips is a narrow liquid intrusion where little solidification is taking place; the flow velocities that are induced by the density change upon solidification are correspondingly small in this region.
A Phase-Field
4
Model with Convection:
Numerical Simulations
143
Conclusions
In this paper we have shown that computations based on a simplified form of a recent phase-field model that includes convection [1] exhibits numerical solutions that show the expected physical behavior. In particular, we considered the growth of a dendrite in a thin gap between two thermally insulated plates and allowed the density of the solid and liquid phases to be different. We found that the flow was directed towards or away from the dendrite depending on whether the density of the solid phase was greater or less than that of the liquid phase, respectively. Our model for solidification with convection is derived using the formalism of irreversible thermodynamics, and allows the systematic incorporation of a consistent thermodynamic description of the two-phase system. It allows a unified treatment of both equilibrium and non-equilibrium effects in a single set of governing equations. Sharp-interface limits of the diffuse-interface description then lead to boundary conditions of the solid-liquid interface that recover the usual conditions at equilibrium, and provide thermodynamically-consistent generalizations of these conditions under non-equilibrium conditions [56]. The detailed nature of the nonequilibrium conditions at the interface can be sensitive to the specific forms that are assumed to describe the variation of the thermophysical parameters through the interfacial region; for example, the non-equilibrium solute-trapping behavior of a diffuse-interface model of a binary alloy depends quantitatively on the exact form that is assumed for the variation of solute diffusivity, D(), near the interface [59]. In this model, the solid is treated as a liquid with high viscosity. This allows residual convection to occur in the solid, with a magnitude that is determined by the viscosity ratio. The consideration of extreme viscosity ratios tends to eliminate velocity gradients in the solid, but allows states of uniform convection that correspond to rigid-body motion. This is an attractive feature for dealing with such issues as fragmentation and subsequent motion of sidearms through the melt. Such transported fragments can serve as sites for the growth of independent grains when the fragments are incorporated into the growing phase, which is a problem of considerable technological importance. It is thus beneficial to have a model that allows for both topological changes in the interface as well as possible rigid-body motion of the solid phase. 5
Acknowledgments
The authors dedicate this work to Professor S. H. Davis in honor of his sixtieth birthday. We would like to express our most sincere gratitude to Steve for the wide range of contributions he has made to our careers as teacher, colleague, and friend. The authors are also grateful for helpful discussions with W. J. Boettinger, R. J. Braun, T. J. Burns, S. R. Coriell, B. T. Murray, and R. F. Sekerka during the preparation of this manuscript. This research was conducted with the support of the Physical Sciences Research Division of NASA.
144
D. M. Anderson,
G. B. McFadden & A. A.
Wheeler
References 1. D. M. Anderson, G. B. McFadden, and A. A. Wheeler, Physica D 135, 175 (2000). 2. S. D. Poisson, Paris: Bachelier (1831). 3. J. W. Gibbs, Scientific Papers of J. Willard Gibbs, ed. H. A. Bumstead and R. G. Van Name, Longmans, Green, and Co., London, UK, pp. 55-371 (1928). 4. J. C. Maxwell, in Scientific Papers of James Clerk Maxwell, Vol. 2, Dover, New York, pp. 541-591 (1952). 5. Lord Rayleigh, Phil. Mag. 33, 209 (1892). 6. J. D. van der Waals, Verhandel. Konink. Akad. Weten. Amsterdam, Sec. 1, Vol. 1, No. 8 (1893); Transl. from Dutch: J. S. Rowlinson, J. Stat. Phys. 20, 197 (1979). 7. D. J. Korteweg, Arch. Need. Sci. Exactes Nat. Ser. II 6, 1 (1901). 8. J. S. Rowlinson and B. Widom, Molecular Theory of Capillarity, Clarendon, Oxford, UK (1989). 9. S. H. Davis, J. Appl. Mech. 50, 977 (1983). 10. E. B. Dussan V. and S. H. Davis, J. Fluid Mech. 65, 71 (1974). 11. D. Jacqmin, J. Comput. Phys. 155, 96 (1999). 12. D. Jacqmin, J. Fluid Mech. 402, 57 (2000). 13. P. Seppecher, Int. J. Engng. Sci. 34, 977 (1996). 14. R. Kobayashi, Physica D 63, 410 (1993). 15. A. A. Wheeler, B. T. Murray, and R. J. Schaefer, Physica D 66, 243 (1993). 16. J. A. Warren and W. J. Boettinger, Acta. Metall. Mater. 43, 689 (1995). 17. S.-L. Wang and R. F. Sekerka, Phys. Rev. E 53, 3760 (1996). 18. A. Karma and W. J. Rappel, Phys. Rev. Lett. 77, 4050 (1996). 19. A. Karma and W. J. Rappel, Phys. Rev. E 57, 4323 (1998). 20. N. Provatas, N. Goldenfeld, and J. A. Dantzig, Phys. Rev. Lett. 80, 3308 (1998). 21. N. Provatas, N. Goldenfeld, and J. A. Dantzig, J. Comput. Phys. 148, 265 (1999). 22. V. L. Ginzburg and L. D. Landau, Soviet Phys. JETP 20, 1064 (1950). 23. P. G. de Gennes, Mol. Cryst. Liq. Cryst. 12, 193 (1971). 24. J. W. Cahn and J. E. Hilliard, J. Chem. Phys. 28, 258 (1958). 25. J. W. Cahn, Acta Metall. 9, 795 (1961). 26. J. W. Cahn and S. M. Allen, J. Phys. (Paris) 38, c7 (1977). 27. S. M. Allen and J. W. Cahn, Acta Metall. Mater. 27, 1085 (1979). 28. R. J. Braun, J. W. Cahn, G. B. McFadden, and A. A. Wheeler, Phil. Trans. Roy. Soc. London A 355, 1787 (1997). 29. D. M. Anderson, G. B. McFadden, and A. A. Wheeler, Ann. Rev. Fluid Mech. 30, 139 (1998). 30. J. S. Langer, unpublished notes (1978). 31. J. S. Langer, in Directions in Condensed Matter Physics, ed. G. Grinstein and G. Mazenko, World Scientific, Philadelphia, PA, pp. 165-186 (1986). 32. G. Caginalp, in Applications of Field Theory to Statistical Mechanics, ed. L. Garrido, Springer-Verlag, Berlin, pp. 216-226 (1985).
A Phase-Field
33. 34. 35. 36. 37. 38. 39.
40.
41. 42.
43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59.
Model with Convection:
Numerical Simulations
145
J. B. Collins and H. Levine, Phys. Rev. B 31, 6119 (1985). G. Caginalp and P. C. Fife, Phys. Rev. B 33, 7792 (1986). G. Caginalp, Arch. Rat. Mech. Anal. 92, 205 (1986). G. Caginalp, Phys. Rev. A 39, 5887 (1989). O. Penrose and P. C. Fife, Physica D 43, 44 (1990). G. Caginalp and J. Jones, Appl. Math. Lett. 4, 97 (1991). G. Caginalp and J. Jones, in On the Evolution of Phase Boundaries, ed. M. E. Gurtin and G. B. McFadden, The IMA Series in Mathematics and Its Applications, Vol. 43, Springer-Verlag, New York, pp. 27-50 (1992). H. J. Diepers, C. Beckermann, and I. Steinbach, in Solidification Processing 1997, ed. J. Beech and H. Jones, Proc. 4th Decennial Int. Conf. on Solid. Process., University of Sheffield, Sheffield, UK, pp. 426-430 (1997). C. Beckermann, H. J. Diepers, I. Steinbach, A. Karma, and X. Tong, J. Cornput. Phys. 154, 468 (1999). R. Tonhardt, Convective Effects on Dendritic Solidification, Ph.D. Thesis, Department of Mechanics, Royal Institute of Technology, S-100 44 Stockholm, Sweden (1998). R. Tonhardt and G. Amberg, J. Cryst. Growth 194, 406 (1998). A. A. Wheeler and G. B. McFadden, Eur. J. Appl. Math. 7, 367 (1996). J. Lowengrub and J. Truskinovsky, Proc. Roy. Soc. Ser. A 454, 2617 (1998). S. R. de Groot and P. Mazur, Non-Equilibrium Thermodynamics, Dover, New York (1984). A. P. Umantsev, J. Chem. Phys. 96, 605 (1992). S.-L. Wang, R. F. Sekerka, A. A. Wheeler, B. T. Murray, S. R. Coriell, R. J. Braun, and G. B. McFadden, Physical) 69, 189 (1993). D. W. Hoffman and J. W. Cahn, Surface Science 31, 368 (1972). J. W. Cahn and D. W. Hoffman, Acta Metall. 22, 1205 (1974). A. A. Wheeler and G. B. McFadden, Proc. Roy. Soc. Lond. A 453, 1611 (1997). G. B. McFadden, A. A. Wheeler, and D. M. Anderson, Physica D 144, 154 (2000). C. Charach and P. C. Fife, Open Sys. and Infor. Dyn. 5, 99 (1998). G. Caginalp, Phys. Rev. A 39, 5887 (1989). A. Karma and W. J. Rappel, Phys. Rev. E 53, 3017 (1996). D. M. Anderson, G. B. McFadden, and A. A. Wheeler, Physica D 151, 305 (2001). J. G. Blom, R. A. Trompert, and J. G. Verwer, ACM Trans. Math. Software 22, 302 (1996). G. K. Batchelor, An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge, UK (1970). N. A. Ahmad, A. A. Wheeler, W. J. Boettinger, and G. B. McFadden, Phys. Rev. E 58, 3436 (1998).
This page is intentionally left blank
P H A S E FIELD MODEL OF M U L T I C O M P O N E N T ALLOY SOLIDIFICATION W I T H H Y D R O D Y N A M I C S ROBERT F. SEKERKA+ AND ZHIQIANG BI Department of Physics, Carnegie Mellon University, Pittsburgh, PA 15213-3890, U.S.A. We develop a thermodynamically consistent phase field model for solidification of a multicomponent alloy, including hydrodynamics. The solid is treated as a very viscous fluid. The model is based on an entropy functional that includes gradiententropy corrections for internal-energy density, partial-mass densities, and phase field. It allows for external forces and chemical reactions, and incorporates nonclassical temperature and chemical potentials, defined as functional derivatives. Korteweg stresses related to inhomogeneities in all of these quantities appear in momentum and energy balances.
1
Introduction
The phase field model is based on a diffuse interface model of solidification [1,2] (crystallization from the melt). It was introduced originally [3-6] to study the dynamics of solidification of a pure material. The model incorporates an auxiliary variable, the phase field,
147
148
R. F. Sekerka & Z. Bi
wald ripening and coalescence [53], recoalescence during dendritic solidification [54], and cell to plane front transitions during directional solidification [55]. Phase field models for solidification of eutectic alloys were developed by Karma [56], Wheeler, McFadden, and Boettinger [57], and Elder, Gunton, and Grant [58]. Phase field models that include hydrodynamics have also been formulated. Early versions by Caginalp and Jones [59,60] in which some inviscid hydrodynamic equations were simply added, a finite-element version by Tonhardt and Amberg [61,62], and a hybrid version based on mixture theory by Beckermann et al. [63, 64] incorporated some aspects of fluid flow with phase field models. Gurtin, Polignone, and Vifials [65] have used a phase field model to treat convection in two-phase binary fluids and immiscible fluids. Thermodynamically-consistent diffuse interface models that include hydrodynamics were formulated for a pure material and for a binary alloy by Anderson, McFadden, and Wheeler [66,67]; the reader is referred, to [66] for an excellent review of the literature. Such models include Korteweg stresses [68] that have been related to Noether's theorem [69] in the context of diffuse interface models by Gurtin and Fried [16,17] and Wheeler and McFadden [70]. Subsequently, Anderson, McFadden, and Wheeler [71] developed a phase field model, including convection and anisotropy, for solidification of a pure material and performed numerical computations for dendritic growth [72]. In the present paper, we develop a phase field model for solidification of a multicomponent alloy, including hydrodynamics. The solid is treated as a fluid having high viscosity, as in [71], in order to avoid the more complicated thermodynamics of a crystalline solid [73-76]. This model is based on an entropy functional and includes gradient-entropy corrections in all dependent variables (internal-energy density, partial-mass densities, and phase field). Because of these gradient-entropy corrections, the theory necessarily incorporates non-classical temperature and chemical potentials [40,77-80]. Korteweg stresses related to inhomogeneities in all dependent variables appear, and different contributions of conserved versus non-conserved quantities become evident. In the interest of generality, we include external forces and chemical reactions in order to see how they interact with non-classical temperature and chemical potentials. Our results reduce to those of Anderson, McFadden, and Wheeler for a compatible case [71]. 2
Theory
We treat a multicomponent fluid system of K components. Our model is based on an entropy functional of the form [ sNCd3x, Jv where V is an arbitrary subvolume of the system and S=
SNC = s(u,pi,p2,...pK,
~ - eu\Vu\2 + Y,£i\Vp^\2
(1)
+ e*>|V
(2)
i=l
is a non-classical entropy density. Here, s(u,pi, p2, • • • pK,
Phase Field Model of Multicomponent
Alloy Solidification
149
mass density of component i, ip is the phase-field variable, and eu, e^, and ev are gradient-entropy coefficients. For simplicity, and in order to focus on aspects of a multicomponent system including fluid flow, we assume isotropy and treat the gradient-entropy coefficients as constants. Generalization to the anisotropic case can be accomplished along the lines indicated by Wheeler and McFadden [70] and carried out for a pure material for gradients only in
_ dS_
f
Sprod '- dt+J 'J , A
-hd2x
> 0,
(3)
where A is the bounding area of the fixed but arbitrary volume V, n is the unit outward normal to V, and J s is the total entropy flux due to both diffusive and convective transport. The inequality in Eq. (3) relates to the second law of thermodynamics, which we require to hold in any subvolume of the system as an inequality in order for an irreversible process to occur. Taking the time derivative of Eq. (1) and integrating by parts with the aid of Gauss's theorem, we obtain
Vod = j
(jj
* + S^
" ~ £ (Y)
dAx
K
L
• hd2x,
(4)
1=1
where the dots above u, pi, and ip indicate partial differentiation with respect to time. The functional derivatives NC
-m
ds ou
NC
ds_
dpi
_2 eiV 2 p/,
(5) (6)
define TNC as a non-classical temperature and ^ c as a non-classical chemical potential" of species i, expressed per unit mass. This notation is motivated by the fact that the classical temperature T and the classical chemical potentials /ij are defined by \/T = ds/du and —fii/T = ds/dpi and also because TNC and fi^c are uniform at equilibrium in the absence of external forces, as will be seen subsequently. The functional derivative ds
^_0
(7)
that multiplies
150
R. F. Sekerka & Z. Bi
For the total masses Mi, the total momentum P , and the total internal energy U, we assume that pi d3x,
Mi= P = / Jv
(8)
pvd3x,
(9)
ud x,
(10)
1 u + -pv2] d x,
(11)
U so that the total energy E
where p = Yli=i Pi *s t n e total density and v is the bary centric velocity. We could include excess quantities in the form of gradient corrections to the integrands in Eqs. (8-11), but we omit this complication in the interest of simplicity and also because it would introduce many new and unknown parameters. The reader is referred to the work of Anderson, McFadden, and Wheeler [66,67] for a treatment of a pure material in which the only excess quantity is a gradient correction to the internal-energy density of the form (1/2)KE\'VP\2, where p is the total density, as well as a treatment [71,72] of a pure material in which the only excess quantities are (anisotropic) gradient entropies and energies of the form —(1/2) e | [r(Vy)] 2 and (1/2) e2E [r(V?)]2, respectively. In [66,67], they also treat a binary alloy in which the only excess quantity is a correction to the internal-energy density of the form (1/2).?CE|VC| 2 , where c is the mass fraction of a component. 2.1
Equilibrium
Before treating the dynamics of our system, we examine the conditions for equilibrium for a system contained in a rigid volume Vo having area AQ, a fixed total internal energy U, and fixed total masses Mi of each species. We exclude chemical reactions and external forces in this section, but include them in subsequent sections. For equilibrium, we have the variational condition 0=
S(S-\uUNC
6u +
E
Vi\
NC
6pi + Sv6ip > d?x
i=l u6uVu
+X
iSpiVpi + e^S/pX/ip • hd2a
e
(12)
i=\
where the constants \u and A; are Lagrange multipliers. Since the variations 8u, 6Pi, and Sip are independent within the volume Vo and on the surface Ao, we deduce the equilibrium conditions 1\NC f)
=A U )
(13)
Phase Field Model of Multicomponent
(EL Y)
Alloy Solidification
151
= A.
(14)
Sv = 0,
(15)
as well as natural boundary conditions on AQ, such as fixed values of u, Pi, and
equilibrium, no external forces,
(16)
where N =
1 + X,
(17)
X = euVu ® Vu + ^2 ei^Pi ® ^Pi + e v V< £ ® V(/?.
(18)
x
'
i=l
and
i=l
Here, the quantity 1 is the unit tensor and A A denotes a tensor product having Cartesian components Aa and Ap. For the special case of a classical system, N = (p/T)l, where p is the pressure and T is constant, so Eq. (16) becomes Vp = 0. 2.2
Dynamical Balance Laws
We write integral balance laws for species, momentum, and total energy for a fixed but arbitrary subvolume V having area A as follows: —
/ pvd3x+ JV
Pid3x+
piVi-nd2x=
I
/ pvv -hd2x = I &-hd2x+ JA JA
nd3x,
/ V^p,g,d3 ^vi^t
(19)
152
R. F. Sekerka & Z. Bi
— / {u+-pv2)d3x+
{(u+-pv2)v+jh]-nd2x
= / v-&-hd2x+
^pivi-gid3x,
(21) where v* is the velocity of species i, Ti is the rate of production of species i by chemical reaction, gj is the external force per unit mass acting on species i, and & is a stress tensor.6 The quantities ji are diffusive fluxes of the species relative to the local center of mass. Since mass is conserved during chemical reactions, YHt=i r i = 0 and summation of Eq. (19) over all species leads to an overall conservation of mass with the barycentric velocity defined by v := ^ * = 1 ( p i / p ) v j . The quantity j ^ is an energy flux relative to the local center of mass. It is given a subscript h rather than u because in a classical treatment (no gradient corrections to the local densities) it is actually an enthalpy flux given by j / , = q c + X}f=i # t j i i where q c is the heat flux due to conduction and Hi are partial specific enthalpies. 0 The classical internalenergy flux would be j M = q c + Y2i=i Uiiii where Ui are partial specific internal energies. For a pure material, ji = 0, and there is no distinction between j ^ and j u . These integral balance equations can be converted to differential equations in the usual way by taking the time derivatives inside the integrals as partial derivatives, using Gauss's divergence theorem, and noting that the volume is arbitrary. The results, in Eulerian form, are: ^ + V - ( / 9 l v + j I ) = ri)
^) \pv2)
d(u+
di
+ V
.(pvv-
1
+ V'
f t g i
(22)
,
(23)
2^
One can take the dot product of Eq. (23) with v to obtain an equation for the mechanical energy balance that can be used to simplify Eq. (24), leading to the energy balance equation du + V - ( u v + j f t ) = o-:Vv + ^ j i . g i . ~dt
(25)
i=i
The notation for the first term on the right-hand-side of Eq. (25) is defined as
a,/3=l
X0
where a and /? are Cartesian indices. fc The sign convention for a is such that a = —pi, with 1 the unit tensor, for hydrostatic pressure Pc de Groot and Mazur [81] use the notation J 9 for j ^ and call it (page 18) a "heat flow," and then introduce (page 26) an alternative heat flux J'q equal to our q c . Fitts [82] uses the notation j e for jh and q for our q c , which he calls a "second law heat flux." The semantics are very confusing, but there is agreement regarding the relationships among corresponding variables.
Phase Field Model of Multicomponent
2.3
Alloy Solidification
153
Entropy Production
We substitute Eq. (25) for u and Eq. (22) for pi into Eq. (4) and integrate by parts using Gauss's divergence theorem to obtain 6 >
+
+
i[(r(^ &-)-g(?)"'
+
L
J s - euiNu - ^ tipiVpi - e^ipVip
hd2x
•
i=l / 1 \ NC
ft
NC
hd2x.
+/
(27)
This equation can be simplified by using the equations -j
ft
O
(28) <9
j=i
(29) i=l
and
"(?)-&'(f)>^© N
/
w
i=l
that would apply to a homogeneous system. In Eq. (28), T and pi are classical temperature and chemical potentials, respectively, defined as partial derivatives. Eq. (29) is an Euler equation that incorporates the classical pressure p, and Eq. (30) is a Gibbs-Duhem equation that can be obtained by comparing the differential of Eq. (29) with Eq. (30). As in classical non-equilibrium thermodynamics, we assume that Eqs. (28-30) can be applied locally for conditions that are not too far from equilibrium. We could proceed to use Eqs. (28-30) directly to simplify Eq. (27), but it is more expedient and useful for other purposes to transform them to other forms that relate to sNC, (l/T)NC, and (pi/T)NC. This can be done by rewriting them with the differential operator d replaced by the gradient operator V and then adding and subtracting appropriate terms to obtain
£) NC aNC
Vu
E
t=l
-E(f)
/Hi\NC
\~T)
VPl + V^-V-X, P
?
i+
T~
„•>
6u
V""V
u
e
T-r2
i = l iPiV Pi ~ I-*
(31)
154
R. F. Sekerka & Z. Bi
e u |Vu| 2 + J ^ eilVftl 2 + e ^ V y f
(32)
and NC
«v(I) - 5 > v ( f ) " c - W = -v(f) K
K
+ e u uV 2 (Vw) + Y,
£
ift V 2 (VpO - ev(X?2
(33)
i=l
Under equilibrium conditions such that Eqs. (13-15) hold, Eq. (31) becomes simply Eq. (16) obtained from Noether's theorem. Eq. (33) is a non-classical version of the Gibbs-Duhem equation. In Eqs. (32) and (33), we note the asymmetry between the non-classical (e u and ej) terms involving the conserved quantities, u and pi, and the e^ terms containing the nonconserved order parameter ip. This asymmetry persists and is reflected in the structure of the tensor Y (see Eq. (38)) to be defined later. Examination of Eq. (27) shows that parts of the expressions contained in Eqs. (32) and (33) occur there multiplied by v. Simplification of Eq. (27) can be accomplished most easily by reasoning that the total entropy flux will be given by
NC
s ^(~)
NC
i*-E(f)
NC
Du„ 3i + eu — V u
Dt
ft
+
1
DPi Vft + Dt
^Wtv^ (34)
where D
d
D~t
•v,
(35)
is the substantial derivative. This form was motivated in part by noting that it reduces to the classical result K
Js = sv + 7:7 + 2 , SiJi
classical.
(36)
Here, we have used j ^ = q c + Yli=i -^*J* an<^ Hi = Hi ~ TSi, where Si are partial specific entropies. The purely convective term s v is the obvious generalization of sv, and the terms involving D/Dt are generalizations of the non-classical surface terms containing d/dt in Eq. (27). This last generalization was suggested by the work of Anderson, McFadden, and Wheeler [66,71] who used a similar identification in order to obtain an entropy flux that had Galilean invariance in the presence of fluid flow. The physical interpretation of Wang et al. [19] of a non-classical entropy flux e^ifVtp for a rigid material was related to excess entropy carried into the control volume by motion of a diffuse interface with respect to that volume. With fluid flow, such entropy motion with respect to the control volume would require replacement by ev(Dp/Dt)Vif. We substitute Eq. (34) into Eq. (27), convert the remaining surface terms to volume terms by Gauss's theorem, and simplify (which requires considerable algebra)
Phase Field Model of Multicomponent
Alloy Solidification
155
to obtain the entropy production NC
v
+
NC
+ £;
\k-y[~
S,prod
+
XM? ®
tfx *)
NC
NC
Y: Vv
o?x,
(37)
where NC
1
Y :=
NC
f
P + < - - euuV2u - ] P tiPiV2Pi i=i
{
e u |Vu| 2 + j S i | V , 0 i | 2 + ejV¥>| 2
(38)
In the classical limit, this entropy production is in agreement with that calculated by de Groot and Mazur [81], Fitts [82], and Landau and Lifshitz [83]. Note in Eq. (38) the asymmetry in the non-classical terms involving u and pi (which relate to conserved quantities) as opposed to ip which is essentially a non-conserved order parameter. 2-4
Constitutive Relations and Equilibrium
The terms in the first line of Eq. (37) represent the entropy production due to the energy flux j ^ and the mass fluxes j , . Since the mass fluxes obey 52i=i Ji = 0> w e can eliminate one of them, say j K , and write the volumetric entropy production due to fluxes in the form NC
Jh-V
re-1
+ i=l£
T
_ It
-V
JVC
/ ,
\NC
+ (yj
(gi-g«)
(39)
According to Curie's theorem [82], these vector fluxes jo := jh and jj can couple to the vector driving forces f0 := V(1/T) J V C and f; := -V[(/z» - [iK)/T}NC + (l/T)NC(gi — gK). If we assume linear isotropic constitutive relations, then «-i
i = 0 , 1 , 2 , . . . ,K — 1.
(40)
3=0
Here, B is a positive definite matrix with elements Bij, where i,j = 0 , 1 , 2 , . . . , K—1. For the classical case, this matrix has been shown [84] by Onsager's method to be symmetric. At equilibrium, these driving forces must vanish, which requires TNC to be uniform throughout the system, in agreement with Eq. (13), and
c c c wc -v(§r fir ^-v(|f (i) + gK rV + T
(41)
156
R. F. Sekerka & Z. Bi
Further simplification of Eq. (41) can be made after we consider mechanical equilibrium. The terms in the second line of Eq. (37) represent the entropy production due to phase transformation, viscous dissipation, and chemical reaction, and Curie's theorem allows their constitutive laws to be coupled [82]. For simplicity, we shall assume them to be uncoupled. The chemical reaction term is the same as in the classical case except that (pi/T)NC replaces pi/T. It can be further rewritten in terms of progress variables for chemical reactions [81] but we omit details because they are the same as in the classical case. If the phase field is assumed to obey a linear constitutive equation, it evolves according to Dip
-^
= MVSV,
(42)
where Mv is positive. For equilibrium, Sv = 0, in agreement with Eq. (15). This leaves the entropy production due to viscous dissipation, which can be made positive by choosing Y to have the form
^-(fe + S) + ( K -l"'^ V v '
(43)
where p > 0 is the shear viscosity and K > 0 is the bulk viscosity [85], which is frequently ignored. Then
a.0
'dva dxfj
dvp\ dxa)
n
_ 2 3
+
K(V-v)z,
(44)
which is a sum of squares, and therefore positive. In equilibrium, Y = 0 and from the momentum equation (23), V<x =
-^pigi.
(45)
i=l
The quantity Yli=i PiSi —'• f i s the total external force per unit volume. Thus, for equilibrium, the divergence of Eq. (38) leads to NC
v(f)+v. / 'it \
z=l
v
'
V ^ euuV2u + jNiftVV i=l K
K
V ( I ) - euuV\Vu)
- ^ e i P i V 2 ( V p 8 ) + e „ ( V V ) V ^ (46) i=i
Moreover, for equilibrium, Eq. (33) becomes NC
I > v ( ^ ) i=l
= - v ( | ) + e „ U V 2 ( V U ) + E ^ V 2 ( V P l ) - e v ( V V ) V ^ . (47) t=l
Phase Field Model of Multicomponent Alloy Solidification 157
Adding Eqs. (46) and (47), we obtain JVC
(~)
n^NC + ( 7F I Si = 0. T
(48)
In view of Eq. (41), every coefficient of pi in Eq. (48) is the same and can be factored out of the sum; then, since YLl=i Pi = P 7^ 0, we obtain
-V(|)
+^-J
g, = 0
* = 1,2
«.
(49)
Thus, for equilibrium, the individual diffusion driving forces f\ = 0, for all i — 1,2,..., K, even though their conjugate fluxes ji are not independent. This is the generalization of Eq. (14) when external forces are included. Finally, Eq. (46) for equilibrium can be written
(
1 \
NC
K
r)
ftot + euUV2(Vu) + ^ ejPi V 2 (V Pl ) + ~VV,
i J
(50)
where we have used S^ = 0 to eliminate e^V 2 ^. For no external forces, eu = 0 and 6; = 0, T is constant and Eq. (50) reduces to Vp = Tds/dtp = —p(dum/d
Discussion
The driving forces for diffusive transport of energy and mass associated with Eqs. (39) and (40) are f0 = V(1/T) J V C ,
(51) NC
fi = - V [ ( M i - ^)/T}
NC
+ (l/T) (gi
- g K ).
(52)
These resemble the classical driving forces except that the temperature and chemical potentials are now functional derivatives instead of ordinary derivatives. This correspondence becomes even more interesting when the species specific external forces per unit mass, gi are conservative and can be expressed as gradients of a potential Vi(x)> i-e-i Si = ~^i}i- Then Eq. (49) for equilibrium becomes (since TNC is constant) -V(/xf C + Vi) = 0
i = l,2,...,K,
(53)
which integrates to give ^TOT
;= rfc + ipi = constant
i = 1,2,..., K. OT
(54)
Thus, one can define total chemical potentials, nf , that are uniform at equilibrium; they are the sum of the non-classical intrinsic chemical potential and the external potential. The equilibrium condition (54) can also be obtained directly from a variational principle, as in section 2.1, by considering the problem 6[S — \U{U + W) - Y11=\ \pi\ — 0) where the variation in the work is 8W = J ipi 8pi d3x. For a uniform gravitational field, ipi = gz independent of i,
158
R. F. Sekerka
& Z.
Bi
where z is measured vertically upward. In this case, g; — gK = 0 so that gravity does not appear explicitly in the forces f^ given by Eq. (52). It does, however, appear explicitly in the equilibrium conditions expressed by Eq. (54). This paradox is resolved by noting that gravity affects the pressure (see Eq. (50) for equilibrium) on which the p^c depend. For electrical forces due to an external electric field E = — V>, we would have ipi = — {zi/m^J7^, where J- = 96500 coulombs/mol is the Faraday constant, Zi is the valence (signed number of electronic charges |e|) for the species i, and mi is the molecular weight of species i. In this case, gj — gK would usually be nonzero for different ions, and there would be a direct effect of external forces on diffusion. These considerations can be used to extend the Nernst equation [86], used extensively in the neurobiological study of membranes, to the realm of phase field models. We can partially isolate the effect of TNC on diffusive transport by rewriting the associated entropy production given by Eq. (39) in the form t , \ JVC
qs y
' {f)
,
+
1
{f)
v NC
K-1
EM- v (^-^) T O r ]>
(55)
where K
NC
qs=h-J2v ii>
(56)
is a heat flux [83]. Then linear constitutive relations, similar to Eq. (40), could be postulated to relate qs and jj to the driving forces fo and — V(/ij — pK)TOT. In the classical limit, q s = q c + TSiji, which relates closely to the diffusive part of the classical entropy flux given by Eq. (36). Note, however, that the chemical potentials in Eq. (56) still depend on temperature. For our multicomponent model, is also possible to identify the surface free energy 7 by considering variations of a system having a planar interface. This is taken up in Appendix B where it is shown (see Eq. (85)) that 7 is the surface excess of the Kramers potential, in agreement with its classical definition. This allows one to obtain a simple general formula (see Eq. (88)) for 7 in terms of integrals over the squares of the gradients of interface profiles. In comparing the present model to other alloy phase field models, it is important to bear in mind that even without anisotropy, the form of the gradient entropies can lead to subtle differences in results. In a binary alloy, for example, the portion of our gradient entropy associated with the partial-mass densities contains e1\VPl\2 + e2\Vp2\2.
(57)
If we express this in terms of the mass fractions ui and w2 we obtain (ei + e 2 )p 2 |Vw 2 | 2 + 2/o(e2w2 - eiWi)Vp • Vw2 + (eiw2 + e 2 o;|)|Vp| 2 .
(58)
If p were a constant, the latter two terms would vanish and we would just have a constant times | Vw 2 | 2 , as we used in [40] (although there in terms of mole fractions). But if p is not a constant, as in the binary-alloy theory of Anderson, McFadden, and Wheeler [66], we can eliminate the cross term in Eq. (58) only by allowing e\
Phase Field Model of Multicomponent
and e2 to be variables. Indeed, if we choose t\ = u2e/p2 Eq. (58) becomes e|V W2 | 2
+
^|Vp| P
2
Alloy Solidification
159
and e2 = cuie/p2, then
,
(59)
which contains a mandatory coefficient of |Vp| 2 . Therefore, one cannot pass easily from one theory to the other unless care is taken about the precise form of the gradient-entropy corrections, including non-constant coefficients. Since phenomenologically there is no way to decide which of the many possibilities is correct or more meaningful (e.g., constant coefficients on a per unit mass basis versus a per unit volume basis, partial densities versus compositions, cross terms or not) one must ultimately appeal to microscopic considerations or to experiments. Acknowledgments The authors are grateful for discussions with G. B. McFadden. This work was done under support from NSF grant DMR9634056. Appendix A We proceed to derive Eqs. (16-18), which are a consequence of Noether's theorem [69] in three spatial dimensions (three independent variables) but for several dependent variables. In the interest of a compact notation, we denote the spatial variables x, y, z by xa, where a = 1,2,3, and the independent variables u, pi, ip by u1, where i = 0,1, • • • K, K + 1. Then the equilibrium problem of Eq. (12) can be written in the schematic form 8 J = 0, where J=
/ I{xa,u\1—). JV
OX
(60) a
In our particular case, K
I = SNC -XuU-^XiPJ, i=\
(61)
and there is no explicit dependence on xa, but we retain the more general notation until we specialize to this case. We employ the summation convention on repeated indices until further notice. The variational problem becomes d 9I dI f dI , u 3 , f x i J2 dul — dxp dul \r- bu a x + / —T r-dung ax. / TT-^ — Q (7 Jv du% dx0 Q (dv>\ JA d (&ui) The Euler-Lagrange equations are therefore
(62)
d{§£)
dl du
l
d
dl
dxg a I diS \
(63)
160 R. F. Sekerka & Z. Bi
Multiplying Eq. (63) by dul/dxa,
summing on i, and simplifying the result by using
dl —
=
)
. 9xa I ,
dl dv} •
dxa
dl
d2u
d
dxadxp'
•
l
du
dxa
(%)
:
(64)
where the partial derivative on the left is with respect to the variable set x, y, z, we obtain dl dxa
( dl \dxa
dul dl dxn« / d \dx0
_8_ dxp
If I is explicitly independent of Eq. (65) vanishes and we obtain
(65) )
the case for Eq. (61), the first term in
^—Nap dxp
(66)
= 0,
where the tensor K+l
Nap — I5ap - 22
dv?
dl
^o^xjg^y
(67)
and we have restored the summation sign for emphasis. Eqs. (66) and (67) can also be written in the form V • N = 0,
(68)
where K+ l
N = Jl-J^A* i=0
dl dAi
(69)
where 1 is the unit tensor, ® denotes the tensor product, and A 1 = V u \ Eqs. (68) and (69) are the desired generalization of Noether's theorem to several dependent variables. For i" given by Eq. (61) we obtain N
sNC - Xuu - ^
XiPi 1 + X,
(70)
i=l
where X = e u Vu ® Vu + ^2 CiVpi <S> Vft + evVy ® Vtp.
(71)
Replacing the Lagrange multipliers by their values given by Eqs. (13) and (14), we obtain Eqs. (16-18) of the text. The tensor
X s " := X - 1 I eu\Vu\2 + ^ £ l | V f t | 2 + ^|V^| 2 J 1
(72)
is traceless and so represents a pure shear, so X itself contains a non-classical hydrostatic pressure.
Phase Field Model of Multicomponent
Alloy Solidification
161
Appendix B In this Appendix, we examine variations of a planar system at equilibrium to identify and obtain a formula for 7, the surface free energy per unit area (surface tension) associated with a diffuse interface. We exclude chemical reactions and external forces and focus on a system with properties that depend on a single spatial variable x, where — L < x < L. We assume that the system is in equilibrium and contains a diffuse interface of width ~ w « L, localized near x = 0. We therefore focus on the limit L —* 00, but retain finite L for intermediate steps. We begin with the entropy functional given by Eq. (1) and with total masses and total internal energy given by Eqs. (8) and (10), but we allow the volume, which we now call V, to vary in a manner such as to preserve the planar nature of the system. These variations include: t y p e 1 The cross section of the system having area A remains unchanged and the positions of its ends, originally at x = ±L, are varied. type 2 The ends remain fixed at x = ±L and A is changed to A + 6A uniformly in every cross section, independent of x. In a variation of type 1, the system can do work by changing its volume; whereas, in a variation of type 2, it can do work by changing its volume and by changing its cross sectional area, A. The equilibrium condition for such variations is K
6{S -\UU-Y,
KMi - XVV - XAA) = 0,
(73)
i=l
where Au, A^, Ay, and A^ are Lagrange multipliers. As we saw in connection with Eq. (12), \ u = (l/T)NC and A; = (p-i/T)NC will be uniform at equilibrium, and Sv = 0, as given by Eqs (13-15). We shall relate Ay to pressure and XA to 7. Under the above equilibrium conditions, 6S = ( I ) N° Jv Sud3x - £
(^)NC
^ 6Pl d3x
(74)
i=l
+ f sNC6Nd2x-
J eu6uVu + ^2 CiSpiVpi + e^SipVf hd x, JA JA »=i where 6N is the distance of variation along the outward normal and where we have factored out the constants Au = (1/T)NC and Aj = (fii/T)NC. For variations of type 1 and type 2, the second surface integral vanishes; for type 1, the ends of the system are so far from the diffuse interface that inhomogeneities in u, pi, and ip are negligible (this becomes rigorous as L —> 00) and for type 2 variations, Vw, Vp», and V
I u6Nd2x,
(75)
JA
and 6Mi = I 6pid3x+ I pt8Nd2x, Jv JA
(76)
162
R. F. Sekerka & Z. Bi
Eq. (74) becomes (77) 8Nd2i v
'
i=i
Eq. (73) therefore becomes JVC
1*°-$)
^%{%r^
SNd2x-XASA
= 0.
(78)
For a variation of type 1, 6A = 0 and Eq. (78) yields 7VC
WC
(79)
= AV. x=±L
From Eq. (32) for a planar system, we obtain NC
d2u dx2
NC
T
2=1
du
v^ i=i
d2p. dx 2
(80)
2
dpi
+£< i=l
• ^
As L —• oo, only the term p/T on the right-hand-side of Eq. (80) survives. We denote this pressure by p^, and noting also that T —» TNC as L —* oo, Eq. (79) for L —> oo yields NC
\v = (4
(81)
Poo-
We now identify the remaining Lagrange multiplier as XA — -ry/TNC, Eq. (73) to be rewritten in the familiar form SU = TNC 8S-Voo8V
+ Yl ^ °
6M
i +7
8A
-
which allows
(82)
Then for a variation of type 2, Eq. (78) for L —> oo yields
c
/:Hr-ip~ *-(r+-(r^ (83) Eq. (83) can be recast into a more familiar form by defining a non-classical Kramers potential per unit volume .NC ..
u-TNCsNC-Y.^CPi=\
(84)
Phase Field Model of Multicomponent
Alloy Solidification
163
Then [coNC - ^c]
7 =
dx,
(85)
where we have used Eqs. (80) and (84) to deduce that w ^ c = UJNG{X —> ±oo) = —Poo. Eq. (85) resembles its classical counterpart, i.e., 7 is the surface excess Kramers potential. A more useful formula for evaluating 7 can be obtained by using the equation for mechanical equilibrium which we can get from Eq. (38) by setting Y — 0, going to one dimension," and noting that axx must be uniform throughout x, and therefore a constant. We obtain JVC
IV rp J
V_- e„,u^^ T "uddx*
°~XX
+
-
1
+
Tx)
(86)
^^iP
^
dpi dx
(
Mt
where the sign reversal on the second line, relative to Eq. (80), arises from the xx component of X. By means of Eqs. (80), (84), and (86), we obtain -.JVC
NC
du dx
+£•
dpi
'dip
dx
^^Tx
(87)
Substitution of Eq. (87) into Eq. (85) then yields 2
7
K
/
,
"°Q-.{£)+p@
= r-
\ 2
rfpiV
J
[dip +€lp
\dx
dx.
Eq. (88) shows that for the present model, 7 is entirely entropic in origin and that the gradient entropies contribute additively. It turns out to be -TNC times twice the surface excess of S. References 1. 2. 3. 4. 5. 6. 7. 8.
J. W. Cahn, Acta. Met. 8, 554 (1960). B. I. Halperin, P. C. Hohenberg, and S.-K. Ma, Phys. Rev. B 10, 139 (1974). J. S. Langer, unpublished notes (August 1978). G. J. Fix, in Free Boundary Problems: Theory and Applications, Vol. II, eds. A. Fasano and M. Primicerio, Pitman, Boston, p. 580 (1983). J. B. Collins and H. Levine, Phys. Rev. B 31, 6119 (1985). J. S. Langer, Directions in Condensed Matter Physics, eds. G. Grinstein and G. Mazenko, World Scientific, Singapore, p. 165 (1986). G. Caginalp, in Applications of Field Theory to Statistical Mechanics, ed. L. Garrido, Lecture Notes in Physics, Vol. 216, Springer, Berlin, p. 216 (1985). G. Caginalp, in Material Instabilities in Continuum Problems and Related Mathematical Problems, ed. J. M. Ball, Oxford University Press, Oxford, p. 35 (1988).
164
9. 10. 11. 12. 13. 14. 15.
16. 17. 18. 19. 20. 21. 22. 23. 24.
25. 26. 27. 28. 29. 30. 31. 32.
33. 34. 35. 36.
37. 38.
R. F. Sekerka & Z. Bi
G. Caginalp and P. C. Fife, Phys. Rev. B 33, 7792 (1986). G. Caginalp, Arch. Rat. Mech. Anal. 92, 205 (1986). G. Caginalp, Ann. Physics (NY) 172, 136 (1986). G. Caginalp and P. C. Fife, SIAM J. Appl. Math. 48, 506 (1988). G. Caginalp, Phys. Rev. A 39, 5887 (1989). G. Caginalp and E. A. Socolovsky, J. Comp. Phys. 95, 85 (1991). G. Caginalp and X. Chen, in On the Evolution of Phase Boundaries, eds. M. E. Gurtin and G. B. McFadden, The IMA Volumes in Mathematics and Its Applications, Vol. 43, Springer-Verlag, Berlin, p. 1 (1992). E. Fried and M. E. Gurtin, Physica D 68, 326 (1994). E. Fried and M. E. Gurtin, Physica D 72, 287 (1994). 0 . Penrose and P. C. Fife, Physica D 43, 44 (1990). S.-L. Wang, R. F. Sekerka, A. A. Wheeler, B. T. Murray, S. R. Coriell, R. J. Braun, and G. B. McFadden, Physica D 69, 189 (1993). R. Kobayashi, Bull. Jpn. Soc. Ind. Appl. Math. 1, 22 (1991). R. Kobayashi, Physica D 63, 410 (1993). R. Kobayashi, in Pattern Formation in Complex Dissipative Systems, ed. S. Kai, World Scientific, Singapore, p. 121 (1992). A. A. Wheeler, B. T. Murray, and R. J. Schaefer, Physica D 66, 243 (1993). B. T. Murray, W. J. Boettinger, G. B. McFadden, and A. A. Wheeler, in Heat Transfer in Melting, Solidification, and Crystal Growth, eds. I. S. Habib and S. Thynell, ASME HTD-234, New York, p. 67 (1993). B. T. Murray, A. A. Wheeler, and M. E. Glicksman, J. Cryst. Growth 154, 386 (1995). S.-L. Wang and R. F. Sekerka, Phys. Rev. E 53, 3760 (1996). S.-L. Wang and R. F. Sekerka, J. Comp. Phys. 127, 110 (1996). N. Provatas, N. Goldenfeld, and J. A. Dantzig, Phys. Rev. Lett. 80, 3306 (1998). N. Provatas, N. Goldenfeld, and J. A. Dantzig, J. Comp. Phys. 148, 265 (1999). Y. T. Kim, N. Provatas, N. Goldenfeld, and J. A. Dantzig, Phys. Rev. E 59, 2549 (1999). N. Provatas, N. Goldenfeld, J. A. Dantzig, J. C. LaCombe, A. Lupulescu, M. B. Koss, M. E. Glicksman, and R. Almgren, Phys. Rev. Lett. 82, 4496 (1999). N. Provatas, N. Goldenfeld, and J. A. Dantzig, in Modeling of Casting, Welding, and Advanced Solidification Processes, eds. B. G. Thomas and C. Beckermann, TMS, San Diego, p. 533 (1998). A. Karma and W.-J. Rappel, Phys. Rev. E 53, 3017 (1996). A. Karma and W.-J. Rappel, Phys. Rev. Lett. 78, 4050 (1996). • A. Karma and W.-J. Rappel, Phys. Rev. E 57, 4323 (1998). A. A. Wheeler and W. J. Boettinger, in On the Evolution of Phase Boundaries, eds. M. E. Gurtin and G. B. McFadden, The IMA Volumes in Mathematics and Its Applications, Vol. 43, Springer-Verlag, Berlin, p. 127 (1992). A. A. Wheeler, W. J. Boettinger, and G. B. McFadden, Phys. Rev. A 45, 7424 (1992). A. A. Wheeler, W. J. Boettinger, and G. B. McFadden, Phys. Rev. E 47, 1893
Phase Field Model of Multicomponent
Alloy Solidification
165
(1993). 39. W. J. Boettinger, A. A. Wheeler, G. B. McFadden, and R. Kobayashi, in Modeling of Coarsening and Grain Growth, eds. C. S. Pande and S. P. Marsh, TMS, Warrendale, PA, p. 45 (1993). 40. Z. Bi and R. F. Sekerka, Physica A 261, 95 (1998). 41. Ch. Charach and P. C. Fife, SIAM J. Appl. Math. 58, 1826 (1998). 42. Ch. Charach and P. C. Fife, Open Systems, Information Dynamics 5, 99 (1998). 43. W. J. Boettinger, A. A. Wheeler, B. T. Murray, and G. B. McFadden, Mater. Sci. and Eng. A 178, 217 (1994). 44. M. Conti, Phys. Rev. E 55, 701 (1997). 45. M. Conti, Phys. Rev. E 55, 765 (1997). 46. N. A. Ahamad, A. A. Wheeler, B. J. Boettinger, and G. B. Mcfadden, Phys. Rev. E 58, 3436 (1998). 47. Ch. Charach, C. K. Chen, and P. C. Fife, J. Stat. Phys. 95, 1141 (1999). 48. Ch. Charach and P. C. Fife, J. Cryst. Growth 198/199, 1267 (1999). 49. J. A. Warren and W. J. Boettinger, Acta Metall. Mater. 43, 689 (1995). 50. M.. Conti, Phys. Rev. E 56, 3197 (1997). 51. J. A. Warren and W. J. Boettinger, in Modeling of Casting, Welding, and Advanced Solidification Processes VII, eds. M. Cross and J. Campbell, TMS, Warrendale, PA, p. 601 (1995). 52. J. A. Warren and W. J. Boettinger, in Solidification Processing 1997, eds. J. Beach and H. Jones, Department of Engineering Materials, University of Sheffeld, UK, p. 422 (1997). 53. J. A. Warren and B. T. Murray, Mod. Simul. Mater. Sci. Eng. 4, 215 (1996). 54. W. J. Boettinger and J. A. Warren, Met. Trans. A 27, 657 (1996). 55. W. J. Boettinger and J. A. Warren, J. Cryst. Growth 200, 583 (1999). 56. A. Karma, Phys. Rev. E 49, 2245 (1993). 57. A. A. Wheeler, G. B. McFadden, and W. J. Boettinger, Proc. R. Soc. Lond. A 452, 495 (1996). 58. K. R. Elder, J. D. Gunton, and M. Grant, Phys. Rev. E 54, 6476 (1996). 59. G. Caginalp and J. Jones, App. Math Lett. 4, 97 (1991). 60. G. Caginalp and J. Jones, in On the Evolution of Phase Boundaries, eds. M. E. Gurtin and G. B. McFadden, The IMA Volumes in Mathematics and Its Applications, Vol. 43, Springer-Verlag, Berlin, p. 27 (1992). 61. R. Tonhardt, Doctoral Thesis, Department of Mechanics, Royal Institute of Technology, Stockholm, Sweden (1998). 62. R. Tonhardt and G. Amberg, J. Cryst. Growth 194, 406 (1998). 63. C. Beckermann, H.-J. Diepers, I. Steinbach, A. Karma, and X. Tong, J. Comp. Phys. 154, 468 (1999). 64. H. J. Diepers, C. Beckermann, and I. Steinbach, in Solidification Processing 1997, eds. J. Beach and H. Jones, Department of Engineering Materials, University of Sheffeld, UK, p. 426 (1997). 65. M. E. Gurtin, D. Polignone, and J. Vihals, Math. Models. Methods Appl. Sci. 6, 815 (1996). 66. D. M. Anderson, G. B. McFadden, and A. A. Wheeler, Ann. Rev. Fluid Mech. 30, 139 (1998).
166
R. F. Sekerka & Z. Bi
67. D. M. Anderson and G. B. McFadden, Phys. Fluids 9, 1870 (1997). 68. D. J. Korteweg, Arch. Need. Sci. Exactes Nat. Ser. II 6, 1 (1901). 69. R. Courant and D. Hilbert, Methods of Mathematical Physics, Vol. I, Interscience, New York, p. 262 (1953). 70. A. A. Wheeler and G. B. McFadden, Proc. R. Soc. Lond. A 453, 1611 (1997). 71. D. M. Anderson, G. B. McFadden, and A. A. Wheeler, Physica D 135, 175 (2000). 72. D. M. Anderson, G. B. McFadden, and A. A. Wheeler, "A Phase-Field Model of Solidification with Convection: Numerical Simulations," in this volume. 73. F. C. Larche and J. W. Cahn, Acta. Metall. 21, 1051 (1973). 74. F. C. Larche and J. W. Cahn, Acta. Metall. 26, 53 (1978). 75. F. C. Larche and J. W. Cahn, Acta. Metall. 26, 1579 (1978). 76. W. W. Mullins and R. F. Sekerka, J. Chem. Phys. 82, 5192 (1985). 77. Z. Bi and R. F. Sekerka, "Model for Solidification of a Pure Material with Energy as the Phase-Field Variable," unpublished notes based on seminars at Philadelphia and Edinburgh (1997). 78. H. W. Alt and I. Pawlow, in International Series on Applied Mathematics, Vol. 95, Birkhauser Verlag, Basel, p. 1 (1990). 79. H. W. Alt and I. Pawlow, Physica D 59, 389 (1992). 80. H. W. Alt and I. Pawlow, Adv. Math. Sci. Appl. 6, 291 (1996). 81. S. R. de Groot and P. Mazur, Non-Equilibrium Thermodynamics, Dover Edition, New York (1984). 82. D. D. Fitts, Non-Equilibrium Thermodynamics, McGraw-Hill, New York (1962). 83. L. D. Landau and E. M. Lifshitz, Fluid Mechanics, Pergamon Press, London (1959). 84. R. F. Sekerka and W. W. Mullins, J. Chem. Phys. 73, 1413 (1980). 85. R. B. Bird, W. E. Stewart, and E. W. Lightfoot, Transport Phenomena, John Wiley & Sons, New York, pp. 79-82 (1960). 86. G. C. Matthews, Cellular Physiology of Nerve and Muscle, Blackwell Scientific, Boston, p. 27 (1991).
T H E EFFECTS OF A S T R E S S - D E P E N D E N T MOBILITY ON INTERFACIAL STABILITY P. W . V O O R H E E S Department Northwestern
of Materials Science and Engineering, University, Evanston, IL 60093, U.S.A. M. J. AZIZ
Division of Engineering and Applied Sciences, Harvard University, Cambridge MA, 02138, U.S.A. A linear morphological stability analysis of a planar interface in interface reactioncontrolled growth is presented. We allow the mobility of the interface to be a function of the elastic stress, and find that the stress-dependent mobility can be either stabilizing or destabilizing. We find that a stress-dependent mobility can be the dominant factor leading to instability at small stresses. The predictions are compared to the recent experiments by Barvosa-Carter et al.ln agreement with their conclusions, we find that the experiments were performed in a parameter range where the stress-dependent mobility is the dominant cause of instability. The prediction of the critical wavenumber and amplification rate of the instability is consistent with the experimental results.
1
Introduction
The stability of interfaces in elastically-stressed solids has been the subject of great interest recently. A reason for this is the connection between the stability of a surface of a thin film that is deposited on a substrate and the subsequent evolution of the film morphology. It is clear that an elastic-stress-driven morphological instability results in a planar film undergoing a transition to islands; for a review, see [1]. Controlling this instability is thus central to the production of novel electronic and optoelectronic devices because the properties of such devices depend upon the morphology of the film. For example, under some circumstances nanoscale islands called "quantum dots" form that can exhibit unique electronic properties as a result of carrier confinement within the small volume of these dots. However, devices with well-characterized properties require that the size of the dots be relatively uniform. To date, much of the work has been focused on the interfacial instability that is driven by the elastic energy in the system. This instability, identified by Asaro and Tiller [2] (AT) and many others [3-6], follows from the dependence of the chemical potential on the elastic energy density. Consider a solid under uniaxial stress in contact with a fluid. Although the stress in the unperturbed or basic state is uniform, this is not the case when the interface is nonplanar. When the surface is perturbed the stress is concentrated at the troughs, in analogy with the stress concentration of a crack, and relieved at the peaks. Because the chemical potential is linearly proportional to the elastic energy density, the chemical potential of an atom is higher at the trough than at the peak and atoms thus move from the trough to the peak tending to make the peaks higher and the troughs deeper. Opposing this, of course, is the surface energy that tends to move atoms in the opposite direction and, thus, leads to a planar surface. The stress concentration in the
167
168
P. W. Voorhees & M. J. Aziz
trough appears in both compression and tension, and thus the AT instability is independent of the sign of the applied stress. Given the basic mechanism, the development of the elastic-energy-driven, or AT, instability must also depend on the mode of mass transport. Thus, the elastic-energy-driven instability has been examined in the context of bulk diffusion [2], surface diffusion [4-6], and evaporation-condensation [4, 7]. The evaporationcondensation model is, essentially, a completely kinetically-controlled interface in which the rate-limiting mass transport process is atom motion across an interface between a solid and another solid, a vapor, or even a liquid. In this case the velocity of the interface is linearly proportional to the difference between the chemical potential at the surface and that in the parent phase. The type of mass transport process does not alter the critical wavenumber of the instability, it alters only the dependence of the growth rate of the instability on the wavenumber. The elastic-energy-driven instability in alloys has also been examined. In an alloy the chemical potential couples to both the stress and the composition. Moreover, as stresses can result from gradients in composition in the solid, the composition and stress are also coupled through the continuum mechanics. As before, however, the evolution of the surface occurs by gradients in the chemical potential for surfacediffusion-driven mass transport [8-11], or by the difference between the chemical potential and that in a fluid for evaporation and condensation mass transport [12]. These compositionally-generated stresses can act to either enhance or prevent the usual AT instability [9,11,12]. In addition, compositionally-generated stresses can cause an instability even in the absence of an applied stress [10-13]. All of this work, however, assumes that the relevant transport coefficient, the surface diffusion coefficient for surface-diffusion-limited interface motion, or the interface mobility for interface-reaction-limited interface motion, is independent of elastic stress. Atomistic simulations have shown that the surface diffusion coefficient can be linearly proportional to the elastic stress on the surface. However, recent experiments by Barvosa-Carter et al. (BC) have shown that a stress-dependent interfacial mobility alone can generate an interfacial instability when the transformation is interface limited [14]. The mechanism for the kinetic instability follows from the observation that the change in the mobility with stress in certain cases can be linearly proportional to the stress. Consider a basic state consisting of a moving planar interface under uniform uniaxial stress. Assume that the stress is compressive, i.e., negative, and that the mobility increases with increasingly positive stress. Then at the troughs, where the stress is more negative than that of the basic state, the mobility is smaller than that of the planar interface. Conversely at the peaks, where the stress is larger than that of the basic state, the mobility is larger than that of the planar interface. Thus, troughs grow slower and peaks grow faster than the moving planar interface. In this case the stress-dependent mobility is destabilizing. Moreover, experiments have shown that the amorphous to crystalline (a-c) transformation in Si is an ideal system in which to explore this kinetic instability. The stress dependence of the interfacial mobility has been measured [15]. The mobility, M, was found to depend upon stress, er, as
M(aij) =
Mexp(Vi*aij/kT),
(1)
The Effects of a Stress-Dependent
Mobility on Interfacial
Stability
169
where V* is the activation strain [15], M is the mobility of the interface in the absence of stress, k is Boltzmann's constant and T is temperature. For the Si[001] a-c interface, V^ = V2*2 = 0.14Q and V33 = — 0.35f2, where fi is the atomic volume, and the off-diagonal elements are zero. The transformation is completely interface-reaction limited, and the other relevant material parameters have been measured. It is possible to carefully control the morphology of the initial interface using lithography. Finally, because, V^ > 0 a compressive uniaxial stress <m < 0, should result in an interfacial instability. Such an instability was observed by BC. BC analyzed the instability using a fully nonlinear numerical approach. This approach has the advantage of being able to follow the instability into the strongly nonlinear, large-amplitude regime and confirmed that the instability was, in fact, responsible for the nonplanar morphologies observed in the experiments. However, because the analysis was numerical, it was difficult to establish the relationship between the various material and processing parameters that control the instability. Grinstein, Tu, and Tersoff [16] examined analytically the effects of a stressdependent mobility in the context of thin-film oxidation. The motion of the interface in this case is controlled by both interface mobility and long-range diffusion. Given the complexity of the problem, however, they did not focus on the reactionlimited case, and their treatment of long-range diffusion is controversial [17,18]. Yu and Suo have recently considered the reaction-limited case as well, focusing on the two-dimensional patterns that might be formed on the interface [19]. We present a linear stability analysis of a planar interface where the motion of the interface is interface-reaction limited. We allow for a stress-dependent interfacial mobility and examine its effect on the stability of the interface. Using the material and processing parameters of the BC experiments, we make a comparison between theory and experiment with no adjustable parameters. However, we assume that the amorphous phase into which the crystal grows is stress free. Extrapolation of the known [20] viscous-blow behavior of a-Si indicates that although there is a significant amount of relaxation in the BC experiments, it does not go to completion in the experimental time scales. We also assume M to be independent of crystallographic orientation. Although the mobility of the [001] interface is known to be 20 times faster on [001] than on [111], M at [001] is an extremum due to crystal symmetry. Nevertheless, we shall neglect this orientation dependence. The effects of growth kinetic anisotropy and flow in the a-Si layer on the instability are currently under investigation. 2
Theory
The crystal grows into the amorphous phase at a constant velocity v in the positive ^-direction, where X3 is parallel with the [001] direction. All quantities except stress and activation strain are assumed to be isotropic. The motion of the a-c interface is kinetically limited, thus in the limit of small driving forces, v is given by [14,15] v = -M
(a)
[AG°V + WpV
+ W (e) +
7K]
,
(2)
where M (er) is the stress-dependent interface mobility, cr is the stress in the small-
170
P. W. Voorhees & M. J. Aziz
strain approximation, the chemical driving force for crystallization AG° is the Gibbs-free-energy change (AG° < 0) per unit volume of growing phase upon crystallization of stress-free material at a planar interface, Wpv is the pV work done on the system by external forces per unit volume crystallized, W (e) is the elastic strain energy, e is the small-strain tensor, 7 is the interfacial energy, and K is the mean curvature of the interface (reckoned positive for a spherical crystal). The elastic strain energy is assumed to be negligible in the parent phase and we will neglect Wpv- Since the transformation is isothermal and occurs in a pure material, the only field equation that must be solved is that for the elastic stress. The stress field is found by requiring that the crystal is in elastic equilibrium, ^ijd
(3)
~ ^!
where summation of repeated indices from 1 to 3 is assumed. The stress is linearly related to the strain, E £kk&ij%] +' 1 + v \ 1 - 2v
et j j
: ,
(4)
where E is Young's modulus and v is Poisson's ratio. The strain is related to gradients in the displacement Uj, e
ij = iui,j + u3,i) / 2 -
(5)
Using Eqs. (4) and (5) in Eq. (3) yields Navier's equations for the displacement field in the crystal, (1 - 2v) uiM + ukM = 0.
(6)
The a-phase is assumed to be stress-free, and thus along the a-c interface, (TijUj = 0,
(7)
where rij is the normal to the interface pointing into the a-phase. Finally, the elastic energy density is given by, W =
aijCij/2.
(8)
In the unperturbed or basic state, the interface is planar and moving at a constant velocity into the amorphous phase. The applied stress in the crystal is biaxial and of magnitude cr°x = a\2 = a a- The stress-free boundary condition (7), implies CTJ3 = 0 throughout.
To examine the linear stability of the moving planar interface located at £3 = 0, we perturb the height of the interface in the a?3-direction, h = h(x\,X2), and all other quantities in normal modes,
I 0 \ v Ui
M W K
\ J
M° W°
V 0 /
( h \ v1
+ 6 M1 l
W 1
V* /
*,
(9)
The Effects of a Stress-Dependent
Mobility on Interfacial Stability
171
where $ = exp (iq -x + rt), q and x are the wavevector and position vector, respectively, in the plane of the interface, <5 0 the perturbation grows, and if r < 0 the perturbation decays. If r is imaginary then an oscillatory instability is possible. Using Eq. (9) in Eq. (2) and setting 6 = 0, we obtain the basic-state, interfacial velocity, v° = -M°
(AG°V + W°) .
(10)
vl = -{AG°v + W°)Ml-M°Wl-M^^K1.
(11)
The perturbed velocity is
1
1
l
1
We now need to find M°, M , W°, W ,V , and K . The elastic-strain-energy density in the basic state follows from the stresses given above, W° = al/Y,
(12)
where Y = E/(l — v). The strain-energy density in the perturbed state is determined by solving Eq. (6) with the boundary condition (7), and the condition that the «H X 3) ""* 0 as x 3 —> —oo [21]. Using these displacements in Eq. (5), substituting into Eq. (4), and using Eq. (8) yields the magnitude of the perturbed elastic-strain-energy density along the interface [12], W1 = -2h{l
+ v)qal/Y,
(13)
where q = |q|. In addition, the magnitude of the trace of the stress is 4k
= -2h(l
+ u)qaa.
(14)
The mobility of the interface in the basic and perturbed states, M° and M1 respectively, can be found by using Eq. (1), M(tr)=M(l + ^ y
(15)
where we have assumed that aijV*j/kT <^C 1. Thus,
M«^( 1 + f2), crl.V* Ml = - ^ ,
(1„ (17)
where of- and a\, are the stresses in the basic and perturbed states, respectively. To determine a\^V*^ we first rewrite it as,
4 - ^ i = olkVn + (VS - Vx\) 4 , ,
(18)
because for an [001] surface of a cubic crystal, V{x = V2*2 [15]. From Eq. (7), we findCT33= 0. Using this and Eq. (14) in Eq. (18) yields o\jV% = -2h(l
+ u)V?1(raq.
(19)
172
P. W. Voorhees & M. J. Aziz
The stress in a system with a planar interface is 4Vii
=
2 y
( 2 °)
i>-
Using these stresses in Eqs. (16) and (17) yields M0 = M ( l + ^
)
,
(21)
and M 1 = -M2h{\
+ v)V^aaq/kT.
(22)
Thus, if V{i > 0, a compressive stress decreases the mobility of the planar interface. For compressive stresses the perturbed mobility M1 is in phase with the interface shape. Therefore, the mobility is the highest at the peaks and the lowest at the troughs. The remaining terms in Eq. (11) are functions only of the shape of the interface. Using the definition of the curvature in Cartesian coordinates, the form of the perturbed interface, and by taking the derivative with respect to time we obtain, K1
= hq2,
v1 = hr.
(23)
Using Eqns. (12-14) and (21-23) in Eq. (11) yields the dispersion relation, j j j = [(1 + ") aAG°vaa + 2pa2a + 3a/3a3a] q -{l
+ aaa)iq2,
(24)
where a = 2V1*1/feT and j3 = (1 + v)/Y. The dispersion relation is linear in r and thus the instability cannot be oscillatory. The wavenumber at which the growth rate of the instability is zero, qct is found by setting r = 0, (l + is)aAG0vaa + 2pa2a + 3aPa3a Qc=
77r ( l + QCT a )7
.
(25)
For q < qc, r > 0; for q > qc, r < 0, see Fig. 1. The wavenumber with the maximum growth rate, qm, is given by qm — qc/2. Thus, qc determines the length scale of the instability. 3
Discussion
The dispersion relation is quadratic in the wavenumber, q, with the terms related to the elastic stress being linear in q and the interfacial-energy-related terms going as q2. Interfacial energy is always stabilizing, because by assumption ctaV^/kT
The Effects of a Stress-Dependent
Mobility on Interfacial
Stability
173
0.002
-0.004
0.2
0.4
0.6
1.0
wavenumber (m~ ) Figure 1. Amplification rate of the instability as a function of the wave number of the disturbance for the a-c Si transformation in the BC experiments.
the usual case where the instability develops by surface diffusion. The term that is linear in the applied stress can be either stabilizing or destabilizing depending on the sign of the product aaa. If aaa < 0, as in the a-c-Si BC experiments, the stress dependence of the mobility itself can drive an interfacial instability. The term that is cubic in the applied stress is the AT instability modified by the stress-dependent mobility. Because aaa < 0 under compressive stresses for the Si[001] a-c interface, both the stress-dependent mobility and the elastic energy promote instability. As shown in Eq. (24), both effects have the same scaling with the wavenumber and thus it can be difficult to determine which phenomenon is responsible for an experimentallyobserved interfacial instability. However, the elastic-energy-driven instability scales with stress in a different manner than the mobility-driven instability. It is clear that for a sufficiently small applied stress, the stress-dependent mobility will dominate. When aaa < 0, this will occur when (1 + A g » ! (26)
^
g
Using the values for the a-c Si interface under the conditions employed in the BC experiments we find that the left-hand-side of the inequality (26) is approximately 12, and thus the instability of the interface is predicted to be due to the kinetic instability. Conversely, if aaa > 0, then the stress-dependent mobility is stabilizing. The critical stress, aca below which the stress-dependent mobility will stabilize the AT instability is found by setting the first term in the brackets of Eq. (24) to zero, a\ = - 1 / (3a) + [1 - 3a 2 (1 + u) AG°v/(3]1/2 / (3a).
(27)
For the a-c Si system, this yields a stress of 9.6 GPa. All tensile stresses below this value result in an interface that is stable against the AT instability. Recent exper-
174
P. W. Voorhees & M. J. Aziz
imental results [22] confirm this prediction. Thus, in the a-c Si system very large (perhaps impossibly large) tensile stresses are required to overcome the stabilization due to the stress-dependent mobility. The dispersion curve for the a-c transformation in Si is shown in Fig. 1, where we have used aa = -0.5 GPa [14], Y = 100 GPa, v = 0.25, T = 520°C, M = 4.52 x 10" 2 0 m 2 / J s [24], AG°V = -6.27 x 108 J / m 3 [25], and 7 = 0.4 J/m 2 [26,27]. The length scale of the interfacial instability is provided by the critical wavenumber, Eq. (25), and can be also be read directly from Fig. 1. The critical wavelength, Ac = 2ir/qc, of the instability is 8.5 nm. Thus, perturbations with wavelengths in excess of 8.5 nm will amplify and those with wavelengths less than 8.5 nm will decay. The critical wavenumber of the AT instability, q^T is, qf
= 2a2J/r
(28)
For the BC experiment, this yields \^T = 2000 nm. It is thus clear that the length scale of the kinetically-driven instability can be quite different from that driven solely by elastic strain energy. The wavelength of the kinetically-driven instability with the maximum growth rate is 17 nm. Unfortunately, the BC experiments were unable to determine qc because, for reasons of experimental practicality, they started not with a planar interface, but rather with an interface with a large initial sinusoidal perturbation with a wavelength of 400 nm. They found that a perturbation of this wavelength is unstable, and thus is consistent with this prediction. The experimentally-observed amplification rate of such a perturbation is 6 x 10~ 5 /s, whereas the predicted amplification rate is 1.6 x 10 _ 4 /s. This is very good agreement considering the uncertainties in the material parameters. It has been suggested that the stress-dependent mobility may also affect the evolution of elastically-stressed interfaces when growth is not fully reaction limited [14]. To investigate this possibility, we consider an interface evolving in a fully surface-diffusion-limited regime in which the atomic mobility for diffusion, Mp, is stress dependent. In this case, the velocity of the interface is related to a surface gradient in the chemical or diffusion potential at the surface, v oc V s • MD (
7K)
,
(29)
where V s denotes the surface gradient operator. Consider the case of a planar surface where the basic-state stress is constant. Perturbing about this state gives v° + Sv1® oc - V s • (M°D + dM1^)
V s (6W1® + 6-yK1®).
(30)
Because the elastic energy density and the chemical driving force in the basic state are constants, they are not present in Eq. (30) and thus the effects of the perturbed atomic mobility, M^, are on the order 62. Therefore, the stress-dependent mobility, or a stress-dependent, surface diffusion coefficient, will not affect the linear stability of a uniformly-stressed planar surface in purely surface-diffusion-limited growth. However, a stress-dependent atomic mobility will clearly affect the nonlinear, or large amplitude, evolution of the surface. In addition, a stress-dependent interfacial mobility will continue to affect the linear stability in a situation of mixed diffusion and interface-reaction, or step-edge attachment, control [23].
The Effects of a Stress-Dependent Mobility on Interfacial Stability
4
175
Conclusions
The linear stability analysis for interface-reaction-limited growth shows that when the mobility depends upon stress it can either lead to morphological instability or act to stabilize the planar interface. In addition, the more classical elasticenergy-driven AT instability is also present. However, because the kinetically-driven instability scales linearly with the applied stress and the AT instability is quadratic in the applied stress, the kinetically-driven instability should dominate at low levels of stress. When the theory is applied to the a-c transformation in Si, we find that for the parameters used in the BC experiments the observed interfacial instability is a result of the stress-dependent mobility. Finally, the predicted critical wavelength and amplification rate of the kinetically-driven instability is consistent with the experiments. Acknowledgments This research was supported by the National Science Foundation through DMR9707073 (PWV) and DMR 98-13803 (MJA). References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22.
V. A. Shchukin and D. Bimberg, Rev. Modern Phys. 71, 1125 (1999). R. J. Asaro and W. A. Tiller, Metall. Trans. 3, 1789 (1972). M. A. Grinfeld, Dot Akad. Nauk SSSR 290, 1358 (1986). D. J. Srolovitz, Acta Metall. 37, 621 (1989). J. Gao, Int. J. Solids Structures 28, 703 (1991). B. J. Spencer, P. W. Voorhees, and S. H. Davis, Phys. Rev. Lett. 67, 3696 (1991). K.-S. Kim, J. A. Hurtado, and H. Tan, Phys. Rev. Lett. 83, 3872 (1999). J. E. Guyer and P. W. Voorhees, Phys. Rev. Lett. 74, 4031 (1995). F. Leonard and R. Desai, Phys. Rev. B 56, 4955 (1997). V. G. Malyshkin and V. A. Shchukin, Semiconductors 27, 1062 (1993). B. J. Spencer, P. W. Voorhees, and J. Tersoff, in press. J. E. Guyer and P. W. Voorhees, Phys. Rev. B 54, 10710 (1996). B. J. Spencer, P. W. Voorhees, S. H. Davis, and G. B. McFadden, Acta Metall. Mater. 40, 1599 (1992). W. Baravosa-Carter and M. J. Aziz, Phys. Rev. Lett. 8 1 , 1445 (1998). M. J. Aziz, P. C. Sabin, and G.-Q. Lu, Phys. Rev. B 44, 9812 (1991). G. Grinstein, Y. Tu, and J. Tersoff, Phys. Rev. Lett. 8 1 , 2490 (1998). V. P. Zhdanov, Phys. Rev. Lett. 83, 656 (1999). G. Grinstein, Y. Tu, and J. Tersoff, Phys. Rev. Lett. 83, 657 (1999). H. H. Yu and Z. Suo, J. Appl. Phys. 87, 1211 (2000). A. Witvrouw and F. Spaepen, J. Appl. Phys. 74, 7154 (1993). B. J. Spencer, P." W. Voorhees, and S. H. Davis, J. Appl. Phys. 73, 4955 (1993). J. F. Sage, W. Barvosa-Carter, and M. J. Aziz, Appl. Phys. Lett. 77, 516
176
P. W. Voorhees & M. J. Aziz
(2000). 23. M. J. Aziz, Mater. Res. Soc. Symp. Proc. 618, 233 (2000). 24. G. L. Olson and J. A. Roth, Mater. Sci. Rep. 3, 1 (1988). 25. E. P. Donovan, F. Spaepen, D. Turnbull, J. M. Poate, and D. C. Jacobson, J. Appl. Phys. 57, 1795 (1985). 26. N. Bernstein, M. J. Aziz, and E. Kaxiras, Phys. Rev. B 58, 4579 (1998). 27. C. M. Yang, Ph.D. Thesis, California Institute of Technology, Pasadena, CA (1997).
NON-CONSTANT GROWTH CHARACTERISTICS OF PIVALIC ACID D E N D R I T E S IN MICRO G R A V I T Y J. C. L A C O M B E , M. B . K O S S , A. O. L U P U L E S C U , J. E. F R E I , A N D M. E . G L I C K S M A N Materials Science and Engineering Department, Rensselaer Polytechnic Institute, Troy, New York 12180-3590, U.S.A. E-mail: [email protected] Dendritic-growth experiments conducted aboard the space shuttle Columbia's USMP-4 mission as part of the Isothermal Dendritic Growth Experiment (IDGE) are described. Video data reveals that pivalic acid dendrites growing in the diffusion-controlled environment of low-earth orbit exhibit a range of transient or non-steady-state behaviors. The observed transient features of the growth process continue to be studied, with the objective of elucidating the mechanisms responsible for these behaviors. While not yet clear, experimental factors may be responsible such as the long-range thermal interactions between the dendrite and its neighbors or container. It is also considered that the transient behaviors may arise from fundamental characteristics of isothermal, diffusion-limited, dendritic growth.
1
Introduction
The formation of dendritic microstructures is common during the solidification of most metals and alloys. It is important to understand the process by which these dendrites form because their microstructures can persist through subsequent material processing stages and affect the properties of the finished product. Over the past five decades, the scientific community has produced a large body of theoretical and experimental work describing dendritic growth. Glicksman and Marsh [1] discuss this research in their 1993 review article, and Bisang and Bilgram [2] also include a detailed review as part of their 1996 article on the subject. As these reviews and others make clear, the speed of an advancing dendrite tip is a critical characteristic of dendritic growth. In 1947, Ivantsov [3] introduced two simplifying assumptions in the theory describing thermal dendrites growing in a supercooled melt: 1) that a dendrite can be represented as a shape-preserving paraboloidal interface with a tip radius, R; 2) that the dendrite grows at a constant rate, V. The notion that real dendrites (as opposed to paraboloidal, needle crystals) also grow in a steady-state manner, with constant velocity, is generally considered to be supported by numerous experimental observations of isolated dendrites (see Huang [4] for an example). Most experimental and simulation studies of dendritic growth thus attempt to extract a constant velocity measurement as a parameterization of the kinetics, whereas most theoretical studies assume constant-velocity behavior. Recent work has focused on studying the thermal interactions between a dendrite and its surroundings, and quantifying what is required for a dendrite to grow in a truly "isolated" manner. Thermal interactions may exist between a dendrite and its neighboring tips, container walls [5], or even its own initial structure [6]
177
178
J. C. LaCombe et al.
and trailing side branches [7,8]. Such interactions are potential explanations for experimentally-observed growth rates that do not match the predictions for an isolated diffusion-limited dendrite, with the thermal boundary conditions set at infinity. If the strength of these thermal interactions changes over time, the velocity of the dendrite should change as well, and in general, there should be deviations from the conventionally-prescribed behavior predicted by an Ivantsov-like solution for an isolated dendrite. In particular, at low supercoolings, the length scale over which the thermal field extends is known to be large compared to the morphological length scales relevant to the dendritic growth process. With such long thermal diffusion lengths, proximate dendrite arms can mutually interact to generate an operating state that is not steady. This type of behavior is evident in the recent phase-field simulations by Provatas et al. [9], which suggest the existence of an early-time transient in two-dimensional dendritic growth prior to approaching a steady-state growth rate. 2
Experiment
The data presented here derives from an analysis of dendritic growth data obtained from the Isothermal Dendritic Growth Experiment (IDGE). This project compiled data on the growth of pivalic acid (PVA) dendrites in microgravity. Additional detailed information describing earlier IDGE experiments can be found elsewhere [10]. 2.1
Apparatus
The experimental apparatus used for the IDGE experiments was located in the payload bay of the space shuttle Columbia. This equipment, depicted in the schematic of Fig. 1, was specially designed for use on earth as well as the environment of lowearth orbit. The sample material was contained by a rigid quartz growth chamber located within a temperature-controlled bath. The growth chamber interior volume measured approximately 31 mm square by 50 mm long. Nucleation was achieved through the use of a hollow stinger tube that penetrated the wall of the growth chamber. The exterior end of the stinger tube was closed and surrounded by a thermoelectric cooler. The interior end was open, allowing the PVA sample material in the chamber to also fill the stinger. The use of a quartz growth chamber was necessitated by the sample material. Pivalic acid was found to be excessively reactive with the stainless steel and borosilicate glass growth chambers used in earlier IDGE space flight experiments where the sample material was succinonitrile [10]. The PVA that filled stainless-steel chambers rapidly degraded in purity and would not supercool enough to conduct the necessary experiments. Additionally, for reasons that are not completely understood, the PVA-filled stainless-steel chambers were also prone to structural failure. During the operation of the experiment, each dendritic growth cycle began by completely melting the PVA, followed by lowering the melt's temperature to the desired supercooling. After the supercooled-melt's temperature reached steady state, the thermoelectric cooler was activated. This nucleated a small crystal in the end of the stinger, which then propagated down the stinger tube to emerge
Non-Constant
Growth Characteristics
of Pivalic Acid Dendrites
179
Uniform Temperature Bath
35mm Camera
Video Camera
—v
Xenon Flash
Dendrite
Figure 1. A schematic representation of the experimental apparatus. The sample material is located inside the quartz growth chamber. This is placed within a temperature-controlled bath where the experiment is conducted with in-situ monitoring of process parameters (temperature, etc.) and imaging takes place.
into the chamber as a freely growing dendrite. Once one of these "growth cycles" was completed, a new growth cycle was initiated by re-melting the sample and proceeding as described above. This arrangement, combined with the microgravity conditions, produced dendritic crystals grown under diffusion-limited conditions, with the bath temperature controlled to within 0.002 K (spatially and temporally). The same apparatus was also operated on the ground to acquire an appropriate "baseline" data set describing dendritic growth under the convective conditions associated with earth gravity. During the growth cycles, once a crystal emerged from the stinger, images of the dendrites were obtained from two perpendicular views using both electronic and film cameras. While the film images constituted the primary data source for the IDGE experiments, the observations described here are derived primarily from data obtained from the electronic video cameras. These cameras provided the spatial and temporal resolution that is necessary to study the transient aspects of the growth process. Specifically, an imaging chip array was used of 640 x 480 pixels (256 grayscale levels) and an imaging rate of 29.963 frames per second (~ 30 Hz). 2.2
Analysis Techniques
The data presented here are primarily based upon measurements of dendrite-tip locations as a function of time obtained from the 30 fps video cameras. The base optical resolution of the system is related to the size of an individual pixel in the video camera's imaging array. Each pixel, after correcting for the magnification of the optics system, images a region of the chamber that is approximately 22 /xm by 22 fira. These values also constitute the raw measurement precision for the tip position data. It is beneficial to improve upon this precision by applying a sub-pixel
180
J. C. LaCombe et al.
Figure 2. Video image of a dendrite from Cycle 04 grown at 0.407 K supercooling during the USMP-4 experiment. This and other similar images were used to produce the results presented here.
resolution image analysis method to each image in the growth cycle (see Figs. 2 and 3 for sample images). The method used for the data presented here begins by first examining the first and last image frames of a growth cycle. The dendrite tips are found in these images using a row-by- row search to locate the lowest pixel in the field of view that is darker than a specified threshold value. If there are more than one pixel in a row satisfying this criteria, an average is calculated to obtain the horizontal coordinate. With these crude estimates of the tip positions at the beginning and end of the growth, it is possible to construct a vector used to predict where the tip will be in all frames during the growth, provided the frame number and frame rate are incorporated. This tip-location prediction scheme improves processing efficiency, though it is only capable of resolving information on the order of 22 /xm (one pixel). Using this predicted tip location, a second stage of refinement is added to the tip-location process for each image. This is achieved by overlaying a "sampling line" along the predicted vector and determining the point along this vector where the image intensity crosses a selected threshold value. However, this time the sampling line is "thick" in that it also comprises several pixels on either side of the mathematical line, creating an averaging effect. Additionally, interpolation is applied to determine the threshold location more precisely. By incorporating a statistically larger number of pixels in this second-stage, tip-locating method, resolution is improved to approximately 7 /im (~ 1/3 pixel). However, the horizontal coordinate
Non-Constant
Growth Characteristics
35mm film
of Pivalic Acid Dendrites
181
Video
™ R «20/m 1DD tin
1 1
R «20 ^m a 1 pixel Figure 3. Images of the tip region of the dendrite grown in Cycle 04. a) A 35 mm film image with an approximate tip radius of curvature of 20 [im. b) A 30 fps video image, revealing considerably less detail. The centroid of the tip is calculated using the pixel intensities located within the box enclosing the tip.
is constrained to fall along the predicted vector line. In practice, this is not usually a significant issue, since these dendrites tended to grow within about 10 degrees of vertical, and, once started, do not deviate in direction under nominal growth conditions. The final stage of resolution enhancement is achieved by a somewhat unorthodox approach, which has the advantage of incorporating still more (statistically speaking) information concerning the tip location. Fig. 3b shows a typical video image of a dendrite tip. Using the refined tip location (stage 2, described above) as a reference location, a box is created around the tip. Next, the centroid of the pixel intensity within this box is calculated. In practice, the coordinates of this centroid exhibit a resolution of approximately 2 /xm (~ 1/10 pixel). This approach does not actually locate the tip's interface. Instead, it uses more data to obtain a more consistent reference point, which serves to track the tip's movement over time. The size of the sampling box used in the centroid calculation is somewhat arbitrary. The concerns in its selection are primarily in obtaining a balance between the desire to have a large number of data points contributing to the measurement, and avoiding the inclusion of side branches. When information that is obtained from regions further removed from the tip is used, nascent side branches can contribute to the centroid calculation in a periodic manner. Once extracted from the images, the tip positions are then converted into a displacement vs. time data set that is used for subsequent analysis (see Fig. 4). The displacement is calculated relative to the tip position in the first available frame of video data, which is assigned to time t = 0. The measures of resolution quoted herein are characterized by examining the residuals produced by a linear regression of the displacement data (Fig. 5a). The standard deviation of the spread in these residuals is calculated on a moving basis, with a two-second window. These
182
J. C. LaCombe et al.
100
ISO
T i m e (seconds) Figure 4. Tip displacement vs. time for dendrite Cycle 09, grown at 0.376 K supercooling. Linear regression was performed using data from the more steady-state portion of the growth (after ~ 125 seconds). The growth rate, 22.9 fim/s results from the slope generated by the regression. Note: 1 of every 30 data points plotted.
a)
Time (seconds)
b)
Time (seconds)
Figure 5. a) Residuals resulting from a linear fit to the tip displacement vs. time data in Fig. 4 for the latter growth stage, i.e., after ~ 125 seconds, b) The standard deviation of the spread in the residuals is below 2 (im, representing the uncertainty in the tip-position measurements.
measurements are shown in Fig. 5b, revealing the 2 /im (at most) uncertainty stated above.
Non-Constant
3
Growth Characteristics
of Pivalic Acid Dendrites
183
Results
Earlier observations by the authors [11] indicated that dendritic growth rates are not constant over the time scale of observation. Specifically, the dendrite tip, after completing an initial transient, continues to experience a small acceleration. It is not yet clear whether or not the dendrite approaches a truly constant velocity. The displacement data shown in Fig. 4 are derived from a supercooling of 0.376 K. They are qualitatively representative of the entire body of data analyzed to date. The gaps in the data resulted from the hardware necessity to adjust lighting during 35 mm camera exposures. As with the data presented in [11], this cycle shows an initial transient clearly accelerating, and then evolves toward a more constant-slope regime from which one can extract the velocity. Despite the relative linearity of the latter part of the displacement plot, a leastsquare regression of the data from the portion of the growth between the hash-marks of Fig. 4 (after ~ 125 seconds) reveals that the dendrite is still accelerating late in its growth. The residuals resulting from the linear regression show a systematic deviation from steady-state growth (Fig. 5a). A dendrite growing at constant velocity would be expected to exhibit random residuals forming a Gaussian distribution centered about zero. Instead, as shown, a distinct non-random residual occurs, with monotonic upward curvature in the graph. As mentioned earlier, the residuals also quantify the uncertainty in the measurements, because their spread is a good measure of the resolution resulting from the sub-pixel interpolation measurement of the tip location. A convenient measure of the asymptotic behavior of the tip speed, as evidenced by the non-random residual, is the growth-rate exponent, K, calculated by comparing the data to a power law for the displacement proportional to tK. An exponent of K — 1 indicates linear displacement in time, i.e., constant velocity. Although at first this method may seem less intuitive than simply comparing the instantaneous velocity versus time, the slope of which is the acceleration, the power law helps elucidate the variations from, and approach to, constant-velocity (K = 1) behavior. The measured values of K (Fig. 6) calculated after first smoothing the displacement data using a two-second moving average, reveals a slightly different trend than seen for the dendrite growth cycle presented in [11]. In the previous report, K started out large and approached 1 from above. For this starts out close to 1 and increases slightly (still an accelerating trend). Aside from this initial transient, other dendrites have been seen to behave qualitatively similar in terms of K. 4
Discussion
The data presented here indicates that there are transient aspects of the dendritic growth process which steady-state theory does not describe. Examination of these experimental data suggests that within experimental practice, the establishment of strict isothermal conditions at the dendritic interface may be illusory. We will discuss briefly several potential sources for non-steady-state behavior. In doing so, it is beneficial to describe the time-dependent process as occurring in two distinct periods. The first of these is the initial transient, which is most evident in Fig. 6
184
J. C. LaCombe et al.
Time ( s e c o n d s ) Figure 6. The growth-rate exponent, K, as a function of time for the dendrite data in Figs. 4 and 5. A value of K = 1 corresponds to steady-state, constant-velocity growth.
prior to approximately 50 seconds. The second distinct time-dependent period is beyond ~ 50 seconds, where the growth-rate exponent continues to increase, though more gradually. To explain the early-growth behavior, the authors [11] suggested two related mechanisms that may induce a transient effect. Summarizing, we note that during the early phase of growth, a dendrite evolves from some pre-dendritic morphological structure to eventually develop an interface shape and thermal field which are commensurate with steady- state growth. This developmental stage takes time. Additionally, in the nascent stages of a dendrite's development, neighboring dendrites (or other branches of an equiaxed dendrite) are located close to one another. As long as these separation distances are less than a few thermal lengths, where the characteristic thermal length is defined here as the thermal diffusivity of the liquid (7 x 10 4 /im 2 /s) divided by the growth rate, thermal interactions are expected to occur between neighboring dendrites causing a localized drop in supercooling. As the dendrites grow, their tip-separation increases, creating a progressive increase in the local effective supercooling. This would manifest itself as an accelerating growth rate. Similar interactions also exist for proximate dendrites growing in the same direction. Eventually, if the interactions are strong enough, one of the dendrites will experience favorable thermal conditions and will "pull away" from the slower dendrite. The experimental (and simulated) observation of an initial transient preceding steady-state growth is not new. In fact, it has been generally assumed that this phase, although easily distinguishable from the steady-state that emerges, is less germane and fundamental to the kinetics than the steady-state itself. However,
Non-Constant
Growth Characteristics
of Pivalic Acid Dendrites
185
as is revealed here, the growth phase following the initial transient might not ever reach a constant velocity. Thus, the more important issue for such investigations is whether the time-dependent state is fundamental to the physics of isothermal dendritic growth, or appears as an artifact of experiments carried out in finite volumes of supercooled melts. The observation of the effect of neighboring dendrites on growth velocity [11] suggests possible long-range thermal interactions with proximate dendrites or arms growing in the same general direction as the tip in question. It is conceivable that similar interactions may contribute to a second stage of time-dependent growth seen after the initial transient is completed. However, many of the analyzed dendrites were isolated growths (noting that 'isolation" is a strong function of time and supercooling), where the closest neighbor was beyond three thermal lengths. In order to conclusively examine this issue, it will be necessary to analyze a significant number of dendrites growing under a variety of conditions. Additionally, one must firmly establish that the thermal interaction between dendrites separated by more than three thermal lengths are still enough to produce the measured changes in growth velocity. The second transient phase, also described in [11], can also potentially be explained by the finite size of the growth chamber. Pines, Chait, and Zlatkowski [5] showed that a dendrite may interact with the walls of a container as growth proceeds. However, it remains to be determined whether the closing rate between the dendrite and the wall is commensurate with the measured time variations in velocity. Given these considerations, it is our hypothesis that it is more likely that the second stage of non-steady- state behavior is fundamental to isothermal dendritic growth. Since dendrites are not truly parabolic bodies of revolution [4,12], there is no compelling phenomenological reason that dendrites should grow strictly at a constant rate. As the data set from which these observations and conclusions were drawn is the only one we know of that is both diffusion-controlled, and measured at the necessary temporal resolution to make the described measurements, we plan additional work in data reduction, analysis, and modeling, to explore this hypothesis further. 5
Conclusions
In summary, a method was developed for evaluating dendritic growth rates that discriminates fine distinctions in the non-steady-state behavior. This method is applied here to PVA dendrites, seen to grow in microgravity at non-steady- state velocities. The mechanism responsible for this behavior may be related to thermal interactions between a dendrite and its surroundings, or it may be intrinsic to the dendritic solidification process. Efforts to further discriminate among the possible causes are currently under way. Acknowledgements This work was supported under NASA Contract Nos. NAS3-25368 and NAG8-1488. The authors wish to thank D. P. Corrigan for discussions about image processing,
186
J. C. LaCombe et al.
and L. A. Tennenhouse, and the many Rensselaer undergraduate and graduate student volunteers for assistance in operating the flight experiment. Additional thanks to IDGE Project Manager D. C. Malarik and her team at (or associated with) NASA's Glenn Research Center at Lewis Field for their engineering support, and the staff of the POCC/HOSC at NASA's Marshall Space Flight Center for flight support. Finally, we thank the crew of the STS-87 for bringing our experiment to orbit and back, and for their extra effort in providing us with the additional best available quality video data for our post-flight analysis. References 1. M. E. Glicksman and S. P. Marsh, in Handbook of Crystal Growth, edited by D. T. J. Hurle, Elsevier Science Publishers, Amsterdam, p. 1077 (1993). 2. U. Bisang and J. H. Bilgram, Phys. Rev. E 54, 5309 (1996). 3. G. P. Ivantsov, Dokl. Akad. Nauk, SSSR 58, 567 (1947). 4. S.-C. Huang and M. E. Glicksman, Acta Metall. 29, 701 (1981). 5. V. Pines, A. Chait, and M. Zlatkowski, J. Cryst. Growth 167, 383 (1996). 6. V. Pines, A. Chait, and M. Zlatkowski, J. Cryst. Growth 182, 219 (1997). 7. J. C. LaCombe, M. B. Koss, D. C. Corrigan, A. O. Lupulescu, L. A. Tennenhouse, and M. E. Glicksman, J. Cryst. Growth 206, 331 (1999). 8. J. C. LaCombe, M. B. Koss, D. C. Corrigan, A. O. Lupulescu, J. E. Frei, and M. E. Glicksman, in Solidification 1999, edited by W. H. Hofmeister, et al, TMS, Warrendale, PA, p. 121 (1999). 9. N. Provatas, N. Goldenfeld, J. Dantzig, J. C. LaCombe, A. Lupulescu, M. B. Koss, M. E. Glicksman, and R. Almgren, Phys. Rev. Lett. 82, 4496 (1999). 10. M. B. Koss, L. A. Tennenhouse, J. C. LaCombe, M. E. Glicksman, and E. A. Winsa, Met. Trans. A, (2000) (accepted). 11. J. C. LaCombe, M. B. Koss, and M. E. Glicksman, Phys. Rev. Lett. 83, 2997 (1999). 12. J. C. LaCombe, M. B. Koss, V. E. Fradkov, and M. E. Glicksman, Phys. Rev. E 52, 2778 (1995).
INTERFACES ON ALL SCALES D U R I N G SOLIDIFICATION A N D MELTING M. G. W O R S T E R Institute of Theoretical Geophysics, Department of Applied Mathematics and Theoretical Silver Street, Cambridge CBS 9EW, U.K.
Physics,
One of the challenges in predicting the solidification and melting of multicomponent materials is to describe the evolution of mushy layers: regions of mixed phase in which solid and liquid coexist in intimate contact. If mush is treated as a distinct phase of matter then it is necessary to determine the conditions that apply at, and control, its interfaces with adjacent solid and liquid regions. At a mush-liquid interface, different conditions apply depending on the direction of fluid flow, with which solute is advected, relative to the direction of advance or retreat of the phase boundary. A mathematically complete set of interfacial conditions is presented for the case of steady solidification or melting and is used to determine states of buoyancy-driven convection in a mushy layer.
1
Introduction
Solidification, the transformation of liquid into solid, involves the molecular rearrangement of matter, yet has consequences for global-scale phenomena such as the growth of the Earth's inner core and the freezing of the polar oceans. At every level of description, it is convenient to identify interfaces that separate regions of different phase or physical property. At the smallest scale usually considered, there is an interface separating solid from liquid. On planetary scales, it may be convenient to identify the layer of sea ice on the surface of the ocean as an interface between ocean and atmosphere, for example. Let me be more precise at this point and define an interface as a singular surface, a mathematical construction of infinitesimal thickness, separating two regions in which different field equations apply. The task for the mathematical modeller is to obtain or prescribe properties of the interface that relate the field variables on either side of it. Intrinsic or internal to the interface are physical processes and interactions that are unresolved at the level of mathematical description being used but may need to be explicitly investigated in order to derive the appropriate conditions to be applied at or across the interface. A concrete example is provided by the solid-liquid interface of a pure substance. Its temperature is given by T = Tm-7V-n,
(1)
where Tm is the equilibrium freezing temperature of the substance, 7 is related to the surface energy, and n is the unit normal to the interface pointing into the liquid. The temperature fields on either side of the solid-liquid interface are related by kan-VT\a-kin-VT\l=p3LVn,
(2)
where k is the thermal conductivity, p is the density, L is the latent heat of fusion per unit mass, Vn is the local rate of solidification, and subscripts I and s denote
187
188
M. G.
Worster
properties of the liquid and solid respectively. Two of the parameters in these equations, 7 and L, are properties of the interface resulting from molecular processes. The surface energy results from long-range, intermolecular, van-der-Waals forces, which tend to bind molecules more strongly to the solid when the interface is concave towards the liquid. The latent heat parameter is related to the change (decrease) in entropy as molecules become ordered within the crystalline lattice of the solid. Yet the molecular-scale description is unnecessary when calculating the macroscopic evolution of a solid-liquid interface: the continuum equations (1) and (2) suffice, and the parameters can be determined from thermodynamic measurements. Can sea ice be treated as an interface in the sense described above? This is a question to motivate the discussions presented herein, though the theories proposed have much wider application within a range of geophysical and industrial systems. Sea ice occupies a region no more than a few meters thick. Even though, as we shall see, many complex physical interactions take place within the interior of sea ice, its thickness is many orders of magnitude smaller than the scales of atmospheric and oceanic processes on either side of it. It would therefore be convenient in numerical models of ocean-atmosphere interactions, for example, to be able to relate properties of the ocean to those of the atmosphere at just the other side of the layer of sea ice in a simple and predictive way that does not involve detailed evaluation of the internal dynamics of the sea ice. As we try to address the question posed above, we shall be led to consider interfaces at a number of intermediate scales between the simple, solid-liquid interface and the complex sea-ice interface.
2
Sea Ice — Some Experimental Observations
In experiments designed to investigate the formation and growth of sea ice, Wettlaufer et al. [1] cooled aqueous solutions of sodium chloride (NaCl) through the roof of a rectangular container (Fig. la). In consequence of the fact that salt molecules cannot be readily incorporated within the cystalline lattice of solid ice, almost pure ice crystals grown from solution form a matrix whose interstices are filled with brine enriched by salt rejected by the growing ice. The two-phase region of ice crystals and interstitial brine is an example of a mushy layer, which form very commonly during the solidification of multi-component melts. Measurements of the salt concentration in the solution below the mushy layer (Fig. lb) show that the interstitial brine remains within the mushy layer until the layer has reached a critical thickness hc. Thereafter, brine drains from the mushy layer, increasing the concentration of the underlying liquid. The graph in Fig. l b delineates three stages of evolution: a period of relative quiescence when h < hc; a period of active salt transfer from the mushy layer to the liquid region; and a transition between these periods. The initial period and the transition have been studied quite extensively (see reviews [2] and [3]), while the period of active transfer, effected by compositional convection of the dense brine, is the subject of current research.
Interfaces
on All Scales During Solidification and Melting
189
(b)
10
12
14
16
Thickness (cm) Figure 1. (a) The experimental apparatus used to investigate the growth of sea ice. (b) The concentration of NaCl in the liquid region C in weight percent NaCl as a function of the thickness of the mushy layer h. The initial concentration is labelled Co. Rejected brine remains within the interstices of the mushy layer until its thickness exceeds a critical value hc •
2.1
The Quiescent Mushy Layer
The thickness of a mushy layer is controlled principally by a balance of heat fluxes [4]. Additionally, the thermal field within a mushy layer can be approximated as having a quasi-steady, linear profile. Grossly, therefore, the evolution of a quiescent mushy layer is governed by the equations
PsL-{H)
FA-F0
TQ-TA
= k{)
h
'
(3)
where To and TA are the temperatures in the ocean and atmosphere, Fo and FA are the heat fluxes from the ocean and into the atmosphere respectively, h is the thickness of the sea ice, and k((j>) is its mean thermal conductivity, which depends significantly on its mean solid fraction 4> (Fig. 2). The mean solid fraction can be determined from the lever rule
T0-TA
^-ibTTV
(4)
provided T is measured in degrees Celsius. Note that, during growth TA < To < 0 and is positive. This rather simple approach needs to be modified in the case of melting. We see, therefore, that during this period sea ice can indeed be treated as an interface: the field variables To,TA, Fo, and FA in the ocean and atmosphere are related by Eqs. (3) in terms of the interfacial parameters h and cf>, which themselves can be determined from values of the field variables. 2.2
The Transition
The onset of significant brine drainage from the mushy layer is a consequence of a buoyancy-driven instability. Linear theory [5] shows that the critical conditions for stability are governed by the magnitude of a Rayleigh number *^m —
p(T0-TA)gIlh mKV
(5)
190
M. G. WoTster
FA
Atmosphere TA Sea ice
Thickness h
Ocean
Solid fraction (,> °
i F0
Fr
Figure 2. The gross features of sea ice, treated as an interface between ocean and atmosphere. The aim is to relate the heat fluxes Fo and FA and the salt flux FQ to the external parameters To and T&, the temperatures of the ocean and atmosphere, and the internal parameters of the sea ice h and >.
in which (3 is the constant of proportionality between the density of brine and its salinity, m is the slope of the liquidus curve (the concentration-dependent freezing temperature), g is the acceleration due to gravity, K is the thermal diffusivity, v is the kinematic viscosity, and II = II(>) is the permeability of the mushy layer, treated as a porous medium [6]. The permeability is difficult to measure directly but, given that the transition to brine drainage occurs at a critical value Rc of i? m , Eq. (5) shows that [n^r
1
= ( ~ ^ )
(To - TA)hc ex {T0 - TA)hc .
(6)
The data from laboratory experiments using a wide range of operating conditions [1] collapse to a single curve by relation (6) as shown in Fig. 3. This figure can be used as a phase diagram for brine convection from sea ice according to the external parameters TO,TA, and the internal (interfacial) parameters (h,
Brine Drainage
Once the critical depth hc is exceeded, brine is observed to drain from the mushy layer in the form of isolated plumes emanating from brine channels, which are narrow, essentially vertical, cylindrical channels in the porous matrix. Brine channels, also called chimneys in the metallurgical literature, form by dissolution of the solid phase of a mushy layer whenever there is flow of the interstitial liquid with a component parallel to the temperature gradient that exceeds the speed of the isotherms [9]. It will be convenient in the remainder of this article to consider situations in which a mushy layer grows upwards from a cooled lower boundary and the residual (interstitial) liquid is less dense than the original melt or solution. This is the case in the commonly studied system of ammonium chloride crystals growing from
Interfaces
on All Scales During Solidification
and Melting
191
200 r
150
hc(T0-TA) m
Strong compositional convection through chimneys
100
(cm wt.%) 50 Weak convection :Xi.
0.0
0.2
,». 0.4
0.6
0.8
1.0
<$>C
Figure 3. The critical conditions for brine drainage from sea ice. The crosses are experimental data. The curve is hand-drawn through the data.
aqueous solution, as illustrated in Fig. 4, for example. The solution is depleted of the salt (NH4C1) forming the solid phase of the mushy layer and therefore has a tendency to rise convectively upwards through and out of the mushy layer. Once the convection is sufficiently strong, chimneys form, as shown in Fig. 4. The top of the mushy layer and the chimney wall are examples of mush-liquid interfaces. Our aim is first to determine appropriate conditions to apply at these interfaces and then to explore the nature of convection in mushy layers, through chimneys, with the aim of describing the whole convecting mushy layer as an interface. The convection in this system is similar to that in sea ice, with the sign of gravity reversed. 3
Interfaces W i t h Mushy Layers
Fig. 5 shows a close-up of the top of a mushy layer of ammonium-chloride crystals showing some of the scales at which different features can be resolved. As previously noted, mushy layers are porous media. Typical sizes of the interstices are in the range 6 ~ 0.1 — 1 mm, which are very much less than the macroscopic dimensions of the mushy layer. Theories of mushy layers do not resolve these scales and can be developed formally in the limit 6 —> 0 [10]. The interface between the mushy layer and the liquid region is ambiguous. It could be considered as the envelope of the tips of the crystals, but that leads to fractal uncertainties: the area of such an envelope, for example, is dependent on the yardstick used to define it. It is more convenient to think of the interface initially as a region of finite thickness Si, as shown in Fig. 5, and then let 6j —» 0. Provided 61 » 8, i.e., the limit 6 —* 0 is taken before the limit 8j —> 0, then the ambiguity is removed. We shall not formally take these limits here, but these ideas will direct our thinking. We are concerned with the nature of mush-liquid interfaces in the presence of flow, and it will be convenient to consider two different frames of reference for different purposes. I shall denote the dynamical fluid flow, flow relative to the solid phase
192
M. G.
Worster
Fi gure 4. A photograph showing two chimneys in a mushy layer of ammonium-chloride crystals grown from aqueous solution.
Liquid
Figure 5. The interfacial region of a mushy layer of ammonium chloride crystals growing from aqueous solution. The inter-dendritic spacing has a characteristic length scale <5, while the interface has a characteristic thickness <5/, which can be chosen to encompass any boundary layers in the solute or temperature fields, if necessary.
of the mushy layer, by u, and the motion of the liquid relative to the mush-liquid interface by q.
Interfaces on All Scales During Solidification and Melting 193
3.1
Conservation of Heat
Some interfacial conditions can be determined from conservation laws. Mathematically this can be achieved by applying the divergence theorem to the field equations over a control volume that spans the interface. Conservation of heat is governed in the liquid region by the advection-diffusion equation PiCpl(^
+ u-VT]=V-(k^T),
(7)
and in the mushy region by ^C;~+plCplu-VT
= V-(kS/T)+PsL^,
(8)
[2], where Cp is the specific heat capacity and an overbar denotes a volume-fractionweighted average. The solid fraction is now a local field variable of space and time. Eq. (8) applies equally in the liquid region (where it reduces to Eq. (7) as <j) = 0). It can be integrated (in enthalpy form) across the interface to give Pl
Cpln-q
[T}lm +psLVnl4>}lm = [kn. VT}lm ,
(9)
where n is the unit normal to the interface pointing into the liquid, and Vn is the normal rate of advance of the interface. Eq. (9) can be applied across any thermal boundary layer that might exist, which can be particularly convenient when there is vigorous thermal convection in the liquid region [11]. More commonly, the thermal field is fully resolved on the macroscale, in which case the temperature field is continuous and Eq. (9) reduces to Ps Li Ym
Vn = kn-VTm-kin-VTi , (10) which is similar to Eq. (2). But note that the interfacial value of (ft = <j>m is so far undetermined. 3.2
Conservation of Solute
Similar conservation laws apply to the solute field, governed in the mushy layer by ( l - ^ ) ^ + u - V C = V . (DVC) + (C-CS)^,
(11)
where Cs is the concentration of salt in the solid phase and D is the difFusivity of salt. Again, this equation reduces to the more familiar advection-diffusion equation in the liquid region. Integration of this equation across the interface gives - n • q [ C L + (Cm - Cs)4>m Vn = [-Dn • VC]lm .
(12)
At this point there enters a certain amount of controversy or, at least, uncertainty. Whereas the macroscopic dimensions of mushy layers are controlled by thermal balances, the microscale 6 is determined in part by the solute field. One view is that the scale of solute variations 6c, including any variations within the interfacial region of the liquid, is comparable to 6. In this case, the diffusive flux of solute is unresolved when the homogenizing limit 8 —> 0 is taken, and Eq. (12) reduces to -n-q[C}lm
+ (Cm-Cs)4>mVn
= 0.
(13)
194
M. G.
Worster
Note that the jump in the macroscopic solute concentration [C] cannot be determined without invoking further physical principles, and may take on any value, just as discontinuous shear layers can exist in an inviscid fluid. An alternative view is that 6c is larger than 6, at least in the interfacial region of the liquid, where a compositional boundary layer may exist of a thickness larger than the inter-crystalline spacing, and perhaps too in the mushy region, where morphological instabilities giving rise to side branches on the primary dendrites are limited only by surface energy on much smaller scales. In this case, and if one chooses to resolve any compositional boundary layer on the macroscale, the macroscopic solute field is continuous and Eq. (12) becomes (Cm - Cs)cj>m Vn = \-Dn • WCfm .
(14)
If we subsequently choose our smallest resolved scale to be the scale of thermal variations ST, then the limit 6c —* 0 can be taken appropriately given that the ratio of diffusivities D/K
Melting Interfaces
The conservation laws described above are sufficient to determine the evolution of mush-liquid interfaces that are melting, V„ < 0. Although the equations are coupled, it is convenient to think of the thermal conservation equation (10) as determining the rate of advance (retreat) of the mush-liquid interface once the interfacial value of the solid fraction is determined by solute conservation. It is instructive to consider separately what happens in two cases: when the flow is from liquid to mush (q • n < 0); and when the flow is from mush to liquid (q • n > 0). In the first case (Fig. 6a) advection balances diffusion in the interfacial region of the liquid, creating a compositional boundary layer there, which is governed locally by dC
nd2C nc\ (15) ^ ^ where q = n • q < 0 and n is the normal coordinate pointing into the liquid. This has the solution q
= D
C = C0 + {Cm-
C0)e«n/D
,
(16)
where Co is the concentration in the liquid far from the interface and Cm is the interfacial value of the liquid concentration. This can be used in Eq. (14) to determine that Co — Cm\ n • q Cm ~ Cs ) Vn
,
The same conclusion is reached from Eq. (13) if one chooses not to resolve the boundary layer, but recognizes that in this case the solute field may be discontinuous: [C] =C0 -CmWhen the flow is from mush to liquid (q • n > 0) no compositional boundary layer can be sustained at the interface (Fig. 6b). Provided that q • n 3> d6c/dt,
Interfaces
on All Scales During Solidification
and Melting
195
Figure 6. A mushy layer melting into a hot liquid, (a) When flow is from liquid to mush, advection can balance diffusion of solute on the liquid side of the interface and give rise to a compositional boundary layer there, (b) When flow is from mush to liquid, any compositional boundary layer is advected away from the interface.
the compositional gradient at the interface is vanishingly small and in consequence <> / = 0 there. This is typically the case for a melting interface (one at which the phase change is rate limited by heat transfer), but may not be for a dissolving interface (one at which the phase change is rate limited by solute transfer). Note in all these discussions that because the temperature and composition are tied by the liquidus relation T = TL(C) in the interior of a mushy layer, no compositional boundary layer can exist there, the compositional gradient is proportional to the temperature gradient, and the solute flux on the mush side of the interface is negligible when D/K -C 1. 3.4
Solidifying Interfaces
It needs to be recognized that the differential operators acting on the solid fraction in a mushy layer are hyperbolic. During melting the characteristics of the hyperbolic operator flow towards the interface, whereas they flow away from the interface, into the mushy layer, during solidification. Consequently, an additional boundary condition is required during solidification over and above the conservation laws already described. A common suggestion is arbitrarily to impose < 0 in the mushy layer), and we shall see later that in some circumstances this
196
M. G.
Worster
iM' 1
&
(a,) 4-
Marginal Equilibrium
Figure 7. If the length scale over which the solute field varies in the liquid is greater than the dendrite spacing (ai), then the solute field is resolved after the homogenizing limit 5/ —> 0 is taken. Marginal equilibrium in this picture requires that the temperature gradient is equal to the gradient of liquidus temperature at the interface (bi). If the limit of negligible solutal diffusivity is subsequently taken then picture (c) is achieved, in which the solute field is continuous across the, now unresolved, compositional boundary layer. Alternatively, if the dendrite spacing is comparable to the thickness of the compositional boundary layer (3,2), then the homogenizing limit leads to picture (b2), in which the boundary layer is unresolved, but the possibility exists of a j u m p in the solute field. The principle of marginal equilibrium applied to this picture leads again to (c).
condition is not independent from the conservation laws and therefore leads to an under-determined system of equations. An alternative is to apply a principle of marginal equilibrium which says simply that the liquid should not be supercooled (its temperature should be at or above its liquidus temperature) adjacent to the interface. This principle has a physical basis in that mushy layers are caused by supercooling and will certainly relieve if not eliminate it [12]. The principle has so far always led to well-posed mathematical problems with physically-allowable solutions, though these properties are unproven in general.
Interfaces
on All Scales During Solidification
and Melting
197
When the flow is from liquid to mush and the possibility of a compositional boundary layer exists, the principle of marginal equilibrium is expressed by
&T-"^'
(18)
where TL(C) is the local liquidus temperature [12]. This is illustrated on the lefthand side of Fig. 7. If the compositional boundary layer is unresolved, illustrated on the right-hand side of Fig. 7, the principle of marginal equilibrium is expressed simply by T = Ti(C) on the liquid side of the interface. This in turn implies that the potential jump in solute concentrate [C] — 0 across the interface. Once the limit D/K —> 0 is taken in the first case, both approaches lead to the same result, namely that = 0 at the interface. A rather different situation applies when the liquid flows from mush to liquid [13]. Solute is then advected away from the interface into the liquid region, no compositional boundary layer can exist, and [C] = 0. The fact that 0), then solute is conserved along streamlines in the liquid region. Therefore, during steady solidification, such as might occur in a crystal-pulling configuration or during continuous casting, q-VC = 0
(19)
in the liquid region, where q is measured relative to the steady frame of reference. The principle of marginal equilibrium in this case is expressed by q-VT>q-VTL(C)=0
(20)
on the liquid side of the interface, the final equality being a consequence of Eq. (19). On the other hand, on the mush side of the interface, q . VT = m{C - CS)VS • V4> < 0 ,
(21)
where V s is the velocity of the solid matrix in the steady frame. This equation follows from Eq. (11), and the inequality is required in order that <j> be non-negative. Since <> / = 0 at the interface, Eq. (10) shows that the temperature gradient is continuous. On the assumption that the velocity field is also continuous, the two inequalities (20) and (21) combine to show that q • VT = 0
(22)
at the mush-liquid interface. The interfacial conditions that apply between liquid and mushy regions during steady solidification and melting are summarized in Fig. 8. 4
Steady Convection in a Mushy Layer
Armed with these interface conditions we can investigate steady convection in a mushy layer and follow the evolving morphology of the mush-liquid interfaces as the
198
M. G.
Worster
Melting Solute conservation Flow from mush to liquid
0=0
Solidifying Solute conservation
Marginal equilibrium
<> / = 0
q • VT = 0
[C] = 0
Flow from liquid to mush
0^0 see equation (17)
0 =0
<=
[C}=0
Figure 8. Interfacial conditions between the mushy and liquid regions during steady melting or solidification. When flow is from mush to liquid there is no compositional boundary layer, the solute field C is continuous, and solute conservation implies that the solid fraction is zero at the interface. Marginal equilibrium at a solidifying interface has a different expression depending on the direction of flow. When flow is from liquid to mush it is this condition in combination with solute conservation that implies that rp is zero.
2.5
2
Streamlines of q Solid-fraction contours
Figure 9. Convection in a mushy layer just after the formation of a region of negative solid fraction, shown shaded near the centre of the figure. The solid curves are streamlines of q. The dashed curves are contours of solid fraction.
amplitude of convection increases and chimneys are formed. At small amplitudes of the dynamic flow u, the streamlines of q all enter the mushy layer from the overlying liquid region. At some larger amplitude, after the vertical component of u has exceeded the solidification rate, an internal region of the mushy layer is completely dissolved to form a liquid inclusion [13], as shown in Fig. 9. Computation of the subsequent evolution of such a liquid inclusion requires the application of all four types of interface condition described above.
Interfaces
on All Scales During Solidification
and Melting
(a)
199
(b)
Figure 10. (a) Streamlines of q (solid curves) and contours of solid fraction (dashed curves) once a chimney has formed. The chimney is modelled as a singular interface (shown with a thick vertical line) along which boundary conditions are applied that are derived from a lubrication analysis of the flow internal to the chimney, (b) The computed width of the chimney a. Note the greatly exaggerated horizontal scale. These calculations were made with a Darcy number of Da = 0.005 and a Rayleigh number of 12.9.
4-1
Singular Chimneys
Recent numerical calculations [14] have followed the evolution of the liquid inclusion until it bursts through the upper mush-liquid interface to form a chimney. Such computations, in which the flow and structure of the chimney are resolved, are timeconsuming. Significant savings can be made by exploiting the fact that chimneys are long and narrow. Lubrication theory can be used to analyse the flow in the chimney [15-18], the results of which provide boundary conditions for the mushy layer at the chimney wall. By this means, the entire chimney is treated as an interface with a single intrinsic parameter (its width, a) that can be determined as part of the solution. In brief, conservation of heat and mass determine the conditions at the chimney wall [18]
i,: RaC
Ipx (23) 20 v 3RaU ~Da where ip is the stream function such that q = (—ipz,ipx ~ l)i To is the temperature at the bottom of the chimney, and Da = HV2/n2 is the Darcy number. These apply along the chimney wall, indicated by a thick vertical line in Fig. 10. Below the chimney, the boundary conditions at x = 0 are simply that tp = Tx — 0. The width of the chimney a can be determined from the interfacial conditions
Tx=lPTz,
q • V T = 0,
where ipz > 0,
>(C - Cs)Vn = (d - C)q • n,
where i>z < 0.
(24) (25)
The first of these is simply Eq. (22) applied where the flow enters the chimney. The second is Eq. (17) applied where the flow exits the chimney (if anywhere), recognizing that the solute concentration is constant along streamlines in the chimney, so that C\ = C\ (ip) is the value of the concentration where the exiting streamline entered the chimney. These conditions have recently been applied successfully [19] to determine the structure and flow within a convecting mushy layer including a determination of the width and shape of the chimney (Fig. 10). Note that in this model the chimney width is determined locally by thermodynamic balances.
200
5
M. G.
Worster
Epilogue
The theory of mushy layers allows the gross properties (solid fraction, bulk composition) of a solidifying multi-component melt to be determined without computing the microstructure of the internal solid-liquid interfaces. We have seen how the mush-liquid interface can be determined in a variety of circumstances and, further, how chimneys within a mushy layer can be treated themselves as singular interfaces between a convecting mushy layer and the adjacent liquid region. This approach provides a relatively inexpensive way to compute the state of a convecting mushy layer and the flux of solute emanating from it. The results of such calculations should direct the search for simple ways to predict the convective heat and salt fluxes from mushy layers in terms of their gross properties. Such results will not only be of significant benefit to climate modellers, for whom sea ice remains an important and inadequately understood interface between ocean and atmosphere, it will enable simple predictions to be made of macrosegregation in cast alloys and mineral segregation in igneous rocks. Thus, in terms of both natural philosophy and advanced materials processing, theories of solidification, with all their attendant interfaces, pose important and challenging problems for the 21st century. Acknowledgments The recent advances in our understanding of solidification in mushy layers reported in this paper have been made in collaboration with post-doctoral colleagues D. M. Anderson, T. P. Schulze, and C. A. Chung. The first two are among the many 'academic children' of Stephen Davis. I count myself fortunate to have benefitted thus from Steve's nurture of young scientists and for the inspiration he has given me more directly in the study of solidifying systems. I am grateful to Joseph Chung, Herbert Huppert, and Tim Schulze for their helpful comments on an earlier draft of this article. References 1. J. S. Wettlaufer, M. G. Worster, and H. E. Huppert, J. Fluid Mech. 344, 291 (1997). 2. M. G. Worster, in Interactive Dynamics of Convection and Solidification, eds. S. H. Davis, H. E. Huppert, U. Muller, and M. G. Worster, Kluwer (1992). 3. M. G. Worster, Ann. Rev. Fluid. Mech. 29, 91 (1997). 4. H. E. Huppert and M. G. Worster, Nature 314, 703 (1985). 5. M. G. Worster, J. Fluid Mech. 237, 649 (1992). 6. 0 . E. Phillips, Flow and Reactions in Permeable Rocks, Cambridge University Press, Cambridge, U.K. (1991). 7. G. Amberg and G. M. Homsy, J. Fluid Mech. 252, 79 (1993). 8. D. M. Anderson and M. G. Worster, J. Fluid Mech. 302, 307 (1995). 9. M. C. Flemings, Solidification Processing, McGraw Hill (1974). 10. P. W. Emms, D.Phil. Thesis, Oxford University, Oxford, U.K. (1993). 11. R. C. Kerr, A. W. Woods, M. G. Worster, and H. E. Huppert, J. Fluid Mech.
Interfaces
on All Scales During Solidification
and Melting
201
216, 323 (1990). 12. M. G. Worster, J. Fluid Mech. 167, 481 (1986). 13. T. P. Schulze and M. G. Worster, J. Fluid Mech. 388, 197 (1999). 14. T. P. Schulze and M. G. Worster, in Interactive Dynamics of Convection and Solidification, eds. P. Ehrhard, D. S. Riley, and P. H. Steen, Kluwer (2001). 15. P. H. Roberts and D. E. Loper, in Stellar and Planetary Magnetism, ed. A. M. Soward, Gordon k Breach (1983). 16. M. G. Worster, J. Fluid Mech. 224, 335 (1991). 17. A. C. Fowler, Mathematical Models in the Applied Sciences, Cambridge University Press, Cambridge, U.K. (1997). 18. T. P. Schulze and M. G. Worster, J. Fluid Mech. 356, 199 (1998). 19. C. A. Chung and M. G. Worster, J. Fluid Mech., in press.
This page is intentionally left blank
PHASE A N D MICROSTRUCTURE SELECTION IN PERITECTICS
WILFRIED KURZ AND STEPHANE DOBLER Swiss Federal Institute of Technology Lausanne, 1015 Lausanne EPFL, Switzerland Peritectic alloys show a great variety of solidification microstructures. In recent years analytical theories of dendritic growth and of transient growth with nucleation ahead of the moving interface have been applied to describe the essentials of phase and microstructure selection. Coupled peritectic two-phase growth was also recently observed. Among the manyfold microstructures; banded structures, controlled by nucleation, convection, and solute trapping are very important in peritectics. Finally, consecutive phase transformations (liquid/solid and solid/solid) take place and lead to two transformation fronts, each one having its own growth behaviour. This multitude of structures constitutes a great challenge for theoretical analysis.
1
Introduction
Peritectic alloys are technologically very important. Alloys such as steels, bronzes, Al-alloys, permanent-magnet materials, High-Tc superconductors, etc., play a major role in present and future devices and machines. During the solidification and solid-state transformation of peritectic alloys a great variety of microstructures are formed. These can be steady-state microstructures or oscillating structures, singlephase microstructures or two-phase in-situ composites from stable or metastable phases. In certain compositional ranges, consecutive phase transformations over a large temperature interval, including liquid/solid and solid/solid transformations, can be observed. Despite the complex interface response of this technically very important class of alloys and their great scientific interest, many of the abovementioned phenomena have not received the necessary attention. An overview of the major observations and physical reasons behind the formation of the various interface morphologies will be given. It is hoped that this will stimulate future research into the physics and mathematics of peritectic phase transformations. The following subjects (limited to directional growth) will be treated: growth of steady-state interfaces, such as dendrites, cells, and in-situ composites; phase competition (stable to metastable phase transition); oscillating interfaces (banding controlled by nucleation, convection, and solute trapping); and consecutive liquid/solid and solid/solid transformations. Finally, some open questions will be presented. 2
S t e a d y - S t a t e Interfaces
There are several steady-state interface morphologies that have to be considered for the discussion in the following sections: • single-phase plane fronts, • cells,
203
204
W. Kurz & S. Dobler
• dendrites, • in-situ composites. The first three growth morphologies are single phase and are the result of different growth conditions. For a given temperature gradient, G, an increase of the interface velocity, V, will change the growth morphology from planar to cellular to dendritic. At higher rates, the growth morphology will again produce cellular structures and, at rates where the atom attachment kinetics become dominant, high-velocity bands form. At even higher interface velocities, a plane front is stabilized again. The stability of solidification fronts has been analysed in much depth by Davis and co-workers. For example, Huntley and Davis [1] show the limits of morphological and oscillatory instabilities for directionally-solidified Al-Fe alloys (Fig. la). Their results compare well with observations obtained in rapid laser solidification experiments. From this diagram, it can be seen that for one composition, four intersections with the solutions of the stability analysis are produced. These correspond to i) the transition from a morphologically stable to an unstable plane front, Vc, ii) beginning of (not observable) solute-trapping-controlled plane-front oscillation, V , hi) absolute stability, Va, and iv) the upper limit of the oscillating regime, VMIt is interesting to compare these results with dendrite growth theory. Fig. l b shows the interface response in the form of a T\ — V curve as calculated with the IMS (Ivantsov-Marginal-Stability) model [3,4]. In this figure, the above mentioned transition velocities are also indicated. From a comparison of both approaches, stability analysis and dendrite modelling, the complementarity of the results of these models can be seen. The interface response shown in Fig. l b allows semiquantitative modelling of the different phase and microstructure transitions (see below) that are for the moment too difficult to be analysed quantitatively, although Karma and coworkers [5] have produced some interesting first results. Coupled, two-phase growth of in-situ composites is another interesting phenomenon of peritectic alloys (see Fig. 2). This has been clearly shown only recently [6-8] and was a result of the fact that the alloys used had a very narrow solidusliquidus interval and the low-velocity limit of morphological stability was still experimentally attainable. These structures closely resemble eutectic composites. 3
Phase Competition
In peritectic alloys, generally two solid phases, one stable, the other metastable, compete with each other. The transition from one phase to the other is controlled by nucleation and transient growth phenomena. Present day modelling of the complex phase-selection processes uses: (i) steady-state growth theory to determine the interface response; and, (ii) for directional growth, a maximum temperature criterion, assuming that nucleation is easy enough not to limit the process. One proceeds as follows [4,9]. 1. Determine the interface response (Fig. lb) of all possible phases (i = a, /3, 7, ...) for a fixed temperature gradient and composition T*=T(V)i\G,Co-
(1)
Phase and Microstructure
Selection in Peritectics
205
(a) banding - 9xp€rimtnts • (Crimaud tt ul.) 0.01
0.001 L - — — 10* 10"'
10"4
10"'
10 - '
10*
10°
10 1
10;
950 C =lat% 0
^vJjiL,^
900 • •
Ad(v) r
/
850
/ a
800
/
750
1
(b)
T p (V) /
H
/ . /
V r
700 10"4
10-5
\<3*
10'3
10'2
10-'
10°
10'
^ m
^M
102
Velocity (m/s)
C
D
C
Figure 1. Stability ranges for various microstructures (plane front, P; cells/dendrites, C/D; and high-V bands, B) for Al-Fe alloys and a thermal gradient of 5000 K / m m . (a) Instability diagram from Huntley and Davis [1] showing morphological and oscillatory instability windows including results from laser experiments by Gremaud et al. [2]. (b) Interface response for plane front, T P (V), and dendritic/cellular growth, T^{V) for 1 at.% Fe.
2. Apply a selection criterion between the competing phases and growth morphologies assuming abundant nucleation of all phases so that the observed, phase-growth morphology, P, is given by the maximum of its interface temperature. P = max {T?},i =
a,/3,-y,....
(2)
206
W. Kurz & S. Dobler
Ss
•I?
x»
US
&o
u
Figure 2. Longitudinal section of a coupled, lamellar microstructure in Fe-4.22 at.% Ni at the quenched interface, where the S phase is dark and the 7 phase is a light colour (V = 10 /xm/s, G = 1.2 x 10 4 K/m).
Fig. 3 shows the phase selection procedure for the peritectic Fe-4.3 at.% Ni alloy where the stable 6 phase competes with the metastable 7 phase. The corresponding phase diagram is shown in Fig. 3a with stable and metastable phase equilibria. The interface response is shown in Fig. 3b. It can be seen that, depending on the velocity, the phase and the morphology having the maximum growth temperature varies. Considering only the maximum growth temperature criterion, the growing phase/microstructure is 7 plane front at low solidification velocities. At intermediate velocities 6 cells/dendrites are observed, then 7 dendrites/cells are seen at slightly higher velocities. Finally, the 7 plane front returns at the highest solidification velocities. The selected phase is not necessarily the thermodynamically most stable phase, but is a result of both equilibrium and growth kinetics. This approach is however incomplete and does not always lead to the right answer. A Nucleation-Constitutional-Undercooling (NCU) criterion was therefore developed by Hunziker et al. [10]. Due to nucleation of 8 phase ahead of the interface, 7 plane front is not stable at low velocity, even though it has the maximum growth temperature. The NCU criterion considers the solute concentration profile in front of the solidification front and deduces the equilibrium liquidus temperature profile of the metastable phase. The growing phase may be destabilized by nucleation of the metastable phase when there is a constitutionally undercooled zone, and the nucleation undercooling, ATn, is small enough. This may also lead to an oscillating interface (see the next section).
Phase and Microstructure
Selection in Peritectics
207
1794 1792 & 1790 1788
1 7 8 6
„
3
3.5
c* 4
^
cr45
5
55
6
Nt [at%] 1794
•i
(b)
1792 & 1790 H 1788 1786 10"*
t0" s
10*4
I0a 102 VJm/s]
10-'
10°
I0 1
Figure 3. Microstructure/phase selection in Fe-4.3 at.% Ni alloy, (a) Fe—Ni phase diagram (solid curves are stable, dashed curves are metastable). (b) Interface temperature as a function of the solidification velocity for 6 and 7 phases for G = 1.75 x 10 4 K / m .
4
Oscillating Interfaces
In peritectic alloys oscillating interfaces often produce alternating microstructures or phases. Three kinds of such interfaces have been observed forming banded structures by control through: nucleation, convection, and solute trapping. 4-1
Nucleation-Controlled
Banding
Such banding can be obtained at low velocities (V < Vc) when the initial composition, Co, is in the range delimited by both solid phases C* and C 7 in Fig. 3a. The 6 phase will nucleate first and develop a plane front. An initial transient is established and the solid-liquid interface temperature approaches progressively T*(Co). Below the peritectic temperature Tp > J"/(Co), the 7 phase will nucleate. Then 7 overgrows 6. During the new transient to T 7 (Co), the solid-liquid interface temperature rises above the peritectic temperature, and the 6 phase may nucleate again [11]. This cycle of alternating 6 and 7 phases forms a banded microstructure that is shown in Fig. 4a [12,13]. These bands appear at low velocities and therefore convection is generally present except in cases where the crucible has a very small
208
W. Kurz & S. Dobler
0.6 mm
(5 mm
(a)
(IJ)
IUU
nm
(C)
Figure 4. Banded microstructures (growth direction up), (a) Nucleation-controlled banding in Sn-0.9 wt.% Cd alloy, V = 3 /im/s and G = 2.3 x 10 4 K / m . Alpha phase is dark and the beta phase is a light colour, (b) Convection-controlled banding in a Sn-0.075 wt% Cd alloy, V = 4 / a n / s and G = 2.3 x 10 4 K / m [13]. (c) Solute-trapping-controlled banding where cellular morphology is dark and plane front is a light colour [2,19].
diameter (as is the case in Fig. 4a) [14].
4-2
Convection-Controlled Banding
In directional solidification experiments with radial heating, convection is induced by lateral thermal gradients. The direction of flow close to the solid-liquid interface depends on the position and the number of convection rolls formed, which is related to the dimension and geometry of the crucible. For the simplest case of two convection rolls, the rejected solute is transported from the center to the edge inducing a lateral chemical gradient. Thus, the liquid in the center becomes poorer in solute whereas the edge is enriched. If 6 phase is growing first, then 7 phase can nucleate at the edge. A tree-like structure is obtained by the repeated lateral movement of the three-phase junction (6, 7, liquid) due to convection effects (see Fig. 4b) [13,14].
Phase and Microstructure
4-3
Solute-Trapping-Controlled
Selection in Peritectics
209
Banding
In this case there is no phase competition, but competition of two growth morphologies such as dendrites/cells and plane front of the same phase. This can happen in all alloys including peritectics. In Fig. lb, the T(V) relation has a positive slope between the interface temperature minimum, Vm and the interface temperature maximum, VM- The positive slope is due to solute trapping. Coriell and Sekerka [15] and later Davis and coworkers [1,16,17] showed that for plane-front growth, solute trapping leads to oscillatory instabilities. Karma and Sarkissian [18] showed that the oscillatory instability induces large-amplitude relaxation oscillations at a critical velocity. These planar perturbations increase the solidification interface velocity as less solute is rejected and therefore the front accelerates to a velocity where the plane front is morphologically stable. To conform with the thermal field the solidification velocity decreases, developing again a cellular morphology. An oscillating structure of plane front and dendrites/cells is then obtained as shown in Fig. 4c [2,19]. 5
Consecutive Transformations
For alloys having a chemical composition lower than C$, two consecutive phase transformations are observed in directional transformation. Fig. 5 shows both transformation fronts where the upper front is formed by S plane-front solidification, and the lower front is formed by cellular 6 to 7 solid-state transformation. As discussed above, the growth morphology for liquid/6 solidification evolves with increasing solidification velocity from plane front, to cells, to dendrites, to cells, and finally again to plane front at very high solidification velocities. Fig. 6 shows the growth temperature for the various structures as a function of the solidification velocity. At the solidification velocity of 5 /zm/s, the expected morphology is plane front, which is actually observed (see Fig. 5). In the case of the 6/7 solid-state transformation, the diffusion coefficient of solute (Ni) is much lower in 7 iron than in 6 iron. This means that during the solidstate transformation there is a solute pile-up at the transformation front similar to solidification, i.e., volume diffusion in the parent phase. As the solute diffusion coefficient is much smaller in S than in liquid, the interface response function for solid-state transformation is displaced to smaller velocities. Therefore a "high velocity" cellular solid-state transformation front is observed behind a low solidification velocity plane front [20]. This is shown in Fig. 6, where the interface responses for solidification and for solid-state transformation are given. Due to the low absolute stability velocity in solid-state transformations, this phenomenon can be easily followed. Comparison with theory shows reasonable agreement [20]. 6
Conclusions
In summary, analysis of solidification in peritectic alloys is difficult due to the fact that two phases compete with each other. Thus, the microstructure selection
210
W. Kurz & S. Dobler
3
Liquid
\
/
^
^
<»)
Composkiois
.L/5 (b) &7
Figure 5. Consecutive transformation, (a) A schematic of the liquid/6 plane front followed by the cellular 6/7 solid-state transformation. The related phase diagram is also shown, (b) The micrograph is for V = 5 /um/s and Co = 2.75 at.% Ni and shows the two transformation fronts growing with G = 1.15 x 10 4 K / m at a distance of 2.25 mm [20].
depends on the phase and the growth morphology. Four cases can be encountered in the phase-selection process: i) single phase, ii) alternately-growing phases, iii) simultaneously-growing phases, and iv) solid-state transformation. • Single phase: The domain of stable growth of a phase has to be treated by considering nucleation in a constitutionally undercooled zone. Single phase bands of two oscillating growth morphologies can appear at high solidification velocities. • Alternating growth of two phases: A nucleation-controlled, banded two-phase microstructure is observed when neither of the two phases are stable relative to the nucleation of the other phase during transient solidification. Each phase nucleates alternately in front of the other growing phase. Convection-controlled banding occurs when large local variations of chemical composition are induced by fluid flow. • Simultaneous two-phase growth: Simultaneous two-phase growth has been observed with different morphologies: isothermal lamellar and fibrous growth, and coupled growth of cells and plane front. This growth requires very high G/V values to stabilize plane front growth of both phases.
Phase and Micro structure Selection in Peritectics
211
1800 Plane Front |
1799
- \
' Cell
1
Dendrite Cell • Plane Frori
: '
; Needle
:
<
1798
1« 1
1797
'/
\
' / 1
'5 •
\
/
\
"'/
.
1 >•
\
!
1796
L-->5
^1
\
:
1795 1770
to II 1
:
Plate '
j
:
1766 1762
L 1758 1754
:
V8^
1
1 V8^
\
:
1750 10-9
lO-7
10"5
10"3
10"'
101
V [m/s] Figure 6. Interface responses for liquid/6 and 6/7 transformations for Fe-2.9 at.% Ni and (?liqUid/6 = 1.2 x 10 4 K/m, G^z-f = 1.5 x 10 4 K/m. The unstable solidification front forms needle-like cells while the cells in the solid state are plate-like due to crystal anisotropy effects
[20].
• Solid-state transformation: In consecutive transformation experiments two interfaces can be observed. Analysis of cell morphology, cell spacing and of absolute stability have been undertaken and lead to new insight in solid-state transformation mechanisms. Despite recent progress in the analysis of peritectic solidification and solid-state transformation, there are some important open questions that are worth further detailed study. Solidification • Nucleation-controlled phase selection, including convection
212
W. Kurz & S. Dobler
• Isothermal and non-isothermal coupled in-situ composite growth • Coupled growth instabilities • Lamella to fiber transition in coupled growth • Transition from banding to coupled growth • Nucleation- and convection-controlled interface oscillations Solid-State Growth • Interface-energy anisotropy effects • Stress effects • Transformation with interface diffusion, with product-phase diffusion • Solute trapping at solid/solid interfaces • Solid-state interface oscillations • Massive transformations References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20.
D. A. Huntley and S. H. Davis, Acta Metall. Mater. 41, 2025 (1993). M. Gremaud, M. Carrard, and W. Kurz, Acta Metall. Mater. 39, 1431 (1991). W. Kurz, B. Giovanola, and R. Trivedi, Acta Metall. Mater. 34, 823 (1986). R. Trivedi and W. Kurz, Inter. Mater. Rev. 39, 611 (1994). A. Karma and A. Sarkissian, Metall. Mater. Trans. A 27, 635 (1996). J. H. Lee and J. D. Verhoeven, J. Cryst. Growth 144, 353 (1994). M. Vandyoussefi, H. W. Kerr, and W. Kurz, Acta Mater. 48, 2297 (2000). T. Okane and T. Umeda, submitted to Acta Mater. W. Kurz and P. Gilgien, Mater. Sci. Eng. A 178, 171 (1994). O. Hunziker, M. Vansyoussefi, and W. Kurz, Acta Mater. 46, 6325 (1998). R. Trivedi, Metall. Mater. Trans. A 26, 1583 (1995). W. Kurz and R. Trivedi, Metall. Mater. Trans. A 27, 625 (1996). W. J. Boettinger, S. R. Coriell, A. L. Greer, A. Karma, W. Kurz, M. Rappaz, and R. Trivedi, Acta Mater. 48, 43 (2000). A. Karma, W.-J. Rappel, B. C. Fuh, and R. Trivedi, Metall. Mater. Trans. A 29, 1457 (1998). S. R. Coriell and R. F. Sekerka, J. Cryst. Growth 61, 499 (1983). D. A. Huntley and S. H. Davis, J. Cryst. Growth 132, 141 (1993). D. A. Huntley and S. H. Davis, Phys. Rev. B 53, 3132 (1996), A. Karma and A. Sarkissian, Phys. Rev. E 47, 513 (1993). M. Carrard, M. Gremaud, M. Zimmermann, and W. Kurz, Acta Metall. Mater. 40, 983 (1992). M. Vandyoussefi, H. W. Kerr, and W. Kurz, Acta Mater. 45, 4093 (1997).
MODEL P H A S E D I A G R A M S FOR A N FCC ALLOY R. J. BRAUN Department
of Mathematical Sciences, Newark, DE 19716,
University U.S.A.
of
Delaware,
J. Z H A N G Department of Engineering Sciences and Applied Northwestern University, Evanston, IL 60208,
Mathematics, U.S.A.
J. W . C A H N A N D G. B . M C F A D D E N National
Institute of Standards and Technology, Gaithersburg, MD 20899, U.S.A. A. A. W H E E L E R
Faculty
of Mathematical Studies, Highfield, Southampton,
University of Southampton, S017 1BJ, U.K.
We describe a mean field, free-energy model that allows the computation of realistic phase diagrams for a particular set of ordering transitions in face-centered-cubic (FCC) binary alloys. The model is based on a sixth-order Landau expansion of the free-energy function in terms of the three nonconserved order parameters that describe ordering on the underlying FCC lattice. When combined with appropriate gradient-energy terms, this model will allow the self-consistent calculation of energetic and kinetic anisotropies of phase boundaries in future work.
1
Introduction
Equilibrium states of crystalline solids can undergo a variety of phase transitions as the temperature and concentration of the material is varied [1]. In this paper, we will be concerned with order-disorder transitions for a face-centered-cubic (FCC) binary alloy that is composed of A and B atoms. In such a transition, the structure and spacing of the underlying lattice remains unchanged, but the arrangements of the A and B atoms on the lattice sites can vary with temperature and concentration. For this purpose, the FCC lattice can be regarded as consisting of four interpenetrating simple cubic sublattices. The origin of one sublattice can be taken as the corner site of a unit cell and that of each of the other three sublattices is centered on one of the faces (see Fig. 1). At high temperatures, but below the melting point, the equilibrium state is typically a disordered phase, where the ratio of A and B atoms on each sublattice is the same. The structure is FCC and the four sublattices are equivalent by the symmetry of the FCC lattice. At lower temperatures, ordered phases can be favored, in which case there is a symmetry breaking. For the bulk-ordered phases we consider here, the ratio of A and B atoms are uniform on each sublattice, but the ratio varies from sublattice to sublattice. An example of such an ordered phase occurs in the copper-gold system, in which the CU3AU phase at concentrations near 25% Au consists of gold atoms preferentially occupying one sublattice, which for convenience is taken to be the one at the corner of the unit cell. Since the gold enrichment could have occurred on any of the sublattices, there
213
214
R. J. Braun et al.
are four variants. The symmetry of the ordered structure has become that of a primitive cubic lattice. Then, the copper atoms preferentially occupy the remaining three sublattices, a lattice complex composed of the cell faces, in which these three sublattices remain equivalent by symmetry. The CU3AU phase is an example of LI2 ordering in the Strukturbericht notation system [2]. Another example that we consider here is the Llo phase, which represents an ordered phase such as CuAu at concentrations near 50% Au; in this case the symmetry breaking splits the four sublattice into two pairs of two equivalent sublattices, and the symmetry of a primitive-tetragonal lattice. The ordering results in alternating planes of Cu and Au normal to one of the original cube axes, with six possible variants. In the original FCC unit cell the Cu preferentially occupies the corners and one of the sets of faces, and Au the other two sets of faces. Interfaces can occur between any two variants of the same phase over the range of temperature and concentration where this phase is stable. Interfaces can also occur between any two phases at temperatures where these phases coexist, with the concentrations of the phases dictated by equilibrium conditions as depicted by phase diagrams. In previous work, we have examined gradient energy formulations of the diffuse interfaces produced by such order-disorder transitions in FCC alloys [3-5] and HCP alloys [6], the interphase boundaries (IPBs) between an LI2 phase and the disordered FCC phase, and antiphase boundaries (APBs) that separate different variants of the same ordered phase. The formulation permitted the examination of the orientation dependences of the free energy of these interfaces and their mobility. The derivation of these models focused on obtaining a continuum, free-energy functional with a form of the gradient-energy terms that is consistent with the underlying symmetry of the atomic lattice. However, we employed a simplified bulk-free-energy density that, while it permits an easy calculation of the interface properties, has long been known to give an unrealistic phase diagram [7]. Efforts to compute a phase diagram for an FCC binary alloy date back to Shockley and coworkers [7]; the diagram that resulted from the Bragg-Williams approximation used there was topologically different from, and a poor approximation to, the experimental phase diagram for the prototype Cu-Au system [8]. Many efforts to improve upon the situation followed [2,9,10]; realistic phase diagrams for the Cu-Au system were obtained by using the cluster variation method (CVM) and asymmetric multiparticle interactions [11]. The CVM and Monte Carlo (MC) methods have had a number of successes modeling alloy systems of varying complexity; the reader is referred to recent reviews [2,12] for discussion of those methods. Thus, in the last twenty years it has been possible to obtain a realistic phase diagram from statistical mechanical models, particularly for the CVM [2,9,11,13]. However, CVM models are cumbersome and their existence is often ignored [14]. They have been used to make discrete calculations of the energy of the interfaces (e.g., [13,15]). But such discrete calculations of interfaces do not lend themselves to studying the orientation-dependent properties. In this paper, we describe an improved model for the bulk-free-energy density that allows a more realistic description of the equilibrium phase diagram of the system, but still simple enough to be used for describing the orientation dependence of the energy and motion kinetics of the interfaces. It is our hope that its simplicity will find a wide variety of
Model Phase Diagrams for an FCC Alloy
215
Figure 1. A schematic diagram of an FCC lattice. To describe ordering in this model there are four distinguished sites corresponding to a corner and one each for the faces intersecting at that corner.
applications. Our method has elements in common with the sublattice models developed by Sundman, Dupin and others [16-19]. These models have been used successfully to produce realistic phase diagrams for FCC alloy systems. Their method features many more parameters than our model; in practice, these parameters are determined by fitting experimental measurements of the phase diagram and other thermodynamic quantities, such as the latent heat and heat capacity. Here, we show that realistic phase diagrams may also be obtained using our simple model; we anticipate that this simplicity will ultimately be advantageous in the self-consistent calculation of energetic and kinetic anisotropies of phase boundaries. The paper is organized as follows. The formulation is given in Sec. 2; the method for finding phase diagrams and some results are given in Sec. 3. Finally, some discussion and conclusions are presented in the last two sections. 2 2.1
Formulation The Concentration and Order Parameters
We briefly recall the mean-field description of the order-disorder transitions on an FCC lattice given in Ref. [4]. A binary alloy (denoted A-B) on an FCC crystal lattice is described geometrically by four interpenetrating cubic sublattices, with sublattice occupation densities pj defined at the lattice points shown in Fig. 1. The four densities represent the local atomic fraction of species A on each sublattice; their specification is assumed to characterize the overall state of the crystal. It is convenient to introduce four new parameters W, X, Y, and Z defined by W
= - ( p i + p2 + P3 + PA) ,
X = ~(pl+P2~P3-
Pi) ,
(1) (2)
216
R. J. Braun et al.
Y=-(Pl-p2+P3-
p4),
Z = - (pi - Pi - P3 + PA) •
(3) (4)
These relations can be inverted to give px = W + X + Y + Z,
(5)
p2 = W + X - Y - Z,
(6)
p3 = W-X
(7)
+ Y-Z,
p4 = W - X - Y + Z,
(8)
which can be interpreted in terms of the expression [14] p = W + X cos 2-Kx/a + Y cos 2iry/a + Z cos 2irz/a,
(9)
which gives the relations (5-8) upon evaluation at the corners and face centers of the conventional unit cell in Fig. 1; here a is the cubic-lattice constant. Thus, parameter W represents the atomic fraction of the system as a whole and is a conserved order parameter. X, Y, and Z are non-conserved order parameters that can vary between plus and minus one half, and are Fourier coefficients representing density variations along the directions of the crystal axes. In this model the disordered state is represented by px = P2 = Pz = PA = W, which implies that X = Y — Z = 0. Ordered states are characterized by non-zero values for the non-conserved order parameters. For instance, the four equivalent variants of A 3 B L l 2 ordering are |X| = \Y\ = \Z\ ^ 0 with XYZ < 0, and the four variants of AB 3 Ll 2 ordering are \X\ = \Y\ = \Z\ ^ 0 with XYZ > 0. The six variants of L l 0 ordering are X = Y = 0 with \Z\ ^ 0, Y = Z = 0 with \X\ ^ 0, and Z — X = 0 with \Y\ 7^ 0, with layering respectively in the xy, yz, and zx planes. A bulk equilibrium state is characterized by constant values of the densities and order parameters. 2.2
The Thermodynamic
Description
A thermodynamic description of the crystal for the case of an isothermal system is based on the bulk Helmholtz free energy density (per mole) F(X,Y,Z,W,T). In Ref. [4], we considered an energy model based on pair-wise interactions that leads to an unrealistic phase diagram with a multiphase critical point. The problem lies in the positive definiteness of the fourth-order terms in the expansion of the entropic part of F. Pair-wise interaction energies make no contribution to these fourth-order terms. In the present work, we develop an improved model for the freeenergy density, by introducing negative fourth-order contributions to the energy to make the fourth-order free-energy terms no longer positive definite. We are then able to obtain a more realistic description of the binary-alloy phase diagram. The idea behind our modification can be motivated by familiar examples from the bifurcation theory of scalar systems [20]. For a fourth-degree, free-energy density at W = 1/2, the Llo ordering is a second-order phase transition in which the order parameter changes continuously from zero to non-zero values as the temperature passes through its critical value Tc. A scalar model with Y — Z — 0 for the
Model Phase Diagrams for an FCC Alloy
(a)
217
(b)
Figure 2. Schematic bifurcation curves for the order parameter X versus the temperature T near a critical point Tc. Stable and unstable solutions are indicated by solid and dashed curves, respectively, (a) A fourth-degree theory results in a second-order phase transition: the disordered state (X = 0) changes continuously into the ordered state (X ^ 0) at Tc. (b) A sixth-degree theory results in a first-order phase transition: locally stable ordered and disordered states exist on either side of To and have equal energies at To (dotted curve). The order parameter X(T) for the minimum energy state jumps discontinuously as the temperature passes through To.
free-energy density of the Llo transition is F(W, X, T) = f0(W, T) + f2(W, T)X2 + f4{W,
T)X\
(10)
with / 4 (1/2,T) > 0 and / 2 (1/2,T) ~ (T - Tc). For T > Tc the system is in stable equilibrium for the disordered state X = 0, whereas for T < Tc the lowest energy state is an ordered phase with X2 = —f2/(2/4)- The ordered and disordered states are never simultaneously stable, as illustrated in Fig. 2a. A first-order phase transition can be obtained from a sixth-order model such as F(W, X, T) = f0(W, T) + f2(W, T)X2 + f4(W, T)X4 + f6(W,
T)Xe
(11)
where / 2 (1/2,T) ~ (T - Tc) as before, but now / 4 (1/2,T) < 0 and / 6 (1/2,T) > 0. In this illustrated in Fig. 2b, the minimum energy state changes discontinuously from a disordered phase to an ordered phase with X2 = —f4/(2fe) a t the temperature T0 > Tc for which f2 = ft/(^fo)A simplified model [4] that includes the LI2 transition with |X| = \Y\ = \Z\ ^ 0 involves, in addition, a cubic term fo(W,T)XYZ. The coefficient fz(W,T) of the third-order term in the free-energy expansion depends strongly on the composition, and in a symmetric model is antisymmetric about W = 1/2 and equals 0 there. At compositions on either side of W = 1/2, the sign of XYZ can always be chosen to make the term fa(W,T)XYZ negative, with X = ±Y = ±Z. Thus, in a fourthdegree theory, LI2 is always favored over Llo near the critical point Tc, except exactly at W = 1/2. In a sixth-degree theory with negative-definite fourth-order terms, the Llo transition is first order and has a critical point at To > Tc. The
218
R. J. Braun et al.
energy differences between the Llo and LI2 phases at the lower temperature Tc are then unimportant, and the Llo phase can be favored over the LI2 phase near W = 1/2 and T = T0. The free-energy model used in our previous work featured a fourth-degree polynomial expansion with positive definite fourth-order coefficients. This expansion was adequate for the description of the LI2 transition near W = 1/4, but necessarily lead to a multicritical point at W = 1/2 and a second-order phase transition for the Llo structure, whose temperature dependence is similar to Fig. 2a. Our improved models in this paper are based on four-atom interactions in the energy which add negative fourth-order terms of a large enough magnitude to make the fourth-order term negative in a polynomial expansion at W = 1/2. As before the sixth-degree terms result entirely from the entropy and are positive. The main goal of the present work is to extend the model to higher degree in order to obtain a more realistic treatment of all the transitions, and to find appropriate choices for the dependence of the free-energy density on temperature and concentration so that realistic phase diagrams for these order-disorder transitions can be obtained. There are enough parameters in a four-atom interaction model to affect the behavior when W ^ 1/2. In the next two subsections, we describe the models of the internal-energy density E and entropy S that are used to obtain the improved free-energy density F — E — TS that we use in this work. 2.3
The Internal Energy
The internal energy per mole, E, of the system is assumed to be characterized by four-atom cluster energies E40, E31, E22, £13, and £04 of the various occupations of the near-neighbor basic tetrahedral form by A4, A3B, A2B2, AB3, and B4, respectively, that allow for four-particle interactions according to the scheme [11]: E = £4O01020304 + E31 [p!p2P3(l - Pi) + 01/02(1 - 03)04 + 0l(l - 02)0304 + (1 - 0l)020304] + #22 [0102(1 - p 3 )(l - p4) + pi(l - 02)03(1 - 04) + 0l(l - 02)(1 - 03)04 + (1 - 0l)(l - 02)0304 + (1 - 0l)02(l - p3)04 + (1 - 0l)0203(l - p4)] + #13 [0l(l " 02)(l - 03)(l " 04) + (1 - 0l)02(l - p 3 )(l " 04) + (1 " 0l)(l - 02)03(1 " 04) + (1 - 0l)(l " 0 2 )(1 - 03)04] + (12)
£O4(l-0l)(l-02)(l-03)(l-04).
By substituting Eqs. (5-8) into the above expression, we obtain an internal energy of the form E = e 0 + e2(X2 + Y2 + Z2) + e3XYZ e42(X2Y2
+ X2Z2 + Y2Z2),
+ e 4 i(X 4 + F 4 + Z4) + (13)
where the e^ are given in the appendix in terms of £ 4 0 , E31, E22, £13, and E04.
Model Phase Diagrams for an FCC Alloy
2-4 Entropy
219
Approximations
The point approximation to the molar entropy S of the system is given by the expression R 4 S(p1,p2,P3,PA} = --^2l{pj),
(14)
3= 1
where J(a;)=xln(x) + ( l - a ; ) l n ( l - a ; ) ,
(15)
and R is the universal gas constant. This expression for the entropy is the sum of the contribution from each sublattice. We explore two choices based on Eq. (14). One option is to use the expression as is, which is somewhat complicated by the presence of the logarithmic terms, which generally forces one to compute numerical solutions to the problem. Another option is to expand this expression in a power series about the points pj = 1/2, and truncate the series after sixth degree. The result allows more progress to be made analytically. We obtain
(16) Inserting the definitions (5-8) into this expression results in polynomial expressions involving U — W — 1/2 and the order parameters X, Y, Z up to sixth degree. 2.5
Free-Energy Approximations
The form of the Helmholtz free-energy density, F = E — TS, follows from Eq. (13) and is given by an expression of the form F(X, Y, Z, W, T) = e0 + e2(X2 + Y2 + Z2) + e3XYZ e42(X2Y2
+ eAl(X4 + YA + ZA) + +X2Z2
2 2
+Y
Z)
+ ^Yjl{p3),
(17)
where we must use Eqs. (5-8) to get the entropy contribution in terms of the order parameters. Using Eq. (16), the resulting approximation for the free energy takes the form F(X, Y, Z, W, T) = f0 + f2(X2 + Y2 + Z2) + f3XYZ + f41(X4 + Y4 + Z4) + f42(X2Y2 + X2Z2 + Y2Z2) + f51XYZ(X2 +Y2 + Z2) + f61(X6 + Y6 + Z6) + 4 2 f62{X (Y + Z2) + Y4(X2 + Z2) + Z4(X2 + Y2)} + f63X2Y2Z2, (18)
220
R. J. Braun et al.
where the coefficients /;, for parameter choices suggested by those of Kikuchi and de Fontaine [11] discussed above, are given in the appendix. In Sec. 4.1 we shall give the results for a phase diagram using this polynomial approximation. 3
Phase Diagrams
A phase diagram consists of curves in the temperature-concentration plane that describe conditions for equilibrium between various bulk phases at the same temperature, but not necessarily at the same concentration. Mathematically, this is a minimization of the free energy subject to a fixed amount of the total amount of concentration W in the system, which results in a common tangent construction in terms of the free energy of the system [21]. The free energy is also an unconstrained minimum with respect to the non-conserved order parameters Xj. This procedure produces several sets of equations to solve for the concentration between stable bulk phases that delineate regions where two or more phases can coexist. To describe the computation of phase diagrams, we first discuss the bulk equilibrium states that are supported by the model. 3.1
Bulk Phases
In the case of the disordered FCC phase, we have X = Y = Z = 0; for high enough temperatures this will occur for any overall concentration W. We denote the resulting free energy for the bulk FCC phase by FFcc(W,T); for the case in Eq. (17), we then have FFcc{W, T) = F(0,0,0, W,T) = e0(W) + RT [1{W)\.
(19)
For the L l 0 variant with Z / 0, the corresponding free energy becomes
FLI0(Z,W,T)
FLl0(Z,W,T)
= F(0,0,Z,W,T) = e0(W) + e2(W)Z2 + e41Z4 R: +^-mW + Z)+I(W-Z)} ~2
(20)
for this phase. Finally, for the Ll 2 variant with X = Y = Z ^ 0 the free energy FLy2 {Z, W, T) becomes FLl2{Z, W,T) = F(Z,Z,Z,W,T) = e0(W)+ 3 e3(W)Z + 3(e4l+2eA2)Z4
3e2(W)Z2 +
^ p [X{W + 3Z) + 3X(W - Z)\. 3,2
FCC-LIQ
+
(21)
Transition
Conditions under which an equilibrium between the disordered FCC phase and the ordered Llo phase can coexist are described by the system of nonlinear equations c)F ^ ( Z 0 , W o ) = 0,
(22)
Model Phase Diagrams for an FCC Alloy 221 d
^ ( Z o , Wo) =
-^{WFCC)
= Mo,
(23)
and FFCC{WFCC)
- FLlo(Z0,
Wo) - no {WFCc - W0) = 0,
(24)
for a given T. Eq. (22) results from minimizing the free energy with respect to Z, and Eq. (23) is the result of minimizing with respect to W subject to a specified overall concentration, leading to the appearance of the Lagrange multiplier /j,0. Eq. (24) completes the common tangent construction and expresses the fact that the energy is also stationary with respect to variations in the interface position. The unknowns in these equations are the concentration of the disordered phase Wpcc a n d the concentration and order parameter of the Llo phase, WQ and Zo, respectively. These equations were solved using DNSQ [22,23] using suitable initial guesses. These guesses were generated from the To curve, which is, in turn, found by setting the free energies of the two phases equal and requiring dF/dZ — 0. This strategy was used in all cases where there were coexisting ordered and disordered states. 3.3
FCC-L12
Transition
In this case, we must satisfy the system of nonlinear equations ^(Z2,W2)
(25)
= 0,
9FLi2
dFFCc
m
,
^
(26)
and FFcc(WFcc)
- FLU (Z2, W2) - ^2
(WFCC
- W2) = 0,
(27)
for a given T. The unknowns are the concentration of the disordered phase WFcc and the concentration and order parameter of the LI2 phase, W2 and Z2 respectively. 3.4
LI0-LI2
Transition
In this final case, we must satisfy the system of nonlinear equations ^L(Z0,Wo)=0,
(28)
^(Z2,W2)=0,
(29)
^(Z0,W0)
= ^(Z2,W2)
= „lf
(30)
and FLl0(Z0,W0)
- FLl2(Z2,W2)
- /*! (Wo - W2) = 0,
(31)
222
R. J. Brawn et al.
for a given T. The unknowns are the concentrations and order parameters for the respective ordered phases. 4
Results
We next present results of numerically-computed phase diagrams using these equations. The results are given in non-dimensional form by using the energy scale — w to nondimensionalize the free energies and the temperature scale —u>/R to nondimensionalize T. 4-1
Results Using the Sixth-Degree Landau Expansion
We first describe a phase diagram based on the free-energy function given in Eq. (18), using the parameters /o(tf,T) = ~ + 6 C / 2 + T ( l n i + 2 i r 2 ) ,
(32)
h (U, T) = -4 + 40t/ 2 + 2T,
(33)
3
f3(U) = -250(U-aU ),
(34)
/4i = - 4 ,
(35)
/42 = 0,
(36)
/ei = 4,
(37)
/62 - - 3 ,
(38)
fes = 6,
(39)
where U = W - 1/2. With the choice a = 2 and (3 « 27.735894, we will have equal dimensionless temperatures of T = 2.5 at the congruent points of the FCC-L1 2 and FCC-Llo order-disorder transitions near W = 1/4 and W = 1/2, respectively. The choice of the last two parameters, f^ and /g3, in the free-energy function (18) was made so that the sixth-degree terms would drop out of the problem when X = Y = Z and leave the analysis of the bulk LI2 phase unaffected. By solving the required nonlinear equations using the sixth-degree polynomial form and these coefficients, we arrive at the phase diagram shown in Fig. 3. This phase diagram is an idealized, symmetric approximation to the Cu-Au binary system [8]. The choice —LO/R « 265 K would locate the congruent points at about the right temperatures for Cu-Au. The qualitative appearance of the phase diagram for temperatures in the vicinity of the congruent points is satisfactory, but the behavior of the system in the dilute limit is unrealistic. The situation is improved by retaining the logarithmic terms in the entropy contribution to the free-energy density, and we focus our attention on this model in the remainder of the paper. 4-2
Results Using the Free Energy with Logarithmic Terms
We next discuss examples of phase diagrams obtained by using the free-energy model given by Eq. (17) with coefficients given in Table 1. The first case is a qualitative model for the Cu-Au system, and the last two are for demonstration
Model Phase Diagrams for an FCC Alloy
2.6
223
1
fee fee /
2.4
L1 0
1*1 2.2
/
2.0 0.0
0.2
0.4
W Figure 3. A model phase diagram for the Cu-Au system from the parameters listed in Eq. (32). Note that it is symmetric about W = 1/2.
Table 1. Coefficients used for the internal-energy contribution to the free-energy density; here U = W- 1/2. Case eo{U) e2{U) e3(U) e-41
e42
I 6(7 2
- 4 + t/2
200£/(l - 2U2) -12 15
II 6(7* -A + U2 200(7(1 - 2U2) -5 10
III W'2 -4 + U2 100(7(1 - 2U2) -12 15
purposes. We note that with these coefficients, the limit of metastability for the disordered phase for W = 1/2 (corresponding to the temperature Tc in Fig. 2b) occurs at a dimensionless temperature T = 2 in all three cases. With the coefficients of case I, the phase diagram in Fig. 4 is obtained. This diagram is qualitatively similar to that for the Cu-Au system [8], but with symmetry about W = 1/2. This model diagram should be sufficient for the purposes of studying the surface-tension anisotropy of IPBs in our future work. a For the parameters of Case II, we find that the phase diagram has a lowered and dramatically smaller region where the FCC-L1 0 transition occurs; see Fig. 5 and compare to Fig. 4 for Case I. As en rises toward the value —8/3, the F C C - L I Q °In order to obtain quantitative agreement with the congruent point of the FCC-LI2 transition, we would choose —ui/R RS 248 K.
224
R. J. Braun et al.
H
2.3
0.10
0.20
0.30 W
0.40
0.50
Figure 4. A model phase diagram for the Cu-Au system for the parameters of Case I. Note that it is symmetric about W = 1/2.
transition region disappears and becomes a multicritical point at e^i = —8/3. The phase diagram for the parameters of case III is given in Fig. 6. In this case, the congruent point for the FCC-LI2 transition is just outside the FCC-L1 0 coexistence region; further decrease in e% will result in a peritectoid phase diagram. We can summarize the effects of the parameters as follows. The size of e% controls the location of the congruent point for the FCC-LI2 transition. If it is sufficiently large, then the congruent point is shifted to values symmetrically located about W — 1/2 with temperature greater than 2. The value of e^x controls the location of the congruent point for the FCC-Llo transition. If e^x < —8/3, the congruent point remains at W = 1/2 but occurs at T > 2. If e 41 > —8/3, the congruent point is at T = 2 and the transition is second order. Using these parameters we have been able to obtain satisfactory phase diagrams that would allow the description of the variation of concentration across both interphase and antiphase boundaries. 5
Discussion
The solutions to the nonlinear equations that give the curves on the phase diagram become difficult to compute when the dimensionless temperature drops below about 2 for the case with logarithmic terms in the entropy. The difficulty appears to be that the values of the order parameters and concentration are approaching the
Model Phase Diagrams for an FCC Alloy
3.0
!
'
225
1
fee 2.5
fee
-
yT
L12
1.0
0.00
V\
/
h- 2.0 -
1.5
•
I
1
0.10
0.20
\
0.30
0.40
0.50
W Figure 5. The phase diagram for the parameters of Case II. The region where Llo is the sole stable phase is near (W,T) = (0.5,2). The multicritical point at e,ji = —8/3 is being approached and the FCC—Llo transition is disappearing.
0.50
Figure 6. The phase diagram for the parameters of Case III.
226
R. J. Braun et al. 0.50
0.40
0.30 N
0.20
0.10
0.00
Figure 7. The dashed lines show the limits of the allowed ranges for the arguments of the In functions. The solid curves are plots of the order parameters found along the coexistence region boundaries for the parameters of Case I. The lower curves near the middle of the plot are for the LI2 boundaries and the curves in the upper right are for the Llo boundaries.
limits of the ranges allowed by the logarithmic terms. For example, in the freeenergy density for the LI2 phase, there is a term in the argument of the In function that requires 1-W-3Z>0;
(40)
solving for Z gives (41) From the Llo phase, we can also conclude that we must have W>Z.
(42)
We may plot the boundaries of these inequalities along with the solutions to the nonlinear equations that give the phase diagram in the (W, Z)-plane; this has been done in Fig. 7. The temperature is decreasing as the solid curves (representing the coexistence-region boundaries) approach the edges of their allowed ranges (dashed lines). The jumps in slope occur at the ends of the tie line at the eutectoid temperature for Case I. We hypothesize that when the solutions get too close to these boundaries, the iteration procedure breaks down because the procedures allow iterates beyond the boundaries. We believe that we can recast the equations to eliminate the logarithmic terms and thus alleviate this problem; this is allowed by the special form of the free-energy density.
Model Phase Diagrams for an FCC Alloy
6
227
Conclusion
We have been able to compute a simple model phase diagram with an approach that is a modification to the quasi-chemical description. The modifications were rooted in the choices of Kikuchi and coworkers [11,15] in their successful CVM approaches, but we have modified the coefficients to suit our needs in drawing the phase diagram. We found that two coefficients (e$ and e 4 i) in the internal-energy contribution to the free energy made the biggest contributions to controlling the resulting phase diagram. The coefficient e 4 i of the terms involving Xf strongly affects the temperature of the congruent point for the FCC-Llo transition; there is a multicritical point at e4i = —8/3, at which the first-order nature of the transition disappears. Elsewhere, we will show that as one approaches the multicritical point for the FCC-Llo phase transition, there is a quadratic decay in the interfacial energy of the interphase boundaries for any orientation [24]. The size of the coefficient e3 plays a crucial role in setting the temperature and concentration of the congruent point in the FCC-L1 2 transition. When it is larger, the congruent point occurs at larger temperatures. When it decreases, the congruent point disappears inside the FCC-Llo coexistence region and a peritectoid phase diagram occurs. We have not yet computed the peritectoid phase diagram due to numerical difficulties with the logarithmic terms; recasting the root-finding procedure to eliminate these terms should alleviate this difficulty. Acknowledgments It is a pleasure to dedicate this paper to S. H. Davis on the occasion of his sixtieth birthday, and to acknowledge the tremendous influence of his research and his friendship on the many students and colleagues that have had the good fortune of associating with him. The authors are also grateful for helpful discussions with S. Langer and W. J. Boettinger during the preparation of this manuscript. This research was conducted with the support of the Physical Sciences Research Division of NASA and the University of Delaware Research Foundation. Appendix If we begin with Eq. (12), and substitute the expressions (5-8) to eliminate the occupation densities pi, we obtain the expression E = ^ [ 4 ^ 3 1 + 6^22 + 4£i3 + £ 0 4 + £4o] + U[E3i - E13 - EM/2 + Em/2\ 2
+
3
[/ [3£ 4 0 /2 - 3-E22 + 3£W2] + t / [ - 4 £ 3 i + 4 £ i 3 - 2E0A + 2E40] + UA[-AEZ1 + 6£ 2 2 - 4 £ i 3 + £ 0 4 + £40] + (X 2 + Y2 + Z2){E22 - £ 4 0 / 2 - £ 0 4 / 2 + U{4E31 - 4£i3 + 2£ 0 4 - 2E40] + U [SE31 — \2E22 + 8E13 — 2E04 — 2.E40]} + XYZ{-8E31 + 8£i3 - 4E 0 4 + 4£ 4 0 +
228
R. J. Braun et al.
U[-2,2E31 + 48£ 2 2 - 32£ 1 3 + &E04 + 8E40)} + (X2Y2 + X2Z2 4
4
+ Y2Z2){8E31
4
(X + Y + Z ){-AE31
- 12E22 + SE13 - 2E04 - 2E40} +
+ 6£ 2 2 - 4E13 + E04 + E40},
(43)
where U = W - 1/2. We note that if E40 = E31 = E22 = E13 = E04 = 1, then E=\. A simple bond counting argument is based on the bond energies EAA, EAB and EBB between the A-A, A-B, and B-B pairs of atoms, respectively. If we assume that EAA = EBB and use a reference energy corresponding to EAA, then the energies of the different configurations according to this scheme are E40 = 0,
E31 = 3w,
£22 = 4w,
E13 = 3w,
E04 = 0,
(44)
where u> = EAB — EAA, SO that the configurational energies are thus characterized in terms of the energy u. Substituting these expressions into Eq. (43) gives E = EQ + AIO[X2+Y2
+ Z2].
(45)
This expression for the internal-energy density is inadequate for the purposes of drawing phase diagrams [7]. A better model for the multiparticle interactions in the internal energy, as given by Kikuchi and de Fontaine [11] and used in Kikuchi and Cahn [15], is to take E40 = 0,
E31 = 3w(l + a),
E22 =
4LJ,
E13 = 3w(l + 6),
EM = 0, (46)
where the constants a, b, and u) were determined from a best fit to the phase diagram for the case of the Cu-Au system: a = 0.01,
b = -0.08,
^ = -663 K.
(47)
With this scheme, we obtain more terms in the internal-energy density of Eq. (13), and the coefficients e* in that expression are given by e 0 = ^to(U2 - 1/4) [4 + a + b + 4(a - 6)17 + 4(o + b)U2} , e2 = 2w + QwU[a -b + 2(a + b)U], e3 = 12u(a + b)[b - a - 4(a + b)U], e4i = -6w(a + b),
e 42 = 12a;(a + b).
(48)
These expressions are used to motivate the choices made for the energy coefficients shown in Table 1. For the polynomial free energy (18), we have the coefficients /o = \w {(U2 - 1/4) [4 + a + b + 4(a - b)U + 4(a + b)U2]} +
RT Uu2 + ^U4 + ^U6 + ln(l/2)^ , h = 2u + toU[a - b + 2(a + 6)17] + RT [2 + 8U2 + 32C/4] , h = 12w[6 - a - 4(a + b)U] + RT [32C7 + 256C/3] ,
Model Phase Diagrams for an FCC Alloy
/ 4 i = -6u)(a + b) + RT
229
| + 32[/ 2
/ 4 2 = 12w(a + b) + RT [8 + 192C/2] , f51=256RTU,
/ 6 1 = 32/2T/15,
f62 = 32RT,
/ 6 3 = 192i?T.
(49)
These expressions are used to motivate the choices made for the energy coefficients shown in Eqs. (32-39). References 1. J. W. Christian, The Theory of Transformations in Metals and Solids, Part I, Pergamon Press, Oxford, U. K. (1975). 2. F. Ducastelle, Order and Phase Stability in Alloys, North-Holland, New York (1991). 3. R. J. Braun, J. W. Cahn, J. Hagedorn, G. B. McFadden, and A. A. Wheeler, in Mathematics of Micro structure, Evolution, ed. L.-Q. Chen et al., TMS/SIAM, Philadelphia, PA, p. 225 (1996). 4. R. J. Braun, J. W. Cahn, G. B. McFadden, and A. A. Wheeler, Trans. Roy. Soc. London A 355, 1787 (1997). 5. R. J. Braun, J. W. Cahn, G. B. McFadden, H. E. Rushmeier, and A. A. Wheeler, Acta Mater. 46, 1 (1997). 6. J. W. Cahn, S. Han, and G. B. McFadden, J. Stat. Phys. 95, 1337 (1999). 7. F. C. Nix and W. Shockley, Rev. Mod. Phys. 10, 1 (1938). 8. H. Okamoto, D. J. Chakrabarti, D. E. Laughlin, and T. B. Massalski, Bull. Alloy Phase Diagrams 8, 454 (1987). 9. R. Kikuchi, Prog. Theo. Phys. Supp. 87, 69 (1986). 10. R. Kikuchi and H. Sato, Acta Metall. 22, 1099 (1974). 11. R. Kikuchi and D. de Fontaine, Application of Phase Diagrams in Metallurgy and Ceramics, ed. by C. G. Carter, NBS Special Publication 496, 967 (1978). 12. D. de Fontaine, Solid State Phys. 47, 33 (1994). 13. A. Finel, in Ordering and Disordering in Alloys, ed. A. Yavari, Elsevier Applied Science, New York, p. 182 (1992). 14. A. G. Khachaturyan, Prog. Mater. Sci. 22, 1 (1978). 15. R. Kikuchi, and J. W. Cahn, Acta Metall. 27, 1337 (1979). 16. I. Ansara, B. Sundman, and P. Wilemin, Acta Metall. Mater. 36, 977 (1988). 17. N. Dupin, Contribution a devaluation thermodynamique des alliages polyconstitues a base de nickel, Ph.D. Thesis, Laboratoire de Thermodynamique et de Physico-Chimie Metallurgiques de Grenoble, Institut National Polytechnique de Grenoble (1995). 18. I. Ansara, N. Dupin, H. L. Lukas, and B. Sundman, J. Alloys and Compounds 247, 20 (1997). 19. B. Sundman, S. G. Fries, and W. A. Oates, Calphad 22, 335 (1998). 20. S.-N. Chow and J. K. Hale, Methods of Bifurcation Theory, Springer-Verlag, Berlin (1982). 21. P. Haasen, Physical Metallurgy, 3rd ed, Cambridge University Press, Cambridge, U. K. (1996).
230
R. J. Braun et al.
22. SLATEC Common Math Library, National Energy Software Center, Argonne National Laboratory, Argonne, IL. The program Snsq was written by K. L. Hiebert and is based on an algorithm of Powell [23]. 23. M. J. D. Powell, "A hybrid method for nonlinear equations," in Numerical Methods for Nonlinear Algebraic Equations, P. Rabinowitz, ed. Gordon and Breach (1988). 24. G. Tanoglu, R. J. Braun, J. W. Cahn, G. B. McFadden, and A. A. Wheeler, in preparation.
PART 2
C O N T R I B U T E D ABSTRACTS
This page is intentionally left blank
233
I N F L U E N C E OF CONTACT-ANGLE C O N D I T I O N S O N EVOLUTION OF SOLIDIFICATION F R O N T S VLADIMIR S. AJAEV AND STEPHEN H. DAVIS Department of Engineering Sciences and Applied Mathematics, Northwestern University, Evanston, IL 60208
We consider contact-line effects in directional solidification of a dilute binary alloy in a Hele-Shaw cell. We find steady basic-state solutions that satisfy the contact-angle conditions at the walls and then study their stability. Perturbation methods are used for values of the contact angle that are close to ir/2, when the front is almost planar, while a numerical boundary integral method is used for arbitrary contact angles. We find that under physically realistic conditions the curved interface is less stable than the planar front, which is in agreement with experimental observations. The numerical solution is in agreement with the perturbation result in the region where the latter is valid. A weakly nonlinear analysis is carried out in the framework of the same perturbation approach. As the deviation from the planar interface becomes more significant, the region of subcritical bifurcation is expanded. For values of thickness near one-half of the wavelength of the Mullins-Sekerka instability the three-dimensional steady-state structures can be observed even when the interface is normal to the wall; these structures are anti-symmetric with respect to the center plane of the Hele-Shaw cell. As the contact angle deviates from 7r/2, these shapes become unstable and the steady-state interface is symmetric with respect to the center plane. For values of gap thickness much larger than the characteristic diffusion length we derive a long-wave evolution equation with appropriate boundary conditions at the walls and use it for numerical studies of strongly nonlinear evolution of the system. We find that formation of deep roots and secondary bifurcations are promoted due to contact-line effects for concave down (toward the solid) interfaces.
234
C R E E P I N G S T E A D Y T H I N FILM ON A N INCLINED P L A N E WITH A N EDGE NURI AKSEL Department of Applied Mechanics and Fluid Dynamics, University of Bayreuth, D-95440 Bayreuth, Germany
Film flows occur in various technical processes, in environmental sciences, and everyday life. Hence, they were the subject of many investigations, e.g., [1-5], to name a few. Here, we study the influence of a sharp edge on the surface shape and the velocity profile of a creeping steady thin film with high surface tension on an inclined plane. Within the framework of the lubrication assumption an asymptotic solution is constructed [6]. A parameter-free direct relation between the surface shape and the wall friction at the bottom is found that gives rise to an indirect detection of the wall friction by measuring the surface shape. It is shown that capillary effects make the surface shape rise to form a maximum in the near front of the edge similar to the problems with moving film fronts. The region of validity and the sensitivity of the asymptotic solution on the upstream and downstream boundary conditions and on the inclination angle are found by comparing with exact numerical results. A numerical parameter study with respect to the capillary number and inclination angle is performed. Depending on the parameter combinations various surface shapes can be obtained. The theoretical (analytical and numerical) results are verified by experiments. References 1. E. B. Dussan V. and S. H. Davis, J. Fluid Mech. 65, 71 (1974). 2. B. G. Higgins, W. J. Silliman, R. A. Brown, and L. E. Scriven, Ind. Eng. Chem. Fundam. 16, 393 (1977). 3. B. G. Higgins and L. E. Scriven, Ind. Eng. Chem. Fundam. 18, 208 (1979). 4. K. N. Christodoulou and L. E. Scriven, J. Fluid Mech. 208, 321 (1989). 5. P. M. Schweizer, J. Fluid Mech. 193, 285 (1988). 6. N. Aksel, Arch. Appl. Mech. 70, 81 (2000).
235
PHASE-FIELD SIMULATION OF C O N V E C T I V E EFFECTS ON DENDRITIC GROWTH GUSTAV AMBERG AND ROBERT TONHARDT Department of Mechanics, KTH, S-100 44 Stockholm, Sweden
Phase-field simulations have been used to study the evolution from a small nucleus to a dendrite in an undercooled melt, in the presence of convection. As a generic case for growth from solid boundaries, the nucleus is in one case assumed to be attached to a solid wall, and a shear flow parallel to the wall is assumed in the melt. The interaction between the melt flow and the growing dendrite results in a changed growth rate and morphology. One aspect that has been investigated is how the effective tilting of the main stem of the dendrite depends on the flow strength. It is shown that the major factor that determines the tilt angle is the selection of a nucleus with an optimal initial crystal orientation. We have also studied natural convection around a nucleus held in an undercooled melt. The phase-field equations are solved together with the equations for viscous fluid flow using an adaptive finite element method. This method allows a large computional domain, so that a single dendrite in an effectively unbounded region can be studied.
236
A MODEL FOR A S P R E A D I N G A N D MELTING D R O P L E T ON A HEATED S U B S T R A T E DANIEL MANDERSON Department of Mathematical Sciences, George Mason University, Fairfax, VA 22030 M. GREGORY FOREST Department of Mathematics, University of North Carolina, Chapel Hill, NC 27599 RICHARD SUPERFINE Department of Physics and Astronomy, University of North Carolina, Chapel Hill, NC 27599
We develop a model that describes the dynamics of a spreading and melting droplet on a heated substrate. The model, developed in the capillary-dominated limit, is geometrical in nature and couples the contact line, tri-junction, and phase-change dynamics. In this model the competition between spreading and melting is characterized by a single parameter KT that represents the ratio of the characteristic contact-line velocity to the characteristic melting velocity. A key component of the model for the spreading and melting droplet is an equation of motion for the solid. This equation of motion, which accounts for global effects through a balance of forces over the entire solid-liquid interface, including capillary forces at the trijunction, acts in a natural way as the tri-junction condition. This is in contrast to models of tri-junction dynamics during solidification, where it is common to specify a tri-junction condition based on local physics alone. The tri-junction dynamics, as well as the contact angle, contact-line position, and other dynamic quantities for the spreading and melting droplet are predicted by the model and are compared to an isothermally spreading liquid droplet whose dynamics are controlled exclusively by the contact line. Our interest in this comparison has been motivated by experiments performed by Glick (1998) in which polystyrene spheres show different spreading characteristics when subject to these two different thermal configurations. We find that in general the differences between the dynamics of a spreading and melting droplet compared with that of an isothermally spreading droplet are increased as KT increases, that is, when the characteristic contact-line speed is greater than the characteristic melting speed. The presence of the solid phase in the spreading and melting configuration inhibits spreading relative to an isothermally spreading drop of the same initial geometry. Finally, we find that increasing the effect of spreading promotes melting.
237
INSTABILITIES OF A THREE-DIMENSIONAL LIQUID D R O P L E T ON A HEATED SOLID SURFACE STEVEN W. BENINTENDI Department of Mechanical and Aerospace Engineering, University of Dayton, Dayton, OH 45469 MARC K. SMITH The George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405
The behavior of thin liquid films and liquid droplets is of fundamental importance in many natural and technological processes, such as the coating of surfaces, the recovery of oil from porous media, the development of microfluidic devices, and a host of applications in material processing. It is well known that the behavior of fluid systems can be greatly affected by forces such as capillarity, thermocapillarity, and the wetting characteristics of the material system when in a negligible or a reduced gravitational environment. Although the control of liquid droplets and films has been attained through many methods, we seek an approach whereby thermocapillarity effects are used to prescribe the motion of liquid droplets. We examine the behavior of a three-dimensional liquid droplet on a nonuniformly heated or cooled horizontal solid surface. For a thin viscous droplet, lubrication theory was used to derive a two-dimensional evolution equation that includes the effects of viscosity, gravity, surface tension, slip at the contact line, and thermocapillarity. The evolution equation was coupled to a dynamic contactline condition, relating the contact-line speed to the apparent contact angles, to describe the bulk motion of the droplet. The behavior of the droplet was examined in terms of the imposed thermal field on the solid surface, which includes a mean temperature and a constant temperature gradient superimposed on the mean value. This non-uniformity in the thermal conditions initiates a thermocapillary flow within the droplet that drives bulk migration down the temperature gradient. However, for certain values of the heating parameters, the system exhibits an instability characterized by a slight elongation of the contact line and highlighted by the formation of a dimple at the center of the droplet. When both thermal parameters are active, we show that the instability can be stabilized by a sufficiently large imposed thermal gradient. The critical value at which the interface shape transitions back to a non-dimpled steady state increases with uniform heating of the solid surface. Hysteresis behavior in the critical values in forward and backward traces along the steady-state solution branches is demonstrated and suggests a subcritical instability. The instability also appears under axisymmetric or uniform heating conditions. The results indicate that a critical uniform heating value exists above which the droplet transitions from an axisymmetric configuration to a non-axisymmetric state with the interface shape again characterized by a dimple at the droplet center. For the axisymmetric instability, there was no transition back to a non-dimpled
238
state at larger heating values. A stability analysis for the axisymmetric and threedimensional droplet will be conducted to confirm the behavior observed in this study.
239
A LABORATORY MODEL FOR T H E SOLIDIFICATION OF T H E EARTH'S I N N E R CORE A N D T H E I N N E R CORE'S SEISMIC A N I S O T R O P Y MICHAEL I. BERGMAN Physics Department, Simon's Rock College, Great Barrington, MA 01230 and Department of Earth and Planetary Sciences, Harvard University, Cambridge, MA 02138
Although seismologists do not yet agree on the details of the depth dependence, the longitudinal variations, and perhaps even the strength, to a first approximation the inner core has been inferred to be elastically anisotropic, with the direction parallel to the rotation axis fast. Many hypotheses have been suggested as a cause for the anisotropy, all involving a lattice-preferred orientation, but the physical reality of many of them has not yet been demonstrated. Using ice as an analog material for high-pressure, hexagonal-closest-packed iron, a series of laboratory experiments have been carried out in which salt-water is solidified from the center of a rotating, hemispherical shell. The experiments reveal more rapid growth of the solid near the pole than the equator, presumably because the Coriolis force inhibits convection within the tangent cylinder parallel to the rotation axis and circumscribing the inner core. Unlike in the Earth where pressure effects are important, in the experiments this leads to colder temperatures within the tangent cylinder. Because of the small length and time scales of the experiments it is not possible to study subsequent solid-state flow, but the experiments do demonstrate that the equilibrium solidification surface is not spherical. Using polarized light on thin sections of the ice, the columnar nature of the dendritic crystals is apparent, with the a-axes lying in the direction of growth. The experiments show how fluid flow in the melt may affect the solidification texture. The laboratory model suggests how a solidification texture is frozen in, from which further textural changes may develop.
240
S U P P R E S S I O N OF C H A N N E L C O N V E C T I O N I N SOLIDIFYING P B - S N ALLOYS V I A A N APPLIED M A G N E T I C FIELD MICHAEL I. BERGMAN Physics Department, Simon's Rock College, Great Barrinton, MA 01230 and Department of Earth and Planetary Sciences, Harvard Cambridge, MA 02138
University,
DAVID R. FEARN Department of Mathematics, University of Glasgow, Glasgow G12 8QQ, Scotland, UK JEREMY BLOXHAM Department of Earth and Planetary Sciences, Harvard Cambridge, MA 02138
University,
Channel convection through the porous, dendritic mushy zone in solidifying alloys results from a nonlinear focusing mechanism, whereby liquid enriched in the solute melts dendrites as it convects away from the solid. The local melting reduces the Darcy friction and increases the flow speed to form a convective channel. However, it has been predicted that an applied magnetic field might prevent channels from forming because, as the Lorentz force replaces the Darcy friction as the primary resistance to flow, the focusing mechanism no longer operates. We find that an applied magnetic field can suppress channel convection when Qm, the mushy zone Chandrasekhar number, exceeds order one. Qm is a measure of the ratio of the Lorentz force to the Darcy friction. The longitudinal macrosegregation is not affected by the absence of channels, suggesting such channels are not always primarily responsible for the mass flux between the mushy zone and the melt and/or that convection in the mushy zone can occur without channels forming.
241
DYNAMICS AND STABILITY OF VAN-DER-WAALSDRIVEN THIN FILM RUPTURE ANDREW J. BERNOFF Department of Mathematics, Harvey Mudd College, Claremont, CA 91711 THOMAS P. WITELSKI Department of Mathematics, Duke University, Durham, NC 27708-0320
Williams and Davis (1982) derived a lubrication model for the evolution of thin films under the influence of van der Waals forces and surface tension. More recently, Zhang and Lister (1999) showed that this model has an infinite set of similarity solutions corresponding to finite-time film rupture. In this poster we discuss the dynamics and stability of thin-film rupture in planar and axisymmetric geometries. As the film thickness decreases, the planar state loses stability; we discuss the structure of the bifurcating solutions in both geometries. Loss of stability can lead to finite-time rupture of the thin film. We have developed a systematic technique for calculating self-similar rupture solutions and determining their linear stability. The dynamically stable similarity solutions produce observable rupture behavior as localized, finite-time singularities in the models of the flow. For the problem of axisymmetric van-der-Waals-driven rupture, we identify a unique stable similarity solution for point rupture of a thin film and identify a new mode of rupture - annular "ring rupture."
242
N E W APPROACHES TO FRONT-TRACKING A N D FRONT-CAPTURING M E T H O D S JERRY BRACKBILL Los Alamos National Laboratory, Los Alamos, NM 87545 DIDIER JAMET Los Alamos National Laboratory, Los Alamos, NM 87545 and Commissariat a lEnergie Atomique, France OLIVIER LEBAIGUE Commissariat a lEnergie Atomique, France DAVID TORRES Los Alamos National Laboratory, Los Alamos, NM 87545
A front-capturing method for liquid-vapor flows with phase-change. The second-gradient theory (also called the Cahn-Hilliard theory or the van der Waals theory) is a thermodynamic model of three-dimensional continuous liquidvapor interfaces. The surface tension is therefore the integral of an energy concentrated in the interfacial zone, proportional to (Vp)2 in the classical form of the theory. From the second-gradient theory, one can derive the equations of motion, which are valid everywhere. Therefore, one solves one system of three-dimensional partial differential equations everywhere: the interfacial motion and the phase-change at the interfaces are just a part of the solution. Moreover, a boundary condition for the density gradient appears naturally: n • Vp must be imposed (n is the unit normal to the boundary), whose value is directly related to the contact angle. The motion of the contact line is therefore a natural part of the solution of the equations of motion of such a fluid. This particularly allows one to account for the contact-angle hysteresis phenomenon is a very simple way: n • Vp has to vary spatially. Simulations of film boiling are performed using this model. Front-tracking without connectivity. Front-tracking is a powerful technique for modeling surface tension. However, it has difficulty in handling topological changes in an interface. This defect is due to the current reliance of front tracking on the connectivity between interfacial points. We have developed a technique in which the surface-tension force can be calculated from interfacial points without connectivity, thus allowing front tracking to model topological changes in an interface naturally much like front-capturing models. The technique is based on determining an indicator function I(x) for the interface. The interfacial normals and curvatures can then be calculated by dif-
243
ferentiating the indicator function. We have also developed a projection technique that allows us to dramatically reduce parasitic currents. We present numerical simulations of coalescence and parasitic-current reduction in 2D and 3D.
244
M A N I P U L A T I O N OF INTRAVASCULAR GAS EMBOLISM D Y N A M I C S W I T H EXOGENOUS SURFACTANTS A. B. BRANGER Department of Biomedical Engineering, Northwestern University, Evanston, IL 60208 DAVID M. ECKMANN Department of Anesthesia and Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA 19104-4283
B A C K G R O U N D : Intravascular gas embolism (IGE) is the entrapment of gas bubbles in the circulation and can occur in decompression sickness or during cardiopulmonary bypass. These IGE are harmful because they temporarily block blood flow causing local ischemia in the surrounding tissue. Altering the interfacial tension and contact-line mechanics at the bubble/blood/vessel interface with a surfaceactive agent changes the bubble conformation and dynamics. To determine the potential therapeutic benefit of manipulating interfacial forces, we investigated the effects of the surfactant, Antifoam 1510-US, on the dynamics of bubble entrapment and break-up in the intact rat cremaster circulation. METHODS: The animals used in the experiments were handled according to NIH guidelines. Adult male Wistar rats (n = 6) were anesthetized, instrumented, and divided evenly into two groups, control, and those receiving a 2.5 ml pretreatment bolus of diluted Antifoam, raising the surfactant concentration to 1.5% of the total blood volume. Air bubbles (4 (A) were injected through the femoral artery and observed in the cremaster circulation over time with video-microscopy. Dimensions of the bubbles (0.2-10.0 nl), as well as other parameters relating to embolism dynamics, were measured from the video-taped recording. RESULTS: The arterial vessels in the cremaster circulation of rats receiving the pre-treatment of Antifoam showed a much greater degree of constriction (<~ 65%) than the arterioles of the control rats. This increased vasoconstriction in the pretreated animals, coupled with surfactant at the interface, lead to bubble elongation and caused bubble pinch-off into several smaller bubbles, not seen in the controls. CONCLUSIONS: Addition of exogenous surfactant increased deformability of the IGE. This permitted a smaller radius of curvature at the ends of the bubble and allowed for bubble narrowing, elongation, and eventual bubble pinch-off. The pinch-off phenomena induced by surfactant may clinically create multiple smaller bubbles more apt to traverse the microcirculation and lodge in smaller vessels. This might maintain vital organ blood flow and reduce tissue damage. This work was supported by NIH Grant R01 HL60230.
245
ADIABATIC HYPERCOOLING OF B I N A R Y MELTS K. BRATTKUS Department of Mathematics, Southern Methodist University, Dallas, TX 75275
A binary melt is hypercooled when it is cooled to a temperature below its solidus. In the isothermal limit planar solidification fronts propagate at a constant velocity determined by the kinetic undercooling and are subject to a long-wavelength morphological instability if speeds fall below a critical value. In this letter we examine the adiabatic limit where the accumulation of a small latent heat release causes the velocity of the interface to slowly decrease through its critical value. The evolution of the hypercooled interface is governed by a damped Kuramoto-Sivashinsky (dKS) equation with coefficients that vary as the interface decelerates. Using this equation we show that morphological transitions are delayed by an amount that reflects both the time the system spends in a stable state and the magnitude of the damping. For a sufficiently large latent heat of fusion the long-wavelength morphological instability is annihilated. Finally, the adiabatic dKS equation predicts late-stage coarsening of the microstructure with length scales that increase as i 1//2 . In finite systems this coarsening removes the morphological instability.
246
A N INSOLUBLE SURFACTANT MODEL FOR A D R A I N I N G VERTICAL LIQUID FILM RICHARD J. BRAUN Department of Mathematical Sciences, University of Delaware, Newark, DE 19716
The drainage of a thick non-aqueous film with a surfactant will be studied theoretically and compared with experimental results. Although the film is thin physically, the film is thick enough so that intermolecular forces acting across the film are not important. In the experiment, a thin film is suspended vertically from a wire frame over a bath and gravity drives the drainage of the film back into the bath. Lubrication theory is applied to the situation and the thin film is patched to the bath. Surfactant-transport, Marangoni, and surface-viscous effects are included in theoretical models. Models have been formulated that span the range of drainage regimes, which correspond to the range from rigid to mobile films. Computed and analytical results show that films may be made rigid by surface-viscous or Marangoni effects. Similarity solutions are found in some cases. This is joint work with S. A. Snow and U. C. Pernisz of Dow Corning Corporation and S. Naire of the University of Delaware.
247
T H E EFFECT OF TIME-PERIODIC AIRWAY WALL S T R E T C H ON SURFACTANT A N D LIQUID T R A N S P O R T IN T H E L U N G J. L. BULL Biomedical Engineering Department, The University of Michigan, Ann Arbor, MI 48109 D. HALPERN Department of Mathematics, University of Alabama, Tuscaloosa, AL 35487-0350 J. B. GROTBERG Biomedical Engineering Department, The University of Michigan, Ann Arbor, MI 48109 Understanding the transport of surfactant and liquid in the lung is of great importance in designing treatment for respiratory distress syndrome (RDS), and in understanding clearance from the lung. The standard method, surfactant replacement therapy (SRT), for treating this deficiency is to instill exogenous surfactant into the lung. It spreads in the small airways due to Marangoni flows. After the initial transient spreading, the surfactant concentrations at generation 7 (the approximate start of the Marangoni region) and generation 18 (the approximate start of the alveolated region) are essentially constant, so that a steady background surface tension gradient is established. Most of the surfactant transport in SRT takes place once this background gradient is established. On the other hand, liquid and surfactant clearance in normal lungs occur because production of surfactant in the alveoli results in lower surface tensions there than in the airways. This surface tension gradient causes flows towards the trachea. We consider a thin film lining a stretchable tube with a radius that depends on axial position to model airway branching and examine the effects of axial and radial wall oscillations on surfactant and liquid transport. Evolution equations for the film thickness, surface-surfactant concentration, and bulk-surfactant concentration are derived from conservation of momentum, conservation of fluid mass, and conservation of surfactant mass. The model is general and allows for choices of (1) interfacial linear sorption kinetics or squeeze-out phenomena; (2) uniform or linear membrane wall strain; (3) higher (SRT) or lower (liquid and surfactant clearance) concentration at the proximal end of the domain; (4) various strain amplitudes; and (5) various breathing-cycle periods. The evolution equations are solved using the methods of lines, and the cycleaverage transport of both surfactant and liquid is computed. While the transport of surfactant on the surface and in the bulk varies greatly with strain amplitude, the transport does not vary as much with cycling period at a given strain amplitude in our model of SRT. As the transport of surfactant into the alveolar region is considerably higher when the strain amplitude is large, it may be advantageous to use large tidal volumes and large breathing-cycle periods with SRT in order to gain larger surfactant (both surface and bulk) transport. Our model of liquid and surfactant clearance predicts that large breathing-cycle periods and strain amplitudes result in greater clearance.
248
T H E D Y N A M I C EFFECTS OF SURFACTANTS O N STATIONARY GAS B U B B L E S IN LIQUID FLOWS DANIEL P. CAVANAGH Department of Chemical Engineering, Bucknell University, Lewisburg, PA 17837 DAVID M. ECKMANN Department of Anesthesia and Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA 19104-^283
We have experimentally examined the effects of common soluble surfactants on gas bubbles in liquid flows in inclined tubes. Air bubbles of known size (A = 0.8, 1.0, 1.5) are held stationary under minimum flow conditions in tubes oriented at specific inclination angles (a = 45°, 90°). Sodium Dodecyl Sulfate (SDS) or Triton X-100 (TX100) is infused into the bulk flow at specific bulk concentrations (T = 10% CMC, 100% CMC). In addition to recording pressure and flow waveforms, we capture images of the bubbles before and during exposure to the surfactant. The modification of the interfacial properties by the surfactant results in bubble behavior ranging from interfacial deformation to deformation plus axial translation to bubble detachment from the wall plus axial translation. Correspondingly, we observe modification of the measured pressure gradient within the test section of the tube. These surfactant-mediated responses are dependent upon the interrelated effects of r , A, a, and the surfactant characteristics. Specifically, a high bulk concentration of surfactant produces more rapid bubble modification and increases the potential for bubble detachment. The potential for detachment increases further as bubble volume is increased. In both vertical tubes where contact forces are not important and in non-vertical tubes the infusion of surfactant may lead to axial translation either with or against the bulk flow. Finally, surfactants that are more effective at modifying contact properties produce bubble-shape and pressure-gradient changes more rapidly. This investigation demonstrates the ability of surfactants to potentially dislodge dried gas bubbles through the manipulation of interfacial properties. This work was supported by the Whitaker Foundation and NIH Grant R01 HL60230.
249
B U C K L I N G INSTABILITIES IN THIN VISCOUS SHEETS SAHRAOUI CHAIEB, R. DA SILIVEIRA, L. MAHADEVAN, AND G. H. MCKINLEY Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139
Buckling instabilities are not restricted to solids; they can occur in thin layers of Newtonian fluids. Here we investigate two types of buckling. The first experiment is an instability in a thin viscous layer that lies in a cylindrical annular geometry. As the shear rate becomes larger than a critical threshold, the viscous film deforms sinuously without changing its thickness leading to ripples arranged like the spoke of a bicycle wheel. The ripples appear because the shear gives rise to tensile and compressive stress; the latter are responsible for out-of-plane motions of the sheet. Linear stability analysis predicts the critical shear rate at which the instability occurs as well as the wavelength and angular frequency of the instability. This analysis is compared to the results of our experiments. The second scenario of a buckling of a viscous sheet occurs when a bubble of air rises to the top of a highly viscous liquid, it forms a dome-shaped protuberance on the free surface. Unlike a soap bubble it bursts so slowly as to collapse under its own weight simultaneously, and folds into a wavy structure. This rippling effect occurs for both elastic and viscous sheets, and a theory for its onset is formulated. The growth of the corrugation is governed by the competition between gravitational and bending forces (shearing).
250
FLUID-FLUID INTERFACE E X P E R I M E N T S AT T H E U N I V E R S I T Y OF CHICAGO ITAI COHEN, SIDNEY R. NAGEL James Franck Institute, University of Chicago, Chicago IL 60637 MICHAEL P. BRENNER Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139 JENS EGGERS Department of Theoretical Physics, University of Essen, Essen, Germany ROMAN O. GRIGORIEV School of Physics, Georgia Institute of Technology, Atlanta, GA 30332-0430 TODD F. DUPONT James Franck Institute, University of Chicago, Chicago IL 60637
Understanding and controlling how a liquid interface changes its topology from being singly connected to being multiply connected (as is the case with a drop dripping from a faucet) or from being bounded to being unbounded in a particular direction (as is the case in the selective-withdrawal problem) is crucial for the control of many manufacturing processes including the creation of emulsions and mono-dispersed sprays. Furthermore, the mathematics describing these types of topological changes is not understood very well. In my poster I will present two beautiful phenomena that explore and shed light on these issues. First, I will show results from experiments on the two fluid drop snap-off problem where fluid is dripped through an outside liquid medium that is viscous. I will then show results from experiments on the selective-withdrawal problem. Here, we lower a straw so that its orifice rests above a water-oil interface. We then withdraw the oil through the straw. By changing the rate of withdrawal we control a transition between having only oil being withdrawn and having water being entrained along with the oil.
251
A S Y M P T O T I C ESTIMATES FOR 2-D SLOSHING MODES: THEORY A N D E X P E R I M E N T ANTHONY M. J. DAVIS Department of Mathematics, University of Alabama, Tuscaloosa, AL 35487-0350 PATRICK D. WEIDMAN Department of Mechanical Engineering, University of Colorado, Boulder, CO 80309-0427
A classic problem in surface gravity waves is that of two-dimensional sloshing modes in a channel of arbitrary shape. Exact solutions are known for vertical walls and for triangular containers with a, the angle subtended at the free surface, equal to 45° or 30°. At higher frequencies, the wave motion is confined near the free surface and only the shape of the container at the corners is important. Fox and Kuttler (ZAMP, 1983) conjectured the asymptotic form 1 2*
n+i(l-M)
»=Ta
for the dimensionless frequency parameter. This estimate can be anticipated by simple use of sloping-beach potentials. The next correction for straight-sided containers (fi = 1,2,3) is exponentially small in the exact solutions, 0(l/n) for the infinite dock with gap (a — n), and is now shown to be 0(n~2^) in general. Modifications for asymmetric containers are included. Experiments designed to verify the theory have been successfully completed for odd values of n up to 13. Five geometries, including angles 60° and 135° for which fi is not an integer, were used and surface tension in the fluid minimized. The measured data was adjusted to account for the remaining surface tension and good agreement with the theory was obtained.
252
B U O Y A N C Y - D R I V E N I N T E R A C T I O N S OF VISCOUS D R O P S W I T H DEFORMING INTERFACES ROBERT H. DAVIS Department of Chemical Engineering, University of Colorado, Boulder, CO 80309-0424 JOSEPH KUSHNER Department of Chemical Engineering, Arizona State Tempe, AZ 85287-6006
University,
MICHAEL A. ROTHER Department of Chemical Engineering, University of Colorado, Boulder, CO 80309-0424
Two different-sized drops of one liquid immersed in a second, immiscible liquid will move relative to one another under the influence of gravity. When the drops come close together, they interact owing to hydrodynamic disturbances, and various outcomes are possible: separation of the drops, capture with or without coalescence of the drops, breakup of the smaller drop into two or more drops, or even a combination of capture and breakup phenomena. Interfacial deformation may promote any of the results, depending on the drop-to-medium viscosity ratio, the ratio of the smaller drop radius to the larger drop radius, the gravitational Bond number, which is a measure of how much the drops will deform, and the initial horizontal offset between the drops. This poster presents an experimental investigation into the parameter space governing two-drop interactions to compare with theoretical results obtained previously (Zinchenko, Rother, and Davis, J. Fluid Mech. 391, 249 (1999); Davis, Phys. Fluids 11, 1016 (1999)). Trajectories of two drops consisting of a mixture of glycerol and water moving through an external medium of castor oil are photographed and analyzed to check quantitative agreement with computer simulations.
253
A N O M A L Y A N D U N C E R T A I N T Y W H E N LIQUID FILMS FLOW OVER SOLID SURFACES WALTER DEBLER Department of Mechanical Engineering, The University of Michigan, Ann Arbor, MI 48109-2125
Previous work on the flow of thin liquid films over inclined solid surfaces has shown the existence of two forms for the advancing leading edge. In an investigation to determine the speed of advance of the chevron-like form of this disturbance, it was observed that in a single experiment the other pattern, rivulets with sides parallel to the down-slope direction, occurred when using silicone oil on a glass surface. Subsequently an extensive series of tests could not replicate this single result on clean glass. Various contaminants were then applied to the glass, but only chevrons appeared when the silicone oil film flowed. Finally, a stain inhibitor for fabrics, Scotch-gard, was sprayed onto the glass and the silicone film produced rivulets at its leading edge. Hence, it was concluded that some unknown substance that yielded a similar effect must have been deposited on the glass plate that had produced, with but one exception, chevrons. The present tests also showed that clear Teflon film will also form rivulets with silicone oil. Moreover, the author obtained rivulets with glycerin on glass, in contrast to previously-reported results of others. The ultimate lesson that can be drawn is that even when the solid surface is. diligently and well cleaned, molecular thin films might be present to alter the contact angle of the liquid at the solid boundary and, thereby, affect the type of leading-edge pattern that is observed in an experiment.
254
FLOW BEHAVIOR OF L A N G M U I R MONOLAYERS MICHAEL DENNIN AND R. S. GHASKADVI Department of Physics and Astronomy, University of California, Irvine, Irvine, CA 92697-4575
We report on measurements of the viscosity of Langmuir monolayers. Langmuir monolayers are comprised of amphiphilic molecules that are confined to the airwater interface. They exhibit a rich two-dimensional phase behavior that includes the usual gas, liquid, and solid phases, as well as a number of phases that are analogs of three-dimensional smectic liquid crystals. We have built a Couette viscometer that is capable of measuring velocity profiles, viscosity, and the complex shear modulus of monolayer films. The velocity profiles are obtained by imaging the monolayer using a Brewster-angle microscope. The apparatus has been tested on both Newtonian and non-Newtonian phases, and the results for the velocity profiles are consistent with theoretical calculations. It is known that the addition of divalent ions to the water will "stiffen" monolayers of fatty acids. This is due to the binding of two fatty acids by a single divalent molecule. The influence of divalent ions on the phase behavior has been studied, but the effects on the viscosity have not been studied in detail. We have made measurements of both the viscosity and the shear modulus and find that they increase by three orders of magnitude over a 10-hour time scale. Further, the evolution appears to occur with three distinct time scales. We directly measured the percentage of C a + + ions bound to the monolayer as a function of time by repeatedly transferring the monolayer onto a solid substrate. The multilayer formed by this process was rinsed in an acid to release the C a + + ions, and their concentration was measured with an ion meter. We found that the fraction of bound C a + + had a similar time evolution to the viscosity. Also, a simple ad-hoc model for the viscosity in terms of the percentage of bound C a + + is able to explain the time evolution. A more detailed study of this process will provide important insights into the fundamental nature of viscosity.
255
SOLUBLE SURFACTANTS A N D CONTACT-ANGLE D Y N A M I C S DAVID M. ECKMANN Department of Anesthesia and Institute for Medicine and Engineering, University of Pennsylvania, Philadelphi, PA 1910^-4283 Little is detailed about surfactant effects on the contact angle, contact-angle hysteresis, and contact-line motion. Ultimately we are interested in wetting/dewetting phenomena related to intravascular gas embolism and its treatment with exogenous surfactants. We have therefore begun to characterize the relationships between contact angle, contact-angle hysteresis, spreading parameters, and surfactant moieties on various solid substrates. Using a Wilhelmy-plate surface tensiometer, we measured the air-liquid surface tension, a, of aqueous solutions of sodium dodecyl sulfate, (SDS) with and without 5% bovine serum albumin (BSA) present. By video image analysis, we evaluated a drop interfacial shape on a stationary flat horizontal substrate and on a flat slowly-rotating substrate (< l°/second). Drop volumes were 40-80 /A and all measurements were recorded at room temperature in a sealed, humidified chamber 30 minutes after solution instillation into the experimental apparatus. Solid substrates included glass, acrylic, and stainless steel. We measured static, advancing, and receding contact angles, a, and 0 r , respectively, of drops of the same aqueous SDS ± BSA dilutions. We calculated S, the spreading parameter, based on >g. The parameter a cos <j) incorporates both the surface tension and contact-angle effects needed to calculate a spreading coefficient. It is a measure of the adhesion force acting at the contact line. On glass, as SDS concentration increases, a cos s increases to a local maximum and then decreases to its plateau value (~33-34 dyne/cm) at and above the CMC. The variation of a cos (ps with concentration of SDS in the presence of BSA (which is surface active) maintains substrate dependence with a more narrow range of values. Competition between BSA and SDS for interfacial sites likely accounts for this. The contact-angle hysteresis (a — T) is nearly constant (~ 26°-30°) over the range of SDS concentrations. There is an 8-fold change in S, which tends from -19.2 to -2.4 as the SDS bulk concentration increases from 0 to the CMC. The surfactant studied lowered interfacial tension, and reduced contact angles and the spreading coefficient with substrate dependence. These experiments help to explain a potential role that exogenous surfactants may have in the treatment of intravascular gas embolism. A particular surfactant delivered to the interface may significantly change the wetting potential of the liquid onto the solid and thereby increase or decrease the adhesion force acting on a trapped bubble. The exact influence of interactions of individual species (surfactant, protein, substrate) on the interfacial quantities measured in these experiments are not easily inferred. They must be derived by continued careful experimental analysis. This work was supported by NIH Grant ROl HL60230.
256
E X P E R I M E N T A L STUDIES OF T H E H Y D R O D Y N A M I C S N E A R MOVING CONTACT LINES S. GAROFF Department of Physics, Carnegie Mellon University, Pittsburgh, PA 15213 E. RAME National Center for Microgravity Research, NASA Glenn Research Center, Cleveland, OH 44135 Dynamic wetting, spreading of a fluid over a solid surface, controls many natural phenomena and technological processes. When surface tension forces are important, the interface shape and local flow field very near the moving contact line control the macroscopic configuration of the fluid body. However, identifying the correct assumptions needed for predictive models of spreading is not trivial. Theoretically, the central difficulty is the unphysical stress singularity at the contact line arising when classical hydrodynamic assumptions are applied up to and including the moving contact line. The singularity suggests that very near the contact line (in an "inner" region), unique microphysical processes other than the classical hydrodynamic assumptions control the fluid motion. Thus, we must explore: what are these unique processes operating on the microscopic scale? And how can we find a legitimate boundary condition for the macroscopic problem? Asymptotic expansions provide a description of the hydrodynamics near moving contact lines in a limited parameter range. These expansions and very accurate experiments provide a tool for answering the questions central to understanding the moving-contact-line problem. We have found that for a set of Newtonian polymer fluids (PDMS, polydimethylsiloxanes) on various surfaces, these asymptotic expansions describe the hydrodynamics at small capillary number, Ca < 0.1, and negligible Reynolds number, Re <~ 0. At higher Ca (with Re ~ 0) and moderate Re ~ 0.1 (with Ca < 104), our experiments find that the curvature of the interface near the contact line is less than that predicted by the asymptotic expansions. Having established the validity of the asymptotic analysis, we may now use it to probe the microphysical properties controlling the fluid motion on the inner scale. Although many processes have been suggested, no direct experimental identification of these processes exists. We must carefully use the tool provided by the asymptotic expansions to indirectly recover information on the inner scale hydrodynamics from our experiments. Use of static approximations or Tanner's Law in the analysis of experimental data leads to systematic errors that obscure the correct characterization of the inner hydrodynamics. We find that for PDMS, the parameters describing the inner hydrodynamics to 0(1) in Ca, as Ca —> 0, must vary with contact-line speed. This variation is complex and suggests that different mechanisms operate at different contact-line speeds. Further, the inner hydrodynamics vary, even for very closely-related systems. Use of Ca to scale the velocity does not unify the data for the different systems, implying that this is not the correct scaling of the fluid velocity in the inner region.
257
A X I S Y M M E T R Y - B R E A K I N G INSTABILITIES IN A X I S Y M M E T R I C FREEZING OF ICE ALEXANDER YU. GELFGAT, PINHAS Z. BAR-YOSEPH, ALEXANDER SOLAN Computational Mechanics Laboratory, Faculty of Mechanical Engineering, Technion - Israel Institute of Technology, Haifa 32000, Israel TOMASZ A. KOWALEWSKI Polish Academy of Sciences, IPPT PAN, Swietokrzyska 21, PL 00-049, Warszawa, Poland
This work is devoted to a theoretical explanation of an effect observed during the study of freezing of ice in an axisymmetric container [1]. The experimental setup consists of a cylindrical glass container filled with water, which is immersed in a bath with hot water and is closed by a cold steel cover. After the temperature of the cover is reduced below 0°C, an ice front is formed. The solidification front observed on the ice had an asymmetric, circumferentially-periodic surface shape. Likewise, measurements of the temperature field showed a periodic structure. It was assumed that the effect was due to an axisymmetry-breaking instability of the convective flow of water. Detailed numerical analysis of the stability of the basic axisymmetric flow with respect to all possible three-dimensional perturbations showed that an instability sets in with a relatively high azimuthal wavenumber k, which varies between 7 and 10. Details can be found in [2]. The corresponding pattern of the most unstable three-dimensional perturbation of the temperature is similar to the experimentally-observed temperature distribution. It is shown that the calculated instability is caused by the Rayleigh-Benard mechanism, which leads to the appearance of a system of convective rolls distributed along the azimuthal direction inside a relatively thin convective layer. References 1. T. A. Kowalewski and A. Cybulski, IPPT PAN Report 8/97, Warsaw (1997) (in Polish). 2. A. Yu. Gelfgat, P. Z. Bar-Yoseph, A. Solan, and T. A. Kowalewski, I. J. Trans. Phenomena 1, 173 (1999).
258
SEPARATION M E C H A N I C S OF THIN INTERFACIAL LIQUID LAYERS: T H E ROLE OF VISCOUS F I N G E R I N G ASHOK GOPINATH Department of Mechanical Engineering, Naval Postgraduate School, Monterey, CA 93943-5146
The mechanics of separation of a thin interfacial liquid layer trapped between two parallel surfaces was studied in a controlled manner. The liquid of choice was a silicone oil with constant surface tension, but variable viscosity. Different viscosities, layer thicknesses, and separation velocities were used to determine the separation behavior and its dependence on viscous fingering and capillary number. Force, displacement, and time data were recorded for all experimental runs and the plots used to gain preliminary insight into this process. Qualitative flow-visualization data has also been recorded to corroborate the trends in the onset of viscous fingering with the predictions of a simple interfacial stability analysis.
259
LARGE FINITE-ELEMENT MODELING OF AXIALLY S Y M M E T R I C FREE-SURFACE FLOWS ROMAN GRIGORIEV School of Physics, Georgia Institute of Technology, Atlanta, GA 30332-0430 TODD DUPONT James Franck Institute, University of Chicago, Chicago, IL 60637
The dynamics of free-surface flows and in particular the mechanisms for singularity formation at the interface of fluids with different physical properties constitute a problem of high theoretical as well as practical interest. The applications are abundant and include the functioning of such commonplace devices as ink-jet printers and fuel injectors, and processes such as oil extraction and fiber spinning. While considerable theoretical and computational advances have been achieved in our understanding of the problem in the inviscid and Stokes limits, the analyses at intermediate Reynolds numbers remain scarce and inconclusive. We present a general computational algorithm able to describe different kinds of axially symmetric, free-surface flows for arbitrary Reynolds numbers. The distinctive feature of the proposed algorithm is that the interface is treated as a mathematical singularity corresponding to the discontinuous change in the fluid properties, rather than being artificially smeared over a finite region, as is usually done. This should allow one to numerically simulate the singularity formation with precision adequate for comparison with experimental data. The idea of the algorithm is to replace the mathematical problem involving multi-fluid flow with complicated boundary conditions imposed at the free interface^) with a simpler problem involving a single-fluid flow with spatially (and temporally) varying viscosity and density, and with a surface-tension force acting at the interface(s), thus effectively getting rid of the boundary conditions. This is achieved by writing the Navier-Stokes equation and the incompressibility condition in the weak form by integrating these equations with an appropriately chosen weight. The obvious advantage of the finite-element formulation is that it contains no singularities associated with the jump in viscosity at the interface. The viscosity jump can be removed by performing integration by parts, effectively reducing the order of the equations by one. Furthermore, the singularity produced by the jump in pressure can also be removed by using weight functions that are themselves discontinuous at the interface. Present implementation assumes no angular dependence and zero angular velocity, but the algorithm allows one to lift these assumptions at the expense of a slight complication in the formulas. The interface is advanced using a version of the interface-tracking technique based on the advection of massless markers with the velocity field computed on a fixed grid, although other thechniques such as the volume-of-fluid or the level-set method can be easily incorporated. Our preliminary results for the two-fluid, selective-withdrawal problem provide an illustration of the computational method.
260
MOLECULAR SIMULATIONS OF INTERFACE P H E N O M E N A : A N ALTERNATIVE A P P R O A C H NICOLAS HADJICONSTANTINOU Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139
We present molecular simulations of two classical interfacial phenomena in fluid mechanics. The first is the Rayleigh-Taylor instability, arising when a heavier fluid is situated on top (with respect to a gravity field) of a lighter fluid. Our molecular techniques allow a novel approach to this problem. The instability is triggered by molecular thermal fluctuations instead of the artificial perturbations imposed in continuum simulations. Additionally, a direct comparison between miscible and immiscible Rayleigh-Taylor mixing is possible. The second problem presented is the motion of a contact line in two dimensions. Molecular simulations of a fluid displacing a second immiscible fluid in a twodimensional channel are performed. A continuum simulation of the same problem is employed to answer the following question: what continuum boundary conditions are required to reproduce the molecular interface shape? We found that by using the actual slip profile observed in our molecular simulations as a slip boundary condition and by setting the dynamic contact angle equal to the static value, that we could reproduce the molecular results satisfactorily in the range 0.059 < Ca < 0.072 [Phys. Rev. E 59, 2475 (1999)].
261
E X P E R I M E N T A L INVESTIGATION OF E N V I R O N M E N T O X Y G E N C O N T E N T IN SOLDER-JET T E C H N O L O G Y E. HOWELL, S.-Y. LEE, C. M. MEGARIDIS Department of Mechanical Engineering, University of Illinois at Chicago, Chicago, IL 60607 M. MCNALLAN Department of Civil and Materials Engineering, University of Illinois at Chicago, Chicago, IL 60607 D. WALLACE MicroFab Technologies, Piano, TX 75074
Solder-jet technology is targeted for use in high-density-electronics-assembly applications where 63Sn/37Pb solder is utilized as an attachment and/or structural material. Solder jetting relies on ink-jet printing technology to create and precisely place miniature molten-solder droplets (50-100 /xm in diameter) on substrates or pads. These droplets solidify after impact, forming bump deposits that are subsequently used for the flip-chip bonding of electronic components on the substrate. Ink-jet-based solder deposition is low cost (no masks or screens are required), flexible, easily automated using digital technology, suitable for a manufacturing environment (does not operate in a vacuum), highly repeatable, and offers high resolution. Even though solder jetting shows great promise to achieve solder deposition at pitch geometries well below what is feasible with existing techniques, there exists a few formidable challenges that must be overcome before commercialization. Because maximum throughput is an important requirement in electronic assembly, solder jetting is performed openly, but with local environmental control. A sheathing nitrogen-ring flow surrounds the solder jet to minimize solder oxidation that can drastically delay the atomization process and degrade the effective adherence of the solder deposits on the targets. Therefore, successful commercial implementation requires quantification of the influence of the ambient oxygen content on the surface properties of the molten solder jet. The method employed to measure the surface tension of the molten solder jet is the oscillating-jet technique. When a liquid is forced through an elliptical capillary (aspect ratio « 2), an oscillating jet forms. The stationary shape of this jet (wavelength and both semiaxes) is affected by the liquid/gas interface properties and is used to determine the surface tension as a function of distance from the orifice (surface age) using the analysis of Bechtel et al. [J. Fluid Mech. 293, 379 (1995)]. The jetting is performed in a controlled environment containing a known nitrogen/oxygen mixture. A CCD camera with a microscope lens is used to obtain images of the oscillating solder jet along both major axes. From these images, the wavelength and the semiaxes are determined at the beginning, middle, and end of successive wavelengths. This data along with other material properties is used to determine the dynamic surface tension of the molten solder along the direc-
262
tion of jet propagation. The experiment is repeated at oxygen mole concentrations ranging from 5ppm to 21% (air). In addition to the oscillating-jet experiments, a maximum-bubble-pressure apparatus is employed to determine the equilibrium values of surface tension of molten solder in controlled nitrogen/oxygen mixtures.
263
DROPLET SPREADING WITH SURFACTANT: MODELING AND SIMULATION JOHN HUNTER Department of Mathematics, University of California, Davis, Davis, CA 95616 ZHILIN LI Department of Mathematics, North Carolina State Raleigh, NC 27695-8205 HONGKAI ZHAO Department of Computer Science, Stanford Stanford, CA 94305-9040
University,
University,
W h e n a liquid droplet is placed on a flat solid surface, we may observe partial wetting or complete wetting of the surface depending on the static and dynamic contact angles. We present a simple mathematical model for this free b o u n d a r y problem and a new numerical method to simulate this model. Our numerical algorithm is based on the modified immersed-interface method and the level-set formulation. Our numerical results agree with those experiments and analytic results in the literature. Topological changes are handled easily using our methods.
264
AIR E N T R A I N M E N T AT LOW VISCOSITIES ALEXANDRA INDEIKINA, IGOR VERETENNIKOV Department of Physics, University of Notre Dame, Notre Dame, IN 46556 HSUEH-CHIA CHANG Department of Chemical Engineering, University of Notre Dame, Notre Dame, IN 46556
We examine experimentally and theoretically the breakdown of a straight contact line on a roller that rotates into a liquid. This breakdown occurs at a critical dimensionless speed (capillary number Cac) and precedes the industrially important air-entrainment phenomenon. A corn-syrup/water mixture with relatively low viscosity (0.3-0.7 Poise) is used as the test fluid. Our flow-visualization experiments indicate that the liquid interfacial velocity at the nearly critical air cusp above the contact line decelerates rapidly (by two orders of magnitude) over the short length of the cusp (a few capillary lengths). Moreover, Cac is found to be insensitive to agitation in the liquid phase, but extremely sensitive to obstacles placed above the cusp in the gas phase. The evolution of the air flow coupled with changes in the shape of the air-liquid interface is of paramount importance in the stability of the two-dimensional air cusp, since air is dragged into the cusp by both the solid roller and the liquid phase. Both the injection drag forces and the length of the cusp increase with increasing roller speed. The build-up of stagnation pressure, however, saturates as the cusp flattens. Consequently, at a critical Ca, the steady-flow balance breaks down. The required pressure gradient for air ejection becomes impossible in two dimensions — a larger free-surface-curvature gradient is needed to push the air out of the longer flat cusp. Hence, the third dimension should be invoked and the formation of triangular air pockets along the contact line is initiated. We support this physical mechanism for air entrainment via matched asymptotics. Momentum boundary layers on the gas side within the cusp and on the roller and below the cusp interface in the liquid are connected to each other and to inviscid outer solutions. The matching at the contact line requires a molecular model for the gas viscosity when the air-cusp width approaches the mean-free path of gas molecules. Of particular interest is the dependence of Cac on the inception length of the gas momentum boundary layer as specified by the position of a scraper — a potentially useful means of delaying air entrainment in industry.
265
NUMERICAL SIMULATIONS OF VIBRATION-INDUCED DROPLET EJECTION ASHLEY JAMES Department of Aerospace Engineering and Mechanics, University of Minnesota, Minneapolis, MN 55455 MARC K. SMITH AND ARI GLEZER The George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405
Vibration-induced droplet ejection occurs when a liquid drop is placed on a vertically-vibrating surface. T h e vibration leads to the formation of capillary waves on the free surface of this primary drop. W h e n the forcing amplitude is large enough, secondary droplets are ejected from the wave crests. For low-frequency forcing a low-order, axisymmetric wave mode is excited in the primary drop. For large enough forcing amplitudes, a steep depression or crater forms in the center of the drop during the downward phase of the oscillation cycle. During the next phase the crater collapses and liquid flowing inward forms an upward, high-momentum jet at the center of t h e crater. One or more small droplets may fragment from the end of this jet to produce secondary droplet ejection. For large forcing frequencies the wave motion is chaotic. Secondary droplets are ejected from multiple locations on the free surface of the primary drop. At each ejection site, a crater forms before droplet ejection. As in the low-frequency case, the crater then collapses and a jet is formed at the center of the crater. One or more droplets may then be ejected from the end of the jet. Under certain conditions the entire primary drop will rapidly atomize. An axisymmetric, Navier-Stokes solver, based on the M A C method, is used to simulate the low-frequency process. T h e solver includes a piecewise linear, volumeof-fluid m e t h o d and a continuum-surface-force implementation of surface tension. T h e simulations capture the sequence of events discussed above: a crater is formed, which collapses to form a jet, from which secondary droplets are ejected. Comparison of a time sequence of calculated interface shapes to an experimental sequence shows good agreement. T h e simulations illustrate the large velocities t h a t develop in the jet. T h e simulations will be used to investigate the physical mechanism responsible for droplet ejection. T h e effect of the driving parameters and physical parameters on the ejection process, including the size and velocity of the secondary droplets, will be evaluated. This will be used to gain an understanding of high-frequency atomization which, in turn, will be applied to the design of systems involving spray formation. This technology has application in a wide range of aerospace, industrial, and biomedical applications. A heat transfer cell t h a t utilizes this technology is under development for microelectronics-cooling applications. This cell is similar t o a heat pipe, but has a higher liquid flow rate t h a t results in a substantially higher heat transfer rate.
266
INTERFACIAL D Y N A M I C S ASSOCIATED W I T H EVAPORATION OF LNG IN A STORAGE T A N K SANG W. J 0 0 School of Mechanical Engineering, Yeungnam University, Gyongsan 712-749, Korea CHULSOO PARK AND SEONGHO HONG KOGAS, Korea
Liquefied natural gas (LNG) is stored in an insulated tank where heat transfer throught the tank wall into the liquid is released by the evaporative heat loss at the liquid/vapor interface inside the tank. Excess vapor is transported out and the vapor pressure is usually kept constant. When the tank is refilled (usually bottom filled) with new cargo, it has different constituent concentrations from the existing LNG in part due to the preferential evaporation of constituents. As a result the new LNG usually has higher density, and a stable stratification results. The new LNG in the bottom, however, does not release heat efficiently due to the lack of an evaporating interface. As the heat builds up, the density approaches that of the upper layer and a reversal of the layers can occur. In the present study, we examine this multiphase flow of liquid and gas mixtures by stability analysis, a dynamical-systems approach, and finite-difference computations. In particular, the conditions for the reversal and the elapsed time are presented.
267
OSCILLATORY THERMOCAPILLARY C O N V E C T I O N GENERATED B Y A B U B B L E M. KASSEMI AND N. RASHIDNIA National Center for Microgravity Research, NASA Glenn Research Center, Cleveland, OH 44135
In this work, we study steady and oscillatory thermocapillary and natural convective flows generated by a bubble on a heated solid surface. The dynamic characteristics of the time-dependent convection are captured using a combined numericalexperimental approach. The index of refraction fringe distribution patterns constructed numerically by taking an inverse Abel transform of the computed temperature fields are compared directly to the experimental Wollaston Prism (WP) interferograms for both steady state and oscillatory convection. The agreement between numerical predictions and experimental measurements is excellent in all cases. It is shown that below the critical Marangoni number, steady-state conditions are attainable. With increasing Ma number, there is a complete transition from steady state up to a final non-periodic fluctuating flow regime through several complicated symmetric and asymmetric oscillatory states. The most prevalent oscillatory mode corresponds to a symmetric up and down fluctuation of the temperature and flow fields associated with an axially travelling wave. Careful examination of the numerical results reveals that the origin of this class of convective instability is closely related to an intricate temporal coupling between large-scale thermal structures that develop in the fluid in the form of a cold finger and the temperature-sensitive surface of the bubble. Gravity and natural convection play an important role in the formation of these thermal structures and the initiation of the oscillatory convection. Consequently, at low-g, the time evolution of the temperature and flow fields around the bubble are very different from their 1-g counterparts for all Ma numbers.
268 A B O U T COMPUTATIONS OF THIN-FILM FLOWS
L. KONDIC Department of Mathematical Sciences, New Jersey Institute of Technology, Newark, NY 07102 J. DIEZ AND A. BERTOZZI Department of Mathematics, Duke University, Durham, NC 27708-0320
We explore the computational performance of different models used to compute the flow of a thin film down an inclined plane, within the framework of the lubrication approximation. Our fD computations show t h a t the results and computational efficiency strongly depend on the model t h a t is used to deal with the contact-line singularity. This study allowed us to develop efficient fully nonlinear 2D simulations. Preliminary 2D computational results concerning contact-line instability will be presented.
269
VISUALIZATION OF C O N V E C T I O N IN LIQUID METALS JEAN N. KOSTER Department of Aerospace Engineering Sciences, University of Colorado, Boulder, CO 80309-0429
A flow-visualization technique for density fields in liquid and solidifying alloyed metals has been developed. Density fields have been visualized similar to interferometry conventions. Density fields include information on temperature and concentration. In alloys, concentration is the dominant signal, thus chemical segregation in the melt can be revealed. Natural convection and double-diffusive flows have been visualized along with solidification and melting. Some visualized results are challenging theoretical analysis.
270
TIME-EVOLVING INTERFACES IN VISCOUS FLOWS MARY CATHERINE A. KROPINSKI Department of Mathematics and Statistics, Simon Praser University, Burnaby, British Columbia V5A 1S6, Canada
We present numerical methods for computing the motion of two-dimensional bubbles or drops in a slow viscous flow. New methods are presented for both the solution of the governing fluid equations and for the time integration of the evolving interfaces. The interfacial velocity of the bubble or drop is found through the solution to an integral equation whose analytic formulation is based on complexvariable theory for the Stokes equations. The numerical methods are spectrally accurate and employ a fast multipole-based iterative solution procedure, which requires only 0{N) operations where N is the number of nodes in the discretisation of the boundary. It is known that the dynamic equations for the interfacial motion become stiff as the curvature of the interface increases. A small-scale decomposition is performed to extract the dominant term driving the stiffness. By introducing an appropriate tangential velocity into the dynamics, this dominant term becomes linear. This leads to implicit time-integration schemes that are explicit in Fourier space. Examples will include the deformation of drops in an extensional flow and the viscous sintering of glass.
271
I N F L U E N C E OF A N O N L I N E A R EQUATION OF STATE O N C O N T A M I N A T I O N FRONTS AT A I R / W A T E R INTERFACES J. M. LOPEZ Department of Mathematics, Arizona State University, Tempe, AZ 85287-1804 A. H. HIRSA Department of Mechanical Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180-3590 The presence of surface-active materials, ubiquitous at most gas/liquid interfaces and particularly at the air/water interface, has a pronounced effect on the stress balance at the interface, and this in turn is nonlinearly coupled to the bulk flow. Even with minute amounts of surfactants on a liquid free surface, the boundary conditions for the Navier-Stokes equations are functions of the interfacial viscoelastic properties. The Boussinesq-Scriven constitutive relation for stress at a Newtonian interface consists of three intrinsic properties of the interface. These are the surface tension, er, the surface shear viscosity, fist and the surface dilatational viscosity, KS; consistent measurements of the latter have not yet been reported. These viscoelastic properties vary with the surfactant concentration at the interface, c. Here, we present a fundamental description of the interface and its coupling to the bulk flow and develop an axisymmetric Navier-Stokes numerical model. We perform both numerical studies using this model and experimental measurements of the liquid flow in an annular region bounded by stationary inner and outer cylinders and driven by a rotating floor. The difficulties in obtaining a clean free surface even when triply distilled water is used will be reported. DPIV measurements with a clean free surface will be presented. Computational results from our Navier-Stokes model that incorporates measured a(c) and f/,s(c), and modeled KS(C) for hemicyanine (an insoluble surfactant) will be presented. The computations for surfactant cases are compared with the hydrodynamics of a clean interface and a no-slip rigid surface. These computations provide insight into the dynamics that result from the surfactant with a nonlinear equation of state and finite surface viscosity. In this presentation, we focus on low initial surfactant concentration cases where a contamination front is observed and find that its location varies linearly with the initial concentration. Typically, when the Marangoni stress (due to surface-tension gradients) is dominating, the interface is thought of as immobile, acting as a no-slip surface. However, this is only true for the velocity components in the directions that would have lead to surfactant-concentration gradients. In the present axisymmetric swirling flow, this direction is radial. In the azimuthal direction, since the flow is axisymmetric, there are no azimuthal gradients and so there are no Marangoni stresses acting in that direction. Thus, in the radial direction, the Marangoni stress makes the interface act like a no-slip surface, but in the azimuthal direction it is essentially stress-free. This has fundamental consequences for models of contaminated interfaces that are not planar two-dimensional, where surfactant coverage does not simply mean that the interface is no-slip.
272
STABILIZATION OF A N ELECTRICALLY C O N D U C T I N G CAPILLARY B R I D G E FAR B E Y O N D THE RAYLEIGH-PLATEAU LIMIT USING F E E D B A C K CONTROL OF ELECTROSTATIC STRESSES MARK J. MARR-LYON, DAVID B. THIESSEN, FLORIAN J. BLONIGEN, AND PHILIP L. MARSTON Department of Physics, Washington State University, Pullman, WA 99164-2814
A liquid bridge between two solid surfaces is known as a capillary bridge and the stability of such bridges is relevant to materials processing as well as to other aspects of the management of liquids. For a cylindrical bridge in low gravity of radius R and length L, the slenderness S = L/2R has a natural (Rayleigh-Plateau) limit of •K beyond which the bridge breaks. We demonstrate a novel method of suppressing the growth of this mode on an electrically conducting bridge surrounded by an insulating liquid of the same density in a Plateau tank. The shape of the bridge is optically sensed as in our related demonstration of acoustic stabilization [M. J. Marr-Lyon et al., J. Fluid Mech. 351, 345 (1997)]. In the present stabilization method the optical information is used to control the potentials on a pair of ring electrodes concentric with the bridge. Slenderness values can be made to reach 4.46 when the generalized feedback force for the mode of interest is taken to be proportional to the modal amplitude. At S = 4.49, the next higher mode (which is not controlled in our current experiments) is predicted to become unstable. The electrical conductivity of the bridge liquid need not be large. Supported by NASA.
273
STABILIZATION OF CAPILLARY B R I D G E S IN AIR FAR B E Y O N D THE RAYLEIGH-PLATEAU LIMIT IN LOW G R A V I T Y USING ACOUSTIC RADIATION P R E S S U R E MARK J. MARR-LYON, DAVID B. THIESSEN, AND PHILIP L. MARSTON Department of Physics, Washington State University, Pullman, WA 99164-2814
In the absence of gravity, cylindrical capillary bridges consisting of liquid between two circular supports naturally become unstable and break when the length L of the bridge exceeds its circumference. This is the Ray leigh-Plateau limit where the slenderness S = L/2R is -K and R is the bridge radius. In experiments performed aboard NASA's low-gravity KC-135 aircraft, it was found that acoustic radiation pressure can be used to stabilize capillary bridges against breakup. Capillary bridges composed of a mixture of water and glycerol were deployed in a 21 kHz ultrasonic standing wave in air. The bridges were extended to a slenderness as great as S = 4.1 prior to breaking. Bridges extended beyond n broke immediately when the ultrasound was turned off. In contrast with previous work [M. J. Marr-Lyon et al, J. Fluid Mech. 351, 345 (1997)], this stabilization method does not use active feedback; the stabilization is a passive effect of the sound field. The acoustic wavelength is chosen such that the average radiation pressure due to the sound field is a function of the local bridge radius, so that areas of larger radius are squeezed, and areas of smaller radius are expanded. The KC-135 environment, however, is poorly suited for exploring the limitations of this method and the passive acoustic stabilization method cannot be simulated in Plateau tanks. Supported by NASA.
274
I N T E R A C T I O N S B E T W E E N HELE-SHAW FLOWS A N D DIRECTIONAL SOLIDIFICATION: N U M E R I C A L SIMULATIONS ECKART MEIBURG Department of Mechanical and Environmental Engineering, University of California, Santa Barbara, Santa Barbara, CA 93106
High-accuracy numerical simulations are presented for the nonlinear evolution of directional solidification processes, both in the absence and in the presence of potential flow in the melt. To this end, a numerical method is employed that utilizes an analytical boundary-fitted coordinate transformation in conjunction with a combination of spectral and finite-difference discretizations. The flow field is computed by a boundary-element technique. The accuracy of the computational scheme is demonstrated by comparing the growth rates at small interfacial amplitudes with linear stability results. In calculations with random initial perturbations, we observe the emergence of a large-amplitude dominant wavelength in agreement with the predictions by linear stability theory. Furthermore, the simulations demonstrate the stabilization of the solidification process by a uniform flow parallel to the interface. Results are presented as a function of wavenumber, morphological parameter, and parallel-flow velocity.
275
MODELLING T H E CONTACT REGION OF A N E V A P O R A T I N G M E N I S C U S W I T H A V I E W TO APPLICATIONS S. J. S. MORRIS Department of Mechanical Engineering, University of California, Berkeley, Berkeley, CA 94720-1740
The evaporating meniscus of a perfectly wetting liquid exhibits an apparent contact angle 0 owing to small-scale, evaporatively driven flow in the contact region. Existing theory by Wayner predicts © and the heat flow as the solution of a freeboundary problem. Published simulations use that theory to model the behaviour of grooved heat pipes. In that application, the heat flow across the meniscus in a millimetre-sized groove is found by dividing the meniscus in two. The heat flow from the inner contact region is found from the theory of the small-scale flow. That across the outer visible meniscus is found by solving the two-dimensional conduction equation. The simulations cover scales from nanometres to millimetres. At the top end of this range, they include the geometry of the solid from which liquid evaporates. At the other end, they use the existing theory for the small-scale flow. This covers a continuum of behaviours. The extremes include the isothermal meniscus that does not show an apparent contact angle; and two forms of strongly evaporating menisci that do. The first of these occurs in the limit of vanishing flow resistance to evaporation. All evaporation then occurs from a quasi-parallel film at dimensions smaller than those on which 0 is established [1]. The second occurs in the limit of vanishing thermal resistance to evaporation. The contact region is then isothermal, and all heat flow occurs at dimensions larger than those on which 0 is established [2]. As there is no previous analysis of the existing theory, these possibilities are not recognised in the literature. Yet the case of vanishing thermal resistance is common in applications. By emphasising it, one obtains insight, and useful formulae for the heat-pipe problem. Morris [3] simplifies the heat-pipe problem by dividing it into several pieces. To begin, scaling is used to find conditions under which there is negligible heat flow at the dimensions on which 0 is determined. The heat flow is then found from a linear conduction problem in which 0 appears as a parameter. The geometrical complications associated with heat pipes are added to the conduction model in two stages. A simplified treatment of these effects is possible as two conditions typically apply. The solid is highly conductive and 0 is small. Analysis of a model problem (linear meniscus on a conductive slab) shows that for vanishing liquid-solid conductivity ratio, the heat flow is asymptotically independent of solid shape. This is plausible as the heat flow would be independent of the shape of a perfectly conducting, and so isothermal, solid. The predictions of the model thus apply to more complex solid geometries. Next, meniscus curvature is incorporated for it influences the heat flow. Existing simulations include the meniscus shape by solving the two-dimensional conduction equation for the heat flow across the visible meniscus. In that method, the heat flow is a functional of meniscus shape. The new method simplifies the picture. It is shown that for small 0 , the heat flow across
276
an arbitrary meniscus can be found by approximating the curve by its osculating parabola at the contact line. Since this parabola is specified by G and the curvature at the contact line, the heat flow is a function of certain parameters rather than a functional of meniscus shape. The new method gives the heat flow explicitly. The predictions are confirmed by published simulations. Morris [2] extends those results by using the existing theory of the contact region to obtain a formula for G. It is shown that for vanishing thermal resistance, the flow has an inner and outer structure. The inner contact region transfers negligible heat, but defines 0 . Heat is transported by pure conduction through the outer region in a geometry imposed by 0 . This part of the analysis overlaps that in the previous paragraph. Formulae are given for G and the total heat flow q as functions of the model parameters. The predictions are confirmed numerically. The phenomenological variables 0 and q depend weakly on atomic-level wetting physics, whose overall role is to impose a scale, namely the film thickness seen at the apparent contact line by the distant flow. 0 depends weakly on this scale, and q depends on wetting physics only through 0 . The poster presented at this meeting covers this analysis of the small-scale flow. References 1. S. J. S. Morris, "Effect of flow resistance in the contact region of an evaporating meniscus," in prep. (2001). 2. S. J. S. Morris, J. Fluid Mech. 432, 1 (2001). 3. S. J. S. Morris, J. Fluid Mech. 411, 59 (2000).
277
A S P E C T S OF VORTEX D Y N A M I C S AT A FREE SURFACE BILL PECK, LORENZ SIGURDSON University of Alberta, Edmonton, Alberta, Canada PETROS KOUMOUTSAKOS, AND JENS WALTHER ETH Zurich, Switzerland
In our present work we are investigating the formation and subsequent instability of vortex rings formed by drops impacting a pool. We are especially interested in two aspects of the problem: the vorticity-generation mechanism at the free surface and the instability of the vortex ring at early times after impact. This instability leads to shedding of large-scale structures into the vortex-ring wake that we refer to as petals. In some cases, the petals pinch at their tips forming small vortex rings that travel away from the central axis of the primary vortex ring. We refer to this structure as the blooming vortex ring. For a more thorough discussion of the vortex structure, see [1]. The intensity, or level, of tangential vorticity u5t at a free surface can be predicted with the expression ut = 2n x n, where ft is the free-surface normal vector and an overdot denotes the material derivative [2]. After the drop and pool coalesce, the free-surface normal vector of a particle embedded in the free surface rapidly rotates and vorticity must be present to satisfy the shear-free boundary condition. This argument predicts the sign of vorticity observed in experiments. This argument does not explain however, the means by which torques are applied to fluid elements near the interface to create vorticity. An analysis of the vorticity equation accounting for steep viscosity and density gradients across the interface shows that several mechanisms are likely responsible. These include the well-known baroclinic torque, as well as viscous torques and torques that result from terms dependent on viscosity- and density-gradient coupling. Also, surface tension will lead to vorticity creation through surface gradients in the mean curvature. We are also interested in the mechanism that leads to the instability of the vortex ring. A high wave number instability appears on the vortex ring at early times (5-10 ms) after impact. The wavenumber of the instability decreases with time from about 20 to three, four, or five. This instability does not appear to be a straining instability such as the Widnall instability, but rather a secondary instability of filaments on the primary core. The unstable filaments then wrap around the primary core to form a series of counter-rotating pairs. These interact with one another until a preferred wavenumber is found. These larger pairs convect away from the primary core under their mutual self induction and escape the trapped streamlines of the primary vortex ring. This leads to the formation of petals that are seen in the wake. We are presently investigating this instability with a vortex-in-cell algorithm as used by [3] to study the formation of the Widnall instability on an initially perturbed vortex ring. In that case, the peaks of the waves eventually break leading to vortex filaments being wrapped around the primary core and eventually being
278
ejected into the wake. The drop-formed ring instability seems to be substantially different and we hope that the numerical analysis will lead to useful insights. We also hope to determine what part the presence of the interface plays in the onset of this instability. Acknowledgments We thank Karim Shariff for many interesting discussion on vortex rings and acknowledge the financial support of the Natural Sciences and Engineering Research Council of Canada. Also, BP thanks the participant support fund for their assistance in attending this conference. References 1. B. Peck and L. Sigurdson, Phys. Fluids 6, 564 (1994). 2. B. Peck and L. Sigurdson, IMA J. of Appl. Math 6 1 , 1 (1998). 3. P. Koumoutsakos, J. H. Walther, and J. B. Sagredo, "DNS of vortex rings using vortex methods," Phys. Fluids, to appear (2001).
279
INSTABILITIES AT THE "INTERFACE" B E T W E E N MISCIBLE FLUIDS — E M E R G E N C E OF A N EFFECTIVE SURFACE TENSION PHILIPPE PETITJEANS, PASCAL KUROWSKI, AND JUAN FERNANDEZ Laboratoire de Physique et Mecanique des Milieux Heterogenes, UMR CNRS 7636, Ecole Superieure de Physique et de Chimie Industrielles (ESPCI), 10, rue Vauquelin, 75005 Paris, France
Instabilities can develop at the "interface" between two miscible fluids that exhibit patterns very similar to those observed between immiscible fluids. For example, a layer of colored glycerine hanged-up below a glass plate and placed into a water tank gives rise to organized patterns. Also, a thin layer of glycerine falling into water is unstable and produces falling columns with a regular distance between each other. These similarities between miscible and immiscible fluids lead to the possible existance of an "effective surface tension" between miscible fluids. This effective tension has a meaning for short times only, and should tend to zero with time. It can be explained by the strong concentration gradient at the "interface."
280
S E C O N D A R Y INSTABILITIES OF FALLING FILMS USING MODELS CHRISTIAN RUYER-QUIL AND PAUL MANNEVILLE Laboratoire dHydrodynamique (LadHyX), Ecole Polytechnique, F-91128 Palaiseau Cedex, France
Viscous fluids flowing down inclined planes are of special interest in the study of pattern formation and the transition to spatio-temporal chaos. At moderate Reynolds numbers, their dynamics is controlled by surface-tension effects and viscous dissipation. Making use of the slaving principle, one can eliminate most local internal flow variables that are essentially bound to follow the slow evolution of the film thickness h. Combining a systematic expansion of the Navier-Stokes equation in powers of a small dimensionless parameter measuring the amplitude of thickness gradients to a Galerkin approximation method involving a functional basis made of polynoms, we have derived several models of increasing accuracy and complexity In the two-dimensional case, our "best model" is consistent to second order in the gradient expansion and involves four slowly varying quantities: h(x, t), the flow rate q(x,t), and two supplementary fields describing local corrections to the basic semi-parabolic flow profile. Though yielding results in excellent agreement with direct numerical simulations [2] and experiments [3], it is somewhat difficult to use. A simplified version involving just h and q has been obtained from the secondorder equations by neglecting the corrections to the flow profile. Slightly more complicated than Shkadov's model [4], it contains all relevant physical processes, and especially the effects of viscous dispersion. Quantitatively acceptable results are obtained with the simplified model in a domain of flow parameters extending from the onset of the free-surface instability to deep in the strongly nonlinear regime of large-amplitude solitary waves. The critical Reynolds number is correctly predicted, the main characteristics of the waves are faithfully reproduced, e.g., velocity vs. wavelength in the saturated-wave regime, and the simulated film evolution is apparently free of finite-time blow-up. The simplified model can be straightforwardly extended to deal with threedimensional flows by including spanwise modulations. This allows us to tackle the wave-stability problem in the most general context. Floquet analysis in the moving frame has been developed yielding results in general agreement with those obtained by the same method within the boundary-layer approximation [5]. The agreement with experimental results [6] is less good. However, since the extension of the complete model consistent to second order is too complicated to be of use, for the moment it seems difficult to attribute the observed discrepancy to a limitation of the simplified model that would show up only in three dimensions rather than to some sensitivity to the precise choice of physical parameters fitting the experiments. To conclude, the work done "by hand" in deriving models of film flows seems rewarding when compared to computationally expensive direct numerical simulations. The enslaving of hydrodynamic fields to the local thickness and flow rate
281
seems to be an important factor explaining the success of a low-order truncation of the Galerkin approximation mostly relying on the functional dependence of the basic-flow profile. Our results support the hope to "understand" the growth of disorder and the transition to turbulence in open flows. References 1. Ch. Ruyer-Quil and P. Manneville, Eur. Phys. J. B 15, 357 (2000). 2. T. R. Salamon, R. C. Armstrong, and R. A. Brown, Phys. Fluids 6, 2202 (1994). 3. (a) J. Liu, J. D. Paul, and J. P. Gollub, J. Fluid Mech. 250, 69 (1993). (b) J. Liu and J. P. Gollub, Phys. Fluids 6, 1702 (1994). 4. V. Ya. Shkadov, Izv. At Nauk SSSR, Mekh. Zhi Gaza 2, 43 (1967). 5. H. C. Chang, E. A. Demekhin, and D. I. Kopelevitch, J. Fluid Mech. 250, 433 (1993). 6. J. Liu, B. Schneider, and J. P. Gollub, Phys. Fluids 7, 55 (1995).
282
INTERFACIAL P H E N O M E N A IN S U S P E N S I O N S UWE SCHAFLINGER AND GUNTHER MACHU Institut fur Sromungslehre und Warmeubertragung, Technical University Graz, Inffeldgasse 25, A-8010 Graz, Austria
Recent resuspension experiments suggest that a possible interfacial tension exists between pure fluid and a suspension consisting of the same fluid and heavy, small, solid beads of identical size and density. Since little is known about interfacial phenomena in suspensions we experimentally investigated the formation and expansion of a suspension drop in the same fluid. To our surprise, the motion of the droplet exhibits all phenomena demonstrated by the classical experiments in which vortex rings of one liquid are created in another from drops falling from rest under gravity. Membranes formed even when the concentration of particles was smaller than 5%. We also observed the breaking of the torus by the Rayleigh-Taylor instability and the formation of a cascade of new rings. Two drops falling one behind the other penetrate but do not lose their distinct nature for several moments until they finally mix and move as a single drop. Larger drops typically create a long cylindrical tail of particles throughout the vessel that contains the clear liquid. This column is unstable and it breaks and forms small 'capillary droplets.' By making use of earlier investigations we were able to estimate the interfacial tension from the growth of the torus.
283
SESSILE DROP SOLIDIFICATION WILLIAM W. SCHULTZ Department of Mechanical Engineering, The University of Michigan, Ann Arbor, MI 48109-2125 M. GRAE WORSTER Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Silver Street, Cambridge CBS 9EW, England, UK DANIEL M. ANDERSON Department of Mathematical Sciences, George Mason Fairfax, VA 22030
University,
Containerless solidification, confining its melt by surface tension, is an important technique to produce very pure materials. T h e form of the solidified product is sensitive to conditions at the tri-junction between the solid, the melt, and the surrounding vapor. We reconsider the analysis of Anderson, Worster, and Davis (1996) to test whether a more simple tri-junction condition can model experimental behaviour when the flat solidifying interface is no longer imposed. We find t h a t the simpler condition can describe the inflection point and cusp of the axisymmetric drop if the solid-liquid interface is no longer assumed to be flat. T h e solution is now checked asymptotically for large and small time and for short distances from the tri-junction condition.
284
A HOMOGENIZED MONTE-CARLO MODEL FOR FILM G R O W T H TIMOTHY SCHULZE Department of Mathematics, University of Tennessee, Knoxville, TN 37996-1300
We derive a fully continuous, macroscopic model for epitaxial film growth from an analogous, but microscopic, Monte-Carlo procedure. Our approach features an explicit adatom density that evolves subject to deposition, diffusion, nucleation, and edge-adsorption of adatoms and is coupled to a separate height evolution equation. A specific aim of this approach is to incorporate multi-species effects, which are important in many systems (e.g., YBCO). To this end, it is natural to make a distinction between an atom on the film surface but not yet incorporated into the film (i.e., in the surface-density field), and an atom that has become part of a completed unit cell (i.e., in the film-height field). Such a distinction is familiar in mesoscopic models involving diffusion on terraces and explicitly represented ledges, but is absent from existing continuum models. Thus, the multi-species version of this model has an evolution equation for the density of each species on the surface of the film.
285
I N F L U E N C E OF T H E S U R R O U N D I N G C O N D I T I O N S N E A R T H E INTERFACE ON T H E STABILITY OF LIQUID B R I D G E S V. M. SHEVTSOVA, M. MOJAHED, AND J. C. LEGROS MRC, Universite Libre de Bruxelles, CP-165, 50 av. F. D. Roosevelt, 1050 Brussels, Belgium
Time-dependent thermal convection has been investigated experimentally in deformed liquid bridges of silicone oil with a Prandtl number Pr — 105. The temperature oscillations were measured by five thermocouples placed through the upper rod into different azimuthal positions in the liquid bridge some distance from the free surface. The signals from the thermocouples were taken for Fourier analysis and determination of their phase shift. The experimentally obtained stability diagram (ATcr,V) shows that the transition from axisymmetric steady flow to an oscillatory flow is very sensitive to variations in the volume. It consists of two different oscillatory instability branches belonging to different azimuthal wave numbers. Between them, there is a small range of volumes for which the steady flow is stable up to very high values of AT. For high Prandtl fluids, the branch on which ATcr increases with increasing volume has an azimuthal wave number m = 1, and the descending branch has an azimuthal wave number m = 2. It was found that the beginning of the gap is linked to the upper contact angle a. The physical explanation of the stability diagram follows the arguments proposed by Shevtsova and Legros [Phys. Fluids 10, 1621 (1998)]. A detailed analysis of the power spectrum and phase shift of the thermocouple signals reveals that the instability begins as a mixed mode, with wave numbers m = 0 and m — 1, within a narrow interval of AT. This is followed by a nearly standing wave with wave number m = 1 that changes to an m = 1 travelling wave when AT > 1.2ATcr. It has been proven experimentally that if standing (or travelling) waves are established in the liquid bridge, they can exist indefinitely as long as experimental conditions are maintained constant. To the best of our knowledge, there has been no previous experimental study of the influence of the average temperature inside the liquid bridge (the temperature of the cold rod) on the onset of instability. It is shown that the critical Marangoni number and critical wave number are very sensitive to the average temperature in the liquid bridge (e.g., to the temperature of the cold rod). Is this only an effect of temperature-dependent viscosity, or is it some other effect? The reason for such behaviour remains obscure. We have also discovered the possibility of changing the most dangerous mode at the threshold of instability. By surrounding the liquid bridge with another cylindrical volume with a larger internal diameter and kept at a constant temperature, the critical mode m = 2 can be switched to m = 1.
286
CONVECTIVE-DIFFUSIVE LATTICE MODELS OF INTERFACIAL A N D W E T T I N G D Y N A M I C S YITZHAK SHNIDMAN Department of Chemical Engineering and Chemistry, Polytechnic University, Six MetroTech Center, Brooklyn, NY 11201
Many molecular and colloidal systems exhibit coexistence of different phases, separated by interfacial regions with characteristic density/concentration profiles and curvatures. External stresses, temperature gradients, and chemical potentials drive such systems out of thermodynamic equilibrium, giving rise to distinct deformations and flows within and across interfacial boundaries that are coupled to those in the bulk. Dynamic interfacial phenomena occurring in multiphase and thermocapillary flows, and in processes such as wetting, coating, adhesion, friction, and lubrication, play a very important role in many physicochemical, biological, and industrial processes. Molecular dynamics is a realistic tool for simulating these phenomena, but it is limited to small systems and short simulation times, in which statistical fluctuations mask local mean densities and flows. A surface-thermodynamic level of modeling is another approach that has been very useful in the past. It postulates separate constitutive relations between strains and dissipative contributions to the stresses (rheology) at Gibbsian (infinitesimally thin) dividing surfaces. However, these constitutive relations proliferate for interfacial and non-Newtonian systems, and are hard to parametrize experimentally. In addition, the Gibbsian approximation leads to unphysical singularities in some important dynamic processes, such as wetting. We present a novel, convective-diffusive lattice-gas approach for modeling interfacial dynamics in multiphase flows. The equations of motion are expressed in terms of probabilities over microscopic lattice degrees of freedom consisting of molecular occupancies and Eulerian velocities at discrete lattice sites. In contrast to hydrodynamic lattice-gas and lattice-Boltzmann approaches, these velocities are not limited to discrete directions and magnitudes. Our method is based on a local self-consistent mean-field approximation for these probabilities and their moments, which are evolved by alternating discrete convective and dissipative timesteps, both subject to local conservation laws in discrete form. We model dissipation as a Markov-chain process corresponding to diffusion of the molecules and their momenta by short-range hopping that depends explicitly on molecular interactions, as well as on site occupancies and velocities. We demonstrate applications of the isothermal version of this method to computational modeling of interfacial dynamics, wetting, and coating phenomena occurring in multiphase flows that are in contact with moving solid walls, as well as the use of a nonisothermal version of this method for modeling thermocapillary flows.
287 A LEVEL-SET A P P R O A C H TO D O M A I N G R O W T H IN M U L T I C O M P O N E N T FLUIDS KURT A. SMITH, FRANCISCO J. SOLIS, AND MONICA OLVERA DE LA CRUZ Department of Chemical Engineering, Northwestern University, Evanston, IL 60208
During the late stages of spinodal decomposition, domains reach their equilibrium concentrations and sharp interfaces develop. In fluids, surface-tension-induced flow creates a hydrodynamic coarsening mechanism assuming domains are interconnected. For symmetric binary fluids, scaling arguments predict a growth rate of length ~ time at low Reynolds number and length ~ time^1!^ at high Reynolds number. Using a level-set method, we find evidence of a universal curve approaching the length ~ time^2'3' limit at large lengths (high Reynolds number). We find similar behavior in ternary systems. Growth behavior in ternary systems is rich due to the various possible topologies. For example, we find that growth of a disconnected phase may be promoted in the presence of two interconnected phases.
288
MORPHOLOGICAL INSTABILITY IN S T R A I N E D ALLOY FILMS
B. J. SPENCER Department of Mathematics, University at Buffalo, Buffalo, NY 14260-2900 P. W. VOORHEES Department of Materials Science and Engineering, Northwestern Evanston, IL 60208
University,
J. TERSOFF IBM Research Division
We analyze the development of compositional and surface nonuniformities during the growth of strained alloy films. A continuum model is derived for the evolution of the film surface and composition due to the processes of deposition and surface diffusion. T h e compositional stresses arising from compositional variations of different-size alloy components is incorporated into the elasticity problem for the epitaxial film, and the thermodynamics of stressed solids is used t o derive the chemical potentials for surface diffusion. From this coupled free-boundary problem for the morphology, composition, and stress state, the stability characteristics of planar film growth is determined from a linear stability analysis. For the case of equal surface mobilities of the alloy components, we find t h a t compositional stresses make the film more unstable to the formation of stress-driven surface undulations. T h e destabilization is greatest over a moderate range of deposition rates, b u t weak at fast and slow deposition rates. For the case of different surface mobilities of the alloy components, we find t h a t the difference in surface mobilities can completely suppress the stress-driven morphological instability. T h e stabilization is not symmetric with respect to the compositional and misfit strains: it occurs in films with compressive (tensile) misfit, when one of the atomic species in the film is b o t h large and slow (large and fast) relative to the other component. A feature of the stabilization due to mobility differences is the existence of a critical deposition rate for stabilization, above which the film is stable. T h e existence of a critical deposition rate suggests the possibility of growing controlled lateral composition modulations using the small-amplitude steady-state behavior of the system at deposition rates slightly below the critical condition.
289
STABILITY ISSUES IN SPIN-CASTING MOLTEN METALS PAUL H. STEEN School of Chemical Engineering, Cornell Ithaca, NY 14853-5201
University,
Planar-flow casting (PFC) (i.e., melt-spinning or spin-casting) produces flat product by forcing molten metal onto a moving substrate (i.e., a spinning wheel) where it solidifies and leaves as a ribbon, strip, or foil. Heat and momentum fluxes dominate in the gap between the nozzle and the wheel. Success of the planar-flow process depends on the dynamics and stability of a group of two- and three-phase contact regions. The stability of these contacts is crucial in spin-casting, impact-forming, die-casting, spray-forming, extrusion, or any casting process where bulk properties and surface quality are important. They must also be understood in the design of new processes that depend on rapid phase change and high throughput, whether to produce innovative materials or conventional materials more efficiently. The solidifying flow is distinguished by the following features: • Meniscus: deformable molten-metal/gas interface, • Tri-junction: 3-phase common-line with phase change, • Dynamic contact-line: 3-phase common-line with slip, • Solidification front: surface of solid/liquid phase change. Results: Recent experimental results are based on casting aluminum/silicon (Al/Si) alloys with Si content ranging from 0% (0.999 pure Al) to about 12% Si, depending on the cast. • Experiment: The movement of the upstream meniscus is coordinated with the contact exposure of the wheel, suggesting that the puddle length (or contact length) can respond to the heat-up of the wheel [1]. • Analysis: The flow structure within the puddle, predicted by a previous adhoc analysis [2,3] is rederived in a systematic way to clarify the appropriate scalings [4]. The resulting pressure field is tested against observation [5]. • Theory: Vorticity transport in solidification boundary layers has been elucidated [6]. Forming singularities are likely to be important in PFC where 'fresh' surface is created at high rates; contacting and forming singularities are illustrated in a simple context [7]. References 1. T. Ibaraki, Planar-Flow Melt-Spinning: Experimental Investigation on Solidification, Dynamics of the Liquid Puddle, and Process Operability, M.S. Thesis, Cornell University, Ithaca, NY (1996).
290
2. C. Karcher and P. H. Steen, Phys. Fluids 13, 826 (2001). 3. C. Karcher and P. H. Steen, Phys. Fluids 13, 834 (2001). 4. B. L. Reed, Planar-Flow Spin Casting: Momentum Transport, Vorticity Transport, and Texture Formation, Ph.D. Dissertation, Cornell University, Ithaca, NY (2001). 5. P. H. Steen, B. L. Reed, and M. B. Kahn, Solidification-Induced Secondary Flows in Spin-Casting, in Interactive Dynamics of Convection and Solidification, eds. P. Ehrhard, D. Riley, and P. H. Steen, pp. 145-154, Kluwer (2001). 6. B. L. Reed, A. H. Hirsa, and P. H. Steen, J. Fluid Mech. 426, 397 (2001). 7. P. H. Steen and Y.-J. Chen, Chaos 9, 164 (1999).
291
THE STABILITY OF THERMOCAPILLARY CONVECTION IN HALF ZONES WITH DEFORMED FREE-SURFACE PROFILES L. B. S. SUMNER Mercer University, School of Engineering, Macon, GA 31207 G. P. NEITZEL The George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405
Several theoretical analyses have examined the stability of half zones, always assuming that the free surface of the liquid bridge is cylindrical, as it would be in a microgravity environment. However, recent experiments by different groups have shown that significant free-surface deformation can have a pronounced effect on the stability boundary. Here, experiments and a theoretical stability analysis consider the onset of oscillatory thermocapillary convection in half zones with significantly deformed free-surface profiles. Free-surface deformation is quantified by a volume ratio representing the volume of liquid in the bridge relative to the volume of a right circular cylinder of equal height and endwall radius. The experiments mark the boundary for the onset of oscillatory flow in half zones with volume ratios ranging from 0.5 to 1.6. The theoretical analysis applies energy stability theory to steady, axisymmetric flow in the half zone under the conditions of the experiments. A comparison between the theoretical and experimental results show that the energy limits are above the experimental stability boundaries in clear contradiction to theory. An explanation for the exaggerated stability limits has not been determined although the results may provide new insights about the already well-researched instability mechanisms.
292
D I R E C T W R I T E OF PASSIVE CIRCUITRY USING INK-JET T E C H N O L O G Y J. SZCZECH, C. M. MEGARIDIS Department of Mechanical Engineering, University of Illinois at Chicago, Chicago, IL 60607 D. GAMOTA, AND J. ZHANG Motorola Advanced Technology Center, Schaumburg, IL 60196
Ink-jet technology is targeted for use in conformal electronics manufacturing where a homogeneous nano-particle suspension is utilized as the structural material. The suspension is dispensed onto flexible substrates by means of a piezoelectric droplet generator device capable of producing droplets on the order of 30-60 /im in diameter. The droplet generator is driven by a square-pulse waveform in which the rise/fall voltage and dwell time are variable. The substrates containing the dispensed material are then cured at 300°C for 15 minutes to yield a finished product. This method of circuit production is attractive to the electronics industry for it: 1) minimizes material usage and chemical waste production, 2) reduces turnover time, and 3) is capable of rapid prototyping. Ink-jet printing technology offers the potential for highly compact, ultra-lightweight assemblies and, especially, for rapid product introductions and enhancements, which are becoming increasingly critical in today's competitive wireless communications market. Current attempts in producing circuit interconnections at Motorola have proven to be successful. Although improved component tolerances and unique patterning capabilities have been verified using drop-on-demand ink-jet techniques at Motorola, it remains to be demonstrated whether it is capable of the reproducibility, robustness, and accuracy necessary for reliable production of commercial products. The repeatability of the process relies heavily on steady, satellite-free, droplet generations. To attain such optimum droplet generations it is essential to investigate the relation between the device excitation parameters/design and fluid properties. Previous studies show surface tension to be a critical fluid property in the generation of repeatable droplet formations. Therefore, quantification of the influence of surface tension in correlation with the device excitation parameters is required for successful commercial implementation. To quantify the drop-formation process, a short-duration flash videography technique is employed with time-delay capabilities down to 1 ^s. With this technique, it is possible to observe the droplet ejection and formation process as a function of various impulse excitation parameters. Key metrics, such as droplet size and velocity, are measured, and satellite occurrences are recorded to generate an extensive parameter matrix that assists in the optimization of the dispensing process.
293
STABILITY OF THE MENISCOID PARTICLE BAND AT A D V A N C I N G INTERFACES IN HELE-SHAW S U S P E N S I O N FLOWS* H. TANG, W. D. GRIVAS, T. J. SINGLER, J. F. GEER, AND D. HOMENTCOVSCHI Department of Mechanical Engineering, State University of New York, Binghamton, Binghamton, NY 13902-6000 Instability can occur in the displacement of two or more contiguous fluids when a less viscous fluid displaces a more viscous fluid. In Hele-Shaw flow, evolution of the instability leads to fingers of the less viscous fluid growing into the more viscous fluid as displacement proceeds. Viscous fingering has been observed in the displacement of colloidal suspensions. The operative physical principle in suspension fingering is that the effective viscosity of a dense suspension exceeds the viscosity of the pure carrier liquid and is a monotonically increasing function of the local particle volume fraction. In natural or industrial processes, and in laboratory experiments which seek to model such processes, the unfavorable viscosity contrast necessary for viscous fingering typically exists as an initial condition in the displacement process. Here, however, we discuss a physical system in which the conditions for viscous fingering are not initially present, but which evolve concomitantly with displacement. We observe the formation of an effectively more viscous region in Hele-Shaw, free-surface suspension flow. The increase in effective viscosity derives from the autogenous accretion of particles in a region adjacent to the free surface between the suspension and the fluid being displaced. This region can itself be unstable to displacement by the trailing suspension, which is effectively less viscous because of its smaller particle volume fraction. It is well known that the volume fraction of particles adjacent to a meniscus formed by a suspension in contact with another immiscible fluid increases if the meniscus is advancing. The mechanisms responsible for particle accumulation are complex, but a simple kinematic argument shows that a particle can enter the meniscus region along centerline-region streamlines faster than it can exit along wall-region streamlines because of the finite diameter of the particle. The cumulative effect of this differential kinematic particle flux is meniscoid banding. When the band achieves a critical thickness and particle volume fraction, fingers of the less viscous trailing suspension may be observed growing into the band. Observations show that band suspension is shed into finger sidebands, which due to their higher effective viscosity move more slowly than the suspension/air interface and tend to elongate with time. Because the band has finite radial extent, there is limited space for finger evolution and the complex evolution patterns observed in nonlinear growth into unbounded domains cannot occur here. However, sufficient band thickness is maintained to allow simple tip-splitting of the fingers and an increase in wavenumber with displacement. The suppression of viscous fingering in connection with recovery and regeneration processes has been a subject of considerable interest, and success in minimizing growth rates of instability for unfavorable viscosity contrasts has been achieved by
294
injection of a solute resulting in a graded viscosity in the interfacial zone. The account above pertains to the development of conditions favorable to instability of an initially stable flow. Conversely, we have also observed that, for initially unstable conditions, e.g., when an initially uniform suspension displaces a more viscous fluid, the evolution of an increased effective-viscosity band region can stabilize the interface between the suspension and displaced fluid. A weak local fingering instability is still observed in the narrow band region, but the global flow is stable. Thus, at least for length scales accessible by our apparatus, we conclude that autogenous band growth at the interface of an unstable couple leads to delayed growth of global instability, i.e., the evolution of the band region represents a natural stabilization mechanism of an otherwise very unstable interface, with the small price of a weak internal mode disturbance within the band. This research was supported by DARPA, N00164-96-C-0074, SUNY at Binghamton.
and the IEEC at
* Full paper appearing as "Stability considerations associated with the meniscoid particle band at advancing interfaces in Hele-Shaw suspension flows," H. Tang, W. Grivas, D. Homentcovschi, J. Geer, and T. Singler, Phys. Rev. Lett. 85, 2112 (2000).
295
INSTABILITIES IN T H I N FLUID SHEETS B. S. TILLEY Olin College of Engineering, 1735 Great Plains Avenue, Needham, MA 02492-1245 D. T. PAPAGEORGIOU AND R. V. SAMULYAK Department of Mathematical Sciences, New Jersey Institute of Technology, Newark, NY 07102
We investigate the interfacial dynamics of potential flow in a thin fluid sheet for varicose long-wave disturbances when surface forces and inertia are the dominant physical effects. A coupled set of evolutions equations is derived for the interfacial deflection and axial velocity. For sufficiently small amplitudes, spatially periodic disturbances on the film lead to counter-propagating waves whose wavespeeds depend on the initial data. For larger-amplitude initial data below a threshold, these standing-wave-type solutions persist. In this regime, general initial conditions lead to quasi-periodic solutions in time. However, there is a class of initial data for which the transients are simply travelling-wave solutions. Beyond this threshold, the film ruptures according to the similarity scales reported in Pugh and Shelley (1998), independent of the amplitude of the initial data above threshold. Near the rupture criterion, transients suggest a similarity solution that has a bounded velocity but is singular in its spatial gradient. Further, a similarity property of the evolution equations predicts the threshold criterion up to the evaluation of a single constant, which is obtained numerically.
296
T H E R M A L EFFECTS OF I N T E R N A L INTERFACES: EQUILIBRIUM M I C R O S T R U C T U R E A N D KINETICS A. UMANTSEV Department of Chemistry and Physical Sciences, Saint-Xavier University, Chicago, IL 60655
The thermal effects of internal interfaces, like interphase, antiphase, and domain boundaries, stem from their possession of an internal-energy excess. In my presentation, I will analyze the influence of the interfacial thermal effects on the formation of microstructures. These effects have received very little attention in the literature; their study is long overdue. Thermal effects significantly change the phase diagrams of one-component materials, in particular, those of thin films. For very thin films with thicknesses only 5-20 times greater than the interfacial thickness, phase separation does not occur and equilibrium is achieved with homogeneous transition states that can never be obtained in bulk samples because of their absolute instability. The thermodynamic and dynamic explanations will be presented. This type of a thin-film phase diagram is important for electronic devices and integrated circuits, for the theory of amorphization, and for nanophase composite materials where small particles or thin whiskers capable of undergoing a transition are immersed into a poorly conducting matrix. As large systems evolve far from equilibrium, they produce complicated fractal structures with high interfacial-area density, like dendrites during solidification, highly dispersed coexisting phases during martensitic transformation, or selfentangled domains of opposite ordering after continuous symmetry-breaking ordering below the critical temperature. I will elucidate the influence of different thermal effects on the general properties of evolving microstructures. One of them (the heattrapping effect) is caused by temperature gradients across an interphase boundary that change the course of dynamics and allow the emergence of a metastable (superheated) phase in the growth stage of the transformation. Conditions for such a regime may be met in many organic materials, including polymers. Another thermal effect of interphase boundary motion that will be discussed in conjunction with structure formation is the surface creation and dissipation effect. The temperature gradients across antiphase-domain or grain boundaries stem from the transmission of energy across the interface and cause a drag effect, which changes the course of structural coarsening for the later stages of material evolution. An experimental setup designed to reveal this thermal drag is suggested.
297
T H E ROLE OF L O N G - R A N G E FORCES IN T H E STABILITY OF NEMATIC FILMS M. P. VALIGNAT, F. VANDENBROUCK, AND A. M. CAZABAT Laboratoire de Physique de la Matiere Condensee, College de France, Paris, 75231 Cedex 05, France
We present an experimental study of spontaneous and forced wetting of 5CB in the nematic phase on silicon wafers. At room temperature, a droplet of 5CB completely wets the surface and the ellipsometric profile reveals the existence of a characteristic thickness (~ 200 A). As the temperature is increased, this thickness increases and "diverges" at the nematic/isotropic transition. This is the signature of a wetting transition: a 5CB droplet in the isotropic phase does not wet the surface. The question that arises is: why does the wetting transition takes place at the nematic/isotropic transition? In the discussion, we propose a simple model that takes into account long-range interactions and we show that the balance between elastic, anchoring, and van der Waals energies may fix the wetting condition. The knowledge of the energy per unit surface area allows us to predict the stability of a thin nematic film. For 5CB spun-cast onto silicon wafers, thin films are unstable while thick ones are metastable. We then observe experimentally the characteristic spinodal or nucleation dewetting in the two cases.
298
V I B R A T I O N - I N D U C E D D R O P ATOMIZATION BOJAN VUKASINOVIC, MARC K. SMITH, AND ARI GLEZER The George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405
The subject of this research is a process we call vibration-induced drop atomization (VIDA). The process starts with a small water drop (~ 1 cm in diameter) placed at the center of a horizontal, circular metal diaphragm. The diaphragm is fastened to a holder at its periphery and a piezoelectric disk is centrally mounted to its bottom side. Excitation of the piezoelectric disk by a sinusoidal voltage causes the diaphragm to vibrate in the vertical direction in its fundamental axisymmetric mode. When the excitation signal is held at a fixed frequency and the amplitude is slowly increased, different stages of free-surface motion on the drop are clearly observed. First, axisymmetric standing waves appear on the free surface for small values of the excitation amplitude. These waves have the same frequency as the excitation and are present at even very small excitation amplitudes. At a first critical excitation amplitude, an azimuthal instability is triggered along the contact line. It couples with the existing axisymmetric waves so that the free surface undergoes a transition to a non-axisymmetric wave form. Here, the most energetic spatial mode has a frequency that is the first subharmonic of the excitation frequency. This is a characteristic of the parametric instability seen in a standard Faraday-wave experiment. As the excitation amplitude increases further, the three-dimensional, free-surface waves increase in magnitude, spread over the entire free surface of the drop, and exhibit very complex, nonlinear spatial behaviors. At a second critical excitation amplitude, small secondary droplets begin to be ejected from the freesurface wave crests. When droplet ejection occurs, the ejection sites appear in a central region that covers about two-thirds of the free surface. After droplet ejection begins and for large enough excitation frequencies, complete atomization of the drop will suddenly occur as the result of a rapid increase in the rate of secondary droplet ejection — an event we call drop bursting. If the excitation amplitude is large enough and modulated by a step function, atomization occurs immediately. Even so, we still observe all of the different freesurface motions of the drop described earlier. We see the axisymmetric waves, then the growth of the azimuthal instability, and the development of complex wave patterns primarily near the contact line. Droplet ejection begins around the circumference of the primary drop near the contact line due to the large-amplitude growth of the azimuthal instability. The time from the start of the excitation until the first secondary droplets are ejected is about 15 forcing periods. These secondary droplet ejection sites then develop over the entire central part of the free surface of the drop. For this case, the entire primary drop is atomized in less than a second after the initiation of the excitation. The rate of droplet ejection depends on the excitation amplitude, but more interesting is that for a fixed excitation amplitude and frequency, the rate of droplet ejection can increase or decrease with time. This behavior depends entirely on
299
the driving frequency. A droplet can burst or atomize almost instantly, or the bursting can be delayed on the order of seconds to maybe a minute or more after the excitation is applied. In some instances, bursting does not occur at all, even though initial droplet ejection is present; the droplet-ejection process just dies out. All of these facts suggest that the evolution and the rate of droplet ejection depends on an interplay between ejection from the primary drop and the vibrating diaphragm. This dependence is fully understood and experimentally tested. The VIDA process is currently being utilized in a microelectronics cooling application using a self-contained heat transfer cell. It has also proven to be effective in other applications like spray coating, emulsification, encapsulation, and mixing enhancement.
300
CAPILLARITY-DRIVEN INSTABILITIES A N D T H E EVOLUTION OF SOLID T H I N FILMS HARRIS WONG Department of Mechanical Engineering, Louisiana State Baton Rouge, LA 70803
University,
MICHAEL J. MIKSIS Department of Engineering Sciences and Applied Mathematics, Northwestern University, Evanston, IL 60208 PETER W. VOORHEES Department of Material Science and Engineering, Northwestern Evanston, IL 60208
University,
STEPHEN H. DAVIS Department of Engineering Sciences and Applied Mathematics, Northwestern University, Evanston, IL 60208
Thin solid films are the basic structure in most microelectronic and optoelectronic devices. This is a direct result of the layer-by-layer manufacturing technique, which involves roughly a hundred steps of successive material deposition and patterning. During these processes, the device is often heated to high temperatures, and the solid components inside the device can deform. The deformation becomes more disruptive as the size of integrated circuits decreases. To continue this trend of miniaturization, the stability and evolution of thin solid films needs to be understood. In this poster, we will present the morphological instabilities and evolution of solid-film shapes commonly encountered in micro-devices, such as holes, wedges, rods, and steps.
301
INTERFACIAL WAVE THEORY FOR D E N D R I T E G R O W T H JIAN-JUN XU Department of Mathematics and Statistics, McGill University, Montreal, Quebec H3A 2K6, Canada
During the last decade two analytical theories for free-dendrite growth, the microscopic solvability condition (MSC) theory and the interfacial wave (IFW) theory, were proposed in the areas of condensed-matter physics and material science. These theories were attempts to resolve the problem of selection in dendrite growth, and to explain the essence of pattern formation. This article attempts to clarify the differences and commonalities between these two theories and to compare the predictions of these theories with some of the latest numerical evidence and experimental data. Since the MSC theory is the most well-developed for the two-dimensional case, the comparison of the theories with numerical simulations is made mainly by using, but is not restricted to, two-dimensional, numerical solutions for dendrite growth with anisotropy of surface tension. Such kinds of numerical simulations have been carried out by Wheeler et al. (1993), Provatas et al. (1998, 1999), and Karma et al. (1996, 1997) with the phase-field model, and by Ihle and Muller-Krumbhaar (1994) with the free-boundary-problem model. It is seen that if the anisotropy parameter is not too small, the numerical simulations yield steady needle solutions, whose side-branching structures are not selfsustaining. This result supports the conclusions drawn by both the MSC and the IFW theories. However, the numerical simulations also showed that there exists "a smallest value of the anisotropy parameter," less than which "no steady needle solution was found." This numerical evidence appears to be in agreement with the IFW theory and to contradict the MSC theory. The prediction of the IFW theory is also compared with the latest experimental data obtained by Glicksman et al. (private communication) in the microgravity environment and excellent overall agreement between both is found.
302
UNIFORMLY VALID A S Y M P T O T I C SOLUTIONS FOR DENDRITE GROWTH WITH CONVECTION JIAN-JUN XU AND DONG-SHENG YU Department of Mathematics and Statistics, McGill University, Montreal, Quebec H3A 2K6, Canada
The interaction of convection and dendrite growth has been a subject of great interest in the area of material science in recent years. Experimental observations have shown that convective motion in the liquid may have a significant effect on dendrite growth. The existence of convection may significantly change the growth velocity of the tip and the micro-structure pattern. Preliminary investigations of dendrite growth with convection are usually focused on the special case of zero surface tension. The solution for this special case, as in the Ivantsov solution for dendrite growth with no convection, cannot resolve the problem of the selection of the tip velocity of the dendrite or the dynamics of pattern formation on the interface. However, this solution is expected to provide a basis for further investigations in the general case of dendrite growth with non-zero surface tension. In the past several years, theoretical studies of steady dendrite growth in an external flow have been conducted by a number of authors (such as Ananth and Gill (1989, 1991); Benamar, Bouisson, and Pelce (1988); Saville and Beaghton (1988), etc.) both numerically and analytically. An analytical solution was obtained by Ananth and Gill in terms of the Oseen model of fluid dynamics, which led to a paraboloid shape for the interface. Their solution satisfies all the interface conditions, but does not satisfy a proper upstream far-field condition. In fact, their solution yields a non-vanishing perturbed flow in the upstream far field, which is not appropriate from a physical point of view. Xu (1994) made the first attempt to derive an asymptotic-expansion solution in terms of the Navier-Stokes model of fluid dynamics for the case of the Prandtl number, Pr —> oo. Xu derived a matched-asymptotic-expansion solution for the problem. However, for simplicity, the problem formulated in Xu's work neglected the regular condition for the velocity field at the symmetry axis. Furthermore, the same as in Ananth and Gill's work, the upstream far-field condition for the stream function in the formulation was weakened. In this paper, we attempt to reconsider this problem with a fully justified mathematical formulation. We assume that a dendrite is growing with a constant tip velocity U against the flow, which has a uniform velocity U„> in the far field. We study two limiting cases: • the weak-flow case: Uoo U, and derive uniformly valid asymptotic solutions.
303
COMPARISON OF A S Y M P T O T I C SOLUTIONS OF PHASE-FIELD MODELS TO A SHARP-INTERFACE MODEL G. W. YOUNG AND S. I. HARIHARAN Division of Applied Mathematics, Department of Mathematics and Computer Science, The University of Akron, Akron, OH 44325-4002
A one-dimensional directional solidification problem is considered for the purpose of analyzing the relationship between the solution resulting from a phase-field model to that from a sharp-interface model. An asymptotic analysis based upon a large Stefan number is performed on the sharp-interface model. In the phase-field case, the large-Stefan-number expansion is coupled with a small-interface-thickness boundary-layer expansion. The results show agreement at leading order between the two models for the location of the solidification front and the temperature profiles in the solid and liquid phases. However, due to the non-zero interface thickness in the phase-field model, corrections to the sharp-interface location and temperature profiles develop. These corrections result from the conduction of latent heat over the diffuse interface. The magnitude of these corrections increases with the speed of the front due to the corresponding increase in the release of latent heat. By properly selecting the potential coupling the order parameter and temperature in the phase-field model, and by tuning the kinetic parameter, we are able to eliminate the corrections to the outer temperature profiles in the solid and liquid phases of the phase-field model. This in turn eliminates the correction to the location of the solidification front. Hence, the phase-field temperature profiles agree with the sharp-interface profiles, except near the solidification front, where there is smoothing over the diffuse interface and no jump in the temperature gradients.
304
A N E W A P P R O A C H TO M E A S U R E T H E C O N T A C T A N G L E A N D T H E EVAPORATION R A T E W I T H FLOW VISUALIZATION IN A SESSILE D R O P NENGLI ZHANG AND DAVID F. CHAO NASA Glenn Research Center, Cleveland, OH 4-4135
The contact angle and the spreading process of a sessile drop are very crucial in many technological processes, such as painting and coating, material processing, film-cooling applications, lubrication, and boiling. Additionally,' it is well known that the surface free energy of polymers cannot be directly measured for their elastic and viscous restraints. Thus, measurements of the liquid contact angle on polymer surfaces become extremely important in order to evaluate the surface free energy of polymers through indirect methods linked with the contact-angle data. Since the occurrence of liquid evaporation is inevitable, the effects of evaporation on the contact angle and spreading become very important for a more complete understanding of these processes. It is of interest to note that evaporation can induce Marangoni-Benard convection in sessile drops [1]. However, the impact of convection inside the drop on the wetting and spreading processes is not clear. The experimental methods used by previous investigators cannot simultaneously measure the spreading process and visualize the convection inside the drop. Based on the laser shadowgraphic system used by the present authors [2,3], a very simple optical procedure has been developed to measure the contact angle, the spreading speed, the evaporation rate, and to visualize the convection inside of a sessile drop simultaneously. Two CCD cameras were used to synchronously record the realtime diameter of the sessile drop, which is essential for determination of both the spreading speed and the evaporation rate, and the shadowgraphic image magnified by the sessile drop acting as a thin plano-convex lens. From the shadowgraph, the convection inside the drop can be observed, if any, and the image outer diameter, which links to the drop profile, can be measured. Simple equations have been derived to calculate the drop profile, including the instantaneous contact angle, height, and volume of the sessile drop, as well as the evaporation rate. The influence of convection inside the drop on the wetting and spreading processes can be figured out through comparison of the drop profiles with and without convection when the sessile drop is placed at different evaporation conditions. References 1. N. Zhang and W. J. Yang, ASME J. Heat Trans. 105, 908 (1983). 2. N. Zhang and W. J. Yang, ASME J. Heat Trans. 104, 656 (1982). 3. N. Zhang and W. J. Yang, Rev. Sci. Instrum. 54, 93 (1983).
PART 3
PANEL DISCUSSION SESSION
This page is intentionally left blank
N E W RESEARCH DIRECTIONS IN INTERFACIAL SCIENCE M A R C K. S M I T H The George W. Woodruff School of Mechanical Georgia Institute of Technology, Atlanta, GA 30332-0405 U.S.A., E-mail: [email protected]
Engineering,
The last event of the conference was a panel discussion session chaired by one of the conference organizers, Michael J. Miksis [1]. The panelists were distinguished scientists from the fields of fluid mechanics and materials science and each one is a major contributor to the study of interfaces in his respective fields of research. The panel members were Stephen H. Davis (fluid mechanics and materials science) [2], Joel Koplik (fluid mechanics) [3], Wilfried Kurz (materials science) [4], Tony Maxworthy (fluid mechanics) [5], Robert F. Sekerka (materials science) [6], and Gretar Tryggvason (fluid mechanics) [7]. The panelists were chosen to strike a balance between the fields of fluid mechanics and materials science and also between the activities of theoretical, experimental, and numerical research. Professor Miksis asked each panelist to give a short statement on the general problem or problems he would like to see solved or worked on in the future. The intent of the discussion session was to bring out the similarities between interfacial dynamics in fluid mechanics and materials science and to suggest future avenues for research and cooperation. This account of the discussion is paraphrased from the complete transcript and was written by the author to clearly and smoothly describe the main themes raised during the session. A complete transcript of the discussion session was sent to the National Science Foundation in fulfillment of their contract-support requirements [8].
1
Introduction
There have been tremendous successes in understanding the essential physics of interfaces in the fields of fluid mechanics and materials science. These successes have often been made with similar techniques, such as stability theory and lubrication theory, and common numerical techniques such as phase-field, front-tracking, and level-set methods. The talks and poster presentations given during this conference have shown the similarities between the two research areas with respect to the progress that has been made to date. The job of the panel members in this discussion session was to stimulate discussion and further personal interactions in an attempt to help chart an appropriate course for future exploration in interfacial science. This paper is organized as follows. Section 2 is a paraphrased and rewritten account of the entire discussion session. It provides the background and context for the main themes summarized in Section 3. If desired, Section 3 can be read independently of Section 2. 2
The Flow and Micro structure
of the Discussion
Professor Davis began the discussion by saying, "I think predicting the future is hopeless, " Several individuals on the panel and in the audience expressed immediate agreement with this statement. However, the purpose of this discussion
307
308
M. K.
Smith
was not to predict.specific future events, but rather to chart a reasonable path for interfacial research in the future. Professor Davis continued by describing one example of the kind of work that he thought was important and necessary in order to make progress in understanding the behavior of complicated systems in the next century. He mentioned the biological problem presented by Gallez [9] in which she modeled in a general way the adhesion of a biological cell or vesicle to a solid substrate. In her model, the biological membrane was modeled as an interface between two liquid films of different viscosities. Charged, mobile surfactants populate this interface. Her interest was in the general deformation of the interface on length scales much larger than either film thickness. To obtain this, she used lubrication theory to develop a set of evolution equations for the behavior of the system that effectively reduced the microscale structure and effects of the membrane into a set of averaged "properties." Each property is expressed as a term in an evolution equation with an associated parameter or coefficient that governs the magnitude of that property. These equations work remarkably well in predicting the macro-scale dynamics of the system. This idea could be taken much further and more of the microscale features of the membrane could be added to the modeling, such as bending stresses, dilatational stresses, etc. In doing so, one could obtain a better model for the behavior of the cell or vesicle; a model that may truly predict the overall system responses on the larger length scales of interest. This is one avenue for the kind of research that can be done in the next century. There are structures, such as cell membranes, that are inherently thin entities when viewed on the larger length scale of the associated global system. Such structures may be modeled as an interface. This kind of modeling works very well when the global system responds to the presence of the interface in such a way that inessential details of the interfacial structure are forgotten. In effect, the global system responds to an averaged effect of the interface. This marriage of interfacial modeling and classical methods like lubrication theory may not work in every case, but when it does the results often prove to be very useful. [Author's note: Another example of this kind of modeling is the contribution to this volume of Worster [10]. In this work, a sea ice layer was modeled as an interface between the ocean and the atmosphere. By examining the solidification processes occurring in sea ice, including the mushy layer, Worster obtained a set of interfacial conditions that could be used to couple the global physical processes occurring on either side of the ice layer.) Professor Sekerka entered the discussion at this point by bringing up a different theme that was underlying the comments of Prof. Davis. The type of interfacial problems that will be studied in the future will be very complex. There will be problems containing fluid mechanics and materials science coupled with complicated biological, chemical, electrical, and certainly other more esoteric physical phenomena. In such problems, one quickly runs out of the expertise needed to do the necessary modeling, experimentation, and computations. So, this next century presents a great opportunity to create multi-disciplinary teams to work on problems of common interest. Such teams would include a modeler, an experimentalist, and an expert in numerical computations who may not necessarily be at the same
New Research Directions in Interfacial
Science
309
institution, but who get together to solve a real problem cooperatively, in a realistic time frame. Professor Davis agreed with this comment and added that the physical and chemical processes in biology, for instance, were so broad that no one could possibly do realistic modeling without some guidance from experiments. Thus, the use of cooperative teams will be essential in the next century to solve problems of real interest. Anyone involved in computing over the past few decades has seen the rapid increase in speed and capability of computers and a corresponding decrease in their cost. Professor Tryggvason echoed this trend when he said that in the near future it will be a routine matter to do computations of three-dimensional physical problems in complicated geometries that include strong couplings between various physical effects, such as fluid mechanics, solidification, surfactants, etc. All that will be required is to have the correct constitutive relations for the system. This is a very important and sometimes elusive requirement. To get it right, Prof. Tryggvason agreed with Profs. Davis and Sekerka that we will have to change the way we do business and team up with scientists in other fields in order to get a more informed look at the important problems they face. Continuing, Prof. Tryggvason stated two requirements on theory that he felt needed to be met in the next century. The first is to investigate complicated systems at molecular scales and build the information obtained from these studies into useful constitutive relations for continuum models. The second requirement is to go beyond the detailed continuum models that we will be able to compute and build useful engineering models. As an example, he cited a first principles simulation of the solidification of a binary alloy that includes a forest of dendrites with fluid flow and species segregation all coupled together, as would happen in a casting process. For engineering purposes, we must go beyond the details of the simulation and produce a realistic reduced-order model that is able to predict the average physical properties of the final product. This requirement echoes the statement of Prof. Davis when he suggested the use of classical methods like lubrication theory to average out the inessential details of the structure of an interface to produce a model that can faithfully predict the behavior of the corresponding system on the macroscale. In Prof. Tryggvason's requirement, we take the details from a complex and complete simulation of a physical system and find a way to produce a simple reduced-order model that can accurately predict the global properties of interest. At this point Prof. Timothy Pedley [11] commented from the audience that this idea of going from the microscale to the macroscale is especially important in the field of biology. Here, the challenge is first to find out how cells behave on an individual level, and then to use this information to construct an appropriate model for the function of a tissue or an organ in its in vivo environment, for example. In a brief exchange with the panel and the audience, Prof. Grigory Barenblatt [12] posed a question that can be paraphrased as follows. Suppose that you had the capability to determine the behavior of any physical system based on first principles, say, by using an extremely advanced computing system or a new mathematical technique. Then, what would you do with this capability? What questions would you ask? Professor Sekerka commented that this was indeed the critical
310
M. K.
Smith
question. How can we determine the truly important scientific questions or frontiers that need to be addressed? Professor Stephen Morris [13] replied from the audience that this is a question that can not really be answered. If one looks back on the previous century, an important driver for posing scientific questions was technology. Without knowing the technology, some questions would not even exist. For example, the investigation of compressible flow in the Twentieth century would have been rather pointless if humans had not been able to develop a means to travel faster than the speed of sound. {Author's note: While this is certainly a valid argument, we must realize that scientific knowledge also enables technology. An example here would be the development of the first solid-state transistor. That scientific achievement has driven technological development throughout the course of the computer age, and it is still an active driver today. In the end, it will come down to individual human interests. People will examine ideas that are important to them, either from a technological standpoint or purely from scientific curiosity. After examination, our task is to determine what ideas are worth pursuing in order to advance technological development or to promote economic activity.) Professor Joel Koplik changed the direction of the discussion by making some remarks about the treatment of complex fluids. Current continuum models of such fluids leave a lot of information out. Thus, it is extremely important to coordinate continuum models with molecular information obtained from molecular simulations and microscopic measurements from experiments, if possible. This is particularly important at the interface between two complex fluids. One area where improvements would be very useful is in the matching of molecular and continuum simulations. He cited Hadjiconstantinou [14] as one example of such work. Here, Hajiconstantinou simulated the moving contact-line problem by treating the flow in the contact-line region with a molecular model and the flow away from the contact line with the traditional Navier-Stokes continuum model. He developed a good way to match the information from the molecular and the finite-element fluid flow simulations so that they could be properly coupled. The reason this is important is that many large molecular simulations of a moving contact line, for example, use many molecules to simulate the solid surfaces and the liquid far away from the contact line where nothing much happens. The action occurs near the contact line and this region may contain only a small fraction of the total molecules in the simulation. This technique is obviously an inefficient use of available resources. In another remark, Prof. Koplik said that there are not many people doing molecular simulations because of the mistaken belief that they are very difficult computations requiring huge computer resources. This is not true. One major simplification that can be employed is similarity. All Newtonian fluids are similar so the simplest molecular-interaction potential, like Lennard-Jones, is a sufficient model. For surfactants, one need only use a molecule that remains on the interface between two fluids to model the effect of the surfactant on that interface. More complicated surfactant molecules just do the same thing and so their complexity can often be ignored. Molecular simulations of non-Newtonian fluids appear to be another story, because of the great variety of different fluids. However, Prof. Koplik expressed the opinion that there probably exist only a few canonical classes of non-Newtonian fluids and that each of these would have a relatively simple molec-
New Research Directions in Interfacial
Science
311
ular model. If this is true, then molecular models properly coupled to continuum calculations would become a very powerful technique for the future of interfacial research. Thus, more researchers need to become involved in this work in order to provide some healthy competition for future progress. Similar molecular (or atomic) computations can be done in the field of materials science. Professor Martin Glicksman [15] commented from the audience that he has seen some truly amazing progress in this area of ab initio computations taking place in some national laboratories. In these computations, one begins with only a description of the atoms or molecules involved and the algorithms will calculate the molecular structure of a material with details like annealed grain boundaries and interphasic dislocations. From this information, it has been claimed that one can determine bulk material properties even more accurately than they can be measured. Using this work in continuum-based computations of materials is obviously very attractive since material properties are often a source of contention when experimental and computational results disagree. Professor Glicksman suggested that one agenda for the future should be better communication and cooperation between this group of ab initio scientists and people in the area of computational materials science; for example, developers of new phase-field methods. Professor Wifried Kurz agreed that obtaining precise material properties is crucial to the experimental validation of new continuum theories and computational methods. It was here, in the area of new work, that he was very concerned. There are many practical problems that have received little attention, but that are critical to understanding and predicting phase transformations in materials science. In heterogeneous system, interfaces occur between the different materials, and phase transformations imply deviations from equilibrium. He then presented several topics for future work that he considered were of fundamental importance to the determination of the structure and properties of materials. Nucleation: Heterogeneous nucleation is at the core of phase transformation in materials. Here, systems develop moving material interfaces with boundary layers at the interfaces, and the whole system evolves to form a material with complex microstructure. To predict this microstructure, arid hence the properties of the bulk material, good models of heterogeneous nucleation and growth of the second phase are needed. This is particularly important for directionally solidified peritectic materials. Cellular and Dendritic Growth: A better understanding of the transition from cells to dendrites during solidification is needed. Computational models need to do a better job of coupling the solutal and thermal fields during cellular and dendritic growth. In addition, the directional solidification of dendrites involves a coupling to the mechanical stress field in the solid phase. These solid stresses may be the origin of the loss of texture in materials because of their coupling to the formation of inhomogeneous solutal and thermal fields. Convection: Future work on convection in material growth should involve the influence of convection on the three-dimensional growth of dendrites. The flow of the liquid phase through arrays of dendrites with numerous side branches
312
M. K.
Smith
causes the fragmentation of dendrites into small crystals that in turn affect the final microstructure of the material. Accurate predictions of microstructure must deal with this highly coupled fluid-flow and material-growth problem. Multiphase Interfaces: The growth of eutectic and peritectic materials involves multiphase interfaces that exhibit morphological instabilities. These instabilities in isothermal and non-isothermal systems produce interfaces with very complex structure. A better understanding of these instabilities and their coupling to solutal and/or thermal fields is needed, especially for peritectic materials. In addition, the growth of composite material in the coupled zone is a problem needing more work in the future. Solid-State Transformations: Solid-state transformations are extremely important in materials science. In many materials, the structure produced during the liquid-solid transformation is altered or replaced by the structure produced during the solid-state transformation. The final properties of the material can only be determined by an understanding of this final structure. There are interesting aspects to cell and dendrite growth in the solid state. In many materials, the product phase has a higher rate of diffusion resulting in lamellar growth, rather than fibrous (or dendritic) growth. Mechanical stresses can build up in the solid state, disturb the local equilibrium, and change the local atomic attachment kinetics. This kind of interaction is extremely important in order to gain an understanding of martinsite transformations and other transformations in which the solid phase is completely supersaturated. This kind of problem affects the everyday life of a materials scientist interested in designing new materials for industrial or commercial use. Professor Miksis now invited the last member of the panel, Prof. Tony Maxworthy, to make a few remarks. Professor Maxworthy was happy to hear comments from the other panelists that experiments will be a vital part of research in the future. In particular, future experimental research needs to be directed towards the development of finer instrumentation, instruments capable of looking into a fluid-flow or material-transformation process and measuring the fine-scale structure. The development and use of such instrumentation would also benefit collaboration between experimentalists and computational physicists because the scale of the data obtained from each technique would be comparable. This kind of collaboration would in turn stimulate the development of even better theories, numerical schemes, and instrumentation. In terms of collaboration, Prof. Maxworthy was more in favor of an interaction between scientists and engineers in one institution, rather than in geographically dispersed locations. The ease of personal communication in these kinds of collaborations is a great asset and it helps to maintain the vitality of the interaction. Professor Sekerka spoke again saying that this was not always the case. Depending on the people involved, and the ease of travel, long distance collaborations are possible and can be tremendously productive. Professors Sekerka and Maxworthy agreed that collaborations are a highly personal endeavor and either mode is possible and can be productive. It just depends on the personal dynamics of the individuals involved.
New Research Directions in Interfacial
Science
313
Speaking from the audience, Prof. Robert Kohn [16] raised a concern over the tendency of the materials science community to think of microstructure in terms of multi-point correlations functions. Has this representation of microstructure been pushed to its limits? Does it really do what is needed? Has it in some way limited research? Perhaps, he suggested, a new mode of thinking is needed. As an example, he cited the computer-vision community, in which complex structure is being represented using wavelets as basis functions. Here, the randomness in a structure is included in the model by using joint-distribution coefficients for the wavelet basis functions. Examining this idea and seeing how it has performed for the computer-vision community may stimulate a new mode of thinking in materials science that could take the representation of microstructure in materials to new levels of success and prediction. In a question that addressed many of the ideas presented at this conference, Prof. Yitzhak Shnidman [17] asked what really happens when a system is said to breakdown at the microscale? Is it the physics or the mathematical model of the physics that is in trouble? How can a model based on physical principles of conservation ever really breakdown? He sees the challenge as devising a way to represent these conservation laws correctly and hierarchically from the microscale, through the mesoscale, and into the macroscale. Wavelet methods are one way to proceed; adaptive gridding in a numerical computation is another. But the real problem is to start with a probabilistic description of these conservation laws on a molecular or atomic scale, to add forces and moments on this level, and then to formulate evolution equations that represent the effect of such microscale phenomena up to the mesoscale, and then onward to the macroscale to produce the standard continuum equations that are used so widely today. Professor Sekerka responded to this question with an example from his experience with a colleague who developed a correction to the continuity equation in a multiphase flow problem. Professor Sekerka thought his colleague was simply wrong at first, but a closer inspection revealed that the nature of the flow and its representation did indeed lead to an additional term in the continuity equation. So we are the victims of our own mindset. Broadening our vision and opening our minds to new ideas used in other areas may indeed lead to progress in otherwise unconnected fields of science and engineering. To support this view, Prof. Barenblatt commented again that asking very sharp questions about the applicability and/or limitations of an idea is a very effective method to expand the frontiers of science in any discipline. The problem is to formulate the correct questions. In response, Prof. Sandra Troian [18] said that the success of our predecessors in formulating such questions and answering them was partly due to the fact that they had the time to invest in this activity. One of her fears and concerns is that the present-day system of funding for academic research does not really allow people to ask such questions. The pressure in academics of publishing and obtaining funding is preventing many people from having the time and energy to pose such questions and to work on them. She hoped that one of the important trends in academics in the next century would be to reverse this trend. Academic researchers are highly trained people and they need the time to use the tools they have acquired to pose
314
M. K.
Smith
and answer these kinds of fundamental questions. As this point, Prof. Davis took the opportunity to have the last word. He commented on the notion of a linking of scales. Physical systems range in scale from larger than light years to smaller than angstroms. The prospect of a single unified theory that can explain all of the phenomena that may occur on any of these scales is very unlikely. Therefore, theories have evolved to explain phenomena at a particular length scale. They do link to theories at a smaller scale, but mostly in the form of averages. The larger-scale theory "forgets," in a sense, most of the details found at the smaller scale. For example, a microscopic theory of a fluid would be able to compute a great deal about the motion of individual atoms or molecules. However, the information transmitted to a larger-scale continuum theory takes the form of averaged quantities like velocity and pressure, and bulk properties like density and viscosity. The interesting thing is that the continuum theory of fluids was formulated and proved to be very successful years before a microscopic theory of fluids was developed. It was successful because it was expressed in terms of averaged quantities that were measurable at the time. This example shows that the linkage of scales between various theories goes both ways. Small-scale theories can compute quantities of interest for a theory at the next level in scale. Simultaneously, largescale theories can determine and describe the truly important information that is needed from a smaller-scale theory. Noting that the hour had drawn to a close, Prof. Miksis thanked the panel for their contributions, thanked the audience for their attention, and closed the discussion session. 3
Summary
Predicting the future is hopeless. This statement began the panel discussion. After reflecting on it for a few moments, it is amusing to note that our job as scientists and engineers is to do the impossible — predict the future. We do not use a crystal ball for this task, but rather the three tools of the scientific method — experimentation, mathematical analysis, and numerical computation. We use these tools to devise and analyze mathematical models of physical systems and use the results to predict the future for those systems. Throughout the discussion, several comments were made that addressed the use of the scientific method at one level or another. To be successful with this method, one must put forward a clear, testable hypothesis for analysis. In other words, simple, clear, and precise questions need to be asked. There are several motivators that are used to formulate these questions. One is just plain curiosity about an observed phenomenon that may, in turn, lead to the formation of new technology. Another is the development of new technology, or societal demands to improve existing technology. Examples here are micro-electromechanical systems, nano-devices, biotechnology, and ecologically clean technology. All of these technologies need careful development before they can provide any benefit to society in the next century. Successful development will involve asking simple and direct questions with clear answers. These answers will assist in the design, production, and use of these new technologies to the physical and economic benefit of us all.
New Research Directions in Interfacial Science
315
To help out in this task, the panel offered some clear directions for interfacial research in fluid mechanics and materials science in the next century. These are summarized below. Collaboration: The new technologies of the future will be so complex and interdisciplinary that no individual could possibly ask all of the right questions. Thus, collaborations involving individuals from several disciplines will be needed. There are several models for collaborations of this kind and the choice belongs to the individual; a single center of excellence, a multi-university center, a loose collection of several individuals in one or several places, etc. Collaborations between people from very different disciplines is difficult, but it will be essential for progress in existing and new fields of science and technology. Instrumentation: Today, we are dealing with smaller and more complex systems than ever before. Micron-sized devices are common and nanometer-sized devices are being developed. New materials are very complex. Systems or processes under development use non-Newtonian fluids, closely packed suspensions, or substances with complex microstructures resulting from the solidification of mixtures of several materials. Observations in these systems will demand new and more sophisticated instrumentation. Measurements will need to resolve smaller length scales, faster time scales, and provide much more detailed information. The development of such instruments will be essential for future progress. The resulting measurements will help guide the development of new models and then verify and support theoretical and numerical predictions. Material Properties: The correct prediction of a physical system's behavior by a model depends on the values of the material properties that are used. This is indeed a critical aspect of comparing and validating a model with its related experiments. Thus, accurate property measurements or direct numerical computations of properties based on numerical models will be essential to the development and use of systems containing complex materials. Modeling: There will probably never be a universal model of a physical system that is valid at all scales. Even if one existed, it would likely be too unwieldy to use effectively. Thus, the modeling of a complex physical system must use the idea of a linkage of scales to good effect. We must formulate and use large-scale theories to determine what is needed from smaller-scale theories. One can use the properties needed in a macroscale model and the mathematical breakdown of the theory as a guide to what to look for and where to look in a microscale model. We also need to learn how to effectively link microscale models with macroscale models. Mathematical Methods: Another critical aspect of modeling is the question of deciding the most useful mathematical framework with which to pose and solve the model equations. People tend to get wedded to a set of techniques that they are comfortable with and fail to see when a new path is needed. We need to be aware of alternate methods of analysis and be willing to borrow
316
M. K.
Smith
successful mathematical methods from other disciplines that may prove useful in explaining complex structures in fluids and materials science. Molecular Simulations: Molecular models are becoming increasingly important in fluid mechanics (e.g., for the contact-line problem and the interfacial dynamics of complex fluids) and in materials science (e.g., phase transformations in a wide range of material systems leading to complex microstructure). Critical needs for the future will involve the development of techniques to link the results from such simulations to the material properties and other constitutive relations used in continuum models. In particular, where does a particular molecular phenomenon govern or modify the behavior of a macroscale system? How does one construct an effective boundary condition or constitutive relation for use in the related continuum theory? What ways can we devise to simultaneously link numerical simulations done at the molecular scale to those done at the continuum scale? Reduced-Order Models: Future work will be needed on the techniques of building simplified reduced-order engineering models suitable for the design of complex systems. Formulating such models and determining what kind of information is really important will involve asking clear questions about what is truly needed at this larger scale. We will be able to compute giga-bytes or tera-bytes of data related to all of the field variables in a continuum or macroscale model. How do we visualize these data effectively? How do we average these data to produce the quantities of interest for a reduced-order model? How do we integrate these reduced-order models with one another to produce a system model that can be used to design a complex process or device? Complex Interfaces: Interfacial science in the next century will involve the investigation of thin boundaries with complex structure between two different phases of possibly complex materials. Examples of such boundaries include cell and tissue membranes in biotechnology applications, two-fluid interfaces loaded with surfactants for chemical and biological processing, and material interfaces associated with materials-processing technology in which the solidifying liquids themselves have complex structure and the solidified materials undergo complex phase transformations in either the liquid or solid states. The development of appropriate models of these interfaces that include the essential physics will be a challenge. Work will be needed to develop theoretical and numerical tools to analyze these models and extract the relevant data necessary to make accurate predictions. Finally, experimental tools need to be developed to observe these interfaces on the small scale. These data will be used to guide the development of proper models and to validate their predictions so that these tools can be used for the development and design of new technologies and materials in the future. The path into the future for interfacial science in fluid mechanics and materials science, and indeed in any field of science, depends on how we formulate good questions and use our time and resources to answer them. There may be no way to
New Research Directions in Interfacial
Science
317
predict that future or to know what we will find, but we do know how to lay down the bricks that pave this research path into the Twenty-First century. References 1. Michael J. Miksis, Department of Engineering Sciences and Applied Mathematics, Northwestern University, Evanston, IL 60088, U.S.A. 2. Stephen H. Davis, Department of Engineering Sciences and Applied Mathematics, Northwestern University, Evanston, IL 60088, U.S.A. 3. Joel Koplik, Benjamin Levich Institute and Department of Physics, City College of the City University of New York, New York, NY 10031, U.S.A., E-mail: [email protected]. cuny. edu. 4. Wilfried Kurz, Physical Metallurgy Laboratory, Swiss Federal Institute of Technology, 1015 Lausanne EPFL, Switzerland. 5. Tony Maxworthy, Department of Aerospace and Mechanical Engineering, University of Southern California, Los Angeles, CA 90089-1191, U.S.A. 6. Robert F. Sekerka, Department of Physics, Carnegie Mellon University, Pittsburgh, PA 15213-3890, U.S.A. 7. Gretar Tryggvason, Mechanical Engineering Department, Worcester Polytechnic Institute, Worcester, MA 01609-2280, U.S.A., E-mail: [email protected]. 8. "Interfaces for the Twenty-First Century," NSF Grant No. DMS-9974551. 9. D. Gallez and E. Ramos de Souza, in Interfaces for the Twenty-First Century, eds. M. K. Smith, M. J. Miksis, G. B. McFadden, G. P. Neitzel, and D. R. Canright, Imperial College Press, p. 73 (2002). 10. M. G. Worster, in Interfaces for the Twenty-First Century, eds. M. K. Smith, M. J. Miksis, G. B. McFadden, G. P. Neitzel, and D. R. Canright, Imperial College Press, p. 187 (2002). 11. Timothy J. Pedley, Department of Applied Mathematics and Theoretical Physics, Silver Street, University of Cambridge, Cambridge, CB3 9EW, U.K. 12. Grigory Barenblatt, Department of Mathematics, University of California, Berkeley, Berkeley, CA 94720-3840, U.S.A. 13. Stephen J. S. Morris, Department of Mechanical Engineering, University of California, Berkeley, Berkeley, CA 94720-1740, U.S.A. 14. N. Hadjiconstantinou, in Interfaces for the Twenty-First Century, eds. M. K. Smith, M. J. Miksis, G. B. McFadden, G. P. Neitzel, and D. R. Canright, Imperial College Press, p. 260 (2002). 15. Martin E. Glicksman, Materials Science and Engineering Department, Rensselaer Polytechnic Institute, Troy, NY 12180-3590, U.S.A., E-mail: [email protected]. 16. Robert V. Kohn, Courant Institute of Mathematical Sciences, New York University, New York, NY 10012, U.S.A. 17. Yitzhak Shnidman, Department of Chemical Engineering and Chemistry, Polytechnic University, 6 MetroTech Center, Brooklyn, NY 11201, U.S.A. 18. Sandra M. Troian, Department of Chemical Engineering, Princeton University, Princeton, NJ 08544-5263, U.S.A.
Interfaces
for
t h e 21 s t Century New Research Directions in Fluid Mechanics
and Materials Science
''"'1's book highlights some recent advances in interfacial research in the fields of fluid mechanics and materials science at the beginning of the twenty-first century. It is an .extension of the presentations made during the conference "Interfaces for the I \\ eniy-First Century," held on August 16-18, 1999, in Monterey, California. It includes papers by sixteen renowned experts in the field of interfacial mechanics, abstracts contributed by research scientists, and a summary of a panel discussion on future research directions. The book covers experimental and theoretical approaches, With the unifying philosophy being the investigation of new techniques for modeling the dynamics of interfaces. A number of new and ex( iting solution methods and experimental studies. as well a.s the physical problems that initiated them, are presented.
'
P259 he ISBN 1-86094-319-5
Imperial College Press www.icpress.co.uk