UNIVERSITY OF CINCINNATI January 16 04 _____________ , 20 _____ JUNTAO ZHANG I,______________________________________________, hereby submit this as part of the requirements for the degree of: DOCTOR OF PHILOSOPHY (Ph.D.) ________________________________________________ in: Department of Mechanical, Industrial and Nuclear Engineering ________________________________________________ It is entitled: EXPERIMENTAL AND COMPUTATIONAL STUDY OF ________________________________________________ NUCLEATE POOL BOILING HEAT TRANSFER IN AQUEOUS ________________________________________________ SURFACTANT AND POLYMER SOLUTIONS ________________________________________________ ________________________________________________ Approved by: Raj M. Manglik ________________________ Milind A. Jog ________________________ Michael Kazmierczak ________________________ Rupak K. Banerjee ________________________ ________________________ EXPERIMENTAL AND COMPUTATIONAL STUDY OF NUCLEATE POOL BOILING HEAT TRANSFER IN AQUEOUS SURFACTANT AND POLYMER SOLUTIONS A Dissertation Submitted to the Division of Research and Advanced Studies of the University of Cincinnati in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY (Ph.D.) in the Department of Mechanical, Industrial and Nuclear Engineering of the College of Engineering 2004 by Juntao Zhang B.S., University of Science and Technology Beijing, China, 1992 M.S., University of Science and Technology Beijing, China, 1995 Committee Chair: Professor Raj M. Manglik ABSTRACT Saturated, nucleate pool boiling on a horizontal, cylindrical heater and the associated bubble dynamics in aqueous surfactant and polymer solutions are experimentally and computationally investigated. Boiling curves ( qw" vs. ∆Tsat data) for different additive concentrations and photographic records of the salient features of the ebullient behavior are presented, along with a characterization of the interfacial properties of the aqueous solutions. The surfactant additive significantly alters the nucleate boiling in water and enhances the heat transfer. The enhancement increases with concentration, with an optimum obtained in solutions at or near the critical micelle concentration (CMC) of the surfactant. On the other hand, boiling in aqueous polymer solutions shows contrary results. The enhancement is seen to increase with polymer concentration in the aqueous surface-active HEC solutions, while boiling in shear-thinning Carbopol 934 solutions is seen to decrease continuously with increasing concentrations. In order to develop a theoretical model and understand the associated convective mechanisms, the dynamics of a single growing and departing bubble during nucleate boiling from a horizontal heated surface has been numerically simulated. The results highlight the role of the microlayer in nucleate boiling, as well as the effects of altered surface tension and viscosity, apparent contact angle, and wall superheat on the bubble dynamics and boiling behavior in aqueous surfactant and polymer solutions. In multiphase heat transfer, the nature and dynamics of surface contact often plays a dominant role. In fact, almost all the transport processes in nucleate boiling, from the inception of an embryonic bubble to the subsequent phase change, are intricately connected with this interface. To better characterize the molecular dynamics of the additives at the interfaces, an extensive literature review is presented that delineates the surfactant adsorption process and its electrokinetic effects (zeta potential), and the concomitant surface wetting behavior. This fundamental understanding is supported by the measured interfacial properties of the solutions reported in this study. And as shown by the boiling data, the heat transfer performance is changed dramatically when the interfacial transport process is altered. Finally, as documented by the pool boiling experiments with aqueous surfactant solutions, a markedly different ebullient behavior than not only that of water is observed, but between pre- and post-CMC solutions as well. The characteristic bubble dynamics is found to correlate well with the measured surface wettability and the adsorption isotherm, which in turn correlates with the zeta potential. In essence, the unique structure of a surfactant - a hydrophilic head with a hydrophobic tail - provides the means to control of nucleate boiling by altering surface wettability. The advantage of this very effective method is that there is no need to make any changes to any existing boiling equipment, except for the addition of desired amount of surfactant into the solution. The potential application impact extends to any process that involves phase-change (large-scale as well as micro-scale heat exchange devices). Furthermore, the surfactant physisorption and its electrokinetics are not only basic to a fundamental understanding of nucleate boiling control, but also can provide insights into related natural phenomena in other applications as well. There include chemical and biological sensors, microgravity applications, and microfluidic (or lab-on-chip) devices, among others. Also, some nanofluids or nanoparticles with special surface properties based on the different structures of surfactant micelles can be developed to provide even greater interfacial control in these broad spectrum of emerging applications. ACKNOWLEDGEMENTS First, I would like to thank Professor Raj M. Manglik for his excellent guidance, support, encouragement, and inspiration that make this dissertation possible. I am also thankful to the committee members Drs. Milind A. Jog, Michael Kazmierczak and Rupak K. Banerjee for their valuable suggestions, and Mr. Bo Westheider of the Instrumentation Laboratory, MINE department, and Mr. Doug Hurd of the machine shop for their help in the construction of the experimental facility. Also, the valuable communications with Professor Vijay K. Dhir, UCLA, on the microlayer and phase-change modeling, the generous help of Dr. Osamu Tatebe, National Institute of Advanced Industrial Science and Technology, Japan, on the MPCG implementation, and instructive suggestions of Professor Douglas W. Fuerstenau, University of California – Berkeley, on the surfactant adsorption characteristics, are gratefully acknowledged. I specially thank the love and support of my wife, Jia Li, which made my work possible and worthwhile. Also, I would like to thank my parents for their constant encouragement and support through my studies. The financial support provided by National Science Foundation, the University Research Council, and Ohio Board of Reagents is gratefully acknowledged. My fellow graduate students: Manish Bahl, Satish Vishnubhatla, and S. Sethu Raghavan provided much needed assistance in acquiring some rheology and interfacial data. Finally, I would like to thank all the members in the TFTPL laboratory and all my friends in UC for enriching my everyday life. TABLE OF CONTENTS Page ABSTRACT ACKNOWLEDGEMENTS TABLE OF CONTENTS i LIST OF TABLES v LIST OF FIGURES vi NOMENCLATURE xi 1. INTRODUCTION 1 1.1 Surfactants, Colloid Systems, and Interfacial Phenomena 2 1.1.1 Surfactants 5 1.1.2 Colloid systems and interfacial phenomena 5 1.1.3 Surface tension and micelles 7 1.1.4 Electrokinetic effects and zeta potential 10 1.1.5 Surfactant physisorption at the solid-liquid interface 14 1.2 Surface Wettability and Nucleate Boiling 18 1.2.1 Contact angle and surface wettability 18 1.2.2 Surface wetting effects on nucleate boiling 21 1.3 Nucleate Pool Boiling with Surfactants 22 1.4 Nucleate Pool Boiling with Polymers 26 1.5 Computational Fluid Dynamics with Moving Boundaries 30 1.6 Scope of Study 32 i 2 INTERFACIAL PROPERTIES AND RHEOLOGY MEASUREMENTS 2.1 Surface Tension Measurements 34 34 2.1.1 Introduction 34 2.1.2 Viscosity effects on surface tension measurements 38 2.1.3 Results and discussions 39 2.1.3.1 Aqueous surfactant solutions 39 2.1.3.2 Aqueous polymeric solutions 46 2.2 Contact Angle Measurements 52 2.3 Rheology Measurements 55 2.3.1 Aqueous surfactant solutions 56 2.3.2 Aqueous polymer solutions 56 3. POOL BOILING HEAT TRANSFER 61 3.1 Experimental Setup 61 3.2 Nucleate Pool Boiling in Aqueous Surfactant Solutions 65 3.2.1 Pool boiling in aqueous cationic surfactant solutions 65 3.2.2 Optimum heat transfer and critical micelle concentration 71 3.3 Nucleate Pool Boiling in Aqueous Polymer Solutions 72 3.3.1 Pool boiling in aqueous polymer solutions 74 3.3.2 Surface-active and rheological effects 77 4. VISUALIZATION AND CHARACTERIZATION OF NUCLEATE POOL BOILING IN AQUEOUS SUFACTANT SOLUTIONS 81 4.1 Introduction 81 4.2 Zeta Potential and Contact Angle 82 ii 4.3 Ethoxylation Effect 85 4.4 Dynamic Surface Tension – Molecular Weight Effect 86 4.5 Surface Wettability Effect on Nucleate Boiling Heat Transfer 92 4.6 Ebullient Dynamics Visualization 94 4.6.1 Visualization in aqueous surfactant solutions 94 4.6.2 Visualization in aqueous polymer solutions 99 4.7 Characterization of Nucleate Pool Boiling in Aqueous Surfactant Solutions 102 5. SIMULATION OF A SINGLE BUBBLE 112 5.1 Introduction 112 5.2 Mathematical Formulation 114 5.3 The Numerical Method 120 5.4 Solution Validation 122 5.4.1 Efficacy of the MPCG method 122 5.4.2 Efficacy of the level-set method 125 5.4.3 Verification of phase change modeling 127 5.5 Results and Discussion 130 5.5.1 Microlayer 130 5.5.2 Surface tension effect 131 5.5.3 Viscosity effect 138 5.5.4 Temperature, velocity, and pressure fields 141 5.5.5 Apparent contact angle and superheat effect 144 5.6 Significance and Limitations of Nucleate Boiling Simulations iii 146 6. CONCLUSIONS AND RECOMMENDATIONS 149 6.1 Conclusions 149 6.2 Recommendations for Future Research 153 BIBLIOGRAPHY 161 APPENDIX A. SURFACE TENSION σ (mN/m) DATA 177 A.1 Aqueous Surfactant Solutions 177 A.2 Aqueous Polymer Solutions 181 A.3 Dynamic Surface Tension with Time 182 A.4 Surface Tension with Temperature 186 APPENDIX B. CONTACT ANGLE DATA 187 APPENDIX C. POOL BOILING DATA 188 C.1 Aqueous Surfactant Solutions 188 C.2 Aqueous Polymer Solutions 197 APPENDIX D. TEMPORAL AND SPATIAL DISCRETIZATION iv 200 LIST OF TABLES Page 1.1 Typical Colloidal Systems (Birdi, 2003) 1.2 Chronological Listing of Nucleate Pool Boiling Studies of Aqueous Surfactant Solutions 23 Chronological Listing of Nucleate Pool Boiling Studies of Aqueous Polymeric Solutions 27 2.1 Physico-Chemical Properties of Surfactants 37 2.2 Physico-Chemical Properties of Polymers 38 2.3 Increase in Viscosity of Aqueous Polymer Solutions with respect to Water at 23°C as a Function of Concentration at a Shear Rate of 500 s-1 60 1.3 5.1 Combination of Non-Dimensional Parameters Studied (D = 2mm) v 6 138 LIST OF FIGURES Page 1.1 1.2 Typical boiling curve for controlled wall heat flux and schematic representation of the different heat transfer regimes 2 The conjugate problem in modeling nucleate pool boiling with or without additives 4 1.3 Schematic illustration of the primary structure of a surfactant molecule 5 1.4 Different possible micellar structures (Evans and Wennerström, 1999) 9 1.5 Schematic diagram of a typical electrokinetic boundary layer 10 1.6 Four basic electrokinetic phenomena and the relationship between them 13 1.7 Schematic representation of a typical adsorption isotherm and aggregate states of a surfactant on a solid surface in aqueous solutions 15 1.8 Different wetting conditions of a liquid drop on a solid surface 19 1.9 Surface forces involved in spreading of a liquid 20 2.1 Schematic of surface tensiometer and data acquisition system 36 2.2 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous surfactant solutions at 23°C 40 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous surfactant solutions at 80°C 42 Dynamic surface tension relaxation for aqueous solutions of CTAB and Ethoquad 18/28 44 2.5 Equilibrium surface tension measurements as a function of temperature 45 2.6 Dynamic surface tension measurements for aqueous HEC solutions at 23°C 47 2.7 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous polymer solutions at 23°C 50 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous polymer solutions at 80°C 51 2.3 2.4 2.8 vi 2.9 (a) Measured contact angle for aqueous CTAB, Ethoquad 18/25, SDS, Triton X-100, and Triton X-305 solutions; (b) corresponding ionic surfactant adsorption surface state; and (c) EO group effect on surface wettability 53 (a) Contact angle and adsorption isotherms for nonionic surfactants Triton X-100 and Triton X-305 in aqueous solutions; and (b) non-ionic surfactant adsorption. 54 2.11 Relative viscosity changes of aqueous CTAB and Ethoquad 18/25 solutions 58 2.12 Variation of apparent viscosity with shear rate for aqueous HEC QP-300 solutions 59 Variation of apparent viscosity with shear rate for aqueous Carbopol 934 solutions 60 Schematic of experimental facility: (a) pool boiling apparatus, and (b) cross-sectional view of cylindrical heater assembly 63 3.2 Optical microscope images of the roughness characteristics of heater surface 64 3.3 Nucleate pool boiling data for aqueous solutions of DTAC; all data are for decreasing heat flux except as otherwise indicated 67 Nucleate pool boiling data for aqueous solutions of CTAB; all data are for decreasing heat flux except as otherwise indicated 68 Nucleate pool boiling data for aqueous solutions of Ethoquad O/12 PG; all data are for decreasing heat flux except as otherwise indicated 69 Nucleate pool boiling data for aqueous solutions of Ethoquad 18/25; all data are for decreasing heat flux except as otherwise indicated 70 Variation of the relative heat transfer performance of aqueous cationic surfactant solutions with heat flux and additive concentration (decreasing qw" ) 73 3.8 Nucleate pool boiling data for aqueous solutions of HEC-QP300 75 3.9 Nucleate pool boiling data for aqueous solutions of Carbopol 934 76 3.10 Variation of the enhanced boiling heat transfer performance of HEC-QP300 solutions with heat flux and additive concentration 78 Effect of dynamic surface tension on the boiling heat transfer coefficient 80 2.10 2.13 3.1 3.4 3.5 3.6 3.7 3.11 vii 4.1 Measured streaming zeta potential and contact angle for the adsorption of SDS and CTAB in their aqueous solutions 84 4.2 Measured contact angle for aqueous CTAB and Ethoquad 8/25 solutions 85 4.3 Dynamic surface tension relaxation for aqueous anionic SDS and SLES solutions 87 Dynamic surface tension relaxation for aqueous nonionic Triton X-100 and Triton X-305 solutions 88 Effect of surfactant molecular weight and its ethoxylation on the heat transfer coefficient enhancement 90 Surfactant molecular weight dependence of the maximum enhancement in heat transfer coefficient enhancement (Wasekar and Manglik, 2001) 91 (a) Effect surface wettability (or contact angle θ) on nucleate pool boiling (Liaw and Dhir, 1989) (b) Nucleate pool boiling data for Refrigerant of R113 on a copper tube (Jung and Bergles, 1989) 93 Ebullient behavior in nucleate boiling of distilled water, and aqueous cationic CTAB and Ethoquad 18/25 solutions of different concentrations (C/CCMC = 0.5, 1, and 2) at qw" = 20 kW/m2 and 50 kW/m2 97 Ebullient behavior in nucleate boiling of distilled water, and aqueous SDS (anionic) and Triton X-305 (nonionic) solutions of different concentrations (C/CCMC = 0.5, 1, and 2) at qw" = 20 kW/m2 and 50 kW/m2 98 4.4 4.5 4.6 4.7 4.8 4.9 4.10 Ebullient behavior in nucleate boiling of distilled water, and aqueous HEC-QP300 and Carbopol 934 solutions of different concentrations at different heat fluxes ( qw" = 20 kW/m2, 50 kW/m2, and 100 KW/m2) 101 Schematic of interfacial phenomena in aqueous surfactant solutions (not to scale) 104 4.12 Surfactant effects on nucleate boiling heat transfer in its aqueous solutions 106 4.13 (a) Schematic of surfactant transport process during a bubble formation and departure (not to scale); (b) Dynamic surface tension effect on bubble dynamics (evolution of pre-departure shape and size). 108 Physical domain of a boiling bubble decomposed into a macro region and a microlayer 116 4.11 5.1 viii 5.2 MAC-staggered grid to show where the variables u, p, T, and φ are located 117 5.3 (a) Poisson problem with jump diffusion coefficients (T-shape); and (b) test computational results for u(x,y) for the MPCG method (f = 100) 123 (a) Poisson problem with jump diffusion coefficients (arc); and (b) test computational results for u(x,y) for the MPCG method (f = 100) 124 Contours of level-set function φ(x) with the solid line circle representing φ(x) = 0 125 Rising bubble interfaces plotted at different times for U = 0, V = 1 using the mass-preserved level-set method 126 5.7 Schematic of a growing bubble in an extensive superheated liquid pool 127 5.8 Bubble growth in an extensive superheated liquid pool: (a) bubble interfaces plotted at different times; (b) bubble growth with time 129 5.9 Microlayer shape and vapor-liquid interface temperature distribution for a nucleated bubble 131 Temporal evolution of bubble shapes for various Weber numbers (Re = 278) 133 5.11 Change in bubble rising velocity with time for various Weber numbers 133 5.12 Bubble growth and its departure in nucleate pool boiling for ∆T = 10K, ϕ = 45°. (a) σ = 58.86 mN/m (water), (b) σ = 47.0 mN/m (SDS, C = 1000 wppm), 135 (c) σ = 37.5 mN/m (SDS, C = 2500 wppm ) 5.13 Bubble departure diameter vs. surface tension (∆T = 10°C, ϕ = 45°) 137 5.14 Temporal evolution of bubble shapes for various Morton numbers (We = 0.538) 139 5.15 Change in bubble rising velocity with time for various Morton numbers 139 5.16 Comparison of predicted result with the experiment for water: (a) present simulation; (b) experimental result of Bhaga and Weber (1981) 140 Temperature isotherms during different bubble growth stages 142 5.4 5.5 5.6 5.10 5.17 ix 5.18 (a) Velocity field; (b) filled pressure contours; and (c) line pressure contours, in and around a detached isolated bubble for ∆T = 10K, ϕ = 45°, and σ = 37.5 mN/m, t = 0.06s 143 5.19 (a) Bubble shape at departure for different contact angles; and (b) bubble departure diameter vs. apparent contact angle (∆T = 10°C, σ = 58.86 mN/m) 145 5.20 Effect of wall superheat ∆T on bubble growth (ϕ = 45°, σ = 58.86 mN/m): (a) bubble shape at departure; (b) bubble departure time; and (c) bubble departure diameter vs. wall superheat 147 5.21 Adsorption-desorption controlled surfactant interfacial transport process 148 6.1 Schematic representation and AFM detection images of adsorbed layer structures consisting of (A) spherical micelles, (B) cylindrical micelles, and (C) a bilayer (Schulz et al, 2001) 155 Conceptualization of possible surfactant adsorbate layers at the solid-liquid interface 156 Proposed approach to correlation of nucleate boiling in aqueous surfactant solutions 160 Proposed investigations to correlation of nucleate boiling in aqueous surfactant solutions 160 6.2 6.3 6.4 x NOMENCLATURE A Hamaker constant [J] heater surface area (= 2πroL) [m2] C concentration [dimensionless] Ca Capillary number [dimensionless] D bubble diameter [m] f source term bubble frequency [1/s] Fr Froude number [dimensionless] g gravity vector [m/s2] Gr Grashof number [dimensionless] H Heaviside function h grid spacing for macro region [m] h boiling heat transfer coefficient [kW/m2 K] hev evaporative heat transfer coefficient [W/m2 K] hfg latent heat of evaporation [kJ/kg] I unit vector I current [A] Ja Jacob number [dimensionless] k thermal conductivity [W/m2] interface curvature in Eq. (5.7) diffusion coefficient in Eq. (5.24) xi L length of heated cylinder [m] Γ surface concentration [mol/m2] L characteristic length [m] M molecular weight [kg/kmol] Mo Morton number [dimensionless] Na active nucleation site density [m-2] P pressure [Pa] Pe Peclet number [dimensionless] q heat flux [W/m2] q w′′ wall heat flux [W/m2, or kW/m2] r radial coordinate R radius of wall thermocouple location [m] ro cylindrical heater radius [m] R universal gas constant [J/mol K] R radial location of the interface at y=h/2 Re Reynolds number [dimensionless] S sign function T temperature [K] ∆Tw wall superheat [K] t time [s] u characteristic velocity vector [m/s] U x-direction velocity [m/s] V y-direction velocity [m/s] xii Vmicro rate of vapor volume production from the microlayer [m3/s] ∆Vmicro control volume near the micro region [m3] V voltage [V] We Weber number [dimensionless] x horizontal coordinate Y height of computational domain [m] y vertical coordinate Greek α thermal diffusivity [m2/s] β Coefficient of thermal expansion [1/K] D diffusion term in Eq. (5.11) δ liquid film thickness [m] φ level set function [m] γ shear rate [s-1] η apparent viscosity [Pa⋅s] ϕ apparent contact angle [deg] σ surface tension [mN/m] θ dimensionless temperature static contact angle [deg] ρ density [kg/m3] µ dynamic viscosity [N s/m2] xiii ν kinematic viscosity [m2/s] τ surface age [ms] Ω computational domain in Eq. (5.24) Subscripts ′ dimensionless 0 Initial cap Capillary int Interface l,v liquid, vapor poly pertaining to aqueous polymer solution o outer surface r radial location sat Saturation surf pertaining to aqueous surfactant solution s property at the surface w Wall water pertaining to pure water ∞ bulk, far field condition 50 ms pertaining to surface age of 50 ms xiv CHAPTER 1 INTRODUCTION Nucleate boiling is an important and efficient thermal management process with a broad spectrum of applications, because relatively small temperature differences can sustain very high heat transfer rates. Extensive research on numerous facets of boiling heat transfer has been reported in the literature (Carey, 1992; Collier and Thome, 1996; Dhir, 1998; Kandlikar et al. 1999). Pool boiling essentially occurs at a heated surface in quiescent liquid, where its motion near the surface is driven by natural convection and ebullient (bubble inception, growth, and departure) conditions; whereas in forced flow boiling, liquid flow over the heater surface is imposed by external applied pressure gradients. Figure 1.1 shows a typical pool-boiling curve, a plot of wall heat flux qw" vs. the wall superheat ∆T (= Tw – Tsat), for the entire set of regimes that are encountered in the heat transfer process. The complete boiling curve is characterized by the following transport characteristics. As the heat-input rate to the surface is increased, the first heat transfer mode to appear is natural or free convection. Subsequently, at a certain value of the superheat (Point A), vapor bubbles appear on the heater surface, and this is referred to as the onset of boiling (ONB). In liquids that wet the surface well, the onset of nucleation may be delayed. For these liquids, a sudden activation of large number of cavities at an increased superheat causes a reduction in the surface temperature while the heat flux remains constant, and this feature is often referred to as “Temperature overshoot”. This behavior is not observed when the boiling curve is obtained by reducing the heat flux, 1 Log qw″ C E B D A Hysteresis Log ∆T Fig. 1.1 Typical boiling curve for controlled wall heat flux and schematic representation of the different heat transfer regimes and, thus, a temperature hysteresis results. After inception, a dramatic increase in the slope of the boiling curve is observed. In partial nucleate boiling (Region II, A-B), discrete bubbles are released from randomly located active nucleation sites on the heater surface. The transition from isolated bubbles to fully developed nucleate boiling (Region III, B-C) occurs when bubbles at a given site begin to merge in the vertical direction in the form of jets. The maximum or critical heat flux (CHF) sets the upper limit of the 2 fully developed nucleate boiling. After CHF, most of the surface is very rapidly covered with vapor; the surface temperature rises quickly. When the heat input rate is controlled, the heater surface will rapidly pass through Regions IV (C-D) and V (D-E), and stabilize at Point E in the film boiling regime. If the temperature at E exceeds the melting point of the heater material, the heater will be burnt out. Further details about nature of boiling and boiling behaviors in different regimes can be found in Dhir (1998) and Carey (1992), among others. In recent years, enhancement of nucleate boiling heat transfer has received much attention, and different active and passive techniques have been documented in several reviews (Thome, 1990; Bergles, 1997; Manglik, 2003). The use of liquid additives in particular, which include surfactants or surface-active substances that significantly alter the surface tension of the boiling liquid even at very low concentrations, has been the focus of some current research. The reviews by Wasekar and Manglik (1999, 2001) provide an extended discussion of several issues associated with enhanced boiling heat transfer in surfactant and polymeric solutions. Several studies have investigated enhanced pool boiling in aqueous surfactant and polymeric solutions under atmospheric conditions, and a variety of different predictive parameters and mechanisms have been proposed to describe the complex phase-change process. The primary determinants of the general boiling problem, however, can essentially be classified under three broad categories: heater, fluid, and heater-fluid interface (Nelson et al., 1996; Nelson, 2001). For nucleate boiling with additives in aqueous solutions, the associated potential mechanisms that may be involved are depicted as a conjugate problem in Fig. 1.2. 3 HEATER (smooth or structured) Physical properties Geometry Heat flux (wall superheat) Shape: Plate, Orientation cylinder, Thickness, wire, etc. diameter Surface characteristics (fractal dimension) Cavity density Roughness Constant wall temperature or Constant heat flux HEATER-FLUID INTERFACE (with or without additives) Contact angle Additive physi(wettability) sorption characteristics (zeta potential) Active site density Interaction or potential interactions FLUID (with or without additives) Physics properties Thermal conductivity Additive Viscosity surface tension σ Near-surface features Far-surface features Meniscus Vapor stems Vapor columns Liquid in microlayer Discrete Vapor and macrolayer bubbles mushrooms Liquid-vapor interface Foaming Physico-chemical properties Marangoni convection (adsorption/desorption) Dynamic σ Ionic nature Ethoxylation Molecular weight Apparent viscosity Fig. 1.2 The conjugate problem in modeling nucleate pool boiling with or without additives 4 1.1 Surfactants and Interfacial Phenomena 1.1.1 Surfactants Surfactant is a generic term for a surface-active agent, which literally means active at a surface. It is fundamentally characterized by its tendency to adsorb at surfaces and interfaces when added in low concentrations to an aqueous system. Surfactants have a unique long-chain molecular structure that is composed of a hydrophilic head and a hydrophobic tail as illustrated in Fig. 1.3. They have a natural tendency to adsorb at the liquid-vapor interface with their polar head oriented towards the aqueous solution and the hydrocarbon tail directed towards the vapor. Based on the nature of the hydrophilic part of the molecule, which is ionizable, polar, and polarizable, surfactants are generally categorized as anionics, nonionics, cationics, and zwitterionics (Holmberg et al., 2003). Hydrophilic (polar) head Hydrophobic (non polar) tail Fig. 1.3 Schematic illustration of the primary structure of a surfactant molecule 1.1.2 Colloid systems and interfacial phenomena There are three states of matter – gas, liquid, and solid, and when one of these states is finely dispersed in another, then a colloidal system is obtained. This can be in the form of aerosols, emulsions, inverse emulsions, sols or colloidal suspensions, gels, biocolloids, or association colloids, based on the different dispersion state (Hunter, 2001). 5 Colloidal systems are of great practical importance, and Table 1.1 lists various types of colloidal systems and their common examples. Table 1.1 Typical Colloidal Systems (Birdi, 2003) Dispersed Phases Continuous System type Liquid Gas Aerosol fog, spray Gas Liquid Foam, fire extinguisher foam Liquid Liquid Emulsion (milk) Solid Liquid Sols, suspension waste water, cement Corpuscles Serum Blood Hydroxyapatite Collagen Bone Liquid Solid Solid emulsion (toothpaste) Solid Gas Solid aerosol (dust) Gas Solid Solid foam – insulating foam Solid Solid Solid suspension/solids in plastics Biocolloids Colloidal systems often exhibit rather unusual phenomena at their phase boundaries (interfaces), relative to the expected bulk phase interactions, such that the behavior of the entire system is controlled by interfacial processes (Rosen, 1989; Evans and Wennerström, 1999; Hunter, 2001; Birdi, 2003; Holmberg, 2003; Chen, 2003). Interfacial phenomena are important in almost every industrial process, from heterogeneous catalysis to the manufacturing of composite materials, from medical technology to detergency, as well as in numerous processes such as thermal treating, coating, nanofluids, dispersion, flotation, solubilization of chemicals, oil exploitation, crystallization, fabrication of compound systems (reinforced materials and coated 6 materials), and boiling. With nucleate phase-change and ebullience in aqueous surfactant solutions, which are association colloid systems, where molecules of surface-active substances (e.g. surfactants) are associated together to form small aggregates (micelles) in water, the aggregates formed may often adopt an ordered structure. The consequent interfacial changes significantly affect boiling. Thus, it is critical to understand the causes and results of interfacial behavior and variables that affect it in order to predict and control the properties of not only boiling in surfactant solutions but all colloidal systems in general. Altogether five different interfaces can exist: gas-liquid, gas-solid, solid-liquid, liquid-liquid, and solid-solid. In the case of surfactant solutions, the additive may adsorb at all of the five types of interfaces. For nucleate pool boiling in aqueous surfactant solutions, however, there are two primary interfaces that have a dominating influence: (1) vapor-liquid interface, at which the surface tension reduces because of the surfactant adsorption-desorption process, and (2) solid-liquid (or heater-liquid) interface, where the surfactant physisorption occurs and the surface wetting behavior changes. 1.1.3 Surface tension and micelles Surfactant additives in aqueous solutions naturally tend to diffuse towards the vapor-liquid interface and subsequently get adsorbed on it. Depending upon their chemistry (ionic and molecular structure) and orientation at the interface, some desorption may also occur. The primary effect of the surfactant adsorption-desorption process at the vapor-liquid interface is to reduce the surface tension of the solution. This entire process is time-dependent and it manifests in a dynamic surface tension behavior at 7 an evolving vapor-liquid interface (as in ebullience), which eventually reduces to an equilibrium value after a long time span. Surface tension reduction of an aqueous solution decreases continually with increasing concentrations till the critical micelle concentration (CMC) is reached, at which point the surfactant molecules cluster together to form micelles. All surfactants in their solutions show significant changes in adsorption behavior at or around their respective CMC. The CMC is characterized by micelle formation, or micellization, which is the property of surface-active solutes that leads to the formation of colloid-sized clusters, i.e., at a particular concentration, additives form aggregates in the bulk phase or a surfactant cluster in solution that are termed micelles (Edwards et al., 1991). Different shapes and sizes of micelles exist depending upon the surfactant type and its packing, concentration, solution temperature, presence of other ions, and water-soluble organic compounds in the solution, and typical micelles are shown in Fig. 1.4. However, as pointed out by Porter (1994), the micelle is a dynamic entity and its structure and shape can change with time. Further aggregation above a certain temperature referred to as the cloud point, leads to the separation of the liquid phase that gives the solution a turbid appearance. The presence of ethoxy or ethylene oxide (EO) group in surfactants changes the critical packing parameter (CPP)1, which in turn is a key determinant of their micellar structure and CMC in their aqueous solutions (Rosen, 1989; Holmberg et al., 2003). The micelle formation in highly ethoxylated surfactants consistently yields vesicular mesophases, and their formation results in a reduction in CMC at elevated temperatures 1 CPP is the ratio between the cross-sectional area of the hydrocarbon tail part and that of the polar head group of the surfactant molecule. 8 (Holmberg et al., 2003). On the other hand, molecules of CTAB (a non-ethoxylated cationic) typically cannot pack into a cone truncated by surfaces of high and opposite curvature as needed to direct the vesicular mesostructure (Lu, et al., 1999); instead, lamellar mesophases are commonly formed in bulk and thin-film samples. (a) Spherical micelle (d) Reversed micelle (b) Cylindrical micelle (e) Bicontinuous micelle (c) Lamellar phase (d) Vesicle Fig. 1.4 Different possible micellar structures (Evans and Wennerström, 1999) 9 1.1.4 Electrokinetic effects and zeta potential When two phases come in contact, they generally develop a potential difference between them. With the presence of ions, or excess electrons, or ionogenic groups in one or both phases, there is a tendency for the electric charges to distribute themselves in a particular direction at the interface (Hunter, 1981; Lyklema, 1991; Evans and Wennerström, 1999; Hunter, 2001). An electrokinetic boundary layer develops when a solid surface containing immobilized electrical charges comes in contact with an aqueous solution of mobile ions. Also referred to as an electric double layer (EDL) as shown in Fig. 1.5 (Hunter, 2001), it generally consists of the following two layers: (1) an immobile or stern layer of ions opposite in sign to that of the surface, and (2) a diffuse layer or a cloud of hydrated ions, which transition from the significant excess of counterion at the SOLID SURFCAE fixed charge layer to a balance of cations and anions in the bulk solution. Stern Layer + + _ + _ + + _ + _ + _ + _ + _ + + + _ + + + _ _ _ + _ + Diffuse Layer _ _ + _ DISTANCE ζ POTENTIAL 0 Fig. 1.5 Schematic diagram of a typical electrokinetic boundary layer 10 A variety of phenomena or “electrokinetic effects” are observed when one of these phases is caused to move tangentially past the second phase. The consequent forces on the solid or liquid interface can be characterized in terms of either the charge or electrostatic potential (Lyklema, 1991; Evans and Wennerström, 1999; Hunter, 2001). In the later case, the average potential in the surface of shear or zeta potential ζ is the fundamental quantifying parameter, and it provides the basis for explaining a variety of natural phenomena in colloid chemistry and electrochemistry. These include electrode kinetics, electrocatalysis, corrosion, adsorption, and crystal growth, among others, and the concomitant flow behavior and colloid stability cannot be treated without knowledge of charge distribution in the interfacial region (Lyklema, 1991; Evans and Wennerström, 1999; Hunter, 2001; Birdi, 2003). Many of the important properties of colloidal systems are determined directly or indirectly by the potential at the interfaces. Adsorption of ions and molecules, for instance, is determined by the charge and potential distribution, which itself determines the interaction energy between the molecules or particles. There are four distinct electrokinetic or zeta potential effects depending on the way in which motion is induced: electrophoresis, electroosmosis, streaming potential, and sedimentation potential (Hunter, 1981; Delgado, 2002), and Fig. 1.6 shows the relationship between them. For the surfactant adsorption at the solid-liquid interface, the streaming zeta potential will arise when the solution is in contact with the stationary surface. The measurement of streaming potential is relatively straightforward, as has been described by Hunter (1981), Gu and Li (2000), and Gusev and Horváth (2002), and others. In principle, the double layer ions are carried downstream by the flow through a capillary, and their accumulation there generates a field that causes a back conduction. 11 When the forward and the back currents are equal, then the potential difference across the capillary is the streaming potential. Streaming potential measurements are primarily dependent upon the following: (i) the applied pressure, (ii) the conductivity of the liquid in the capillary or plug, and (iii) the streaming potential developed. The fundamental importance of the zeta potential change and physisorption process in characterizing nucleate boiling in aqueous surfactant solutions cannot be overstated. In these association colloids, molecules of the surface-active additive form small aggregates or micelles in water, which tend to adopt an ordered structure. The adsorption of ionic surfactants, in particular, at the solid-liquid interface alters the behavior of the solid surface considerably (Hunter, 2001; Fuerstenau, 2002). Based on this adsorption process, the electrokinetics and wettability behaviors of the solid surface can be explained from the ions exchange in the EDL, and are directly reflected in the change in ζ. 12 Di sp es y v o ar S Li erse E M n LE qui Ph o I aseStati RES CT d P ase h P e O ROhas St e rs has PH e - O e M a t io sp id P RO i SM ov nar D qu T OSes y C i Applied Field L LE Causes Movement IS E ZETA POTENTIAL ST L R IA T Di EA Applied Force Results N n in Potential LiqspersMIN TEotio y e ui d P G PO M nar Ph has POT N in io ase e S E IO ase Stat N t T a in tio T h Mo na IA TAse P hase L N tio r y E per d P n i IMDis qu D Li SE Fig. 1.6 Four basic electrokinetic phenomena and the relationship between them 13 1.1.5 Surfactant physisorption at the solid-liquid interface The adsorption of ionic surfactants at a solid-liquid interface in physically adsorbing systems, i.e., in systems where chemical interactions are absent and adsorption at low concentrations occurs through electrostatic interaction between the surface and surfactant ions charged oppositely to the surface, has been extensively investigated in the literature. Various theories about the structure of the adsorbate layer have also been presented (Hunter, 2001; Dobiáš and Rybinski, 1999). Of these, the hemimicelle or Fuerstenau model has been broadly adopted in colloid science, and the pioneering work on the nature of surfactant adsorption that lead to this model was reported by Somasundaran and Fuerstenau (1966). The typical isotherm in Fig. 1.7 shows four distinct regions of surfactant adsorption that are associated with the aggregation mode of adsorbed ions at the solidliquid interface (Somasundaran and Fuerstenau, 1966; Scamehorn et al. 1981; Dobiáš and Rybinski, 1999; Fuerstenau, 2002). These can be summarized as follows: Region I: At low concentrations, surfactant adsorption takes places as individual ions in the diffuse part of the double layer and in the stern plane, and obeys Henry’s law. There is no chain-chain association, and it is independent of the hydrocarbon chain length. Under these conditions, the zeta potential is almost constant and surface wettability remains relatively unchanged. Region II: There is sharp increase in the slope of the adsorption isotherm due to selfassociation of adsorbed surfactant ions, and the formation of hemimicelles (Bisio, 1980; Fuerstenau, 2002). Adsorption takes place primarily in the stern plane, and the polar heads of the surfactant ions are oriented toward the 14 CMC (critical micelle concentration) Adsorption Density Region IV Plateau adsorption region (Double layer) Region III Reverse Hemimicelles PZR (point of zeta potential reversal ) Region II Hemimicelles Region I Individual Ions start to aggregate Individual Ions Concentration Hydrophilic head Hydrophobic tail Hydrophobic Hydrophilic Fig. 1.7 Schematic representation of a typical adsorption isotherm and aggregate states of a surfactant on a solid surface in aqueous solutions 15 surface. The surface is hydrophobic in this region for most surfactants except for high molecular weight ones, whose bulky polar head would occupy a larger surface area. Region III: Is characterized by a decrease in the slope of the adsorption isotherm due to the reversal of the ζ potential, which becomes increasingly negative for anionic surfactants and increasing positive for cationic surfactants because of the opposite charges they carry. The surfactant ions adsorb as reverse hemimicelles (Fuerstenau, 2002), with their polar heads oriented both toward the surface and liquid, and the surface becomes increasing hydrophilic. Region IV: As the CMC is approached, the adsorption isotherm becomes independent of the surfactant concentration in solution, forming a bilayer or equivalent (Manne and Gaub, 1995; Schulz et al., 2001; Fuerstenau, 2002; Richard et al., 2003). The ζ potential tends to remain constant and the surface becomes strongly hydrophilic. In this characterization, it should be noted that there are two important transition points in the overall adsorption process. The first is at the end of Region II, or the intersection between Regions II and III, where the ζ potential is zero and all of the adsorbed ions must be in the Stern plane. This point is variously referred to in the literature as the isoelectric point (IEP)2 (zero electrokinetic potential), or point of zero charge (PZC), or point of zeta potential reversal (PZR). In colloidal systems, detecting the IEP poses considerable difficulties, and ζ potential seems a more reliable indictor of specific or inner layer adsorption (Hunter, 2001; Birdi, 2003). Under these conditions, one part of the double 2 The concentration of the potential determining ion at which the zeta potential is zero is defined as the isoelectric point (IEP). 16 layer reflects the surface charge on the solid surface and the other part constitutes the oppositely charged surfactant ions. The second transition point is at the end of region III, when CMC is reached. The bilayer is formed at this transition and the contact angle tends to remain constant even when C ≥ CMC. Several variations of this adsorption model have been postulated in the literature. Bisio et al. (1980), Scameborn et al. (1981), and Chandar et al. (1987) modified the Fuerstenau model by postulating the formation of surfactant double layers (bilayers) as well as surfactant monolayers in the second adsorption stage. They called them lamellar layers, where on top of the first layer (with head groups are oriented toward the surface) a second layer (with head groups are oriented toward the bulk) is formed by means of hydrophobic forces. In their “admicelle” concept, Harwell et al. (1985) and Bitting and Harwell (1985) took into account the surface inhomogeneities, and indicated that the monolayer aggregates are never formed, which contrasts with the “bilayer” model. They regard these bilayer sections on the surface as a pseudophase and call them “admicelles” (adsorbed micelles) to differentiate them from Fuerstenau’s hemimicelles. Koopal and Ralston (1986) further provide a quantitative analysis of chain-chain interactions and the influence of chain conformation without reference to the notion of hemimicelle formation. These models are different, however, in almost all the systems that have been investigated, generally there is increasing hydrophilization of the surface with increasing surface coverage when IEP has been exceeded, which is typically expressed by an improvement in dispersibility of the solid particles in the aqueous solutions (Dobiáš et al., 1999). Some exceptions are charged or polarized surfaces, which are going to alter the 17 orientations of the adsorbed ionic surfactants because of the electro-repulsions. Also, with some types of surfactants that may form trilayers or reversed micelles when CMC is approached, different adsorbate layers and surface wetting behaviors may be exhibited (Evans and Wennerström, 1999). 1.2 Nucleate Boiling and Surface Wettability 1.2.1 Contact angle and surface wettability Surface wettability is critical to many applications that are as diverse as power generation, chemical processing, and biological systems. It generally refers to the manifestation of the molecular interactions between liquids and solids in direct contact at the interface (Blake, 1993), and is generally reflected in the contact angles (Lyklema, 1991; Kwok and Neumann, 1999). In multiphase heat transfer, the nature and dynamics of surface liquid-solid contact often plays a dominant role (Chen, 2003). The liquid-solid system can be either completely wetting (θ = 0°), or have different degrees of wetting (0 < θ < 180°), or be complete non-wetting (θ = 180°) as schematically illustrated in Fig. 1.8. Knowledge of the intermolecular interactions, both within the liquids, and across the liquid-vapor (or liquid-gas) and liquid-solid interfaces, is an important part of characterizing surface wettability or contact angle. When a drop of liquid is placed on a solid substrate, it may spread so as to increase the liquid-solid and liquid-gas interfacial areas. Simultaneously, the solid-gas interfacial area decreases and the contact angle θ between the drop and solid is reduced. The value of θ can be seen as a measure of the balance between the tendency of the drop to spread so as to cover the solid surface and to contract in order to minimize its own surface area (Decker et al., 18 1999; Holmberg et al., 2003); spreading would typically continue until the system reaches equilibrium. θ = 0º θ = acute complete wetting (hydrophilic) θ different degrees of θ = obtuse θ wettability θ = 180º complete non-wetting (hydrophobic) Fig. 1.8 Different wetting conditions of a liquid drop on a solid surface 19 The degree of spreading is governed by the surface tension of the liquid σLG, the surface tension of the solid σSG (usually referred to as the surface free energy), and the interfacial tension of the liquid and the solid σSL. These forces that essentially represent the liquid-gas, solid-gas, and solid-liquid interfacial tensions, and their interactions are depicted in Fig. 1.9. σLG Vapor or gas Liquid θ σSG Solid σSL Fig. 1.9 Surface forces involved in spreading of a liquid The surface free energy of the solid σSG, tends to spread the drop, i.e., to shift the threephase point forward or along the surface. Thus, spreading is generally favored on highenergy surfaces. The interfacial tension σSL and the horizontal component of the surface tension force σLGcosθ act in the opposite direction. At equilibrium state, the resultant force is thus zero and σ SG = σ SL + σ LG cosθ (1.1) This expression is very well known as Young’s equation and has become the basis for understanding the phenomenon of contact angle or surface wetting on solid surfaces. 20 Contact angle measurements are typically made using either sessile drops or adhering bubbles, and are usually automated and computerized, which enables values of the contact angles to be determined with a high degree of reproducibility. Also, the contact angle can be influenced by the surface roughness, surface chemical heterogeneity, and impurities, among some other factors (Kwok and Neumann, 1999). 1.2.2 Surface Wetting Effects on Nucleate Boiling The phenomenological modeling of mechanisms for heat and fluid transport in nucleate boiling is still not completely developed, and is the subject of much study. One reason for this is the complexity of interfacial contact at the surface, which is affected by intermolecular forces. Given that the primary heat transfer is by evaporation and its efficiency is directly related to nucleation site density and bubble dynamics, surface wetting becomes an important predictor. To characterize the inter-relationship, insights are often obtained from a visual observation of the ebullience. As a critical determinant of nucleate boiling, active nucleation site density is found to be a function of wettability and heat flux, and it directly accounts for the energy transfer by ebullience at the heater surface (Dhir, 1998; Barthau, 1992). Wang and Dhir (1993), Dhir (1998), and Basu et al. (2002) have systematically studied the effect of pure liquid wettability on the active nucleation site density and onset of nucleate boiling (ONB) in pool boiling. They have correlated their data for active site density as a function of the wall superheat and contact angle, and pointed out that the fraction of the nucleated cavities decreases as the wettability of the surface increases. The wettability of the surface in their collective work was changed by controlling the degree of oxidation of 21 the heater surface. Hibiki and Ishii (2003) have also presented results that map active nucleation site density as a function of contact angle and wall superheat, as well as the critical cavity size. Similarly, as reviewed by Dhir (1998) and Kenning (1999), increased surface wettability, which results in fewer nucleation sites, typically produces larger bubbles with lower departure frequencies. 1.3 Nucleate Pool Boiling with Surfactants Small amounts of surfactant additives in water tend to change and enhance the boiling heat transfer in water by essentially modifying nucleation and the concomitant bubble dynamics. The importance of surfactant-enhanced boiling heat transfer has been widely recognized, and many studies have investigated the pool boiling behavior in aqueous surfactant solutions under atmospheric conditions. A recent review of this body of work was provided by Wasekar and Manglik (1999), and Table 1.2 gives a more comprehensive chronological listing of the available literature on experimental investigations. Boiling with surfactant additives is a very complex process as shown earlier in Fig. 1.2. Besides the effects of heater geometry, its surface characteristics and wall heat flux level, the bulk concentration of additive, surfactant chemistry (ionic nature and molecular weight), dynamic surface tension of the solution, surface wetting and nucleation cavity distribution, marangoni convection, surfactant adsorption and desorption, and foaming are some of the factors that appear to have a significant influence (Wu, et al. 1998; Hetsroni, et al., 2001; Wasekar and Manglik, 2001, 2002). Also, the bubble dynamics (inception and gestation → growth → departure) has been 22 found to be considerably altered with reduced departure diameters, increased frequencies, and decreased coalescence (Wu et al. 1995; Wasekar and Manglik, 2000; Hetsroni, et al., 2001). A direct correlation of the heat transfer with suitable descriptive parameters for all of these effects, however, remains elusive because of the complicated nature of the problem. Table 1.2 Chronological Listing of Nucleate Pool Boiling Studies of Aqueous Surfactant Solutions Author(s) Heater Geometry Surfactants Stroebe et al. (1939) Morgan et al. (1949) Jontz and Myers (1960) Roll and Myers (1964) Cylinder Cylinder Plate Plate Duponol Drene; SDS Tergitol; Aerosol-22 Aerosols: OT, AY, IB, and MA; Hyonics: PE-200 Frost and Kippenhan (1967) Huplik and Raithby (1972) Shah and Darby (1973) Shibayama et al. (1980) Cylinder Ultra Wet 60L Podsushnyy et al. (1980) Cylinder Filippov and Saltonov (1982) Yang and Maa (1983) Saltanov et al. (1986) Chang et al. (1987) Tzan and Yang (1990) Cylinder Octadecylamine Plate Cylinder Cylinder Cylinder SLS and SLBS Octadecylamine SDS SDS Plate FC-176 Plate Plate Joy Sodium oleate; Rapisool B80; Puluronic: F98, F88, F208 PVS-6 polyvinyl alcohol, NP-3 sulfonol, and SV1017 wetting agent 23 Table 1.2 (continued) Liu et al. (1990) Plate Chou and Yang (1991) Wu and Yang (1992) Plate Cylinder BA-1, BA-2, BA-3, BA-4, DPE-1, DPE-3, Gelatine, Oleic acid, Trimethyl octadecyl ammonia chloride, trialkyl methyl ammonia chloride, and polyvinyl alcohol SDS SDS Wang and Hartnett (1992) Wu et al. (1993) Tan and Wang (1994) Lin et al. (1994) Wu et al. (1994) Wang and Hartnett (1994) Wire Tube Cylinder Sphere Sphere Wire SDS SDS WY SDS SDS SDS and Tween-80 Wu et al. (1995) Cylinder SDS, Tergitol, Aerosol-22,DTMAC, Tween-20, 40, 80, n-Octanol, and Triton X-100 SDS Ammerman and You (1996) Qiao and Chandra (1997) Manglik (1998) Wu et al. (1998a) Wu et al. (1998b) Wu et al. (1999) Wire Plate Cylinder Cylinder Cylinder Cylinder Yang et al. (2000) Cylinder Wasekar and Manglik (2000) Hetsroni et al. (2001) Yang et al. (2002) Wen and Wang (2002) Wasekar and Manglik (2002) Cylinder SDS AGS SDS, and Triton X-100 n-octanol in water and LiBr SDS; DTMAC; Triton X-100, Aerosol CPC, SDS, DTMAC, DTMADS, CPDS SDS Plate Cylinder Plate Cylinder Habon G Triton SP-190, SP-175 SDS, Triton-X-100, Octadecylamine SDS, SLES, Triton X-100, X-305 24 A variety of different predictive parameters and mechanisms have been proposed to describe the complex phase-change process, but much of the focus has been on the effect of reduced interfacial tension. Because of the highly dynamic nature of the nucleate boiling, typically in the range of 0-100 ms (Prosperetti and Plesset, 1978), the dynamic surface tension of the aqueous surfactant solution instead of the equilibrium surface tension has been proposed as a predicator of the nucleate boiling process (Wasekar and Manglik, 2002). Furthermore, the addition of a small amount of surfactant to water, not only changes the liquid-vapor interfacial behavior, but, more importantly, it also alters the solid-liquid interfacial characteristics. All the factors related to surface wettability thus get affected, including one of the most important ones in nucleate boiling - active nucleation density. The wetting of aqueous surfactant solutions is further found to be influenced by the additive’s chemical structure. The presence of the ethoxy or ethylene oxide (EO) group in its molecular-chain, in particular, increases the overall size of the polar head and makes the surfactant more hydrophilic (Barry and Wilson, 1978; Evans and Wennerström, 1999; Holmberg et al., 2003). This increases surface wettability due to the adsorption of surfactant molecules on the solid surface (Ashayer et al., 2000). The concomitant influence on the active nucleation site density and dynamic contact angle should therefore be taken into consideration in characterizing nucleate boiling of aqueous surfactant solutions. 25 1.4 Nucleate Pool Boiling with Polymers Thermal processing of fluid media to produce biochemical, pharmaceutical, personal care, and hygiene products is a very complex heat transfer problem. It typically entails heating and drying of aqueous polymeric solutions by boiling, in order to thicken them and make pastes. Rather anomalous phase-change behaviors have been observed in this process (Kotchaphakdee and Williams, 1970; Wang et al., 1992; Shul’man et al., 1993), and the consequent lack of close thermal control often leads to product loss and quality degradation. A variety of factors play a role, and they include the type of polymer, its molecular weight and concentration, solution rheology and interfacial properties (surface tension and wettability), heated surface geometry, and heat flux levels, among others. Several studies have investigated nucleate pool boiling characteristics of polymeric solutions under atmospheric conditions, and Table 1.3 gives a chronological listing of most of the available literature. In one of the earliest studies on the effects of polymer additives on boiling of water on a plate heater, Kotchaphakdee and Williams (1970) found the heat transfer to be enhanced in HEC-H and PA-10 solutions. While both additives make the solution more viscous with a shear-thinning flow behavior, HECH has surface-active properties as well (i.e., it reduces surface tension of the solution appreciably) and a lower molecular weight (M = 2×105); for PA-10, M = 106. The combined effects of nucleation site density and solution concentration on boiling are reported by Ulicny (1984). Appreciable enhancement was observed in aqueous HEC solution on a 600 grit roughened surface, whereas no apparent enhancement was observed for the same solution on a much smoother surface. The experimental data of 26 Table 1.3 Chronological Listing of Nucleate Pool Boiling Studies of Aqueous Polymeric Solutions Author(s) Kotchaphakdee and Heater Geometry Polymers Plate Acrylamide, PA-10, PA-20, HEC- Williams (1970) Miaw (1978) Yang and Maa (1982) Paul and Abdel-Khalik L, HEC-H, HEC-M Plate HEC-H, PA-10, PA-30 Plate and wire HEC-250HR, 300HR, 250GR Wire Separan AP-30, NP-10P, MGL; (1983) PEO; HEC 250MR, 250HR, 250 HHR Ulicny (1984) Plate PA-10 Hu (1989) Wire Separan AP-30, HEC 250HHR Wang and Hartnett Wire SLS and Separan AP-30 Shul’man et al. (1993) Plate PAA, HEC-H, PEO Shul’man and Levitskiy Plate PAA, HEC-H, PEO Plate PAA, HEC-H, PEO (1992) (1996) Levitskiy et al. (1996) Bang, et al. (1997) Sphere PEO Wang and Hartnett (1992), Paul and Abdel-Khalik (1983), and Hu (1989), however, indicate a deterioration in boiling heat transfer for very dilute aqueous polymeric solutions when compared to that of pure water. All of these studies (Paul and AbdelKhalik, 1983; Hu, 1989; Wang and Hartnett, 1992) used platinum wire heaters instead of a plate heater (Kotchaphakdee and Williams, 1970; Ulicny, 1984) and geometry effects may be present. The results of Yang and Maa (1982) are even more contrary, and the same boiling heat transfer performance for dilute aqueous HEC solutions with both a 27 plate and platinum wire heater has been reported. More recently, Shul’man et al. (1993), and Levitskiy et al. (1996) have shown that even with the same kind of polymers of the same molecular mass, the boiling performance can substantially change with concentration, temperature, and external conditions. They report enhanced boiling heat transfer in dilute solutions (C = 15-500 wppm), but a decreased boiling heat transfer in highly concentrated aqueous solutions (C = 1%) of HEC-H on a plate heater, which has a characteristic size that is much larger than the mean size of the bubbles. The changed boiling heat transfer in polymer solutions is also displayed in a markedly different bubbling behavior (shape and size of bubbles, their growth rate, foaming, and nucleation frequency, etc.) compared to that of pure water (Shul’man et al., 1993; Hu, 1989; Levitskiy et al., 1996). Bubbles of smaller sizes and regular shapes are released from the heater with higher frequencies than seen in pure water, and they rise in a more orderly fashion. Adjacent bubbles tend to coalesce less due to the effects of normal stresses and longitudinal viscosity in thin films formed between them. Levitskiy et al. (1996) suggest that the change in the wetting angle along with a reduction in surface tension for HEC-H solutions account for the observed decrease in bubble sizes. The changes in the interfacial characteristic are perhaps a direct consequence of the molecular adsorption dynamics of the additive. Polymers are typically large molecules, macromolecules, or agglomerates of smaller chemical units called monomers, and are broadly classified as biological or nonbiological macromolecules. Their addition in water primarily increases the solution viscosity, which tends to increase with concentration as well as the molecular weight of the polymer, and often display a shear-rate dependent shear-thinning rheology (Chhabra 28 and Richardson, 1999; Carreau et al., 1997). With the exception of some surface-active polymers (or polymeric surfactants) such as hydroxyethyl cellulose (HEC) and polyethylene oxide (PEO), most polymeric solutions do not show any significant change in surface tension σ (Manglik et al., 2001; Hu et al., 1991). The viscosity of the polymer solution, however, can influence σ measurements considerably, especially at higher viscosity and bubble frequency (Manglik et al., 2001; Fainermann et al, 1993; Janule, 1998). The reduced surface tension in solution of surface-active agents is largely brought about by the molecular adsorption of the additives to the vapor-liquid interface (Holmberg et al., 2003). The time scales of this process vary from order of seconds to minutes depending upon the polymer chemistry and its concentration in solution, which is possibly due to the slow processes of diffusion transport of polymer molecules to the interface and their subsequent reorientation (Persson et al., 1996). This dynamic adsorption process, along with time scales of 10-100 ms for boiling bubble dynamics in water (Prosperetti and Plesset, 1978), thus results in a rather complex interfacial behavior, that significantly alters the nucleate pool boiling in polymeric solutions. 29 1.5 Computational Fluid Dynamics with Moving Boundaries In order to develop theoretical constructs for the complex ebullient phase-change behavior, computational modeling provides an attractive tool. However, this poses considerable difficulties, where, for example, a direct calculation of the nucleate boiling flows without empirical correlations essentially requires accurate tracking of the twophase interfaces. These flows are characterized by the discontinuity of many variables across the phase interface. These discontinuities pose several computational challenges requiring special treatment. In addition, the location of the interface is not known a priori and must be found as part of the solution. Under the broad categories of Langrangian and Eulerian methods, several numerical techniques have been developed so far by researchers in the area of moving boundary problems. In the class of finite-difference methods, such as the marker-and-cell (Harlow and Welch, 1965) and Volume-of-Fluid (VOF) (Hirt and Nichols, 1981) methods, the moving interface is traced with marker particles or functions that are advected through the finite difference mesh. In these methods, the physical quantity at a computing cell implying an interface is calculated by volumetric averaging of vapor and liquid phases; the discontinuous interface is therefore likely to be smoothed out. The numerical schemes such as the Arbitrary-Lagrangian-Eulerian (ALE) method (Hirt et al., 1974), Boundary-Element Method (BEM) (Brebbia et al., 1984), Boundary-FittedCoordinates (BFC) method (Ryskin and Leal, 1984), and Finite-Volume-Method (FVM) (Rhie and Chow, 1983), have an advantage in that their mesh can be generated to fit the interface shape. The generation of interface-fitted mesh, however, is troublesome when the interface merges or has large deformation. In a combined Langrangian-Eulerian 30 method or the Front-Tracking Method (FTM) (Unverdi and Tryggvason, 1992) have employed a moving unstructured boundary-confirming grid in combination with a stationary Cartesian grid to track the motion of bubbles undergoing severe deformation, which is in essence the front tacking schemes applied by Glimm et al. (1987). The Unified Coordinate method proposed by Hui et al. (1999) is an innovative scheme to conduct Langrangian and Eulerian calculations on a single unified coordinate by a simple mathematical transformation, in which the Eulerian or Langrangian coordinate is just one of the special cases. More recently, newer approaches of interface evolution problems have been developed. Phase-field model (Kobayashi, 1993; Wheeler et al., 1993), Phase-field Fourier-Spectral Method (Liu and Shen, 2003), Ghost Fluid Method (GFM) (Fedkiw et al. 1999), Level-Set Method (LSM) (Osher and Sethian, 1988; Sethian, 1990), and Segment Projection Method (Front-Tracking + Level Set) (Tornberg, 2000), have made significant improvements on tracking the liquid-vapor interfaces. Among these, the Level-Set Hamilton-Jacobi formulation has been widely used for capturing interface evolution especially when the interface undergoes extreme topological changes, e.g., merging or pinching off. It is also attractive because it admits a convenient description of topologically complex interfaces and is quite simple to implement. Even though the level-set method does not have the same conservation properties as VOF or front-tracking methods, the strengths of the level set method lie in its ability to accurately compute flows with surface tension and changes in topology. 31 1.6 Scope of Study As stated previously boiling of aqueous surfactant or polymer solutions is not only important in thermal processing of biochemical, pharmaceutical, personal care, and hygiene products, but the addition of desired amounts of surfactants to water is also an innovative and promising technique to control the phase change process itself. The presence of a small amount of surfactant or polymer additive in water changes both the liquid-vapor (bulk chemistry) and the solid-liquid (surface chemistry) interfacial characteristics. This, however, along with the fluid rheology change and the complex nature of nucleate boiling, results in a complex two-phase problem, which is investigated in this dissertation. The primary scope of this study in addressing the many facets of the problem is summarized in the following: 1. Characterization of the interfacial phenomena and the associated transport processes in aqueous surfactant and polymeric solutions at both the liquid-vapor (dynamic and equilibrium surface tensions) and the solid-liquid (surface wettability or contact angle) interfaces. 2. Conducting pool boiling experiments to quantify the effects of surfactant concentration, its ethoxylation, solution rheology, surface wettability, dynamic and equilibrium surface tensions, and additive molecular weight on the nucleate boiling performance of water on a horizontal cylindrical heater. 3. Rheology measurement and characterization of the effects of polymerization degree and surface-active properties on nucleate pool boiling heat transfer in aqueous polymer solutions. 32 4. Surface wetting behavior qualification with the delineation of electrokinetic effects or zeta potential due to the physisorption at the solid-liquid interface, which in turn correlates with the corresponding adsorption isotherm. 5. Visualization and characterization of nucleate pool boiling in aqueous surfactant solutions in order to qualitatively relate the heat transfer performance to the nucleation process and ebullience. This facilitates the phenomenological understanding of the rather complex and elusive transport process associated with boiling heat transfer. 6. Computational simulation of the dynamics of a single bubble in pool boiling of water, which includes the microlayer modeling and its role in nucleate boiling, and the effects of altered contact angle, wall superheat, surface tension and viscosity on the bubble dynamics. 33 CHAPTER 2 INTERFACIAL PROPERTIES AND RHEOLOGY MEASUREMENTS 2.1 Surface Tension Measurements 2.1.1 Introduction For boiling in aqueous surfactant solutions, the detailed knowledge of dynamic σ and corresponding adsorption behavior is very crucial. The dynamic surface tension of liquids can be measured by several different techniques, and these are either direct methods (maximum bubble pressure, oscillating jets, and Langmuir through methods) or indirect methods (surface wave, oscillating bubble, and pulsed drop methods). The equilibrium σ, on the other hand, can be measured by either standard static methods (duNüoy, sessile drop, and Wilhelmy plate) or as long time asymptotes of the dynamic surface tension measurements. Joos (1999) and Dukhin et al. (1995) documented several different equilibrium and dynamic surface tension measurement methods along with their respective characteristic time and temperature ranges of operation, and suitability of application. Surface tension measurements in this study were made by the maximum bubble pressure method using a twin orifice computerized surface tensiometer (SensaDyne QC6000, CSC Scientific Company). Figure 2.1 schematically gives the details of the surface tensiometer and the instrumentation employed. Dry air at 3.4 bar is slowly bubbled through a parallel set of small and large glass orifice probes of 0.5 and 4.0 mm diameter, respectively, which are immersed in the test fluid pool in a small beaker to produce a differential pressure signal proportional to the fluid surface tension. The temperature of the test fluid is measured using a well-calibrated thermistor (±0.1°C precision, 0 - 150°C) attached to the orifice probes. The aqueous solution container is 34 immersed in a constant temperature bath in order to control and maintain its desired temperature. The time interval between the newly formed interface and the point of bubble break-off is referred to as “surface age,” and it gives the measure of bubble growth time that corresponds to the dynamic surface tension value at a given operating bubble frequency. Thus, by altering the air-bubble frequencies through the probes, both static or equilibrium and dynamic surface tension can be measured. Detailed descriptions of the solution preparation, instrument calibration, and validation procedures, along with measurement uncertainties can be found in Manglik et al. (2001). The maximum uncertainties in the measurement of concentration, temperature, and surface tension were found to be ±0.4% for powder form additives, ±5% for additives in liquid form, and ± 0.5% and ±0.7%, respectively. Several different water-soluble surfactants that are representative of a wide spectrum of commonly used additives are employed: dodecyltrimethylammonium chloride (DTAC, cationic), cetyltrimethylammonium bromide (CTAB, cationic), oleylmethylbis[2-hydroxyethyl]ammonium chloride (Ethoquad O/12 PG, cationic), octadecylmethyl[15-polyoxyethylene]ammonium chloride (Ethoquad 18/25, cationic), sodium dodecyl sulfate (SDS, anionic), sodium lauryl ether sulfate (SLES, anionic), octylphenol ethoxylate (Triton X-100, nonionic), and octylphenoxypolyethoxyethanol (Triton X-305, nonionic). They have different molecular weights, ionic nature, and number of EO groups or degree of ethoxylation. Their chemical composition and relevant physico-chemical properties are listed in Table 2.1. Two polymers with different degree of polymerization and surface-active properties: hydroxyethyl cellulose (HECQP300), a nonionic compound with surface-active properties, and Carbopol 934, a 35 cationic shear-thinning polymer without any surface-active properties, are also employed. Their chemical composition and relevant physico-chemical properties are listed in Table 2.2. Sensor Package Differential Pressure Gas Gas Transducer Flow Flow Flow Controller Metering Valves Small Orifice Probe Large Orifice Probe Pressure Controller P Interface Card Temperature Probe Test Fluid Computer Constant-temperature bath Air Supply Tank Fig. 2.1 Schematic of surface tensiometer and data acquisition system 36 Table 2.1 Physico-Chemical Properties of Surfactants (Octadecylmethyl [15-polyoxyethylene] ammonium chloride) (Sodium dodecyl sulfate) SLES (Sodium lauryl ether sulfate) RN(CH3)(CH2CH2O H)2Cl, R=oleyl RN(+)(CH3)[(CH2CH 2O)mH][(CH2CH2O)n H]Cl(-), R=C18H37 C12H25SO4 Na Cationic 2 Yellow viscous liquid Cationic 15 Yellow viscous liquid 364.5 403 Sigma-Aldrich DTAC CTAB Ethoquad O/12 PG (Chemical Name) (Dodecyltrimethyl ammonium chloride) (Cetyltrimethyl ammonium bromide) (Oleylmethylbis[2hydroxyethyl] ammonium chloride) Chemical formula C15H34ClN C19H42BrN Ionic nature a EO group Cationic 0 Cationic 0 Appearance White powder White powder Molecular weight 263.9 Manufacturer Sigma-Aldrich Surfactant Purity ≥ 99% Melting point > 246°C Solubility 50 mg/mL (20°C) Specific Gravity Viscosity b (cp) (pure liquid) Surface Tension (mN/m) 25°C) a Ethoxy or ethylene oxide group b Brookfield viscometer Ethoquad 18/25 SDS Triton X-100 Triton X-305 (Octylphenol ethoxylate) (Octylphenoxy plyethoxy ethanol) C12H25(OCH2 CH2)3 SO4Na C14H21(OCH2 CH2)9-10OH C14H21(OCH2CH 2)30OH Anionic 0 White powder Anionic 3 Nonionic 9-10 Nonionic 30 Slightly yellow viscous liquid Clear liquid Clear liquid 994 288.3 422 624(average) 1526 (average) AkzoNobel AkzoNobel Fisher Henkel ≈ 99% > 230°C ≥ 99% - ≥ 99% - ≥ 99% > 206°C ≥ 99% - Union carbide - 10 % (w/v) > 25% (w/v) - 150mg/mL - - - - 0.986 (25°C) 1.058 (25°C) 0.4 1.03 1.065 1.095 - - 1750(23°C), 110(90°C) - 500 (25°C) 240 (25°C) 470 (25°C) - 40.3 (0.1%), 40.7 (1.0%) 50 (0.1 %) - - - - 37 Union carbide - Table 2.2 Physico-Chemical Properties of Polymers Polymer additive (Trade name) Chemical name Ionic nature Appearance Molecular weight Manufacturer Purity Specific gravity Viscositya, cps (25°C) a Brookfield viscometer HEC (NATROSOL QP-300) Hydroxyethyl cellulose Nonionic White to light tan dry powder ~ 4-6 x 105 Amerchol ~ 99% 1.033 300-400 (2% solution) Carbopol (CARBOPOL 934) Polyacrylic acid Cationic White dry powder ~ 3 x 106 BF Goodrich ~ 99% 1.41 ~ 8-9 x 104 (2% solution) 2.1.2 Viscosity effects on surface tension measurements Under dynamic conditions, the higher σ is primarily obtained due to the viscous resistance offered by the fluid against the growing bubble interface, and it has been found to be dependent upon fluid viscosity, capillary radius, and surface age (Fainermann et al., 1993; Janule, 1998), and to predict the apparent increase in σ, Fainermann et al. (1993) give the following correlation based on the Stokes’ flow approximation: ∆σ = 1.5 ( µ rcap τ ) (2.1) Alternatively, for measurements made by the SensaDyne QC6000 surface tensiometer, the viscosity effect can be corrected by adjusting the bubble frequency (or the surface age) of each orifice by employing the following inverse relationship with their respective radius (Janule, 1998): br r g = bτ τ g 1 2 1 2 (2.2) Here r1, r2 and τ1, τ2 are the respective radius and surface age of the small and large orifice. 38 2.1.3 Results and Discussion 2.1.3.1 Aqueous surfactant solutions The surface tension variations with concentration at both the equilibrium and higher bubble frequency f the four cationic surfactant solutions at 23°C are graphed in Fig. 2.2. While σ at higher f (represented by a surface age of 50 ms) is always larger than the corresponding equilibrium value (surface age 17-59 s), both are seen to decrease with increasing surfactant concentration to asymptotically attain a constant value beyond the critical micelle concentration. The CMC for the four surfactants at 23°C (obtained from the asymptotic intersection point of the equilibrium adsorption isotherm) are ~ 6000 wppm for DTAC, ~ 400 wppm for CTAB, ~ 600 wppm for Ethoquad O/12 PG, and ~ 2300 wppm for Ethoquad 18/25. For aqueous CTAB solutions, the σ - C data compare quite well with the Razafindralambo et al. (1995) results at 20°C, and so does the CMC with the 392.5 wppm value at 25°C reported by Holmberg et al. (2003). The σ values at equilibrium and higher f for the different surfactant solutions at an increased bulk temperature of 80°C are presented in Fig. 2.3. The higher bubble frequency measurements are, once again, for a bubble surface age of 50 ms, which is representative of bubble frequencies typically encountered in nucleate boiling of water. The surfactant adsorption at the bubble vapor-liquid interface is a time-dependent process that gives rise to the dynamic surface tension behavior; this, however, eventually reduces to the equilibrium condition after a long time period (Bahl et al., 2003; Manglik et al. 2001; Iliev and Dushkin, 1992). The variation of σ with surface age in Fig. 2.3 clearly illustrates this. Also, a lower molecular weight surfactant diffuses faster than its higher 39 90 80 Equilibrium Surafce age (50 ms) DTAC CTAB Ethoquad O/12 Ethoquad 18/25 σ [mN/m] 70 60 50 40 T = 23°C Razafindralambo et al. (1995), CTAB 30 100 101 102 103 104 C [wppm] Fig. 2.2 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous surfactant solutions at 23°C 40 molecular weight counterpart, and this is seen in the faster σ relaxation of CTAB in comparison with that for Ethoquad 18/25 in Fig. 2.4. The presence of EO groups in Ethoquad 18/25 makes its polar head more bulky and lowers its mobility. A similar trend is obtained at elevated temperature (Bahl, 2003). As such, the dynamic σ isotherms at 80°C are atypical of the additive adsorption-desorption kinetics during atmospheric pressure boiling of aqueous surfactant solutions3. The equilibrium σ at CMC, on the other hand, represents the maximum possible surface tension reduction for the solutions at 80°C. In general, it is observed that the process of micelle formation takes place over a range of concentrations (Rosen, 1989; Adamson, 1976), and in the present measurements, they are found to be ~ 4500 wppm for DTAC, ~ 500 wppm for CTAB, ~ 850 wppm for Ethoquad O/12 PG, and ~ 500 wppm for Ethoquad 18/25. The results in Fig. 2.3, when compared with the respective values at 23°C in Fig. 2.2, indicate overall reductions in σ at the higher temperature (80°C), which are due to increased surfactant diffusivity with increased temperature (Holmberg, et al. 2003; Rosen, 1989). These reductions, however, are not uniform over both the dynamic and equilibrium conditions. The degree of variation depends on the surfactant ionic nature and molecular structure among some other factors (Bahl et al., 2003; Manglik et al, 2001; Holmberg, 2003), and it reflects completely different adsorption-diffusion kinetics at elevated temperature during short and long transients. Another salient feature is that surfactants with EO groups in their hydrocarbon chain show larger reductions in σ with increasing temperature (Ethoquads versus DTAC and CTAB), which is clearly evident in 3 While 80°C is the upper limit for QC-6000 surface tensiometer, the surface tension data at real boiling temperature can be obtained by extrapolation of surface tension data with temperature, which typically has a linear relationship. 41 65 T = 80°C 60 55 σ [mN/m] 50 45 40 35 30 25 100 Equilibrium Surface age (50 ms) DTAC CTAB Ethoquad O/12 Ethoquad 18/25 101 102 103 104 C [wppm] Fig. 2.3 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous surfactant solutions at 80°C 42 Fig. 2.5, where σ - T variations for Ethoquad 18/25 and CTAB are graphed along with the results for pure water, and its shows larger gradients for Ethoquad 18/25 at all concentrations. The surfactant’s molecular-chain geometry and packing essentially determine the aggregate/micelle structure, and it is well known (Holmberg et al., 2003) that the polyoxyethylene chain compresses as the temperature increases. This leads to an increased CPP (critical packing parameter) value, which lowers the CMC as well as the surface tension (Holmberg et al., 2003). The temperature effect on CMC is even stronger for surfactants with larger number of EO groups, as the polar head size increases with increasing number of ethylene oxide units, and because they tend to form vesicle micelles instead of spherical or lamellar micelles (Lu et al., 1999), thereby exhibiting a totally different temperature dependence (Bahl et al, 2003; Manglik et al., 2001; Partearroyo et al. 1996). However, the σ - T variation generally tends to be linear for a surfactant at a given concentration as shown in Fig. 2.5. 43 70 Ethoquad 18/25 CTAB C/Cc.m.c. = 0.5 C/Cc.m.c. = 1 C/Cc.m.c. = 2.0 σ [mN/m] 60 50 40 Typical bubble frequencies in nucleate boiling of water 30 10-2 10-1 100 Surface Age [s] Fig. 2.4 Dynamic surface tension relaxation for aqueous cationic CTAB and Ethoquad 18/28 solutions 44 101 80 Ethoquad 18/25 CTAB 100 wppm 300 wppm 800 wppm 200 wppm 400 wppm 1000 wppm 70 Water σ [mN/m] 60 50 40 30 10 20 30 40 50 60 70 80 90 T [°C ] Fig. 2.5 Equilibrium surface tension measurements as a function of temperature 45 2.1.3.2 Aqueous polymeric solutions The measured surface tension values at different bubble frequencies and polymer concentrations for HEC QP-300 solutions at 23°C are graphed in Fig. 2.6. Similar to the observations made earlier for surfactants, the surface tension is found to increase with increasing bubble frequency and decreasing concentration. A critical polymer concentration (CPC) akin to CMC in surfactants is observed, such that the σ relaxation attains a saturation value at or around CPC. The value of CPC ascertained from the equilibrium σ - C isotherm (lowest bubble frequency of 0.017 Hz) is estimated to be ~ 600 wppm. The experimental data given by Hu et al. (1991), as well as the manufacturer’s (Union Carbide, 1998) reported σ value for 0.1% (concentration by weight) solution are also graphed in Fig. 2.6. Except for Hu et al. (1991) data, the present measurements for equilibrium σ agree well with the other results. The Hu et al. (1991) data appear to fall in the dynamic conditions represented approximately by the bubble frequency of 0.33 Hz, which corresponds to a surface age of 3 seconds4. As noted by Persson et al. (1996) for polymers belonging to the class of nonionic cellulose derivates that include HEC, time required for the complete relaxation of σ to an equilibrium value is of the order of minutes, possibly due to the slow processes of diffusion transport of polymer molecules to the interface and their subsequent reorientation. In concurrence with this, the present results indicate typical relaxation times for HEC to be around 1-2 minutes. The surface tension measurements for Carbopol 934 solutions displayed a similar behavior with varying bubble frequency and hence are not discussed separately. 4 It may be noted that though the data by Hu et al. (1991) were obtained by the maximum bubble pressure method, the bubbling frequency (as surface age) has not been mentioned in the paper. 46 74 Aqueous HEC-QP300 solutions at 23°C 73 72 σ [mN/m] 71 70 69 Bubble frequency: 10 Hz Bubble frequency: 2Hz Bubble frequency: 0.33Hz Bubble frequency: 0.017Hz Hu et al. (1991) at 25°C Manufacturer's Data (Union Carbide, 1998), 20°C 68 67 66 100 101 102 103 104 C [wppm] Fig. 2.6 Dynamic surface tension measurements for aqueous HEC solutions at 23°C 47 The higher bubble frequency and equilibrium σ values for HEC QP-300 and Carbopol 934 at room temperature are shown in Fig. 2.7. At equilibrium conditions, σ is seen to reduce with increasing concentration for both polymers. In comparison to Carbopol solutions, however, HEC solutions show significantly higher σ relaxation both at dynamic and equilibrium conditions, with a rather sharp change in slope or surface tension gradient near CPC. It should be noted that HEC is a surface-active polymer, and thus its adsorption behavior would tend to be similar to those of surfactant solutions. On the other hand, a reversed trend of increasing σ with concentration is observed in dynamic measurements of aqueous Carbopol solutions. This apparent increase in σ, especially under dynamic conditions, is due to the viscous resistance offered by the fluid against the growing bubble interface. Such an increase in σ measurements by the maximum bubble pressure method has been typically observed in highly viscous liquids like aqueous solutions of polymers and glycerol (Hirt et al, 1990; Fainermann, 1993). The dynamic surface tension measured according to Janule’s procedure (1998) is graphed in Fig. 2.7 along with the corrected values of σ using Eq. (2.1). For both polymer solutions, there is an excellent agreement between the corrected experimental σ values obtained from the modified measurement procedure (Eq. (2.2)) and those computed using Eq. (2.1). More notably, there is an appreciable difference between the viscosity compensated and non-compensated values for Carbopol solutions, with a characteristic decrease in σ with concentration (similar to HEC) even under dynamic conditions. HEC-QP300 being a sparingly viscous polymer, contrastingly shows smaller change in σ after viscosity correction. The surface tension data reported by Hu et al. (1991) and Ishiguro and Hartnett (1992) for Carbopol solutions are also graphed in Fig. 48 2.7. The discrepancies between these and the present data set are relatively large and may perhaps be attributed to the differences in methods of measurement, varying sources of polymer samples, and possible non-accounting of viscosity effects. Figure 2.8 depicts the variations in σ at both the equilibrium and higher f for the two polymers at an elevated solution temperature of 80°C. An overall depression in σ is observed with a maximum reduction in equilibrium σ of around 4% and 8%, respectively, for 2000 wppm Carbopol and HEC solutions. The viscosity effect for Carbopol solutions under dynamic conditions is found to be reduced, presumably due to the decrease in the solution’s apparent viscosity from room temperature to 80°C. Also, the dynamic σ values are close to the surface tension of water, and they do not show a significant variation with concentration. For HEC solutions, the viscosity compensation results in a negligible change in dynamic σ, which is quite similar to that observed in the behavior at room temperature. The polymer adsorption process at the bubble vapor-liquid interface is also time dependent, which manifests in a dynamic surface tension behavior that eventually reduces to the equilibrium condition after a long time period. This σ relaxation behavior is essentially the outcome of the molecular kinetics of the additive in water. In solutions with lower than overlap concentrations or CPC, the polymer molecules remain as isolated macromolecules with little intermolecular interactions. At overlap or “semi-dilute” concentrations, the polymer molecules “touch” each other, and with increasing concentration the frequency of collisions between the polymer coils eventually causes overlapping and entanglement of their chains. 49 74 72 70 σ [mN/m] 68 EXPERIMENTAL DATA (polymer solutions at 23°C) 66 64 Carbopol 934 Equilibrium 50ms, w/o viscosity correction 50ms, with viscosity correction by Eq. (2.1) 50ms, with viscosity correction by Eq. (2.2) Hu et al. at 25°C (1991) Ishiguro and Hartnett at 25°C (1992) HEC-QP300 62 60 101 Equilibrium 50 ms, w/o viscosity correction 50ms, with viscosity correction by Eq. (2.1) 50ms, with viscosity correction by Eq. (2.2) 102 103 C [wppm] Fig. 2.7 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous polymer solutions at 23°C 50 65 σ [mN/m] 60 EXPERIMENTAL DATA (polymer solutions at 80°C) Carbopol 934 55 Equilibrium 50ms, w/o viscosity correction 50ms, with viscosity correction by Eq. (2.1) 50ms, with viscosity correction by Eq. (2.2) HEC-QP300 Equilibrium 50ms, w/o viscosity correction 50ms, with viscosity correction by Eq. (2.1) 50ms, with viscosity correction by Eq. (2.2) 50 101 102 103 C [wppm] Fig. 2.8 Surface tension measurements at equilibrium and higher bubble frequency (surface age of 50 ms) for aqueous polymer solutions at 80°C 51 2.2 Contact Angle Measurements The liquid-solid contact angle was measured by the sessile drop method, using a Kernco GI Contact Angle Meter / Wettability Analyzer. The measurement uncertainty in this case is estimated to be a max of ±1.4% for powder form additives and ±5% for additives in liquid form. The change in surface wettability (measured by the contact angle) with concentration in ionic surfactants (SDS-anionic, CTAB and Ethoquad 18/25cationic) and nonionic surfactants (Triton X-100 and X-305) are graphed in Fig. 2.9(a). Ionic surfactants undergo a different adsorption process than that for nonionic surfactants due to the latter’s lack of charge. The adsorption isotherms for ionics (SDS, CTAB, and Ethoquad 18/25) correlate well with the physisorption characterization schematically illustrated in Fig. 2.9(b). The contact angle reaches a lower plateau around the CMC where bilayers start to form on the surface. Wettability of nonionic surfactants in aqueous solutions, on the other hand, shows that the contact angle data attains a constant value much below CMC. Direct interactions of their polar chain are generally weak in nonionics, and it is possible for them to build and rebuild adsorption layers below CMC (Levitz, 2002). The reduced contact angle trough at lower concentrations (C < CMC) can also be attributed to the absence of any electrical repulsion that could oppose molecular aggregation unlike that associated with ionic surfactants (Miller and Neogi, 1985). Furthermore, the continuous decrease in contact angle for Triton solutions prior to reaching a constant value is brought about by the presence of larger EO groups in the surfactant molecular chain. The number of EO groups increases the overall size of the polar head, and controls the hydrophilic/hydrophobic balance on the surfactant molecule (Holmberg et al., 2003). 52 90 CTAB (Cationic) Ethoquad 18/25 (Cationic) SDS (Anionic) Triton X-100 (Nonionic) Triton X-305 (Nonionic) Contact angle [deg] 80 70 CMC 60 50 CMC CMC CMC CMC 40 30 101 102 103 104 Concentration [wppm] (a) Individual Ions Reversed Hemimicelles Hemimicelles Hydrophobic tail Bilayer Hydrophilic head Hydrophobic Hydrophilic (b) Increasing EO groups (c) Fig. 2.9 (a) Measured contact angle for aqueous CTAB, Ethoquad 18/25, SDS, Triton X-100, and Triton X-305 solutions; (b) corresponding ionic surfactant adsorption surface state; and (c) EO group effect on surface wettability 53 The surfactant wetting behavior can be directly related to its adsorption isotherm, figure 2.10(a) shows that measured contact angle correlates well with the adsorption density. For nonionics, their molecules may just get adsorbed randomly at the liquidsolid interface due to lack of charge as illustrated in Fig. 2.10(b). If the molecule is adsorbing on a substrate like a steel sample that exhibits a contact angle, the more EO groups on the molecule should reduce the adsorption because that molecule would then rather be in the aqueous phase for nonionic surfactants (Ottewill, 1967). 80 60 4 3 CMC 50 2 CMC 40 30 101 Adsorption density [10-6mole/m2] Contact angle [deg] 70 5 Adsorption density Contact angle (Levitz, 2002) Triton X-100 Triton X-305 1 102 103 0 Concentration [wppm] (a) (b) Fig. 2.10 (a) Contact angle and adsorption isotherms for nonionic surfactants Triton X-100 and Triton X-305 in aqueous solutions; and (b) non-ionic surfactant adsorption. 54 2.3 Rheology Measurements Viscosity measurements were carried out using a rotating cylinder rheometer (AR-2000; TA Instruments) that can function in both a controlled-stress and controlledshear-rate environment. It contains an electronically controlled induction motor with an air bearing support for all rotating parts. The drive motor has a detachable draw rod arrangement to which the measuring geometry (rotating cylinder, cup and cylinder, parallel plate, and cone and plate) can be attached. The angular displacement is measured by an optical encoder device, which can detect very small movements down to 40 nRad. The encoder consists of a non-contacting light source and a photocell arranged on either side of a disc-shaped diffraction grating attached to the drive shaft. A stationary segment of a similar disc is positioned between the light source and the encoder disc. The interaction of these discs results in diffraction patterns that are detected by the photocell. When the liquid sample strains under stress, the encoder disc moves, and the diffraction patterns change, and the associated digital signals are directly related to the angular deflection, and, therefore, to the strain of the sample. Sample weights of additives in power form were measured using a precision electronic weighing machine of ±0.1 mg accuracy. For the additives available in liquid form, different ranges of precision syringes were used for the measurement of sample volumes, and the corresponding weight calculated by the known value of specific gravity of each additive. Then samples were dissolved in distilled deionized water (at slightly elevated temperatures to aid solubility for polymers) to obtain the test samples. The instrument calibration was validated using a standard oil sample, and pure water. The test liquid viscosity was then measured using a DIN geometry for shear rates less than 55 100 s-1, and a double concentric cylinder geometry system for higher shear rates. Temperature control was attained via a Peltier system that allows for rapid and accurate heating and cooling of the liquid sample. The data reproducibility was further checked using multiple runs for a few fixed-concentration samples. Again, the maximum singlesample, error propagation uncertainty in viscosity and temperature were ±1.4% for powder form additives and ±5% for additives in liquid form, and ±0.5%, respectively. 2.3.1 Aqueous surfactant solutions The viscosity-shear rate data for CTAB and Ethoquad 18/25 are presented in Fig. 2.11, where the relative changes in the apparent viscosity from that of water are graphed for three different concentration (C/CCMC= 0.5, 1, 2). Taylor-Couette instability will take effect after 150/s to show an untrue shear thickening behavior, therefore, the data after ~ 150/s were cut off in Fig. 2.11. It clearly shows that the viscosity of CTAB and Ethoquad 18/25 are close to that of the water in dilute aqueous surfactant solutions, the date scatter is well within ±5%. That dilute solutions of ionic and nonionic surfactants usually behave as Newtonian liquids, and the viscosity of these solutions is always close to that of solvent was also found in other studies (Wang and Hartnett, 1994; Hoffmann and Rehage, 1986). The measured viscosities for both SDS and Tween-80 aqueous solutions at room temperature were also found to be constant and independent of shear rate over a concentration range of 125 to 500 wppm (Wang and Hartnett, 1994). 2.3.2 Aqueous polymer solutions The viscosity-shear rate data for aqueous HEC QP-300 and Carbopol 934 solutions are graphed in Figs. 2.12 and 2.13, respectively. These were obtained in a 56 controlled-rate mode, where the shear rate was ramped and allowed to equilibrate to a steady-state value before the next successive increase. It is seen in Fig. 2.12 that HEC solutions are significantly more viscous than water, and that viscosity increases with concentration. Also, at low concentrations the solutions virtually behave as Newtonian fluids. The Hu et al. (1991) data are also included in Fig. 2.12 for comparison, though these are for a different grade of the HEC family of polymers (HEC 250HHR) that has a higher molecular weight (M = 1.3 x 106) as well as a higher degree of polymerization (NATROSOL-Hercules, 1999), and their shear-thinning behavior in higher concentration solutions is evident. A similar rheological behavior has also been observed by Maestro et al. (2002). Their data for the lower molecular weight HEC9 (M = 9 x 104) show near Newtonian characteristics even at very high concentration (10% by weight), and a nonNewtonian shear-thinning behavior in a 0.75% HEC130 (M = 1.3 x 106) solution. These results are clearly indicative of the role of molecular weight and degree of polymerization in the rheological behavior of polymers. The data for Carbopol 934 solutions in Fig. 2.13, when compared with the respective values for HEC QP-300 in Fig. 2.12, indicate higher viscosity to reflect the increased degree of polymerization (M = 4-6 x 105 and 3 x 106, for HEC and Carbopol, respectively). Also, the shear-thinning behavior for Carbopol 934 solutions is more obvious at higher concentrations. At a shear rate of 500 s-1, atypical of the bubble-fluid motion in nucleate boiling, the relative change in the apparent viscosity of different concentration solutions from that of water for both HEC QP-300 and Carbopol 934 are given in Table 2.3. 57 C/CCMC = 0.5 C/CCMC = 1.0 C/CCMC = 2.0 1.4 1.2 1.0 CTAB η s /η w 0.8 T=23°C 0.6 1.4 C/CCMC = 0.5 C/CCMC = 1.0 C/CCMC = 2.0 1.2 1.0 0.8 Ethoquad 18/25 T=23°C 0.6 101 γ [s-1] 102 Fig. 2.11 Relative viscosity changes of aqueous CTAB and Ethoquad 18/25 solutions 58 10-1 Hu et al. (1991) (HEC 250HHR @22°C) 500 wppm 1000 wppm 2000 wppm η [Pa.s] 10-2 10-3 Present Data (HEC-QP300) @ 23°C 300 wppm 600 wppm 3000 wppm water 10-4 101 102 103 γ [s-1] Fig. 2.12 Variation of apparent viscosity with shear rate for aqueous HEC QP-300 solutions 59 10-1 Present Data (Carbopol 934) @ 23°C 100 wppm 500 wppm 1000 wppm 3000 wppm water η [Pa.s] 10-2 10-3 10-4 101 102 103 γ [s-1] Fig. 2.13 Variation of apparent viscosity with shear rate for aqueous Carbopol 934 solutions Table 2.3 Increase in Viscosity of Aqueous Polymer Solutions with respect to Water at 23°C as a Function of Concentration at a Shear Rate of 500 s-1 HEC Concentration η x 103 [wppm] [Pa.s] Water 0.935 300 1.08 600 1.25 3000 3.65 Carbopol Concentration η x 103 [Pa.s] [wppm] 100 1.03 500 1.43 1000 1.77 3500 4.39 % increase 15.50 33.68 290.37 60 % increase 10.16 52.94 89.30 369.52 CHAPTER 3 POOL BOILING HEAT TRANSFER Nucleate pool boiling experiments and the measured heat transfer performance of aqueous solutions of the four cationic surfactant and two polymer solutions are described in this chapter. The results for different concentration solutions are presented, and the optimum enhancement in heat transfer is identified. 3.1 Experimental Setup The experimental setup used for the pool boiling studies is shown schematically in Fig. 3.1(a). The inner glass tank, which contains the surfactant or polymer solution pool and the cylindrical heater, is encased in an outer glass tank that has circulating mineral oil fed from a constant-temperature recirculating bath (not shown in figure) to maintain the test pool at its saturation temperature. A water-cooled reflux condenser, along with a second coiled-tube water-cooled condenser, helps condense the generated vapor and maintain an atmospheric-pressure pool. A pressure gage (±0.0025 bar precision) is mounted on top of the boiling vessel to monitor the pressure in the pool throughout the experiments. The heating test section (shown in Fig. 3.1b) consists of a horizontal, gold plated, hollow copper cylinder of 22.2 mm outer diameter; the 0.0127 mm thick gold plating mitigates any surface degradation and oxidation from chemicals in the test fluids. A 240 V, 1500 W cartridge heater, with insulated lead wires, is press- fitted in the hollow cylinder with conductive grease to fill any remaining air gaps and provide good heat transfer contact with the inside of the tube. The cartridge heater is centrally located inside the copper tube, and the gaps at each end are filled with silicone 61 rubber to prevent water contact. Also, because the heater surface condition significantly influences the boiling behavior, it was examined using an optical microscope as well as an atomic force microscope (AFM). The optical microscope images of the heated surface in Fig. 3.2 show a random distribution of pits, cavities, and machining grooves of varying shapes and sizes along the heater surface. The overall r.m.s. roughness from AFM scans, measured at four different locations, range from 0.076 µm to 0.347 µm. The heater-wall and pool-bulk temperature measurements were recorded using copper-constantan precision (±0.5ºC) thermocouples, interfaced with a computerized data acquisition system with an in-built ice junction and calibration curve. A variac- controlled AC power supply, a current shunt (0.15 Ω with 1% accuracy), and two highprecision digital multimeters (for current and voltage measurements) provided the controls and measurements of the input electric power. At each incremental value of power input or heat load, the dissipated wall heat flux qw′′ and the wall superheat ∆Tw were computed from the measured values of V, I, the four wall thermocouple readings (Ti,r), and saturation temperature of the pool from the following set of equations: qw′′ = (VI A) (3.1) 4 Tw = ∑ Ti ,r − ( qw′′ ro k ) ln ( ro r ) 4 i =1 (3.2) ∆Tw = (Tw − Tsat ) (3.3) The maximum experimental uncertainties in qw′′ and ∆Tw, based on a propagation of error analysis (Moffatt, 1998), were 1.44% and 0.5% respectively. Details of the experimental procedure, uncertainty analysis, and the extended validation of test measurements with boiling data for water are given by Wasekar and Manglik (2001). 62 Leads for cartridge heater and thermocouple Cooling water Condenser P To circulating bath Leads for Thermocouples Cylindrical heater Inner tank Test fluid pool Mineral oil bath Outer tank Thermocouples for pool temperature (a) (b) Fig. 3.1 Schematic of experimental facility: (a) pool boiling apparatus, and (b) crosssectional view of cylindrical heater assembly 63 100X 200X Fig. 3.2 Optical microscope images of the roughness characteristics of heater surface 64 3.2 Nucleate Pool Boiling in Aqueous Surfactant Solutions Aqueous solutions of DTAC, CTAB, Ethoquad O/12 PG, and Ethoquad 18/25 with different concentrations were prepared by dissolving weighted samples in distilled water. The sample-measurement procedure and corresponding measurement uncertainties are described in detail by Manglik et al. (2001). The boiling of curve for pure distilled water was first established over a period of four months to verify its repeatability and the effects of heater surface aging, and it provides the baseline reference for the surfactant solutions results as well as validates the experimental reliability of the apparatus (Wasekar and Manglik, 2001). 3.2.1 Pool boiling in aqueous cationic surfactant solutions The pool boiling data for different concentration aqueous solutions of DTAC, CTAB, Ethoquad O/12 PG, and Ethoquad 18/25, are presented in Figs. 3.3, 3.4, 3.5 and 3.6, respectively. In general, with the addition of surfactant to water, the nucleate boiling curve shifts to the left indicating enhancement in heat transfer. CTAB and Ethoquad 18/25, typically representative of the four cationics, were analyzed below in details. While CTAB is a higher molecular weight cationic surfactant without EO groups, Ethoquad 18/25 has an even higher molecular weight but with a relatively high ethoxylation of 15 EO groups (Table 2.1). The impact of their different chemistry is clearly seen in the respective nucleate boiling curves for their aqueous solutions. All data graphed in Figs. 3.3 - 3.6 are for decreasing heat flux unless indicated otherwise. The data for CTAB (Fig. 3.4) show considerable heat transfer enhancement with increasing concentration, as represented by the characteristic leftward shift in the boiling 65 curve relative to that for distilled water. Also, there was early incipience or onset of nucleate boiling (ONB) (observed visually with onset of bubbling activity): for C = 400 wppm, ONB was seen at qw′′ ≅ 8.5 kW/m2 or ∆Tw ≅ 3.7 K, as compared to that for distilled water at qw′′ ≅ 12.83 kW/m2 or ∆Tw ≅ 5.13 K. The optimum heat transfer enhancement is seen to be obtained with 400-500 wppm solutions (~ CMC for CTAB at 80°C). But with C > CMC, the enhancement decreases and the heat transfer even deteriorates below that in distilled water in high concentration (≥ 800 wppm) solutions, particularly at low heat fluxes. A similar dependence on C with a different cationic surfactant (Habon G, M = 500) is seen in the Hetsroni et al. (2001) data, as well as those for anionic and nonionic surfactant solutions reported in other studies (Wu, et al. 1998a, Wasekar and Manglik, 2000; Wasekar and Manglik, 2002). The boiling curves for aqueous Ethoquad 18/25 solutions in Fig. 3.6 display a somewhat different behavior, with considerably less enhancement. In fact significant enhancement is seen only at higher heat fluxes, and, once again, the peak performance is with C ~ CMC (~ 500 wppm). With higher concentrations (C > CMC), there is a rightward shift in the boiling curve, and substantially lower heat transfer coefficients than those for distilled water are obtained when C ≥ 3000 wppm. This is also accompanied with delayed incipience and thermal hysteresis or temperature overshoot, as seen in the increasing and decreasing qw′′ - ∆Tw data for 5000 wppm solution; such hysteresis was not seen in lower (C < CMC) concentration solutions. This boiling behavior is akin to that normally observed in highly wetting liquids (Kandlikar et al. 1999; Kenning, 1999; BarCohen, 1992; Bergles, 1988). 66 103 9 8 7 6 5 4 3 q"w (kW/m2) 2 500 w ppm, q"w ⇑ 500 w ppm, q"w ⇓ 2000 w ppm 3000 w ppm 4000 w ppm 5000 w ppm 10000 w ppm, q"w ⇑ 10000 w ppm, q"w ⇓ Distilled Water DTAC Tsat=100°C 102 9 8 7 6 5 4 3 2 Nucleate boiling 101 hysteresis 9 100 4 7 ∆Tw (K) 101 2 Fig. 3.3 Nucleate pool boiling data for aqueous solutions of DTAC; all data are for decreasing heat flux except as otherwise indicated 67 q"w (kW/m2) 103 100 wppm, q"w ⇑ 100 wppm, q"w ⇓ 200 wppm 300 wppm 400 wppm 500 wppm 700 wppm 800 wppm 1000 wppm, q"w ⇑ 1000 wppm, q"w ⇓ CTAB Distilled Water 102 Tsat = 100°C 101 100 4 7 ∆T w (K) 101 2 Fig. 3.4 Nucleate pool boiling data for aqueous solutions of CTAB; all data are for decreasing heat flux except as otherwise indicated 68 1039 8 7 6 5 4 3 2 200 wppm 400 wppm 600 wppm 800 wppm 1000 wppm 1500 wppm 3000 wppm, q"w ⇑ 3000 wppm, q"w ⇓ Ethoquad O/12 PG q"w (kW/m2) Distilled Water Tsat = 100°C 1029 8 7 6 5 4 3 2 1019 100 4 7 ∆T w (K) 101 2 Fig. 3.5 Nucleate pool boiling data for aqueous solutions of Ethoquad O/12 PG; all data are for decreasing heat flux except as otherwise indicated 69 103 200 wppm 500 wppm 700 wppm 1500 wppm 3000 wppm 5000 wppm, q"w ⇑ 5000 wppm, q"w ⇓ Ethoquad 18/25 q"w (kW/m2) Distilled Water Tsat = 100°C 102 101 100 4 7 101 2 ∆T w (K) Fig. 3.6 Nucleate pool boiling data for aqueous solutions of Ethoquad 18/25; all data are for decreasing heat flux except as otherwise indicated 70 3.2.2 Optimum heat transfer and critical micelle concentration The surfactant additive significantly alters the nucleate boiling in water and enhances the heat transfer. A closer inspection of Figs. 3.3 through 3.6 reveals that the enhancement increases with concentration, with an optimum obtained in solutions at or near the critical micelle concentration or CMC of the surfactant. Such an optimum has also been observed by Wasekar and Manglik (2000,2002), Hetsroni et al. (2001), and Wu et al. (1998a). As discussed earlier, the process of micelle formation characterizes this range of concentration (Manglik, et al., 2001; Rosen, 1989; and Tsujii, 1998). The effects of heat flux and surfactant concentration on the nucleate boiling heat transfer are further highlighted in Fig. 3.7, where the relative increase in the heat transfer coefficient from that of water for all four cationic surfactants are graphed. The relative heat transfer coefficient defined as: (h − hwater ) hwater (q ′′ ∆Tw ) − (q w′′ ∆Tw )water = w (q w′′ ∆Tw )water (3.4) Besides depicting the improved heat transfer in solutions with 0 < C ≤ CMC, it clearly shows the decrease in the enhancement in high concentration (C > CMC) solutions. In fact, at low heat fluxes there is even a degradation in heat transfer compared to that for water in all surfactant solutions except in those with DTAC. With a maximum enhancement of 63% in 4000 wppm aqueous DTAC solution, the performance is seen to be dependent upon the wall heat flux, concentration, and surfactant molecular weight and EO group content. The enhancement is significantly greater in aqueous solutions of DTAC and CTAB (non-ethoxylated cationics) as compared to that in Ethoquad O/12 PG 71 and Ethoquad 18/25 (ethoxylated cationics) solutions. As pointed out previously (Wasekar and Manglik, 2001; Wasekar and Manglik, 2002), the process of micelle formation and the molecular dynamics in a concentration sublayer at the vapor-liquid interface characterizes the resultant optimum heat transfer enhancement in surfactant solutions with C ~ CMC. 3.3 Nucleate Pool Boiling in Aqueous Polymer Solutions In order to determine how nucleate pool boiling heat transfer of water is affected by the addition of polymers that have different degrees of polymerization and surfaceactive properties, hydroxyethyl cellulose (HEC-QP300), a nonionic polymer, and Carbopol 934, a cationic polymer, are employed. While both produce viscous aqueous solutions, the former displays significant surface-active properties and the latter renders a shear-thinning rheology in the shear-rate range of interest (10-1000 s-1). Their chemical composition and relevant physico-chemical properties are listed in Table 2.2. Variations in their shear-rate dependent viscosity, along with temperature-dependent equilibrium and dynamic surface tension are recorded and presented in chapter 2, in order to characterize the rheological and interfacial behaviors of the polymeric solutions. Pool boiling curves ( qw" vs. ∆Tsat) for the incipience to fully developed nucleate boiling regimes under atmospheric pressure saturated conditions are presented, which highlight the effects of polymer concentration, dynamic surface tension or surface active, and wall heat flux on boiling and the associated heat transfer coefficients. 72 0.9 0.8 C/Cc.m.c.= 0.5 C/Cc.m.c.= 1 C/Cc.m.c.= 2 DTAC CTAB Ethoquad O/12 PG Ethoquad 18/25 0.7 (hsurf - hwater) / hwater 0.6 Tsat=100°C 0.5 0.4 0.3 0.2 0.1 0.0 -0.1 -0.2 101 q"w (kW/m 2) 102 Fig. 3.7 Variation of the relative heat transfer performance of aqueous cationic surfactant solutions with heat flux and additive concentration (decreasing qw" ) 73 3.3.1 Pool boiling in aqueous polymer solutions The experimental pool boiling data for surface-active HEC solutions of different concentrations as well as that for water are presented in Fig. 3.8. The heat transfer enhancement with increasing surfactant concentration is evident from the leftward shift in the boiling curve relative to that for pure water. This boiling process was further visually observed to have an early incipience or onset of nucleate boiling (ONB). However, the enhanced heat transfer is seen to “peak” with a 600 wppm concentration (~ CPC for HEC-QP300) solution, and then decrease with higher concentrations. The results for 3000 wppm solutions even show a degradation in performance relative to water at lower heat fluxes. Boiling curves for higher concentrations that exhibit a rightward shift were also seen to have delayed incipience as well. A similar performance with HEC-H has been reported by Shul’man et al. (1993), and their results show that the heat transfer coefficient reaches its maximum at C ~ 500 wppm on a plate heater with a size that is much larger than the mean size of the boiling bubbles. This 500 wppm concentration is probably the CPC for HEC-H, which is a different grade of the HEC family of polymers and has lower molecular weight than HEC-QP300. The boiling data for aqueous Carbopol 934 solutions in Fig. 3.9 display a somewhat different behavior, and the heat transfer is seen to continuously continue to decrease with increasing concentrations compared to that for water, with the rightward shift in the boiling curve. Also, delayed incipience, low bubble departure frequency, and some vapor explosions due to higher viscous resistance were observed. Deterioration of boiling heat transfer has also been reported by Paul and Abdel-Khalik (1983) in drag-reducing polyacrylamide (Separan AP30) solutions. 74 103 100 wppm 300 wppm 500 wppm 600 wppm 700 wppm 1000 wppm 3000 wppm HEC-QP300 q"w (kW/m2) Distilled Water Tsat = 100°C 102 101 3 4 5 6 7 ∆T w (K) 8 9 101 2 Fig. 3.8 Nucleate pool boiling data for aqueous solutions of HEC-QP300 75 103 100 wppm 300 wppm 500 wppm 1000 wppm 1500 wppm 3000 wppm Carbopol 934 q"w (kW/m2) Distilled Water Tsat = 100°C 102 101 3 4 5 6 7 8 9 ∆T w (K) 101 2 Fig. 3.9 Nucleate pool boiling data for aqueous solutions of Carbopol 934 76 The effects of heat flux and surfactant concentration on the nucleate boiling heat transfer in HEC-QP300 solutions are further highlighted in Fig. 3.10, where the relative increases in heat transfer coefficients from that of water are graphed for different concentrations. A maximum enhancement of 22.9% in a 600 wppm aqueous solution is seen, and the improved performance tends to be somewhat weakly dependent upon the wall heat flux. Enhanced heat transfer in nucleate boiling of dilute (C < CPC ~ 500 wppm) aqueous HEC-H solutions on a plate heater is also evident from the Kotchaphakdee and Williams (1970) data. Furthermore, Fig. 3.10 clearly shows the decrease in the boiling heat transfer enhancement in HEC solutions with C > CPC (700 wppm, 1000 wppm, and 3000 wppm). In the very high concentration (3000 wppm) solution, up to 7.5% degradation in the heat transfer coefficient when compared to that for pure water is evident for qw′′ < 70 kW/m2; the degradation also tends to be strongly dependent upon wall heat flux. 3.3.2 Surface-active and rheological effects In general, the factors that affect the nucleate boiling performance of polymeric compounds include, among others, the changes in surface tension of the liquid, adsorption of macromolecules on the heating surface, nucleate site density, heater geometry and its surface characteristics, influence of macromolecules on diffusion heat transfer in solvent evaporation, hydrodynamics of convective flows in the boiling boundary layer and bubble motion (micro-convection), thermodynamics features of the polymer-solvent solutions, and rheological effects. At high heat fluxes and a sufficiently large time duration of nucleation boiling, the possibility of macromolecular 77 0.3 HEC-QP300 Tsat = 100°C (hpoly - hwater) / hwater 0.2 0.1 0.0 Present Data -0.1 (Cylindrical heater) Kotchaphakdee & Williams (1970) 100 wppm 300 wppm 500 wppm 600 wppm 700 wppm 1000 wppm 3000 wppm (HEC-H, Plate Heater) -0.2 -0.3 62.5 wppm 125 wppm 250 wppm 101 102 q"w [kW/m2] Fig. 3.10 Variation of the enhanced boiling heat transfer performance of HECQP300 solutions with heat flux and additive concentration 78 thermodestruction (degradation) should also be taken into account (Levitskiy et al., 1996). For surface-active HEC solutions, the reduction in dynamic surface tension σ (which decreases the required superheat for the onset of boiling), and the macromolecular adsorption on the heating surface (which could contribute to the formation of new nucleation sites and increased bubble frequency) are perhaps the two main factors for the boiling heat transfer enhancement in lower concentration (C < CPC) HEC solutions. On the other hand, the decreases in the nucleate boiling heat transfer coefficients in HEC solutions with higher concentrations (C > CPC) and pure shear-thinning Carbopol 934 solutions are possibly associated with the substantial increase in the liquid viscosity that tends to suppress the micro-convection in the bubble boundary layer as well as retard the growth of vapor bubbles. Finally, Fig. 3.11 provides further insights on the role of dynamic surface tension or surface-active effects on the heat transfer performance. The normalized pool boiling heat transfer coefficient data for HEC (300 and 600 wppm) and Carbopol (100 and 300 wppm) solutions are graphed. While their respective concentrations are different, the apparent viscosity of their dilute solutions is comparable (ηHEC, 100wppm, and 300wppm vs. ηCarbopol, ηHEC, 600wppm vs. ηCarbopol, 300wppm). In the measured range of heat fluxes in the nucleate boiling regime, heat transfer enhancement is seen in HEC solutions, while, contrastingly, there is only heat transfer deterioration in Carbopol solutions. Considering that the only drastic physical property change in these four concentration solutions is the dynamic surface tension relaxation5 (dσ/dτ is 0.9 for the 600wppm HEC solution, which 5 The dynamic surface tension gradient (dσ/dτ) used in this study is the value obtained at a surface age of 50 ms, which is representative of bubble frequencies typically encountered in nucleate boiling of water. 79 shows a much larger dynamic surface tension reduction compared to the values for other three solutions), and that the measured surface wettability (represented by the solid-liquid interface contact angle6) for both HEC and Carbopol are close to that of water (77°), these results clearly suggest that the dynamic surface tension is perhaps one of the more significant prediction parameters. 0.6 0.5 (hpoly - hwater) / hwater 0.4 HEC HEC Carbopol Carbopol 300 wppm 600 wppm 100 wppm 300 wppm dσ/dτ η C 1.08e-3 1.25e-3 1.03e-3 1.29e-3 (x 10-2) θ 0.3 0.9 0.1 0.1 76 75 76 77 0.3 0.2 0.1 0.0 -0.1 -0.2 -0.3 -0.4 Cylindrical heater 101 Tsat = 100°C 102 q"w [kW/m2] Fig. 3.11 Effect of dynamic surface tension on the boiling heat transfer coefficient 6 The solid-liquid contact angle for HEC QP-300 and Carbopol 934 measured in this study by the sessile drop method for different concentration solutions, though not presented here, showed insignificant change from those for water (Manglik et al., 2003). 80 CHAPTER 4 VISUALIZATION AND CHARACTERIZATION OF NUCLEATE POOL BOILING IN AQUEOUS SURFACTANT SOLUTIONS 4.1 Introduction Nucleate boiling in aqueous surfactant or polymer solutions is a complex conjugate process as illustrated earlier in Fig. 1.2. It depends on surfactant concentration, wall heat flux level, surfactant chemistry, dynamic surface tension, surface wettability and nucleation cavity distribution, and marangoni convection, heater geometry and its surface characteristics, and rheological properties of the solutions, etc. The nucleation is a process in which finite size clusters of molecules encompassing properties of the second phase appear in the host liquid (Brennen, 1995), and generally subdivided into two categories: homogeneous nucleation and heterogeneous nucleation. This study is only related to heterogeneous nucleation – a process in which bubbles form discretely at pits, scratches, and grooves on a heated surface submerged in a pool of liquid. The onset of nucleate boiling on a heater submerged in a pool of liquid is characterized by the appearance of vapor bubbles at discrete locations on the heater surface (Kenning, 1999; Dhir, 1998). Surface finish, surface wettability, heater geometry, surface contamination, system pressure, liquid subcooling, gravity, heat flux modes (steady state or transient) are considered to have a significant influence. However, the primary heat transfer is by evaporation and its efficiency is directly related to nucleation site density and bubble dynamics, and phenomenological insights can be obtained from a visual observation of the ebullience. A knowledge of nucleate site density as a function of wettability and wall superheat is critical in order to develop a 81 credible model for predication of nucleate boiling. The latter includes the processes of bubble growth, bubble departure, and bubble waiting time (reformation of the thermal layer). For phenomenological study of nucleate boiling in aqueous surfactant or polymer solutions, visualization is an important approach to investigate how the active nucleation density changes due to surfactant adsorption at the solid-liquid interface, and how the reduced dynamic surface tension affects the bubble dynamics. The boiling heat transfer performance can be quantitatively related to its ebullience (nucleation, inception and gestation → growth → departure). The growth of nucleating vapor bubbles and their motion near the cylindrical heater surface were recorded by a PULNiX TMC-7 highspeed color CCD camera with shutter speeds of up to 0.1 micro-second. The CCD camera is interfaced with a PC through a FLASHBUS MV Pro image capture kit that has high-speed PCI-based bus-mastering capabilities (up to 132 Mbytes/s). It delivers consecutive frames of video in real time into the system memory while keeping the CPU free to operate on other applications. Furthermore a FUJI 12.5-75mm micro lens was used on the CCD camera to facilitate high quality close-up photography. 4.2 Zeta Potential and Contact Angle The surface wettability can be characterized by zeta potential ζ (an electrokinetic control parameter for the stability of hydrophobic colloids), which shows distinct regions of change along the adsorption isotherm that are associated with the aggregation mode of adsorbed ions at the solid-water interface. The higher wettability tends to suppress nucleation and bubble growth, thereby weakening the boiling process. 82 Zeta potential and contact angle measurements complement each other and allow the understanding of electrokinetic characteristics and the surface wetting behavior due to the physisorption of surfactant molecules at the solid-liquid interface. The change in zeta potential and measured surface wettability (represented by the contact angle) with concentration in aqueous SDS (anionic) and CTAB (cationic) are graphed in Fig. 4.1, depicting the distinct regions of change in adsorption and the corresponding wettability variation. The change in wettability with the adsorption of ionic surfactants with concentration, and its direct correlation with zeta potential is clearly evident. SDS shows a stronger adsorption than CTAB at the liquid-solid surface, which is reflected in the magnitude of zeta potential and the larger changes in surface wettability. After the point of zeta potential reversal (PZR) (Lyklema et al., 1991; Fuerstenau, 2002), or isoelectric point (IEP) (Hunter, 1981), the slope of the ζ potential curve becomes negative for the anionic surfactant SDS or positive for the cationic surfactant CTAB because of the opposite charges they carry, and suggests that some of the adsorption may start taking place in reverse orientation to form reverse hemimicelles (Fuerstenau, 2002), where the surface becomes increasing hydrophilic, which is evident from the contact angle measurement as shown in Fig. 4.1. As CMC is approached, a bilayer is formed, the contact angle tends to be constant, and surface becomes highly hydrophilic. 83 101 90 102 103 80 -30 70 60 -10 PZR 50 Zeta Potential [mv] Contact angle [deg] -20 0 40 30 20 101 Contact angle, CTAB Contact angle, SDS Zeta potential (Vanjara and Dixit, 1996), CTAB Zeta potential (Sakagami et al. 2002), SDS 102 103 10 20 104 Concentration [wppm] Fig. 4.1 Measured streaming zeta potential and contact angle for the adsorption of SDS and CTAB in their aqueous solutions 84 4.3 Ethoxylation Effect The presence of ethoxy or ethylene oxide (EO) group in their hydrocarbon chain (ethoxylation) increases the overall size of the polar head and makes the surfactant more hydrophilic. Consequently, it will alter the solution’s surface wettability drastically (Barry and Wilson, 1978; Ashayer et al. 2000), and this is clearly seen from the contact angle measurements for CTAB and Ethoquad 18/257 graphed in Fig. 4.2. The later has significantly lower contact angles with increasing C, particularly when C ≥ CMC. This is a direct consequence of the surfactant chemistry and its physisorption dynamics (Fuerstenau, 2002), and the altered surface wettability probably accounts for the boiling deterioration in Ethoquad O/12 PG and Ethoquad 18/25 sloutions at C > CMC concentrations as shown in Fig. 3.7. Contact angle [deg] 101 80 102 103 70 c.m.c. 60 CTAB Ethoquad 18/25 50 101 102 c.m.c. 103 104 Concentration [wppm] Fig. 4.2 Measured contact angle for aqueous CTAB and Ethoquad 8/25 solutions 7 Both CTAB and Ethoquad 18/25 are cationic and comparable. While CTAB has a higher molecular weight, Ethoquad 18/25 has an even higher M but with a relatively high ethoxylation of 15 EO groups. Surfactant with ethoxylation doesn’t necessarily mean it would have higher wettability. Physisorption at the solid-liquid interface is a complex process, which depends on the surfactant structure, chemical and physical properties, and surface characteristics etc. Electrokinetics (zeta potential) and adsorption isotherm are more direct indictors for the surfactant adsorption process. 85 4.4 Dynamic Surface Tension – Molecular Weight Effect When a new surface in a surfactant solution is created, a finite time is required to reach equilibrium state between the surface concentration and the bulk concentration. This non-equilibrium surface tension is called dynamic σ. The surface tension relaxation is a diffusion-rate dependent process, the time required to reach equilibrium varies from a few milliseconds to a few hours, and is typically found to depend on the type of surfactants, its diffusion-adsorption kinetics, micellar dynamics, ethoxylation, and bulk concentration levels (Manglik et al., 2001). The time scale for complete surface tension relaxation tends to be smallest for lower molecular weight compounds (Iliev et al., 1992). The effectiveness of an additive is often judged by its ability to lower a liquid's static surface tension at the lowest possible concentration. Equally important, the additive should not cause undesirable side effects, such as interference with solid interface adhesion, or increased tendency to foam. For boiling applications with small surface age interface (10-100 ms, Prosperetti and Plesset, 1978), however, the dynamic surface tension relaxation process rather than the equilibrium surface tension is perhaps the more critical determinant (Wasekar and Manglik, 2002). A lower molecular weight surfactant diffuses faster than its higher molecular counterpart, and this is seen in the faster σ relaxation of CTAB (cationic) in comparison with that for Ethoquad (cationic) 18/25 in Fig. 2.4. Anionics (SDS and SLES) and nonionics (Triton X-100 and Triton X-305) show the same dynamic surface tension behavior with molecular weight as present in Figs. 4.3 and 4.4. It should be noted that the molecular weight effect might not necessarily be the same for different ionic categories of surfactants. In most surfactant solutions, the time scale to reach the 86 equilibrium value (total relaxation) at a newly-created interface is of the same order as that of bubble formation and departure in nucleate boiling (0-100 ms). 70 SLES SDS C/Cc.m.c. = 0.5 C/Cc.m.c. = 1 C/Cc.m.c. = 2.0 σ [mN/m] 60 50 40 Typical bubble frequencies in nucleate boiling of water 30 10-2 10-1 100 101 Surface Age [s] Fig. 4.3 Dynamic surface tension relaxation for aqueous anionic SDS and SLES solutions 87 70 Trotion X-305 Triton X-100 C/Cc.m.c. = 0.5 C/Cc.m.c. = 1 C/Cc.m.c. = 2.0 σ [mN/m] 60 50 40 30 20 10-2 Typical bubble frequencies in nucleate boiling of water 10-1 100 101 Surface Age [s] Fig. 4.4 Dynamic surface tension relaxation for aqueous nonionic Triton X-100 and Triton X-305 solutions Figure 4.5 provides insights on the role of surfactant molecular weight, ethoxylation, and dynamic surface tension of their solutions on the heat transfer enhancement. The normalized pool boiling heat transfer coefficient data for cationic DTAC (4000 wppm), CTAB (400 wppm), Ethoquad O/12 PG (600 wppm), and Ethoquad 18/25 (700 wppm) solutions are graphed. While their respective concentration is different, the dynamic surface tension value (nominally representative of nucleate boiling bubble frequencies) of their aqueous solutions at 80ºC is the same (~ 47 mN/m) in each case. In the measured range of heat fluxes, the heat transfer enhancement is seen to be in 88 the order of DTAC > CTAB > Ethoquad O/12 PG > Ethoquad 18/25, which is in the reverse order of their respective molecular weights and number of EO groups. It also shows a boiling deterioration at low heat flux ( qw′′ < 50 kW/m2) for Ethoquad 18/25. Within the typical time transients for bubble growth in nucleate boiling of surfactant solutions, the faster diffusion of lower molecular weight surfactants (higher mobility) leads to a larger number of surfactant molecules approaching the growing bubble interface. They, therefore, reduce the surface tension faster (dσ/dτ8) and increase the bubble growth and departure frequencies to yield better heat transfer performance. Also, in this dynamic ebullient and additive adsorption process, a measure of the dynamic surface tension is the more appropriate scaling property instead of a static value at a fixed bubble frequency. The normalized pool boiling heat transfer coefficient data for anionic (SDS and SLES) and nonionic (Triton X-100 and Triton X-305) solutions (Wasekar and Manglik, 2002) are also presented here in Fig. 4.6. The lower qw′′ range, which represents the partially developed nucleate boiling regime, is characterized by the thermal hysterisis and wetting behavior of the surfactant solution. At higher heat fluxes, typically in the fully developed nucleate boiling regime ( qw′′ > 150 kW/m2), however, the maximum heat transfer enhancement is in the order of SDS > SLES > Triton X-1009 > Triton X-305. This is also, in the reverse order of their respective molecular weights and number of EO groups. These results therefore suggest that the surfactant molecular weight, and dynamic surface tension relaxation instead of the static value are critical parameters in predicting 8 9 The dynamic surface tension gradients (dσ/dτ) used here were also obtained at a surface age of 50 ms. Triton X-100 is highly surface active and depresses σ very quickly; however, it also has 9-10 EO groups. It is the combination of dynamic surface tension and surface wettablity effects that determine the heat transfer performance. 89 their enhanced nucleate pool boiling heat transfer performance under the given operating conditions. 1.0 C EO σ50ms,80°C dσ/d-2τ (x 10 ) 0 47.1 3.6 0 46.5 3.1 2 47.0 2.5 15 46.7 1.6 M DTAC 4000 263.9 CTAB 400 364.5 Ethoquad O/12 600 403 Ethoquad 18/25 700 994 0.8 (hsurf - hwater) / hwater 0.6 0.4 0.2 0.0 Fully developed nucleate boiling -0.2 0 50 100 150 " 200 250 2 q w (kW/m ) Fig. 4.5 Effect of surfactant molecular weight and its ethoxylation on the heat transfer coefficient enhancement 90 1.0 C 0.9 EO σ50ms,80°C dσ/d-2τ (x 10 ) SDS 2500 288.3 0 SLES 1750 422 3 Triton X-100 200 624 9 Triton X-305 750 1526 30 0.8 (hsurf - hwater) / hwater M 42.8 43.1 45.1 47.5 3.5 2.1 2.9 2.1 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 50 100 150 200 250 q"w (kW/m2) Fig. 4.6 Surfactant molecular weight dependence of the maximum enhancement in heat transfer coefficient enhancement (Wasekar and Manglik, 2001) 91 4.5 Surface Wettability Effect on Nucleate Boiling Heat Transfer With an increase in surface wettability, the nucleate boiling curves tend to shift to the right. This is seen in Fig. 4.7(a) where the saturated nucleate boiling data of Liaw and Dhir (1989), obtained on vertical surfaces having different static contact angles, are graphed. These data are for the fully developed nucleate boiling regime, and the data of Nishikawa et al. (1984) for a vertical surface are also included. It is noted that although the boiling curves shift to the right with an increase in the wettability (decreasing contact angle), the qw" − ∆T gradient remains nearly unchanged. Another interesting point is that critical heat flux (CHF) increases as the surface wettability improves. For liquids that wet the surface well (contact angle → 0), a temperature overshoot or hysteresis is observed in the ONB region of the boiling curve as shown in Fig. 4.7(b) for nucleate pool boiling of a highly-wetting liquid R113. For these liquids, the convective heat removal process continues to persist up to a higher superheat; upon incepience, a large reduction in superheat is observed, and it only occurs with increasing heat flux (Dhir, 1998, Kenning, 1999). A closer inspection of Figs. 3.3 through 3.6 shows that boiling in aqueous surfactant solutions when C ≥ CMC exhibits a similar behavior to that in Fig. 4.7. The temperature overshoot or hysteresis was also observed when C ≥ CMC, especially in the highly wetting aqueous Ethoquad 18/25 solutions. From this observation, one can deduce that there is indeed a wettability change considering that the viscosity and thermal conductivity remains almost unchanged in dilute surfactant solutions. This is further clearly established by the contact angle measurements presented in Fig. 2.9. 92 105 100 Water CHF q"w ⇓ q"w ⇑ q"w (W/m2) qw" (W/cm2) 104 50 θ = 90° 3.3 θ = 69° 103 1 θ = 27° θ = 38° θ = 14° Nishikawa et al. (1984), θ = 90° 20 10 15 20 30 102 10-1 40 ∆Τ (Κ) 100 ∆T (K) 101 Fig. 4.7 (a) Effect surface wettability (or contact angle θ) on nucleate pool boiling (Liaw and Dhir, 1989) (b) Nucleate pool boiling data for Refrigerant of R113 on a copper tube (Jung and Bergles, 1989) 93 102 4.6 Ebullient Dynamics Visualization The enhanced boiling performance can be related to the ebullient characteristics. The photographic records for different concentration aqueous solutions of CTAB and Ethoquad 18/25, SDS, Triton X-305 (typically representative of the four cationics, anionics, and nonionics), HEC-QP300 (surface-active polymer), and Carbopol 934 (shear-thinning polymer), are presented in Figs. 4.8, 4.9 and 4.10, respectively. 4.6.1 Visualization in aqueous surfactant solutions Figure 4.8 represents the boiling behavior in water, and CTAB and Ethoquad 18/25 solutions (non-ethoxylated and ethoxylated cationic surfactants, with number of EO groups = 0 and 15, respectively) of three different concentrations (C/CCMC = 0.5, 1, and 2) at two different heat flux levels ( qw′′ = 20, and 50 kW/m2). In comparison to that in water, boiling in CTAB solutions is more vigorous and is characterized by clusters of smaller-sized, more regularly shaped bubbles that originate at the underside of the cylindrical heater. These bubbles then slide along the heater surface at departure, thereby knocking off much smaller bubbles growing on the top surface of the heater. This process was observed to increase with heat flux and the consequent higher bubble departure frequency. Also, because of the considerable reduction in σ for CTAB solutions, much smaller-sized bubbles are nucleated in a cluster of active nucleation sites, especially at lower heat fluxes. They have a significantly higher bubble departure frequency, with virtually no coalescence of either the neighboring or sliding bubbles that come in contact with others around the heater’s periphery when C < CMC. However, when C ≥ CMC, some foaming patches begin to occur, the liquid only coverage of the 94 heater surface increases, and slightly larger bubbles are formed, all of which indicate a surface wetting condition change. A foam layer, whose thickness increases with heat flux, is also seen to form at the free surface of the pool. This boiling history is very similar to that also seen in anionic and nonionic surfactant solutions (Wasekar and Manglik, 2002; Zhang and Manglik, 2003). Boiling in Ethoquad 18/25, on the other hand, shows much smaller-sized bubbles in pre-CMC solutions, and considerably fewer and larger-sized bubbles are formed with increasing concentrations. This presents a contrastingly different behavior from not only that of water but also that of CTAB, and is perhaps due to the surfactant’s high degree of ethoxylation. The presence of large number of EO groups in their hydrocarbon chains totally changes the surface wettability at the solid-water interface. This is quite evident from the measured contact angle data presented in Fig. 4.2. The observed ebullience and boiling data cannot be explained by the reduced dynamic surface tension alone. If this were so, then smallest-sized bubbles would be seen in C ≥ CMC solutions, where the surface tension reaches the lowest possible value. Instead, because of the adsorption of surfactants and their different orientations in the adsorption layer (Fig. 2.9b), the heater surface wettability increases with increasing concentration. Fewer bubbles are thus nucleated that have relatively larger departure diameters. As pointed out by Fuerstenau (2002), the adsorption isotherm and corresponding surface state can be divided into four regimes that are associated with the aggregation mode of adsorbed ions at the solid-water interface: (1) at low concentrations, 95 surfactant adsorption takes place as individual ions10; (2) there is a sharp increase in the adsorption density due to self-association of adsorbed surfactant ions and the formation of hemimicelles; (3) the surfactant ions adsorb as reverse hemimicelles, with their polar heads oriented both toward the surface and liquid, and the surface becomes increasingly hydrophilic; and (4) as the CMC is approached, the adsorption becomes independent of the bulk concentration in solution, and the surfactant molecules form a bilayer on the surface to make it strongly hydrophilic. The measured contact angle data in Fig. 2.9(a) correlate well with this characterization schematically illustrated in Fig. 2.9(b). The reason that Ethoquad 18/28 shows more hydrophilic behavior that CTAB at lower concentrations is because of its bulky polar head that occupies larger portion of the heater surface, and the variation in ebullience at the heater surface (Fig. 4.8) also reflects this wettability change. The ebullience of SDS and Triton X-305 (non-ethoxylated anionic and ethoxylated nonionic surfactants, with number of EO groups = 0 and 30, respectively) as presented in Fig. 4.9 displayed a similar behavior with degree of ethoxylation and varying surfactant concentration and hence are not discussed separately. 10 Adsorption takes place with polar heads of the surfactant ions oriented toward the surface that yields a hydrophobic surface with most surfactants except for high molecular weight ones, whose bulky polar head would occupy a larger surface area. 96 Water Water 22.2 mm CTAB CTAB Ethoquad 18/25 Ethoquad 18/25 C/CCMC = 0.5 C/CCMC = 0.5 C/CCMC = 1 C/CCMC = 1 C/CCMC = 2 C/CCMC = 2 kW/m2 50 kW/m2 20 Fig. 4.8 Ebullient behavior in nucleate boiling of distilled water, and aqueous cationic CTAB and Ethoquad 18/25 solutions of different concentrations (C/CCMC = 0.5, 1, and 2) at qw" = 20 kW/m2 and 50 kW/m2 97 Water Water 22.2 mm SDS SDS TRITON X-305 TRITON X-305 C/CCMC = 0.5 C/CCMC = 0.5 C/CCMC = 1 C/CCMC = 1 C/CCMC = 2 C/CCMC = 2 kW/m2 50 kW/m2 20 Fig. 4.9 Ebullient behavior in nucleate boiling of distilled water, and aqueous SDS (anionic) and Triton X-305 (nonionic) solutions of different concentrations (C/CCMC = 0.5, 1, and 2) at qw" = 20 kW/m2 and 50 kW/m2 98 4.6.2 Visualization in aqueous polymer solutions The nucleate boiling performance of HEC and Carbopol solutions can also be qualitatively related to the ebullient characteristics, and typical photographic records are presented in Fig. 4.10. The photographs depict the boiling history with increasing heat flux ( qw′′ = 20, 50, and 100 kW/m2) for aqueous solutions of two different concentrations each (300 wppm and 1000 wppm for HEC, 100 wppm and 300 wppm for Carbopol11), along with that for deionized distilled water. Boiling and the attendant bubbling process in HEC solutions are seen to be distinctly different from that in pure water. It is more vigorous, and is characterized by smaller-sized and more regularly shaped bubbles that have a reduced tendency for coalescence when C < CPC. It is also noted that the HEC additive leads to an early inception of bubbles with a faster covering of the heating surface and a higher bubble departure frequency. This is essentially the outcome of reduced surface tension at the liquid-vapor interface. Also, molecular adsorption on the heating surface may contribute to the formation of new sites (Levitskiy et al., 1996), which in turn would explain the increase in number of bubbles as there was no change in the surface wettability (measured by the contact angle) (Manglik et al., 2003). However, at very large concentration (C > CPC), the bubbles that originate at the underside of the cylindrical heater tend to coalescence and form bigger bubbles as they slide along the cylindrical periphery of the heater surface at departure. At the same time, there are some small patches that are covered by liquid, and no bubbles formed underneath these patches. This phenomenon is totally different from that in boiling of surfactant solutions, in which surface wetting condition significantly changes or foaming begins to occur 11 In Carbopol 934 solutions, the boiling pool became significantly more cloudy with increasing concentration and it was very difficult to get clear photographs when C > 300 wppm. Hence, snapshots for a higher concentration solution are not presented. 99 when C > CMC. Perhaps the significantly increased viscosity of higher concentration polymer solutions tends to suppress the bubble nucleation process and growth of vapor bubbles. As a result, some nuclei do not get activated at all, and this also then leads to the deterioration of boiling performance in aqueous solutions with C > CPC. Boiling in Carbopol 934 solutions, on the other hand, shows a totally different ebullient behavior than that of water, as well as that of surface-active HEC solutions. Considerable bubble suppression is observed in Carbopol 934 solutions, as well as dispersed vapor explosions (bright white spots captured in the picture; Fig. 4.10) in some regions of the heater surface. Same kind of bubbling activity was also observed by Bang et al. (1997) in dilute polyethylene oxide (PEO) solutions. Furthermore, the Carbopol additive in water lead to a delayed inception of bubbles, or ONB, and the sparsely formed bubbles have a slower departure frequency. This is essentially the outcome of increased viscosity in Carbopol solutions, which gives a higher viscous resistance for the bubbles to nucleate and depart. Again, contrasting the boiling ebullience with that of pure water, as well as between HEC and Carbopol solutions, there appear to be more nucleation sites activated in aqueous HEC QP-300 solutions than that with Carbopol 934. The higher active nucleation site density would also explain the increased heat transfer performance in surface-active HEC solutions. 100 Water HEC 300 wppm HEC 1000 wppm Carbopol 100 wppm Carbopol 300 wppm 20 kW/m2 50 kW/m2 100 kW/m2 Fig. 4.10 Ebullient behavior in nucleate boiling of distilled water, and aqueous HECQP300 and Carbopol 934 solutions of different concentrations at different heat fluxes ( qw" = 20 kW/m2, 50 kW/m2, and 100 KW/m2) 101 4.7 Characterization of Nucleate Pool Boiling in Aqueous Surfactant Solutions A successful correlation always depends on the recognition of the problem (Mordecai, 1930). For boiling in pure liquids, a variety of different parameters and mechanisms have been proposed to describe the complex behavior (Dhir, 1998; Nelson et al., 1996). However, the potential mechanisms that affect nucleate boiling in aqueous surfactant solutions are not clearly established; particularly, the adsorption processes at liquid-vapor and solid-liquid interfaces perhaps require an in-depth scaling of the various observed mechanism. A direct correlation of the heat transfer with suitable descriptive parameters for all effects is difficult because of the complex nature of the problem. A few recent attempts have been made (Wen and Wang, 2002; Sher and Hetsroni, 2002; Wasekar and Manglik, 2002) to correlate the nucleate boiling heat transfer for aqueous surfactant solutions, but the adopted methodologies have significant limitations, and their general applicability is not established. Wen and Wang (2002) modified the pure-fluid Mikic-Rohsenow (1969) correlation with the available contact angle and surface tension data to represent their own boiling data. However, equilibrium surface tension at room temperature, rather than the dynamic effects, and that too without stipulating the level of concentration has been used in their analysis. Because a single concentration value seems to be the implicit guiding factor in its applicability, a generalization of the correlation is circumscribed. Sher and Hetsroni (2002) considered a surfactant diffusion mechanism for both the liquid-vapor and solid-liquid surface tensions, and proposed a model based on the Rohsenow (1952) correlation. It gave an acceptable agreement with their own experimental data for pre-CMC solutions, and that too for only one surfactant (Habon G). 102 No data for post-CMC solutions were presented, and it was indicated that the diffusion and adsorption mechanisms might be altered in such a case. It may be noted that at low concentrations the adsorption process is generally in the monolayer region, which will not show drastic wetting changes on the heater surface (Fuerstenau, 2002). Wettability is essentially governed by the surfactant concentration and its adsorption process at the solid-liquid interface. Also, nucleate boiling heat transfer is influenced by variations in the surfactant chemistry. Wasekar and Manglik (2002) considered the effect of surfactant ionic nature and molecular weight (exponent of -0.5 for the ratio of molecular weights for anionics and 0 for the nonionics when C < CMC at fully developed boiling regime) on the pool boiling performance of water, and pointed out that dynamic surface tension is perhaps a critical predictor of the ebullient phase-change behavior. However, rather simplistically, the value of σ at a fixed bubble frequency or surface age τ was adopted to represent the dynamic surface tension effects. This study explores the role of surfactant adsorption and interfacial phenomena as illustrated in Fig. 4.11, manifest in the dynamic surface tension relaxation due to the dynamic adsorption-desorption process at the liquidvapor interface, and the surface wettability changes due to surfactant physisorption at the solid-liquid interface, on nucleate boiling heat transfer. The nucleate boiling heat transfer depends on the nucleation, bubble growth rates, bubble departure frequencies and sizes, and surface wettability. With the addition of small amounts of surfactants, the saturated nucleate pool boiling of water on a cylindrical heater is altered significantly. The heat transfer is enhanced in solutions with C ≤ CMC, but decreases when C > CMC. In general, besides the heat flux (or wall superheat) levels, the relative extent of performance change is seen to be governed by the surfactant 103 C Liquid-vapor interface Liquid-solid interface q" (a) Schematic of a nucleated bubble Individual Ions Reversed Hemimicelles Hemimicelles Hydrophobic tail Bilayer Hydrophilic head Hydrophobic Hydrophilic C (b) Adsorption at the solid-liquid interface C (c) Adsorption at the liquid-vapor interface Fig. 4.11 Schematic of interfacial phenomena in aqueous surfactant solutions (not to scale) 104 interfacial phenomena at both the liquid-vapor and solid-liquid interfaces. These in turn are determined by several additive physico-chemistry-based factors, and their rather complex inter-relationships are characterized in Fig. 4.12. Some key factors are analyzed in details as following: 1. Equilibrium and dynamic surface tension Surface tension is an important variable for nucleate boiling. The nucleate pool boiling heat transfer coefficient h is related to the equilibrium surface tension σ of the pure liquids by the following (Rohsenow, 1952) h ∝ 1/σ 3 (4.1) and the bubble departure diameter was proposed by Fritz (1935) as D ∝ σ 1/2 (4.2) However, equilibrium surface tension could be deceiving when applied to nucleate boiling heat transfer in aqueous surfactant solutions. When a new surface is created in a surfactant solution, a finite time is required for the reagent adsorption to reach an equilibrium state between the surface concentration and bulk concentration. This time-dependent surfactant adsorption at the vapor-liquid interface of a bubble gives rise to the dynamic surface tension (DST) behavior, which, however, eventually reduces to the equilibrium condition after a long time period (Holmberg et al., 2003; Chang and Franses, 1995; Rosen, 1989). For boiling applications, the formation and departure of bubbles is also a highly dynamic process where the vaporliquid interface has a small surface age; typically in the range of 0-100 ms (Prosperetti and Plesset, 1978). The dynamic surface tension relaxation rather than the equilibrium or static surface tension, therefore, becomes the more critical determinant. 105 Bubble growth rate and departure Marangoni convection Bubble Size Nucleation Liquid-vapor interface adsorption/desportion Physico-Chemical Properties Dynamic Surface Tension* •Ionic Nature (Liquid-vapor, bulk chemistry) •Molecular structure •Ethoxylation Contact Angle/Wettability (Equilibrium and dynamic) Microlayer thickness Heat flux removal rate Wetting/spreading Surfactant •Molecular Weight •Micelle •Packing Zeta Potential (Physisorption)* (Solid-liquid, surface chemistry) Other factors •Heater size and its characteristics Active nucleation site density •Heat flux level (superheat) Bubble nucleation rate and size •Fluid transport properties Bubble collapse and merge •Far surface features and foaming Boiling hysteresis •Physical properties Fig. 4.12 Surfactant effects on nucleate boiling heat transfer in its aqueous solutions 106 When a bubble starts to form at a nucleating site, the initial surface tension is that of water and it then reduces continually with time τ as more and more surfactant molecules move to the interface during the adsorption process. At the same time, desorption may also occur till an equilibrium condition is reached and surface tension becomes constant. This time-dependent adsorption/desorption process is schematically illustrated in Fig. 4.13(a), and is reflected in the measured dynamic surface tension behavior seen in Figs. 4.3 and 4.4. The effects of dynamic surface tension on bubble formation and departure are more clearly demonstrated in Fig. 4.13(b), where results of a controlled adiabatic single-bubble experiment are presented. Photographic images of a near-departure air bubble, captured by a high-speed (2000 frames/sec) camera under identical operating conditions, are presented. Constant-pressure bubbling activity in the following three liquids was recorded: water (pure liquid, σeq = 72.3), 2500 wppm SDS solution (surfactant solution, σeq = 37.5 at CMC), and N,N-Dimenthyl Formamide or DMF (pure liquid, σeq = 37.1). The bubble surface age was controlled at ~100ms, which is of the same order as the time scales for the dynamic adsorption/desorption process in aqueous surfactant solutions and ebullience in nucleate boiling. Dynamic σ effects are selfevident in Fig. 4.13(b), where a larger bubble is seen in aqueous SDS solution when compared to that in a pure liquid (DMF) that has the same equilibrium σ value (~ 37 mN/m). This is essentially due to the time-dependent surfactant adsorption/desorption process at the liquid-vapor interface. 107 τw τg (a) Water (pure-liquid) (σ = 72.3) SDS solution (σ = 37.5) DMF (pure-liquid) (σ = 37.1) (b) Fig. 4.13 (a) Schematic of surfactant transport process during a bubble formation and departure (not to scale); (b) Dynamic surface tension effect on bubble dynamics (evolution of pre-departure shape and size). 108 2. Surface wettability (contact angle) One critical determinant of nucleate boiling is the active nucleation density, which is a function of wettability and directly accounts for the energy transfer by ebullience at the heater surface. Surface wettability is usually measured by the solid-liquid equilibrium contact angle. On a typical large heater surface, the fraction of nucleated cavities decreases as the surface wettability increases (Dhir, 1998). Also, the volume of trapped air or vapor in a cavity, incipient superheat, boiling incipience, hysteresis, critical bubble radius, and departure bubble size are strongly influenced by surface wettability. In the case of aqueous surfactant solutions, the reagent will undergo a certain adsorption pattern (different orientation of the surfactant molecule head or tail) with changing concentration at the solid-liquid interface. Consequently, the heater surface will show different wetting behavior at different adsorption stages because of the unique structure of a surfactant – a hydrophilic head with a hydrophobic tail. 3. Microlayer There is increasing evidence with a growing body of literature to support the hypothesis that a thin layer of liquid (the microlayer) forms beneath a vapor bubble, and that it accounts for most of the wall heat transfer during saturated nucleate boiling process (Cooper and Lloyd, 1968; Lee and Nydahl, 1989; Lay and Dhir, 1995; Zhao et al., 2002). With surfactant additives in water, the liquid-vapor and solid-liquid surface tension are altered, which changes the energy balance at the solid-liquid-vapor threephase contact line. Therefore, the dynamic (advancing and receding) contact angle and microlayer thickness will be different in aqueous surfactant solutions compared to that of pure liquid. The consequent heat flux removal mechanism and the role of the microlayer 109 under these conditions are yet to be developed. 4. Marangoni effect Variation of the temperature distribution at a vapor-liquid interface results in local surface tension gradients along the interface. This in turn produces displacements in the interface in the direction of increasing surface tension, and the phenomenon is commonly referred to as Marangoni effect (Named after Italian physicist Carlo Marangoni) (Scriven and Sternling, 1960). Similarly in aqueous surfactant solutions, local variations in the reagent concentration produce a similar convective effect. Marangoni convection is not only important to the bubble incipience and waiting periods, but also plays a role in the reduced bubble coalescence of boiling in aqueous surfactant solutions (Yang, 1990). The effects of surfactant concentration on the initial short-time-scale Marangoni convection around boiling nuclei in aqueous solutions have been computationally investigated by Wasekar and Manglik (2003). Their model consists of an adiabatic, rigid, hemispherical bubble on a downward facing constant temperature heated wall, in a fluid pool with an initial uniform temperature gradient. With a surfactant present in solution, a surface concentration gradient develops at the bubble interface that will promote diffusocapillary flows along with the temperature-gradient induced thermocapillary flows. However, the Wasekar and Manglik (2003) model is rather constrained and simplistic. The complex interactions of both thermo- and diffuso- capillary Marangoni convection, the excess surfactant concentration due to the physisorption at the solid-liquid interface, mass transfer at the liquid-vapor interface and its development, adjacent bubble effect, and the convection in the bulk solution caused by the bubble growth and departure, are yet to be understood and modeled. 110 5. Nuclei formation Trace amounts of foreign particles in liquids known to affect nucleation process. Though there is no convincing evidence, the additives may agglomerate in solutions and form particle large enough to serve as boiling nuclei. This speculation may be especially valid for long chain polymeric surfactants or polymers in aqueous solutions. It is desirable that the results of all surfactant effects on the ebullient heat transfer can be analyzed and formulated into one convenient equation form for practical design applications. A successful correlation of nucleate boiling data for aqueous surfactant solutions, however, should include the complete interfacial phenomena. That is not only the adsorption/desorption process at the liquid-vapor interface, but also the physisorption at the solid-liquid contact. Excluding one or the other would be an oversimplification of this complex process. Clearly further systematic research is necessary to establish these and other surfactant effects on pool boiling heat transfer before a generalized model can be developed. 111 CHAPETR 5 SIMULATION OF A SINGLE BUBBLE DYNAMICS 5.1 Introduction Several mechanistic and semi-mechanistic models have been developed to facilitate the understanding of boiling behavior and provide performance prediction tools. A critical element of this work is the simulation of the single-boiling-bubble dynamics, in order to extend the fundamental understanding of the phase-change process. However, because of the complexity involved in modeling the continuously evolving vapor-liquid interfaces in pure liquids, greatly simplifying assumptions are often made in developing various models (Shyy et al. 1996). This is further complicated by the need to include the adsorption-desorption controlled interfacial surfactant transport for ebullience in surfactant solutions. Over the past decade, some studies have expanded the complexity of nucleating bubble-dynamics models, by implementing a direct numerical tracking of the changing liquid-vapor interface. The finite-volume method (FVM) (Welch, 1998), volume of fluid (VOF) method (Welch and Wilson, 2000), level-set method (LSM) (Son et al., 1999), and arbitrary Lagrangian-Eulerian (ALE) method (Yoon et al., 2001) have been typically used. Some recent work (Takada et al., 2000; Palmer and Rector, 2000) has also considered the Lattice-Boltzmann method (LBM). Plesset and Zwick (1954) were probably the first to present a one-dimensional analytical model for the motion of a spherical bubble. More recently, Lee and Nydahl (1989), and Patil (1991) have numerically simulated the bubble growth in nucleate boiling with an expanded model by including the microlayer, for whose thickness the formulation of Cooper and Lloyd (1969) was employed. Although the heterogeneous 112 temperature field and the hydrodynamics were accounted correctly by solving the momentum and energy equations in the liquid, it was assumed that the bubble remained hemispherical during its growth, and, as such, the bubble departure process could not be demonstrated. For a direct calculation of the moving interface in nucleate pool boiling, Welch (1998) has applied the finite-volume method on a moving unstructured mesh. However, the calculation was limited to a small deformation of the two-phase interface due to numerical instabilities caused by the severe distortion of the computational grid. The VOF method (Welch and Wilson, 2000) has been widely used because of its nice property of preserving the volumes of the two phases in flows, though a very fine grid is needed to capture a smooth interface. Also because it is difficult to compute accurate local curvatures from volume fractions, modeling surface tension forces becomes a problem. A notably newer approach to the solution of interface evolution problems is the level-set Hamilton-Jacobi formulation (Osher and Sethian, 1988; Sethian, 1990), which has been applied to incompressible two-phase flows by Sussman et al. (1994). A modified version that accommodates the liquid-vapor phase change process in nucleate boiling has been employed by Son et al. (1999) to capture the dynamics of a single bubble growing on a horizontal heated surface. It is noted, however, that this application of the level-set method does not have a liquid-vapor volume-preservation property. A fluid particle tracking mesh-free method (MPS-MAFL) has been extended by Yoon et al. (2001) for an ALE simulation of this problem. In this calculation, the particle number density is not constant and particles are allowed to be concentrated locally for higher resolution; the microlayer beneath the bubble, however, has not been modeled. 113 This study presents a complete numerical simulation of the ebullient dynamics of a single bubble in pool boiling of water with effects of apparent contact angle, wall superheat, and altered surface tension and viscosity. The nucleating bubble is decomposed into a microlayer and a macro region in order to account for the peripheral heat transfer. The fluid motion, pressure drop, and heat transfer in the microlayer are modeled by a fourth-order differential equation, similar to that developed by Lay and Dhir (1995). The governing mass, momentum, and energy conservation equations in the macro-vapor and -liquid regions are solved numerically, with the surface tension modeled by a continuum method introduced by Brackbill et al. (1992). The latter eliminates the need for interface reconstruction, thereby simplifying its calculation and fully integrating it into the momentum equation. The vapor-liquid interface in the macro region is captured by a PDE-based fast local level set method (Peng et al, 1999), which easily handles the breaking and merging of interfaces. To overcome the disadvantage posed by the non-conservation form of the level-set equation, an interface-preserving level-set redistancing algorithm developed by Sussman and Fatemi (1999) is used to preserve the mass in both the vapor and liquid phases. 5.2 Mathematical Formulation The coordinate system and the physical domain for the computational model, indicating the microlayer and macro region, similar to the scheme used by Son et al. (1999), are shown in Fig. 5.1. The microlayer envelops the thin film that forms underneath of the bubble, whereas the macro region consists of the bubble and the liquid surrounding it. The flow is assumed to be axisymmetric and laminar, with all liquid 114 physical properties considered to be constant at 100°C (saturated atmospheric condition). In the computational domain, as illustrated in Fig. 5.2, a MAC-staggered grid is used for the discretization of the differential equations. The discrete velocity field uin, j , temperature field Ti ,nj , and the level set function φin, j are located at cell centers, but the pressure pin+−1/1/2,2 j +1/ 2 is located at cell corners. Microlayer The mass, momentum, and energy conservation equations governing the microlayer, based on the Lay and Dhir (1995) treatment, can be respectively stated as ∂δ = vl − q / ρ l h fg ∂t (5.1) ∂pl ∂ 2ul = µl 2 ∂x ∂y (5.2) q = kl (Tw − Tint ) / δ (5.3) vl = − 1 ∂ δ xul dy x ∂x ∫0 (5.4) By employing a modified Clausis-Clayperon equation (Dasgupta et al. 1993), the evaporative heat flux can be written as q = hev [Tint − Tv + ( Pl − Pv )Tv / ρ l h fg ] (5.5) hev = 2( M / 2π RTv )0.5 ρ v h 2fg / Tv ; Tv = Tsat ( Pv ) (5.6) Thus, noting that the pressure in the vapor and liquid phases are related as pv = pl + σ ⋅ k − ρ l g (δ − δ e ) + 115 A δ3 − q2 ρ v h 2fg (5.7) Macro Region Liquid g T=Tsat Vapor y ϕ x=X Heater Surface y δ0 Microlayer r=0 h/2 r=R r Fig. 5.1 Physical domain of a boiling bubble decomposed into a macro region and a microlayer 116 Physical Boundary Liquid T,u, T,u, T,u, φi-1,j+1 φ i,j+1 φi+1,j+1 Pi-3/2,j+1/2 Pi-1/2,j+1/2 Pi+1/2,j+1/2 T,u, T,u, T,u, φi-1,j φi,j φi+1,j Pi-3/2,j-1/2 Pi-1/2,j-1/2 Pi+1/2,j-1/2 T,u, T,u, T,u, φi-1,j-1 φi,j-1 φi+1,j-1 Pi-3/2,j-3/2 Pi-1/2,j-3/2 Pi+1/2,j-3/2 Fig. 5.2 MAC-Staggered grid to show where the variables u, p, T, and φ are located 117 Equations (5.1)-(5.7) can be combined to yield the following final governing equation for the microlayer: δ '''' = f (δ , δ ', δ '', δ ''') (5.8a) which is subject to the following boundary conductions: r = 0 : δ = 1, δ ' = 0, δ ''' = 0 h r = R :δ = , δ '' = 0 2δ 0 (5.8b) Where curvature of the interface k and apparent contact angle ϕ are defined as k= 1 ∂ ∂δ ∂δ (x / 1 + ( )2 ) ∂x x ∂x ∂x tan ϕ = h ( X1 − X 0 ) 2 (5.9) (5.10) Macro Region The dimensionless conservation equations for momentum, energy, and mass in the macro region are, respectively ∂u ∇p y GrL + ui∇u = − + + θ ∂t ρ Fr Re 2 1 1 1 + ( ∇i(2 µD ) − k ∇H ) We ρ Re (5.11) ∂θ 1 + u ⋅ ∇θ = ∇ ⋅ ∇θ Pe L ∂t (5.12) ∇⋅u = − = . 1 ∂ρ ( + u ⋅ ∇ρ ) + V micro ρ ∂t m ρ2 . ⋅ ∇ρ + V micro 118 (5.13a) where m = ρ ( ui − u) = k ∇T / h fg i V micro = ∫ X1 X0 kl (Tw − Ti ) rdr ρ v h fgδ∆Vmicro (5.13b) Also, the level-set function to capture the evolving vapor-liquid interface is advanced as φt + uint ⋅ ∇φ ( x (t ), t ) = 0 (5.14) And φ is initialized to be the signed normal distance from the interface if x ∈ the liquid > 0, φ ( x, t ) = = 0, if x ∈ Γ{x | φ ( x, t ) = 0} < 0, if x ∈the vapor (5.15) Equations (5.11)-(5.15) are bounded by the following set of conditions: y = 0: u = v = 0, T = Tw ,φ = − x cos ϕ ∂v ∂T ∂φ = = =0 ∂x ∂x ∂x ∂u ∂v ∂φ = = = 0, T = Tsat ∂y ∂y ∂y r = 0, r = R : u = y =Y : (5.16) The various dimensionless variables and groups in Eqs. (5.11)-(5.13) are defined as L = σ /[ g ( ρl − ρ v )], U = gL , x = Lx ', u = Uu ', t = ( L / U )t ', P = P ' ρ lU 2 , ρ = ρl ρ ', µ = µl µ ', k = kl k ', θ = (T − Tsat ) /(Tw − Tsat ) U2 ρ l LU 2 LU Fr = , We = , Re = σ νl gL UL g β (Tw − Tsat ) L3 Pe L = , GrL = ν l2 a 119 (5.17a) (5.17b) 5.3 The Numerical Method The fourth-order differential equation governing the microlayer was integrated using an adaptive, step-size control, fifth-order Runge-Kutta method. In order to obtain the governing equation for pressure that achieves mass conservation, the accurate projection method or fractional-step is used (Brown et al., 2001), which is an improved fully second-order accurate projection algorithm over the method used by Bell and Colella (1989). The diffusion terms are treated by a fully implicit scheme, and the convection terms in level-set function are discretized by the third-order essentially nonoscillatory (ENO) schemes given by Fedkiw et al (2002). For other convection terms, a predictor-corrector method (Puckett et al., 1997; Sussman et al., 1999) is adopted. The time-stepping procedure is based on the third-order total time diminishing (TVD) RungeKutta method as φ n +1 = φ n + ∆tL(φ n ) ∆t L(φ n ) + L(φ n +1 ) 4 ∆t φ n +1 = φ n + L(φ n ) + 4 L(φ n +1/ 2 ) + L(φ n +1 ) 6 φ n +1/ 2 = φ n + (5.18) To prevent numerical instability arising from discontinuous physical properties, the smooth Heaviside function in Eq. (5.11) is defined as if φ < −ε 0 1 φ 1 H ε (φ ) = 1 + + sin (πφ / ε ) if φ ≤ ε 2 ε π 1 if φ > ε (5.19) where ε = α∆x , and which will not exhibit a jagged or sharp irregular shape when α > 1 . 120 The fluid density, viscosity, and thermal conductivity are defined in terms of H ρ (φ ) = ρ g + ( ρ l − ρ g ) H (φ ) µ (φ ) = µ g + ( µl − µ g ) H (φ ) (5.20) k (φ ) = kl H (φ ) The key to success of the level-set method is to maintain level set φ a distance function; φ can become irregular after some period of time, which will cause steep gradients in the distance function, and Eq. (5.14) can be reinitialized as follows: d τ + S ( d 0 )( ∆d − 1) = 0 d ( x,0) = d 0 ( x ) = φ ( x, t ) (5.21) Where Sε (φ 0 ) = φ0 φ 20 + ε 2 (5.22) To overcome the disadvantage posed by the level-set equation not being in a conservation form, the following constraint (Sussman et al., 1999) is used to improve the accuracy of Eq. (5.14): φn +1 = φn +1 + ∆tλi , j H ∆' x (φ0 ) ∇φ0 − ∫ Ωi , j H ∆' x (φ0 ) φn +1 − φ0 ∆t λi , j = 2 ' ∫ Ωi , j H ∆x (φ 0 ) ∇φ0 (5.23) In implementing the level-set method, a PDE-based fast local level-set method developed by Peng et al. (1999) is used to track the water-vapor interface, which reduces the computational effort by one order of magnitude when compared to the method used by Sussman et al. (1994). In order to compute the projection of the momentum equation, the multigrid-preconditioned conjugate gradient method (MPCG) developed by Tatebe (1993), which allows the computation of cases that can not be executed at the proper 121 density ratio (~1000:1) by using the standard Gauss-Seidel iterative method or multigrid methods, was implemented. The preconditioner to the conjugate gradient is a multigrid V-cycle, and symmetric multicolor Gauss-Seidel relaxation is used as the smoother at each level of the V-cycle. Details of temporal and spatial discretization are given in Appendix D. 5.4 Solution Validation 5.4.1 Efficacy of the MPCG method In order to ascertain the efficacy of the MPCG method, the results for the Poisson equation with jump diffusion coefficients for different shapes (T-shape and Arc) are graphed in Figs. 5.3(b) and 5.4(b), and the physical problems for these results are depicted in Figs. 5.3(a) and 5.4(a), respectively. This model can be expressed as follows: −∇ ( k ∇u ) = f Ω = (0,1) x (0,1) (5.24) The MPCG method has robust and efficient convergence properties compared to other methods, and it was found to take much less time to converge than using the incomplete Cholesky (IC) as a preconditioner. Moreover, the MPCG method can be efficiently implemented in parallel computations. 122 1 y k=1 0.5 k=1000 0 0 0.5 1 x (a) y 1 0.5 0 0 0.5 1 x (b) Fig. 5.3 (a) Poisson problem with jump diffusion coefficients (T-shape); and (b) test computational results for u(x,y) for the MPCG method (f = 100) 123 1 y k=1 0.5 k=1000 0 0 0.5 1 x (a) y 1 0.5 0 0 0.5 1 x (b) Fig. 5.4 (a) Poisson problem with jump diffusion coefficients (arc); and (b) test computational results for u(x,y) for the MPCG method (f = 100) 124 5.4.2 Efficacy of the level set method In the level-set method, an interface is represented as a zero level set of a continuous function, designed to have a positive sign on one side of the curve and a negative sign on the other site of the curve. In the present application, the level set φ(x) is defined as a signed distance function φ (x) that yields the closest distance to the interface as illustrated in Fig. 5.5. Here the level set is the distance to a circle centered at the origin (0.5,0.5) with a positive sign outside of the circle (r = 0.15). The full level-set function φ(x) is updated with the velocity field determined by the governing equations. To demonstrate the applicability of the level-set method, the axisymmetric rise of a sphere with a constant velocity v(x,y) = {0, 1} was simulated, and these results are shown in Fig. 5.6 as a test problem. In this case, the interface is accurately tracked while the mass is conserved with reinitialization, Eq. (5.21), and the constraint of Eq. (5.23) is applied. y 1 0.5 0 0 0.5 x 1 Fig. 5.5 Contours of level-set function φ(x) with the solid line circle representing for φ(x) = 0 125 y 1 0.5 0 0 0.5 x 1 Fig. 5.6 Rising bubble interfaces plotted at different times for U = 0, V = 1 using the mass-preserved level-set method 126 5.4.3 Verification of phase change modeling The liquid-vapor phase change process in nucleate boiling has been employed by Son et al. (1999) to accommodate the level-set function to capture the dynamics of a single bubble growing on a horizontal heated surface. A simpler case of bubble growth in an extensive uniformly superheated liquid, as shown in Fig. 5.7, is tested before simulating of the bubble growth during nucleate boiling on a heated surface. Tl q uint Tsat Tl > Tsat Fig. 5.7 Schematic of a growing bubble in an extensive superheated liquid pool There are two limiting cases of the bubble growth process that are generally acknowledged in the literature (Carey, 1992): (a) Inertial-controlled growth. Heat transfer to the interface is very fast and is not a limiting factor to growth. The growth rate is therefore governed by the momentum interaction between the bubble and surrounding liquid. These conditions usually exist during the initial stages of bubble growth, just after the embryonic bubble forms and begins to grow. (b) Heat-transfer-controlled growth. In this regime, growth is limited by the relatively slower transport of heat to the interface. These conditions generally 127 correspond to the later life stages of bubble growth when it is larger and the liquid superheat near the interface has been significantly depleted. This study is only related to the heat-transfer-controlled bubble growth period. The governing equations for the transport of heat in the liquid surrounding the bubble are the following: ∂T ∂T α l ∂ 2 ∂T +u = r ∂t ∂r r 2 ∂r ∂r u= kl dR R dt r (5.25) 2 (5.26) ∂T dR ( R, t ) = ρv hlv dt ∂r (5.27) The boundary and initial conditions on these conservation equations are T ( r ,0 ) = T∞ ; T ( R, t ) = Tsat ( Pv ) ; T ( ∞, t ) = T∞ (5.28) Plesset and Zwick (1954) have given an analytical solution as R(t ) = 2CR α l t (5.29) where CR =Ja 3/π , Ja= C p ρ l (Ts − Tsat ) ρ v h fg (5.30) and which provides the reference for testing the computational modeling. Two cases with different Jacob numbers were tested using the level-set method with phase-change modeling. The results are presented in Fig. 5.8, and it is clearly seen that the numerical simulations agree well the analytical solutions. 128 12 12 Numerical 11 Numerical 11 Analytical 9 9 8 8 7 7 y[mm] 10 y[mm] 10 6 5 6 5 4 4 3 3 2 2 1 0 1 Ja = 1.873e-3 0 1 2 3 4 Analytical 5 0 6 Ja = 3.745e-3 0 1 2 r[mm] 3 4 5 6 r[mm] (a) 6 Numerical Analytical 5 Ja r [mm] 4 3 3 5e.74 3 = -3 .873e Ja = 1 2 1 0 0 10 t [ms] 20 30 (b) Fig. 5.8 Bubble growth in an extensive superheated liquid pool: (a) bubble interfaces plotted at different times; (b) bubble growth with time 129 5.5 Results and Discussion 5.5.1 Microlayer Cooper and Lloyd (1969) were probably the first to present an analytical model and experimental results for the formation of a microlayer, and treated it as a liquid wedge underneath the growing bubble. More recently, Wayner and his co-workers (Wayner, 1992; Dasgupta et al. 1993, 1994) have experimentally and theoretically addressed the problem of the shape of the microlayer. Lay and Dhir (1995) subsequently considered a more detailed theoretically analysis of the pressure distribution in the microlayer to predict the shape of the microlayer and the heat flux. The present investigation is based on the Lay and Dhir analytical model; however, different boundary conditions are applied to accommodate the heat transfer and growth of a nucleated boiling bubble. Figure 5.9 depicts the computed liquid-vapor interface shape and the microlayer thickness, as well as the temperature distribution for a bubble of apparent contact angle of 30°. It can be seen that the slope of the interface in the microlayer decreases to zero rapidly near the nonevaporating region (r = 0, δ = δ0). As the thickness of the liquid microlayer increases, the interface temperature decreases from the wall temperature to the liquid saturation temperature, which suggests that the more significant evaporative heat transfer occurs in the microlayer region. Cooper and Lloyd (1969) have suggested that bubble growth rates are of the same order as the evaporation rates from the microlayers. In Lee and Nydahl (1989), it has been shown that microlayer evaporation accounts for nearly 90 percent of the wall heat transfer during saturated boiling of water at 1 atm and 130 8.5 K wall superheat. Also, the total energy removal from the microlayer is coupled with the heat and mass transport from the macro region. 5e-005 384 4e-005 380 Interface Temperature 378 3e-005 Microlayer Shape 2e-005 376 1e-005 374 372 Microlayer Thickness δ (m) Interface Temperature (Κ) 382 0 1e-005 2e-005 3e-005 0e+000 r (m) Fig. 5.9 Microlayer shape and vapor-liquid interface temperature distribution for a nucleated bubble 5.5.2 Surface tension effect The shape and dynamics of a bubble rising in water is governed essentially by the following set of physical quantities: liquid/vapor viscosity, gravity, liquid/vapor density, bubble mass, bubble initial diameter, surface tension, pressure, and velocity of the bubble relative to the surrounding fluid. A bubble immersed in a fluid maintains its shape mainly due to the surface tension forces that help the bubble surface adapt to variations in 131 the external stresses. Surface tension is associated to a surface energy that it tends to minimize. The variations in shape of an adiabatic spherical rising gas bubble in quiescent water, as computed by the level-set method, are depicted in Fig. 5.10. The results are for different surface tension values while keeping the remaining parameters constant, and which are reflected in the changes of Weber number: We = ρU 2 L σ (5.31) With time the spherical shape is seen to change to an oblate ellipsoid that has a flattening or convex underside. The elliptical stretching increases with We, i.e., the aspect ratio (minor axis vs. major axis) is smaller for larger Weber number (smaller surface tension). However, the rising velocity is larger when compared to that for smaller Weber number, and these computations are graphed in Fig. 5.11 for Re = 278. The bubble terminal velocity is reached after a certain time, which is the result of an equilibrium of the following forces: (1) The drag force, which represents the viscous effects that tend to slow the bubble motion as penetrates the fluid; (2) pressure and surface tension forces, where in general, the main component is the hydrostatic pressure gradient under static conditions, but under dynamic conditions, the dynamic pressure gradient balances out the surface tension force, and they are the primary determinants of the changing bubble shape; (3) Gravity and buoyancy forces, where gravitational force is usually negligible on a bubble, and the buoyancy force is generated due to density variation - also referred to as the Archimedes force; (4) The virtual mass force; the fluid in its immediate surrounding flows at the same speed as that of the bubble, and in fact, the bubble is a gas-liquid coherent object. Therefore, bubbles demonstrate a much higher 132 8 8 7 8 7 7 t = 3.5 6 6 6 5 5 5 4 4 t=2 3 3 2 2 2 t=0 0 -2 -1 0 t=0 1 1 2 0 -2 (a) We=0.538 -1 t=2 4 t=2 3 1 t = 3.5 t = 3.5 t=0 1 0 1 0 -2 2 -1 (b) We=0.782 0 1 (c) We=1.03 Fig. 5.10 Temporal evolution of bubble shapes for various Weber numbers (Re = 278) 1.0 0.9 u [dimensionless] 0.8 0.7 0.6 0.5 0.4 Re=278, We=0.538 Re=278, We=0.782 Re=278, We=1.03 0.3 0.2 0.1 0.0 0 1 2 3 4 5 6 t [dimensionless] Fig. 5.11 Change bubble rising velocity with time for various Weber numbers 133 2 inertia than that of their own mass, and the effect of this added mass is to dampen or retard the bubble acceleration; (5) The lift force, by which when a bubble experiences a vertical shear fluid flow, it develops a movement orthogonal to the main flow with the bubble deformation, especially when a surfactant or other reagents are present; (6) The basset history force, which is the resulting effect of the virtual added mass that must adjust itself to Lagrangian acceleration and take into account the recent past of the bubble evolution; (7) wall force that accounts for the wall effect; (8) Marangoni forces; with a surfactant in solution, a surface concentration gradient develops at the bubble interface that promotes diffusocapillary flows, and so also thermocapillary flows are generated due to the temperature-gradient induced density variations. Figure 5.12 shows the time evolution of bubble growth and departure in saturated nucleate pool boiling of both pure water and aqueous solutions of different surface tensions for a wall superheat of ∆T = 10 K and ϕ = 45°. Compared to the behavior in pure water, the ebullience in solutions with reduced surface tension (σ = 47.0 mN/m for SDS at C = 1000 wppm,) is seen to be characterized by a smaller departure diameter, and higher bubble detachment and departure frequency. The reduced surface tension due to the interfacial adsorption-desorption of the additive causes the bubble to become more spherical with time, and this results in a small increase in speed of the detached bubble, before steady state is reached. Focusing on the bubble detachment process in particular, a neck is seen to form near the wall and thereafter the bubble breaks off, and an accelerated necking, break-off, and subsequent nucleation is seen with reduced surface tension. The small portion left after the bubble departure serves as a nucleus for the next bubble. The 134 nucleation of smaller-sized bubbles and their removal at high frequency essentially promote enhanced heat transfer in aqueous surfactant solutions. t=0.03s t=0.03s t=0.03s t=0.04s t=0.04s t=0.04s t=0.08s t=0.08s t=0.08s (a) (b) (c) Fig. 5.12 Bubble growth and its departure in nucleate pool boiling for ∆T = 10 K, ϕ = 45°. (a) σ = 58.86 mN/m (water), (b) σ = 47.0 mN/m (SDS, C = 1000 wppm), (c) σ = 37.5 mN/m (SDS, C = 2500 wppm) 135 The diameter to which a bubble grows before departing is dictated by the balance of various forces acting on the bubble, namely, inertia of the liquid and vapor, liquid drag on the bubble, buoyancy, and surface tension. Fritz (1935) correlated the bubble departure diameter by balancing buoyancy and surface tension as follows: σ Dd = 0.0208ϕ g ( ρl − ρ g ) (5.32) Subsequently, Cole and Rohsenow (1969) also proposed a correlation of bubble departure diameter as Dd = 1.5 × 10−4 ϕ σ g ( ρl − ρ g ) Ja *5 / 4 (5.33) where Ja* = ρl c plTsat ρ v hlg (5.34) While the wall superheat and apparent contact angle effects on the bubble departure diameter were not considered in Eqs. (5.32) and (5.33), they do provide a reasonably good correct length scale for the boiling process. The calculated departure diameters for the fluids with different surface tension values but an apparent contact angle of 45° (~ representative of the nominal contact angle in pure water (Han and Griffith, 1965)) and a wall superheat of 10°C are graphed in Fig. 5.13. Also included are the results from Eqs. (5.32) and (5.33) for comparison, and a fair agreement is seen where the two results enveloping the computational values. 136 101 D [mm] Cole and Rohsenow's correlation (1969) 100 Fritz's correlation (1935) Present simulation 10-1 100 101 102 σ [mN/m] Fig. 5.13 Bubble departure diameter vs. surface tension (∆T = 10°C, ϕ = 45°) 137 5.5.3 Viscosity effect The effect of viscosity (represented by Morton number) on the motion of the bubble is illustrated in Fig. 5.14. The Morton number is defined as Mo = g µ 4 ( ρl − ρ v ) (5.35) ρl2σ 3 As seen from the simulated shapes graphed in Fig. 5.14, the bubble is more spherical or the aspect ratio (minor axis vs. major axis) is larger for a larger Morton number (higher viscosity). Meanwhile, the rising velocity is smaller for a larger Morton number, as shown in Fig. 5.15 due to the higher viscous force. That, also, in the case of water, the bubble reaches a terminal velocity after a relatively longer time compared to that in surfactant solutions. Furthermore, the predicated shape using the modified level-set method in this study is in good agreement with the experimental results of Bhaga and Weber (1981) for water (Mo =2.59e-11, We = 0.538) as seen in Fig. 5.16. All the cases studied are summarized in Table 5.1. Table 5.1 Combination of Non-Dimensional Parameters Studied (D = 2mm) Re = 278 (water at 23°C) Case No. 1 2 3 Case No. 1 2 3 σ We [mN/m] 5.38e-1 72.3 7.82e-1 48.9 1.03 37.5 We = 5.38e-1 (water at 23°C) η x 103 Mo [Pa.s] 2.59e-11 0.935 2.49e-10 1.77 9.47e-09 4.39 138 C (SDS) [wppm] 0 1000 2500 C (Carbopol 934) [wppm] 0 1000 3500 8 7 8 8 7 7 t = 3.5 t = 3.5 6 6 6 5 5 5 4 4 t=2 4 t=2 3 3 3 2 2 2 t=0 1 0 -2 -1 0 t=0 1 1 0 -2 2 -1 0 t=2 t=0 1 1 0 -2 2 (b) Mo=2.49e-10 (a) Mo=2.59e-11 t = 3.5 -1 0 1 2 (c) Mo=9.47e-09 Fig. 5.14 Temporal evolution of bubble shapes for various Morton numbers (We = 0.538) 1.0 0.9 u [dimensionless] 0.8 0.7 0.6 0.5 0.4 Mo=2.59e-11, We=0.538 Mo=2.49e-10, We=0.538 Mo=9.47e-09, We=0.538 0.3 0.2 0.1 0.0 0 1 2 3 4 5 6 t [dimensionless] Fig. 5.15 Change in bubble rising velocity with time for various Morton numbers 139 (a) (b) Fig. 5.16 Comparison of predicted result with the experiment for water: (a) present simulation; (b) experimental result of Bhaga and Weber (1981) 140 5.5.4 Temperature, velocity, and pressure fields Figure 5.17 provides an indication of the changes in thermal field around the bubble during growth. Because the bubble represents an isothermal sink12 in a nonuniform temperature field, the isotherms near the bubble do not uniformly surround the bubble in a boundary-layer-type manner (Lee and Nydahl, 1989). The overall energy for bubble growth comes from both the bubble cap and the microlayer, and the folding of inflexions in the isotherms in a thin layer near the bubble seen in the computations are somewhat difficult to resolve and identify experimentally. The crowding of the isotherms at the bubble base is reflective of the very high heat flux in that region. Initially when the bubble is located inside the thermal boundary layer in Fig. 5.17(a), then it growth is driven by the evaporation all around the vapor-liquid interface and the bubble quickly grows out of the thermal boundary layer. During a major portion of this time span, however, there is no wrapping of the bubble cap and the energy required for evaporation is restricted to a small portion around the bubble base as shown in Fig. 5.17(b). The flow field and pressure contours in and around a detached bubble in pure water under conditions where σ = 58.86 mN/m and ∆T = 10 K are shown in Fig. 5.18. During the early period of the upward motion of this bubble, the liquid around it is seen to be pushed out. A circulatory flow pattern inside the bubble as well as in the liquid outside is also clearly seen for the freely rising detached bubble. The vapor velocity vectors in the bubble are reflective of its bulk movement in the upward direction and the changes in the bubble shape as it rises in the pool. In the present illustrative calculations, 12 Uniform temperature and pressure are assumed for the vapor phase in the simulation. In reality, the bubble may be non-isothermal (Beer, 1979), and further investigation is needed to establish this condition. 141 the apparent contact angle is assumed to be constant (ϕ = 45°). The pressure distribution around the bubble and the velocity field are coupled together, and the crowding of the isotherms around the bubble as shown in Fig. 5.18(b) is reflective of the very high- 2.0 2.0 1.8 1.8 1.6 1.6 1.4 1.4 1.2 1.2 y/L y/L pressure gradients that are balanced out by the surface tension forces. 1.0 1.0 0.8 0.8 0.6 0.6 0.4 0.4 0.2 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.0 0.2 0.4 0.6 0.8 1.0 x/L x/L (a) early growth (b) later growth Fig. 5.17 Temperature isotherms during different bubble growth stages 142 2 y/L 1.5 1 0.5 0 -1 -0.5 0 0.5 1 x/L 2 2 1.5 1.5 y/L y/L (a) 1 0.5 0 -1 1 0.5 -0.5 0 0.5 1 x/L 0 -1 -0.5 0 0.5 x/L (b) (c) Figure 5.18 (a) Velocity field; (b) filled pressure contours; and (c) line pressure contours, in and around a detached isolated bubble for ∆T = 10K, ϕ = 45° and σ = 37.5 mN/m, t = 0.06s 143 1 5.5.5 Apparent contact angle and superheat effect The dependence of bubble growth on apparent contact angle is depicted in Fig. 5.19. Three different contact angles (ϕ = 35°, 45°, and 55°) were selected for the computations with a fixed wall superheat (∆T = 10°C) and surface tension (σ = 58.86 mN/m). It is seen from Fig. 5.19 that the bubble departure diameter becomes larger and the vapor volume required for bubble departure increases as the contact angle increases. This is probably caused by the surface tension forces at the three phase contact line. The increase of bubble departure diameter with contact angle is found to be generally in agreement with the Fritz correlation (1935) as shown in Fig. 5.19(b). In the present calculations, all apparent contact angles are assumed to be constant. In reality, however, the dynamic contact angles (advancing and receding contact angles) may differ from the static contact angle. The consequent apparent contact angle hysteresis on a real surface during nucleate boiling may have some effect on the shape of the liquid-vapor interface at the heater surface during bubble departure. Gorenflo et al. (1986) proposed the following correlation for bubble diameter at departure for the heat flux controlled bubble growth, which includes the wall superheat effect and is a function of the thermal diffusivity α and the Jacob number Ja: 1/ 3 Ja 4 ⋅ α l2 Dd = C1 ⋅ g 1/ 2 2π ⋅ 1 + 1 + 3 ⋅ Ja 4/3 (5.36) where Ja = ρl c pl ∆T k ,αl = l ρ v hlv ρ l c pl 144 (5.37) 4 3 y[mm] ϕ = 55 ϕ = 45 2 ϕ = 35 1 0 -2 -1 0 1 2 r[mm] (a) 3.5 D [mm] 3.0 Present simulation 2.5 2.0 Fritz correlation (1935) 1.5 25 35 45 55 65 ϕ [deg] (b) Fig. 5.19 (a) Bubble shape at departure for different contact angles; and (b) bubble departure diameter vs. apparent contact angle (∆T = 10°C, σ = 58.86 mN/m) 145 Different values of the proportionality constant were suggested for different boiling liquids, and for water, C1 = 2.63. Figure 5.20 illustrates the effect of wall superheat on the bubble growth, and its departure time, along with a comparison between simulation results and predictions from the correlation proposed by Gorenflo et al. (1986) for a fixed apparent contact angle (ϕ = 45°) and surface tension (σ = 58.86 mN/m). The bubble departure diameter is seen to increase with the wall superheat. However, the bubble departure time or growth period decreases when ∆T increases, and it indicates a higher heat transfer rate and vapor production rate as well. 5.6 Significance and Limitations of Nucleate Boiling Simulations The computational results presented in this chapter are insightful in making a good assessment of the extent of the effects of surface tension, viscosity, wall superheat, apparent contact angle, and microlayer when boiling in aqueous surfactant solutions. Nevertheless, there are certain limitations of the present simulations, as a more meaningful and complete model requires the inclusion of the adsorption-desorption controlled interfacial surfactant transport process as illustrated in Fig. 5.21 into the governing equations. When a surfactant is present in an aqueous pool, its adsorption at the interface results in an equilibrium surface concentration Γeq and surface tension σeq, whereby the process solely results in a surface tension reduction. However, at the dynamic interface of a nucleate boiling bubble, the surface concentration rarely remains at its equilibrium value. Surface convection creates concentration gradients that alter the local surface tension, thereby producing Marangoni stresses on the interface, with larger deformations 146 4 3.5 3.5 3.0 y [mm] 2.5 ∆ T = 13K 2.5 ∆ T = 10K 2.0 D [mm] 3 2 ∆ T = 7K 1.5 1.5 1.0 1 0.5 0.5 -1 0 1 0.0 0.00 2 0.01 0.02 0.03 r [mm] t [s] (a) (b) 0.04 101 9 8 7 6 Present simulation 5 4 D [mm] 0 -2 13K 10K 7K 3 2 Correlation by Gorenflo et al. (1986) 100 9 8 5 6 7 8 101 9 ∆Τ [K] (c) Fig. 5.20 Effect of wall superheat ∆T on bubble growth (ϕ = 45°, σ = 58.86 mN/m): (a) bubble shape at departure; (b) bubble departure time; and (c) bubble departure diameter vs. wall superheat 147 0.05 when compared to the surfactant-free case (Cuenot et al, 1997; Eggleton and Stebe, 1998; Wasekar and Manglik, 2003). In fact, physico-chemical processes such as bulk mass transfer, and additive adsorption/desorption have a strong influence on the flow around the bubble, which in turn affects the surfactant transport process. A few recent studies have attempted to model to investigate the complex interactions between surfactant transport process and single bubble dynamics (Wang, 1999; Liao and McLaughlin, 2000; Palaparthi, R., 2001) with encouraging results. However, all these studies have treated the bubble to be adiabatic, and free-slip or non-slip boundary condition was applied at the liquid-vapor interface; this, unavoidably, will alter the transport process at the interface. Further in depth simulation modeling and computational methods are yet to be developed. U Diffusion From Bulk to Surface Kinetic Adsorption Surface Convection to Trailing End Kinetic Desorption Diffusion From Surface to Bulk Fig. 5.21 Adsorption-desorption controlled surfactant interfacial transport process 148 CHAPETR 6 CONCLUSIONS AND RECOMMENDATIONS 6.1 Conclusions With the addition of small amounts of surfactants or polymers, the saturated pool boiling of water on a cylindrical heater is found to be altered significantly. In general, besides the heat flux (or wall superheat) levels, the relative extent of change in boiling performance is seen to be governed by the interfacial phenomena at the surface contact (solid-liquid and vapor-liquid), which in turn are determined by several additive physicochemistry-based factors. The major accomplishments, findings, and salient features of this study are summarized as follows: 1. An extensive literature review is presented that assess the available information on pool boiling heat transfer of aqueous surfactant and polymer solutions, interfacial phenomena and electrokinetic effects, especially the physisorption process at the solid-liquid interface, and its effects on surface wettability and active nucleation density. 2. Extended measurements and data for interfacial properties (dynamic and equilibrium surface tensions, and wettability) are presented, in order to characterize the adsorption behaviors of additives in their aqueous solutions at both the solid-liquid and liquid-vapor interfaces. It was generally found that the surfactant physisorption process tends to follow a specific adsorption isotherm, which in turn is determined by the electrokinetics (zeta potential) at the solidliquid interface that lead to corresponding changes in the surface wetting conditions. The measured contact angle correlates well with this adsorption 149 isotherm and zeta potential variations with concentration. Also, additives have surface-active properties that lower the surface tension of water considerably. The surfactant adsorption-desorption process, however, is time-dependent and it manifests in a dynamic surface tension behavior, which eventually reduces to an equilibrium value after a long time span. For boiling applications with small surface-age interface, it is the dynamic surface tension relaxation process rather than the equilibrium or static surface tension at a fixed bubble frequency that is perhaps the more critical determinant of the phase-change performance. 3. In the case of polymer additives, the measured shear-rate- and temperaturedependent viscosity for dilute surfactant solutions showed insignificant change from that for water. However, the aqueous polymer solutions become significantly more viscous, and those with a much higher degree of polymerization or molecular weight (Carbopol 934 vs. HEC QP-300) exhibit a distinct non-Newtonian shear-thinning rheology. 4. The heat transfer in saturated nucleate boiling of aqueous cationic surfactant solutions is found to be enhanced considerably. The performance is seen to depend upon the dynamic σ reduction of the solution, micellar structure of the surfactant, its degree of ethoxylation (which influences both surface tension and wettability), and its molecular weight and ionic nature (which influence coverage, inter-molecular repulsion or lack thereof at the vapor-liquid and solid-liquid interfaces, and σ relaxation time). The heat transfer generally increases with qw′′ and additive concentration up to a C ≤ CMC. Depending on C and qw′′ , the heat transfer coefficient is found to increase by as much as 63% over that for pure 150 water for DTAC (a low molecular weight cationic) solutions. With C > CMC, the enhancement decreases and the heat transfer can even deteriorate below that for water depending upon qw′′ and the surfactant chemistry. High concentration solutions of the ethoxylated cationic Ethoquad 18/25 (15 EO groups), for example, show considerable heat transfer deterioration as well as incipience thermal hysteresis that is typically found in highly wetting fluids. The boiling process in non-ethoxylated surfactant solutions was observed to be characterized by an early incipience of regularly shaped smaller-sized bubbles, with a reduced tendency for coalescence and relatively higher bubble departure frequencies. The presence of EO groups in the molecular chain of the surfactant, which changes the surface wettability and alters the active nucleation site density and their distribution, tends to promote the inception of smaller-diameter bubbles in premicellar concentrations and suppress the nucleation process in post-micellar solutions. The different boiling mechanisms between non-ethoxylated and ethoxylated surfactants can essentially be related to the different physisorption of surfactant molecules at the solid-water interface, which is characterized by the surface wettability variations as a function of surfactant concentration. 5. Reflecting its surface-active nature and molecular adsorption at the vapor-liquid interface, the polymeric HEC solutions show much greater relaxation of both the dynamic and equilibrium surface tension in comparison with Carbopol. As a consequence, nucleate boiling in polymeric HEC solutions (C < CPC or the critical polymer concentration) is observed to be characterized by significantly larger number of considerably smaller bubbles that have much higher departure 151 frequencies than that in pure water. The reduced surface tension, along with the molecular adsorption on the heating surface (liquid-solid interface) perhaps also contributes to the formation of new nuclei. The combined mechanisms result in considerable enhancement, with up to 22.9% higher heat transfer coefficients, relative to water, in the near CPC or overlap concentration solutions of HECQP300. In post-CPC solutions, however, a decrease in the heat transfer coefficient from the maximum values obtained at CPC is observed. This is perhaps due to the retardation of vapor bubble growth and suppression of microconvection in the boundary layer because of the high viscosity of high concentration solutions. Boiling in aqueous Carbopol 934 solutions, on the other hand, shows a continuous deterioration in heat transfer, relative to that in pure water, at all concentrations because of the viscous suppression of ebullience activity. Some vapor explosions are also observed on the heater surface, akin to the boiling behavior normally found in highly viscous liquids. These results also suggest that the dynamic surface tension and apparent viscosity are the dominant performance predictors, and should perhaps be accounted for in developing any predictive model for nucleate boiling heat transfer in aqueous polymer solutions. 6. An in-depth systematic characterization of nucleate pool boiling in surfactant solutions is delineated, which is based on the extensive deductive analysis boiling heat transfer experiments, coupled with photographic visualization and interfacial property and fluid rheology measurements. The effects of dynamic surface tension and surface wettability (represented by contact angle) on nucleate boiling heat transfer are analyzed. The roles of molecular weight, ionic nature, 152 ethoxylation, and zeta potential are discussed. Because of the highly dynamic nature of nucleate boiling in surfactant solutions, the measure of dynamic surface tension is seen to be an effective scaling property for the heat transfer data. The faster diffusion of lower molecular weight surfactants tends to reduce the surface tension faster in a short period of time, which is reflected in the better heat transfer performance of their solutions. Besides the dynamic surface tension relaxation, the additive physico-chemical properties, which alter the surface wetting of aqueous solutions due to the interfacial physisorption of surfactant molecules, are shown to be critical parameters in predicting their enhanced nucleate pool boiling heat transfer performance. 7. Finally, in an effort to understand the ebullience behavior, the single-bubble dynamics is computationally modeled. To this end, first, a general review of the computational fluid dynamics methods with moving boundaries is presented. Next, a computational model is developed for the complex single bubble dynamics that addresses the effects of surface tension, viscosity, microlayer, wall superheat, and apparent contact angle on the bubble dynamics. The liquid-vapor interface was tracked by a modified fast local level-set method that accommodates the liquid-vapor phase change process, and which also has mass preservation property. 6.2 Recommendations for Future Research Nucleate boiling in aqueous surfactant or polymer solutions is a complex conjugate problem. This, along with the interfacial phenomena at both solid-liquid and 153 liquid-vapor interfaces, results in a rather complicated transport process. Even though an elaborate and controlled investigation has been conducted experimentally and computationally modeled in this dissertation, many aspects of their boiling heat transfer quantification are yet to be resolved before a generalized design correlation can be developed or the nucleate boiling process can be effectively controlled. Therefore, the following areas are recommended for future research: 1. In-depth interfacial phenomena characterization - zeta potential, surface wettability, and adsorption isotherm complement each other and allow a better understanding of the eletrokinetic characteristics, surface wetting, and the concomitant flow behavior. Extensive investigations are necessary for a better understanding of the surfactant physisorption process and its electrokinetic effects at the solid-liquid interface. Furthermore, different adsorption structures may exist (Richard et al., 2003; Schulz et al., 2001; Manne and Gaub, 1995) besides the well-known double-layer model, which is evident from the images in Fig. 6.1 taken by Schulz et. al. (2001) using Atomic Force Microscopy (AFM). AFM, which can detect and characterize surface aggregates only a few nanometers in diameter, and Neutron Reflectometry (NR), which can be used to measure the thickness and concentration profile of an adsorbed surfactant at an interface, are two of the most important techniques to visualize and measure the adsorption layer. Such measurements are critical for characterizing the surfactant adsorption process at the solid-liquid interface, and the different possible adsorption layers based on different micelle structures are illustrated in Fig. 6.2. 154 Fig. 6.1 Schematic representation and AFM detection images of adsorbed layer structures consisting of (A) spherical micelles, (B) cylindrical micelles, and (C) a bilayer (Schulz et al, 2001) 155 Surfactant molecule Hydrophobic tail Hydrophilic head C < HMC HMC < C < CMC Hydrophilic Hydrophobic Hydrophobic Bilayer Monolayer Spherical Vesicle Slightly Hydrophobic Trilayer C > CMC Reversed Fig. 6.2 Conceptualization of possible surfactant adsorbate layers at the solid-liquid interface 156 2. The surfactant physisorption and its electrokinetics are not only basic to a fundamental understanding of nucleate boiling control, but also can provide insights into related natural phenomena in other applications. For example, using the changed bubble dynamics to detect certain substances (surfactants, DNA, or proteins) leads to quick, simple and inexpensive chemical or bio sensing, and microfluidic devices (Thomas et al., 2003; Cornell et al, 1997; Huber et al., 2003). Another interesting phenomenon is to make water flow uphill (Chaudhury and Whitesides, 1992) on a surface with a wettability gradient produced by coating of self-assembled monolayers (SAMS). All these topics are related to the interfacial phenomena. The idea to alter the transport process from the interfacial contact to improve the heat transfer performance drastically would lead to some interesting research directions. For example, some nanofluids or nanoparticles (Wasan and Nikolov, 2003; Chaudhury, 2003, Hentze et al., 2002; Das et al., 2000) with special surface properties based on the different structures of surfactant or micelles can be developed and the potential impact applies to any process that involves phase change or surface contact. Extensions of the present study with such consideration will be able to reach out to new frontiers of research. 3. Experimental investigation of bubble dynamics using controlled single bubble experiments to visualize and quantify the altered transport process with presence of additives. This will make a significant contribution to the fundamental understanding of the altered bubble dynamics with surface-active and other additives in aqueous solutions. 157 4. Computational modeling of a single boiling-bubble dynamics (inception and nucleation→ growth → departure) to investigate the effect of surfactant additives on the nucleation process, bubble growth, Marangoni convection, and surfactant transport process during the bubble lifetime. Effective and innovative numerical methodologies are probably needed to treat the microscale transport process, surfactant adsorption/desorption, and the free interface properly. 5. The unique structure of a surfactant – a hydrophilic head with a hydrophobic tail – provides an effective means to control the surface wettability, which is critical to control of many applications, including nucleate boiling. Surfactant and polymer additives in water can be used to shift the water-boiling curve to the left or right, CHF and the boiling regimes will change accordingly. The ability to control phase-change process is of great importance to thermal processing, which include not only the traditional phase change devices that work in the nucleate boiling regime, but also the transport processes under extreme high heat flux, for instance, quenching. Additional boiling experiments are required to extend the nucleate boiling curve for aqueous surfactant and polymer solutions on different heaters (shape and size) up to CHF and film boiling regime. 6. Phase-change enhancement and control experiments under microgravity conditions. The method shown in this study to alter the transport process at the surface contact is determined by intermolecular electrokinetics, which could be free of gravity effect, and it would provide an effective solution for cooling devices used in the space program applications. 158 7. Correlation and qualification of the heat transfer performance of aqueous surfactant solutions. Nucleate boiling of water with additives is a complex problem as illustrated in Fig. 1.2. There are some encouraging attempts to correlate this process from one or two aspects. However, due to the rather intricate and interdependent nature of the problem, these models are either oversimplified, or only applied to a small concentration range, or focused on a specific surfactant. A thorough understanding of the associated phenomenon and mechanisms is still required, and the development of generalized predictive correlations is yet to be resolved. Using a systematic and controlled methodology is critical to reach out to reliable and robust models when correlating the boiling heat transfer of water with additives. As for boiling in aqueous solutions, the most suitable and fundamental approach to this conjugate problem could extend the work from the well-established basic mechanistic models for pure liquids. For example, the well established Mikic and Rohsenow model (Mikic and Rohsenow, 1969; Judd and Hwang 1976) for nucleate boiling in water could form a good starting reference: i K12 π (λρ c p )l f Dd2 N a ∆T 2 K2 π + 1 − 1 N aπ Dd2 α nc ∆T + α ev ∆TN a Dd2 2 4 q= (6.1) By combining all the factors affected by surfactant additives as shown in Fig. 6.3, the final correlation can then be established eventually. Figure 6.4 proposes the possible future investigations based on the current recognition of this transport process. 159 Basic Model (Pure Liquid) (well-established) Active density Bubble departure diameter Bubble departure frequency Marangoni convection Physical properties Microlayer evaporation … Surfactant (Additives) Fig. 6.3 Proposed approach to correlation of nucleate boiling in aqueous surfactant solutions Second stage correlation Physico-Chemical Properties •Ionic Nature First stage correlation ? Dynamic Surface Tension ? ? Zeta Potential/ Contact Angle •Molecular structure Bubble departure diameter Bubble release frequency dσ/dt or τD* Single bubble experiment •Ethoxylation •Molecular Weight •Micelle •Packing •Others * Diffusion time Active site density Or correlation from literature (θ vs. Na) Dynamic (advancing) Contact Angle High speed video (actual heater) Microlayer thickness (flux removal rate) Fig. 6.4 Proposed investigations to correlation of nucleate boiling in aqueous surfactant solutions 160 BIBLIOGRAPHY Adamson, A.W., 1976, Physical Chemistry of Surfaces, 3rd ed., Wiley, New York. Ammerman, C.N., and You, S.M., 1996, “Determination of Boiling Enhancement Mechanism Caused by Surfactant Addition to Water,” Journal of Heat Transfer, 118, pp. 429-435. Ashayer, R., Grattoni, C.A., and Luckham, P.F., 2000, “Wettability Changes During Surfactant Flooding,” 6th International Symposium on Evaluation of Reservoir Wettability and Its Effect on Oil Recovery, Socorro, New Mexico, USA. Bahl, M., Zhang, J., and Manglik, R.M., 2003, “Measurement of Dynamic and Equilibrium Surface Tension of Surfactant Solutions by the Maximum Bubble Pressure Method,” Paper No. HT2003-47137, Proceedings of 2003 ASME Summer Heat Transfer Conference, ASME, New York, NY. Bang, K.H., Kim, M.H., and Jeun, G.D., 1997, “Boiling Characteristics of Dilute Polymer Solutions and Implications for the Suppression of Vapor Explosions,” Nuclear Engineering and Design, 177, pp. 255-264. Bar-Cohen, A., 1992, “Hysteresis Phenomena at the Onset of Nucleate Boiling,” Pool and External Flow Boiling, ASME, New York, pp. 1-14. Barry, B.W., and Wilson, R., 1978, “C. M. C., Counterion Binding and Thermodynamics of Ethoxylated Anionic and Cationic Surfactants,” Colloid and Polymer Science, 256, pp. 251-260. Barthau, G., 1992, “Active Nucleate Site Density and Pool Boiling Heat Transfer-an Experimental Study,” International Journal of Heat Mass Transfer, 35(2), pp. 271-278. Basu, N., Warrier, G.R., and Dhir, V.K., 2002, “Onset of Nucleate Boiling and Active Nucleation Site Density During Subcooled Flow Boiling,” Journal of Heat Transfer, 124, pp. 717-728. Beer, H., 1979, “Interferometry and Holography in Nucleate Boiling,” in Boiling Phenomena, S. Van Stralen and R. Cole, Eds., Vol.2, pp. 821-843, Hemisphere, NY. Bell J.B., Colella P., 1989, “A Second-Order Projection Method for the Incompressible Navier-Stokes Equations,” Journal of Computational Physics, 85, pp. 257-283. Bergles, A.E., 1988, “Fundamentals of Boiling and Evaporation,” in Two-Phase Flow Heat Exchangers, (Kakac, S., Bergles, A.E., and Fernandes, O., Eds.), Kluwer, Dordrecht, The Netherlands, pp. 159-200. 161 Bergles, A.E., 1997, “Enhancement of Pool Boiling,” International Journal of Refrigeration, 20, pp. 545-551. Bhaga, D., and Weber, M.E., 1981, “Bubbles in Viscous Liquids: Shapes, Wakes and Velocities,” Journal of Fluid Mechanics, 105, pp. 61-85. Birdi, K.S (ed.), 2003, Handbook of Surface and Colloid Chemistry, 2nd ed., CRC Press, New York, NY. Bisio, P.D., Cartledge, J.G., Keesom, W.H., and Radke, C.J., 1980, “Molecular Orientation of Aqueous Surfactants on a Hydrophobic Solid,” Journal of Colloid and Interface Science, 78(1), pp. 225-234. Bitting D., and Harwell, J.H., 1985, “Effects of Counterions on Surfactant Surface Aggregates at the Alumina/Aqueous Solution Interface,” Langmuir, 3, pp.500-511. Blake, T.D., 1993, “Dynamic Contact Angles and Wetting Kinetics,” in Wettability, Berg, J.C., ed., Marcel Dekker, pp. 251-310. Brackbill, J.U., Kothe, D.B., Zemach, C., 1992, “A Continuum Method for Modeling Surface Tension,” Journal of Computational Physics, 100, pp. 335-354. Brebbia, C.A., Telles, J.C.F., and Wrobel, L.C., 1984, Boundary Element Techniques, Springer-Verlag, New York, NY. Brennen, C.E., 1995, Cavitation and Bubble Bubble Dynamics, Oxford University Press, Oxford, UK. Brown D.L., Cortez, R., and Minion L., 2001, “Accurate Projection Methods for the Incompressible Navier-Stokes Equations,” Journal of Computational Physics, 168, pp. 464-499. Carey, V.P., 1992, Liquid-Vapor Phase-Change Phenomena, Hemisphere, Washington, DC. Carreau, P.J., De Kee, D.C.R., and Chhabra, R.P., 1997, Rheology of Polymeric Systems: Principles and Applications, Hanser Gardner Publications, New York, NY. Cellosize Hydroxyethyl Cellulose, Union Carbide Corporation, 1998. Chandar, P., Somasundaran, P., Turro, J., 1987, “Fluorescence Probe Studies on the Structure of the Adsorbed Layer of Dodecyl Sulfate at the Alumina-Water Interface,” Journal of Colloid and Interface, 117(2), pp. 31-46. Chang, C.-H., and Franses, E.I., 1995, “Adsorption Dynamics of Surfactants at the Air/Water Interfaces: A Critical Review of Mathematical Models, Data, and 162 Mechanisms,” Colloids and Surfaces A: Physicochemical and Engineering Aspects, 100, pp. 1-45. Chang, C. S., Maa, J.R., and Yang, Y.M., 1987, “Dynamic Surface Tension and Nucleate Flow Boiling of Dilute Surfactant Solutions,” Journal of the Chinese Institute of Chemical Engineers, 18(1), pp. 125-130. Chaudhury, M.K., 2003, “Spread the Word about Nanofluids,” Nature, 423, pp. 131-132. Chaudhury, M.K., and Whitesides G.M., 1992, “How to Make Water Run Uphill," Science, 256, pp. 1539-1541. Chen, J.C., 2003, “Surface Contact – Its Significance for Multiphase Heat Transfer: Diverse Examples,” Journal of Heat Transfer, 125(4), pp. 549-566. Chhabra, R.P., and Richardson, J.F., 1999, Non-Newtonian Flow in the Process Industries: Fundamentals and Engineering Applications, Butterworth-Heinemann, Oxford, UK. Chou, C.C., and Yang, Y.M., 1991, “Surfactant Effects on the Temperature Profile within the Superheated Boundary Layer and the Mechanism of Nucleate Pool Boiling,” Journal of the Chinese Institute of Chemical Engineers, 22(2), pp. 71-80. Cole, R., and Rohsenow, W.M., 1969, “Correlations for Bubble Departure Diameters for Boiling of Saturated Liquids,” Chemical Engineering Progress, 65, pp.211-213. Colella, P., 1985, “A Direct Eulerian MUSCL Scheme for Gas Dynamics.” SIAM Journal on Scientific and Statistical Computing, 6, pp. 104-117. Colella, P., 1990, “A Multidimensional Second-Order Godunov Scheme for Conservation Laws,” Journal of Computational Physics, 87, pp. 171-200. Collier, J.G., and Thome, J.R., 1996, Convective Boiling and Condensation, 3rd, Oxford, UK. Cooper, M.G., and Lloyd, A.J.P., 1969, “The Microlayer in Nucleate Pool Boiling,” International Journal of Heat and Mass Transfer, 12, pp. 895-913. Cornell, B.A., Braach-Maksvytis, V.L.B., King, L.G., Osman, P.D.J., Raguse, B., Wieczorek, L., and Pace, R.J., 1997, “A Biosensor that Uses Ion-Channel Switches,” Nature, 387, pp. 580-583. Cuenot, B., Magnaudet, J., and Spennato, B., 1997, “The Effects of Slightly Soluble Surfactants on the Flow Around a Spherical Bubble.” Journal of Fluid Mechanics, 339, pp. 25-53. 163 Das, A.K., Kilty, H.P., Marto, P.J., Andeen, G.B., and Kumar, A., 2000, “The Use of an Organic Self-Assembled Monolayer Coating to Promote Dropwise Condensation of Steam on Horizontal Tubes,” Journal of Heat Transfer, 122, pp. 278-286. Dasgupta, S., Schonberg, J.A., Kim, I.Y., and Wayner, P.C. Jr, 1993, “Use of the Augmented Yong-Laplace Equation to Model Equilibrium and Evaporating Extended Menisci,” Journal of Colloid and Interface Science, 157, pp. 332-342. Dasgupta, S., Kim, I.Y., and Wayner, P.C. Jr, 1994, “Use of the Kelvin-Clapeyron Equation to Model Equilibrium and Evaporating Curved microfilm,” Journal of Heat Transfer, 117, pp. 394-401. Decker, E.L., Frank, B., Suo, Y., and Garoff, S., 1999, “Physics of Contact Angle Measurement,” Colloids and Surfaces A: Physicochemical and Engineering Aspects, 156, pp. 177-189. Delgado, Á.V. (Ed.), 2002, Interfacial Electrokinetics and Electrophoresis, Surfactant Science Series, Vol. 106. Marcel Dekker, New York, NY. Dhir, V.K., 1998, “Boiling Heat Transfer,” Annual Reviews of Fluid Mechanics, 30, pp. 365-401. Dobiáš, B., and Rybinski, W.V., 1999, “Adsorption at the Solid-Liquid Interface,” in Solid-Liquid Dispersions, (Dobiáš, B., Qiu, X., and Rybinski, W.V., Eds., Vol. 81, Surfactant Science Series), Marcel Dekker, New York, NY, pp. 318-470. Dukhin, S.S., Kretzschmar, G., and Miller, R., 1995, Dynamics of Adsorption at Liquid Interfaces, Elsevier, Amsterdam. Edwards, D.A., Brenner, H., and Wasan, D.T., 1991, Interfacial Transport Processes and Rheology, Ch.6, Butterworth-Heinemann, Stonheam, MA. Eggleton, C.D., and Stebe, J.S., 1998, “An Adsorption-Desorption-Controlled Surfactant on a Deforming Droplet,” Journal of Colloid and Interface Science, 208, pp. 68-80. Evans, D.F., and Wennerström, H., 1999, The Colloidal Domain - Where Physics, Chemistry, and Biology Meet, 2nd Ed., Wiley-VCH, New York, NY. Fainermann, V.B., Makievski, A.V., and Miller, R., 1993, “The Measurement of Dynamic Surface Tensions of Highly Viscous Liquids by the Maximum Bubble Pressure Method,” Colloids and Surfaces A: Physicochemical and Engineering Aspects, 75, pp. 229-235. Fedkiw, R., Aslam, T., Merriman, B., and Osher, S., 1999, “A Non-Oscillatory Eulerian Approach to Interfaces in Multimaterial Flows (The Ghost Fluid Method),” Journal of Computational Physics, 152, pp. 457-492. 164 Fedkiw, R.P., Merriman, B, Donat, R, Osher, S., 2002, “The Penultimate Scheme for Systems of Conservation Laws: Finite Difference ENO with Marquina’s Flux Splitting,” in Innovative Methods for Numerical Solutions of Partial Differential Equations, pp. 4985. Filippov, G.A. and Saltanov, G.A., 1982, “Steam-Liquid Media Heat-Mass Transfer and Hydrodynamics with Surface-Active Substance Additives,” Heat Transfer 1982, Hemisphere, Washington, Vol. 4, pp. 443-447. Fritz, W., 1935, “Maximum Volume of Vapor Bubbles,” Physik Zeitschr, 26, pp. 379384. Frost, W. and Kippenhan, C.J., 1967, “Bubble Growth and Heat Transfer Mechanisms in the Forced Convection Boiling of Water Containing a Surface Active Agent,” International Journal of Heat and mass Transfer, 10, pp. 931-949. Fuerstenau, D.W., 2002, “Equilibrium and Nonequilibrium Phenomena Associated with the Adsorption of Ionic Surfactants at Solid-Water Interfaces,” Journal of Colloid and Interface Science, 256, pp. 79-90. Glimm, J., McBryan, O., Melnikoff, R., and Sharp, D.H., 1987, “Front Tacking Applied to Rayleigh-Taylor Instability,” SIAM Journal on Scientific and Statistical Computing, 7, pp. 230-251. Gorenflo, D., Knabe, V., and Beiling, V., 1986, “Bubble Density on Surfaces with Nucleate Boiling – Its Influence on Heat Transfer and Burnout Heat Flux at Elevated Saturation Pressures,” Proceedings of the 8th International Heat Transfer Conference, Vol. 4, August 1986, San Francisco, CA, Hemisphere Publishing Corp., Washington, DC. pp. 1995-2000. Gu, Y., and Li, D., 2000, “The ζ - Potential of Glass Surface in Contact with Aqueous Solutions,” Journal of Colloid and Interface Science, 226, pp. 328-339. Gusev, I., and Horváth, C., 2002, “Streaming Potential in Open and Packed Fused-Silica Capillaries,” Journal of Chromatography A, 948, pp. 203-223. Han, C.H., and Griffith, P., 1965, “The Mechanism of Heat Transfer in Nucleate Pool Boiling,” International Journal of Heat and Mass Transfer, 8, pp. 887-914. Harlow, F.H., and Welch, J.E., 1965, “Numerical Calculation of Time-Dependent Viscous Incompressible Flow of Fluids with Free Surface,” The physics of Fluids, 8(12), pp. 2182-2189. 165 Harwell, J.H., Hoskins, J.C., Schechter, R.S., and Wade, W.H., 1985, “Pseudophase Separation Model for Surfactant Adsorption: Isomerically Pure Surfactants,” Langmuir, 1, pp.251-262. Hentze, H.-P., Khrenow, V., Tauer, K., 2002, “A New Approach towards Redispersable Polyelectrolyte-Surfactant Complex Nanoparticles,” Colloid Polymer Science, 280, pp. 1021-1026. Hetsroni, G., Zakin, J.L., Lin, Z., Mosyak, A., Pancallo, E.A., and Rozenblit, R., 2001, “The Effect of Surfactants on Bubble Growth, Wall Thermal Patterns and Heat Transfer in Pool Boiling,” International Journal of Heat and Mass Transfer, 44, pp. 485-497. Hibiki, T., and Ishii, M., 2003, “Active Nucleation Site Density in Boiling Systems,” International Journal of Heat and mass Transfer, 46, pp. 2587-2601. Hirt, C.W., Amsden, A.A., and Cook, J.L., 1974, “An Arbitrary Lagrangian-Eulerian Computing Method for All Flow Speeds,” Journal of Computational Physics, 14, pp. 227-253. Hirt, C.W., and Nichols, B.D., 1981, “Volume of Fluid (VOF) Method for the Dynamics of Free Boundaries,” Journal of Computational Physics, 39, pp. 201-225. Hirt, D.E., Prud’homme, K., Miller, B., and Rebenfeld, L. 1990, “Dynamic surface tension of hydrocarbon and fluorocarbon surfactant solutions using the maximum bubble pressure method,” Colloids and Surfaces, 44, pp. 101-117. Hoffmann, H., and Rehage, H., 1986, “Rheology of Surfactant Solutions,” in Surfactant Solutions-New Methods of Investigation (Zana, R., Ed.), Vol. 22, Marcel Dekker, New York, pp. 209-239. Holmberg, K., Jönsson, B., Kronberg, B., and Lindman, B., 2003, Surfactants and Polymers in Aqueous Solution, 2nd Ed., Wiley, New York, NY. Hu, R.Y.Z., 1989, Nucleate Pool Boiling from a Horizontal Wire in Viscoelastic Fluid, Ph.D. Thesis, University of Illinois at Chicago, Chicago, IL. Hu, R.Y.Z., Wang, T.A.A., and Hartnett, J.P., 1991, “Surface Tension Measurement of Aqueous Polymer Solutions,” Experimental Thermal and Fluid Science, 4, pp. 723-729. Huber, D.L., Manginell, R.P., Samara, M.A., Kim, B.-I., Bunker, B.C., 2003, “Programmed Adsorption and Release of Proteins in a Microfluidic Device,” Science, 301, pp. 352-354. Hui, W.H., Li, P.Y., and Li, Z.W., 1999, “A Unified Coordinate System for Solving the Two-Dimensional Euler Equations,” Journal of Computational Physics, 153, pp. 596637. 166 Hunter, R.J., 1981, Zeta Potential in Colloid Science-Principles and Applications, Academic Press, New York, NY. Hunter, R.J., 2001, Foundations of Colloid Science, 2nd Ed., Oxford University Press, Oxford, UK. Huplik, V. and Raithby, G.D., 1972, “Surface-Tension Effects in Boiling from a Downward-Facing Surface,” Journal of Heat Transfer, 94, pp. 403-409. Iliev, Tz. H., and Dushkin, C.D., 1992, “Dynamic Surface Tension of Micellar Solutions Studied by the Maximum Bubble Pressure Method,” Colloid and Polymer Science, 270, pp. 370-376. Ishiguro, S., and Hartnett, J.P., 1992, “Surface Tension of Aqueous Polymer Solutions,” International Communications in Heat and Mass Transfer, 19, pp. 285-295. Janule, V.P., 1998, “Measuring Surface Tension of Coatings with Automatic Viscosity Compensation,” SensaDyne Corporation Technical Report, Fairfax, VA. Jontz, P.D. and Myers, J.E., 1960, “The Effect of Dynamic Surface Tension on Nucleate Boiling Coefficients,” AIChE Journal, 6(1), pp. 34-38. Joos, P., 1999, Dynamic Surface Phenomena, Utrecht, The Netherlands. Judd, R.L., and Hwang, K.S., 1976, “A Comprehensive Model for Nucleate Heat Transfer Including Microlayer Evaporation,” Journal of Heat Transfer, 98, pp. 623-629. Jung, C., and Bergles, A.E., 1989, Evaluation of Commercial Enhanced Tubes in Pool Boiling, Tropical Report No. DE-FC07-88ID 12772, Rensselaer Polytechnic Institute, Troy, NY. Kandlikar, S.G., Shoji, M., and Dhir, V.K. (Eds.), 1999, Handbook of Phase Change: Boiling and Condensation, Taylor & Francis, Philadelphia, PA. Kenning, D.B.R., 1999, “What Do We really Know about Nucleate Boiling,” in ImechE: Trans 6th UK National Heat Transfer Conference, Edinburgh, 15-16 September, pp. 143167. Kobayashi, R., 1993, “Modeling and Numerical Simulations of Dendritic Crystal Growth,” Physica D, 63, pp. 410-423. Koopal, L.K., and Ralston, J., 1986, “Chain Length Effects in the Adsorption of Surfactants at Aqueous Interfaces: Comparison of Existing Adsorption Models with a New Model,” Journal of Colloid and Interface Science, 112(2), pp. 362-379. 167 Kotchaphakdee, P. and Williams, M.C., 1970, “Enhancement of Nucleate Pool Boiling with Polymeric Additives,” International Journal of Heat and Mass Transfer, 13, pp. 835-848. Kwok, D.Y., and Neumann, A.W., 1999, “Contact Angle Measurement and Contact Angle Interpretation,” Advances in Colloid and Interface Science, 81, pp. 167-249. Lay, J.H., and Dhir, V.K., 1995, “Shape of a Vapor Stem During Nucleate Boiling of Saturated Liquids,” Journal of Heat Transfer, 117, pp. 394-401. Lee, R.C., and Nydahl, J.E., 1989, “Numerical Calculation of Bubble Growth in Nucleate Boiling From Inception through Departure,” Journal of Heat Transfer, 111, pp. 474-479. Levitskiy, S. P., Khusid, B. M., and Shul’man, Z. P., 1996, “Growth of Vapor Bubbles in Boiling Polymer Solutions-II. Nucleate Boiling Heat Transfer,” International Journal of Heat and Mass transfer, 39(3), pp. 639-644. Levitz, P.E., 2002, “Adsorption of Non Ionic Surfactants at the Solid/Water Interface,” Colloids and Surfaces A: Physicochemical and Engineering Aspects, 205, pp. 31-38. Liao, S.P. and Dhir, V.K., 1989, “Void Fraction Measurement During Saturated Pool Boiling of Water on Partially Wetted Surfaces,” Journal of Heat Transfer, 111, pp. 731738. Liao, Y., and McLaughlin, J.B., 2000, “Bubble Motion in Aqueous Surfactant Solutions,” Journal of Colloid Interface Science, 224(200), pp. 297-310. Lin, H.S., Wu, W.T., Yang, Y.M., and Maa, J.R., 1994, “Effect of Surfactant Additive on Boiling in the Case of Quenching in Subcooled Water,” in Multiphase Flow and Heat Transfer (Chen, X.J., Chen, T.K., and Zhou, F.D., Eds.), Begell House, New York, NY, pp. 464-473. Liu, C., and Shen, J., 2003, “A Phase Field Model for the Mixture of Two Incompressible Fluids and its Approximation by a Fourier-Spectral Method,” Physica D, 179, pp. 211228. Liu, T., Cai, Z., and Lin, J., 1990, “Enhancement of Nucleate Boiling Heat Transfer with Additives,” in Heat Transfer Enhancement and Energy Conservation, Hemisphere, New York, pp. 417-424. Lu, Y., Fan, H., Stump, A., Ward, T.L., Rieker, T., and Brinker, C.J., 1999, “AerosolAssisted Self-Assembly of Mesostructured Spherical Nanoparticles,” Nature, 398, pp. 223-226. Lyklema, J., 1991, “Fundamentals of Interface and Colloid Science,” Academic Press, London, UK. 168 Maestro, A., Gonález, C., and Gutiérrez, J.M., 2002, “Shear Thinning and Thixotropy of HMHEC and HEC Water Solutions,” Journal of Rheology, 46(6), pp. 1445-1457. Manglik, R.M., 1998, “Pool Boiling Characteristics of High Concentration Aqueous Surfactant Emulsions”, Heat Transfer 1998, KSME, Seoul, Vol. 2, pp. 449-453. Manglik, R.M., 2003, “Heat Transfer Enhancement,” in Heat Transfer Handbook (Bejan, A. and Klaus, A.D., Eds.), Wiley, New York, Ch. 14. Manglik, R.M., Bahl, M., Vishnubhatla, S., and Zhang, J., 2003, Interfacial and Rheological Characterization of Aqueous Surfactant and Polymer Solutions, ThermalFluids & Thermal Processing Laboratory Report No. TFTPL-9, University of Cincinnati, Cincinnati, OH. Manglik, R.M., Wasekar, V.M., and Zhang, J., 2001, “Dynamic and Equilibrium Surface Tension of Aqueous Surfactant and Polymeric Solutions,” Experimental Thermal and Fluid Science, 25, pp. 55-64. Manne, S., and Gaub, H.E., 1995, “Molecular Organization of Surfactants at Solid-Liquid Interfaces,” Science, 270, pp. 1480-1483. Miaw, C.B., 1978, A Study of Heat Transfer to Dilute Polymer Solutions in Nucleate Pool Boiling, Ph.D. Thesis, University of Michigan, Ann Arbor, MI. Mikic, B.B., and Rohsenow, W.M., 1969, “A New Correlation of Pool-Boiling Data Including Effect of Heating Surface Characteristics,” Journal of Heat Transfer, 9, pp. 245-250. Miller, C.A., and Neogi, P., 1985, Interfacial Phenomena: Equilibrium and Dynamic Effects, Ch. 4, Marcel Dekker, New York, NY. Moffatt, R.J., 1988, “Describing the Uncertainties in Experimental Results,” Experimental Thermal and Fluid Science, 1(1), pp. 3-17. Mordecai, E., 1930, Methods of Correlation Analysis, Wiley & Sons, New York, NY. Morgan, A.I., Bromley, L.A., and Wilke, C.R., 1949, “Effect of Surface Tension on Heat Transfer in Boiling,” Industrial and Engineering Chemistry, 41(12), pp. 2767-2769. NATROSOL- Hydroxyethylcellulose, A Nonionic Water-Soluble Polymer, Physical and Chemical Properties, Hercules Incorporated, 1999. Nelson, R.A., 2001, “Do We Doubt too Little? Examples from the Thermal Sciences,” Experimental Thermal and Fluid Science, 25, pp. 255-267. 169 Nelson, R.A., Kenning, D.B.R., and Shoji, M., 1996, “Nonlinear Dynamics in Boiling Phenomena,” Japan Heat Transfer Society Journal, 35(136), pp. 22-34. Nishikawa, K., Fujita, Y., and Ohta, H., 1984, “Effect of Surface Configuration on Nucleate Heat Transfer,” International Journal of Heat and Mass Transfer, 27, pp. 15591571. Osher, S., and Sethian, J.A., 1988, “Fronts Propagating with Curvature Dependent Speed: Algorithms Based on Hamilton-Jacobi Formulations,” Journal of Computational Physics, 79, pp. 12-49. Ottewill, R.H., 1967, “Effect of Nonionic Surfactants on the Stability of Dispersions,” in Nonionic Surfactants, (Schick, M.J., Ed., Ch. 19, Vol. 2, Surfactant Science Series), Marcel Dekker, New York, NY, pp. 627-681. Palaparthi, R., 2001, Effect of Surfactant Transport on the Mobility of Bubbles in Liquids - an Experimental and Computational Study, Ph.D. Thesis, The City University of New York, New York, NY. Palmer, B.J., and Rector, D.A., 2000, “Lattice-Boltzmann Algorithm for Simulating Thermal Two-Phase Flow,” Physical Review E, 61(5), pp. 5295-5306. Partearroyo, M.A., Alonso, A., Goni, F.M., Tribout, M., and Paredes, S., 1996, “Solubilization of Phospholipid Bilayers by Surfactants Belonging to the Triton X Series: Effect of Polar Group Size,” Journal of Colloid and Interface Science, 178, pp. 156-159. Patil, R.K., 1991, A Numerical Model for a Bubble in Nucleate Pool Boiling, Ph.D. Thesis, Iowa State University, Ames, Iowa. Paul, D.D., and Abdel-Khalik, S.I., 1983, “Nucleate Boiling in Drag-Reducing Polymer Solutions,” Journal of Rheology, 27(1), pp. 59-76. Peng, D., Merriman, B., Osher, S., Zhao, H., Kang, M., 1999, “A PDE-Based Fast Local Level Set Method,” Journal of Computational Physics, 155, pp. 410-438. Persson, B., Nilsson, S., Sundelöf, L.-O., 1996, “On the Characterization Principles of Some Technically Important Water-Soluble Nonionic Cellulose Derivatives. Part II: Surface Tension and Interaction with a Surfactant,” Carbohydrate Polymers, 29, pp. 119127. Plesset, M.S., and Zwick, S.A., 1954, “The Growth of Vapor Bubbles in Superheated Liquids,” Journal of Applied Physics, 25(4), pp. 493-500. Podsushnyy, A.M., Minayev, A.N., Statsenko, V.N., and Yakubovskiy, Yu.V., 1980, “Effect of Surfactants and of Scale Formation on Boiling Heat Transfer to Sea Water,” Heat transfer - Soviet Research, 12(2), pp. 113-114. 170 Porter, M.R., 1994, Handbook of Surfactants, 2nd ed., Blackie, London, UK. Prosperetti, A., Plesset, M., 1978, “Vapor-Bubble Growth in a Superheated Liquid,” Journal of fluid Mechanics, 85(2), pp. 340-368. Puckett, E.G., Almgren, A.S., Bell, J.B., Marcus, D.L., and Rider, W.G., 1997, “A HighOrder Projection Method for Tracking Fluid Interfaces in Variable Density Incompressible Flows,” Journal of Computational Physics, 130, pp. 269-282. Qiao, Y.M. and Chandra, S., 1997. “Experiments on Adding a Surfactant to Water Droplets Boiling on a Hot Surface” Proceedings of the Royal Society of London, A 453, pp. 673-689. Razafindralambo, H., Blecker, C., Delhaye, S., and Paquot, M., 1995, “Application of the Quasi-Static Mode of the Drop Volume Technique to the Determination of Fundamental Surfactant Properties,” Journal of Colloid and Interface, 174, pp. 373-377. Rhie, C.M., and Chow, W.L., 1983, “Numerical Study of Turbulent Flow Past an Airfoil with Trailing Edge Separation,” AIAA Journal, 21, pp. 1525-1532. Richard, C., Balavoine, F., Schultz, P., Ebbesen, T.W., and Mioskowski, “Supramolecular Self-Assembly of Lipid Derivatives on Carbon Nanotubes,” Science, 300, pp. 775-778. Rohsenow, W.M., 1952, “A Method of Correlating Heat Transfer Data for Surface Boiling of Liquids," Trans ASME, 74, pp. 969-976. Roll, J.B. and Myers, J.E., 1964, “The Effect of Surface Tension on Factors in Boiling Heat Transfer,” AIChE Journal, 10(4), pp. 530-534. Rosen, M.J., 1989, Surfactants and Interfacial Phenomena, 2nd Ed., Wiley, New York, NY. Ryskin, G., and Leal, L.G., 1984, “Numerical Solutions of Free-Boundary Problems in Fluid Mechanics,” Journal of Fluid Mechanics, 148, pp. 1-17. Sakagami, K., Yoshimura, T., and Esumi, K., 2002, “Simultaneous Adsorption of Poly(1vinylpyrrolidone-co-acrylic acid) and Sodium Dodecyl Sulphate at Alumina/Water Interface,” Langmuir, 18, pp. 6049-6053. Saltanov, G.A., Kukushkin, A.N., Solodov, A.P., Sotskov, S.A., Jakusheva, E.V., Chempik, E., 1986, “Surfactant Influence on Heat Transfer at Boiling and Condensation,” Heat Transfer 1986, Hemisphere, Washington, Vol. 5, pp. 2245-2250. 171 Scamehorn, J.F., Schechter, R.S., and Wade, W.H., 1981, “Adsorption of Surfactant on Mineral Oxide Surfaces from Aqueous Solutions,” Journal of Colloid and Interface, 85(2), pp. 463-501. Schulz, J.C., Warr, G.G., Butler, P.D., and Hamilton, W.A., 2001, “Adsorbed Layer Structure of Cationic Surfactants on Quantz,” Physical Review E, 63, Paper No. 041604. Scriven, L.E. and Sternling, C.V., 1960, “The Marangoni Effects,” Nature, 187, pp. 186188. Sethian, J.A., 1990, “Numerical Algorithms for Propagating Interfaces: Hamilton-Jacobi Equations and Conservation Laws,” Journal of Differential Geometry, 32, pp. 131-161. Shah, B.H. and Darby, R., 1973, “The Effect of Surfactant on Evaporative Heat Transfer in Vertical Film Flow,” International Journal of Heat and mass Transfer, 16, pp. 18891903. Sher, I., and Hetsroni, G., 2002, “An Analytical Model for Nucleate Pool Boiling with Surfactant Additives,” International Journal of Multiphase Flow, 28, pp. 699-706. Shibayama, S., Katsuta, M., Suzuki, K., Kurose, T., and Hatano, Y., 1980, “A Study on Boiling Heat Transfer in a Thin Liquid Film,” Heat Transfer (Japan Research), 9(4), pp. 12-40. Shul’man Z.P., Khusid, B.M., and Levitskiy, S.P., 1993, “Special Features of Boiling of Macromolecular Polymer Solutions,” Heat Transfer Research, 25(7), pp. 872-878. Shul’man, Z. P., and Levitskiy, S. P., 1996, “Growth of Vapor Bubbles in Boiling Polymer Solutions-I. Rheological and Diffusional Effects,” International Journal of Heat and Mass transfer, 39(3), pp. 631-638. Shyy, W., Udaykumar, H.S., Rao, M.M., and Smith, R.W., 1996, Computational Fluid Dynamics with Moving Boundaries, Taylor & Francis, Philadelphia, PA. Somasundaran, P., and Fuerstenau, D.W., 1966, “Mechanisms of Alkyl Sulfonate Adsorption at the Alumina-Water Interface,” Journal of Physical Chemistry, 70(1), pp. 90-96. Son, G., Dhir, V.K., and Ramanujapu, N., 1999, “Dynamics and Heat Transfer Associated With a Single Bubble During Nucleate Boiling on a Horizontal Surface,” Journal of Heat Transfer, 121, pp. 623-631. Stroebe, G.W., Baker, E.M., and Badger, W.L., 1939, “Boiling-Film Heat Transfer Coefficients in a Long-Tube Vertical Evaporator,” Industry and Engineering Chemistry, 31(2), pp. 200-206. 172 Sussman, M., Almgren, A. S., Bell, J. B., Colella, P., Howell, L. H., and Welcome, M. L., 1999, “ An Adaptive Level Set Approach for Incompressible Two-Phase Flows,” Journal of Computational Physics, 148, pp. 81-124. Sussman, M., and Fatemi, E., 1999, “An Efficient, Interface-Preserving Level Set Redistancing Algorithm and its Application to Interfacial Incompressible Fluid Flow,” SIAM Journal on Scientific Computing, 20(4), pp. 1165-1191. Sussman, M., Smereka, P., Osher, S., 1994, “A Level Set Approach for Computing Solutions to Incompressible Two-Phase Flow,” Journal of Computational Physics, 114, pp. 146-159. Takada, N., Misawa, M., Tomiyama, A., and Fujiwara, S., 2000, “Numerical Simulation of Two- and Three-Dimensional Two-Phase Fluid Motion by Lattice Boltzmann Method,” Computer Physics Communications, 129, pp. 233-246. Tan, Y.K., and Wang, H.T., 1994, “Solid Additives for Enhancing Nucleate Pool Boiling of Binary Mixtures on a Smooth and a Gewa-T Tube,” Institute of Chemistry Engineering Symposium Series, 135(6), pp. 117-122. Tatebe O., 1993, “The Multigrid Preconditioned Conjugate Gradient Method,” In 6th Copper Mountain Conference on Multigrid Methods, Copper Mountain, CO, pp. 621634. Thomas, O.C., Cavicchi, R.E., and Tarlov, M.J., 2003, “Effect of Surface Wettability on Fast Transient Microboiling Behaviour,” Langmuir, 19, pp. 6168-6177. Thome, J.R., 1990, Enhanced Boiling Heat Transfer, Hemisphere, New York, NY. Tornberg, A.-K. 2000, Interface Tracking Methods with Application to Multiphase Flows, PhD Thesis, NADA, KTH, Stockholm, Sweden. Tsujii, K., 1998, Surface Activity: Principles, Phenomena, and Applications, Academic, San Diego, CA. Tzan, Y.L., and Yang, Y.M., 1990, “Experimental Study of Surfactant Effects on Pool Boiling Heat Transfer,” Journal of Heat Transfer, 112, pp. 207-212. Ulicny, J.C., 1984, Nucleate Pool Boiling in Dilute Polymer Solutions, Ph.D. Thesis, University of Michigan, Ann Arbor. Unverdi, S.O., and Tryggvason, G., 1992, “A Front-Tracking Method for Viscous, Incompressible Multi-Fluid Model,” Journal of Computational Physics, 100, pp. 25-37. 173 Vanjara, A.K., and Dixit S.G., 1996, “Adsorption of Alkyltrimethylammonium Bromide and Alkylpyridinium Chloride Surfactant Series on Polytetrafluoroethylene Powder,” Journal of Colloid and Interface Science, 177, pp. 359-363. Wang, Y., 1999, A Theoretical Study of Bubble Motion in Surfactant Solutions, Ph.D. Thesis, New Jersey Institute of Technology, New Jersey. Wang, A.T.A., and Hartnett, J.P., 1992, “Influence of Surfactants on Pool Boiling of Aqueous Polyacrylamide Solutions,” Warme- und Stoffubertragung, 27, pp. 245-248. Wang, C.H., and Dhir, V.K., 1993, “Effect of Surface Wettability on Active Nucleation Site Density During Pool Boiling of Water on A Vertical Surface,” Journal of Heat Transfer, 115, pp. 659-669. Wang, T.A.A., and Hartnett, J.P., 1994, “Pool Boiling Heat Transfer from a Horizontal Wire to Aqueous Surfactant Solutions,” Heat Transfer 1994, I Chem. E, UK, Vol. 5, pp. 177-182. Wasan, D.T., and Nikolov, A.D., 2003, “Spreading of Nanofluids on Solids,” Nature, 423, pp. 156-159. Wasekar, V.M. and Manglik, R.M., 2003, “Short-Time-Transient Surfactant Dynamics and Marangoni Convection Around Boiling Nuclei,” Journal of Heat Transfer, 125(5), pp. 858-866. Wasekar, V.M., and Manglik, R.M., 1999, “A Review of Enhanced Heat Transfer in Nucleate Pool Boiling of Aqueous Surfactant and Polymeric Solutions,” Journal of Enhanced Heat Transfer, 6, pp. 135-150. Wasekar, V.M., and Manglik, R.M., 2000, “Pool Boiling Heat Transfer in Aqueous Solutions of an Anionic Surfactant,” Journal of Heat Transfer, 122, pp. 708-715. Wasekar, V.M., and Manglik, R.M., 2001, Nucleate Pool Boiling Heat Transfer in Aqueous Surfactant Solutions, Thermal-Fluids & Thermal Processing Laboratory Report No. TFTPL-4, University of Cincinnati, Cincinnati, OH. Wasekar, V.M., and Manglik, R.M., 2002, “The Influence of Additive Molecular Weight and Ionic Nature on the Pool Boiling Performance of Aqueous Surfactant Solutions,” International Journal of Heat and Mass Transfer, 45, pp. 483-493. Wayner PC Jr. Evaporation and Stress in the Contact Line Region, Proceedings of the Engineering Foundation Conference on Pool and External Flow Boiling 1992. VK Dhir and AE Bergles eds., Santa Barbara, CA; 251-256. Welch, S.J., 1998, “Direct Simulation of Vapor Bubble Growth,” International Journal of Heat Mass Transfer, 41, pp. 1655-1666. 174 Welch, S.J., and Wilson, J., 2000, “A Volume of Fluid Based Method for Fluid Flows with Phase Change,” Journal of Computational Physics, 160, pp. 662-682. Wen, D.S., and Wang, B.X., 2002, “Effect of Surface Wettability on Nucleate Pool Boiling Heat Transfer for Surfactant Solutions,” International Journal of Heat and mass Transfer, 45, pp. 1739-1747. Wheeler, A.A., Murray, B.T., and Schaefer, R.J., 1993, “Computations of Dendrites Using a Phase Field Model,” Physica D, 66, pp. 243-262. Wu, W.T., and Yang, Y.M., 1992, “Enhanced Boiling Heat Transfer by Surfactant Additives,” Pool and External Flow Boiling, ASME, New York, pp. 361-366. Wu, W.T., Lin, H.S., And Yang, Y.M., 1994, “Critical Heat Flux in Pool Boiling of Aqueous Surfactant Solutions as Determined by the Quenching Method,” International Journal of Heat and mass Transfer, 37(15), pp. 2377-2379. Wu, W.T., Yang, Y.M., and Maa, J.R., 1995, “Enhancement of Nucleate Boiling Heat transfer and Depression of Surface Tension by Surfactant Additives,” Journal of Heat Transfer, 117, pp. 526-529. Wu, W.T., Yang, Y.M., and Maa, J.R., 1998a, “Nucleate Pool Boiling Enhancement by Means of Surfactant Additives,” Experimental Thermal and Fluid Science, 18, pp. 195209. Wu, W.T., Yang, Y.M., and Maa, J.R., 1998b, “Effect of Surfactant Additive on Pool Boiling of Concentrated Lithium Bromide Solution,” International Communications in Heat and Mass Transfer, 25(8), pp. 1127-1134. Wu, W.T., Yang, Y.M., and Maa, J.R., 1999, “Pool Boiling Incipience and Vapor Bubble Growth Dynamics in Surfactant Solutions,” International Journal of Heat and mass Transfer, 42, pp. 2483-2488. Wu,W.T., Hu, C.L., and Yang, Y.M., 1993 “Surfactant Effect on Boiling Incipience and Bubble Growth Dynamics of Surface Boiling in Water,” Journal of the Chinese Institute of Chemical Engineers, 24(2), pp. 111-118. Yang, Y.M., 1990, “Dynamic Surface Effect on Boiling of Aqueous Surfactant Solutions,” International Communications in Heat and Mass Transfer, 17, pp. 711-727. Yang, Y.M., and Maa, J.R., 1982, “Effects of Polymer Additives on Pool Boiling Phenomena,” Letters in Heat and Mass Transfer, 9, pp. 237-244. Yang, Y.M., and Maa, J.R., 1983, “Pool Boiling of Dilute Surfactant Solutions,” Journal of Heat Transfer, 105, pp. 190-192. 175 Yang, Y.M., Lin, C.Y., Liu, M.H., and Maa, J.R., 2002, “Lower Limit of the Possible Nucleate Pool Boiling Enhancement by Surfactant Addition to Water,” Journal of Enhanced Heat Transfer, 9(3-4), pp. 153-160. Yang, Y.M., Lin, C.Y., Liu, M.H., Chien, C.L., and Maa, J.R., 2000, “Surface-Tension Effects on Nucleate Pool Boiling in Aqueous Cationic Surfactant Solutions,” Journal of Chinese Colloid & Interface Society, 22(1), pp. 9-18. Yoon, H.Y., Koshizuka, S., Oka, Y., 2001, “Direct Calculation of Bubble Growth, Departure, and Rise in Nucleate Pool Boiling,” International Journal of Multiphase Flow, 27, pp. 277-298. Zhao, Y-H, Masuoka, T., and Tsuruta, T., 2002, “Unified Theoretical Prediction of Fully Developed Nucleate Boiling and Critical Heat Flux Based on a Dynamic Microlayer Model,” International Journal of Heat and Mass Transfer, 45, pp. 3189-3197. Zhang, J., and Manglik, R.M., 2003, “Visualization of Ebullient Dynamics in Aqueous Surfactant Solutions,” Journal of Heat Transfer, 125(4), p. 547. 176 APPENDIX A SURFACE TENSION σ (mN/m) DATA A.1 AQUEOUS SURFACTANT SOLUTIONS AQUEOUS CTAB SOLUTIONS T = 23°C T = 80°C σ (mN/m) σ (mN/m) (50ms) (Equilibrium) 59.2 58.7 C (wppm) σ (mN/m) 50 (50ms) 69.1 σ (mN/m) (Equilibrium) 67.5 100 67.7 60.1 55.2 52.1 200 65.7 49.9 50.4 43.8 300 62.9 43.1 48.4 40.1 400 60.7 39.2 46.5 37.9 500 57.9 38.4 45.2 36.5 600 - 38.4 - 35.9 700 54.6 38.4 42.6 35.9 800 53.4 38.3 - - 900 - 38.3 40.5 35.8 1100 52.6 38.2 39.1 35.7 1300 52.4 38.2 38.7 35.7 AQUEOUS DTAC SOLUTIONS 50 71.9 71.9 - - 100 - 70.9 - 58.4 200 71.4 69.9 - - 300 71.3 68.9 - - 177 400 - 68.2 - - 500 - 67.5 59.6 53.4 600 - 66.4 - - 700 - 65.6 - - 900 - 63.1 - - 1000 66.5 - 57.8 - 1100 - 60.5 - - 1300 - 59.2 - - 1500 - 56.8 55.6 47.2 2000 59.9 54.5 - - 2500 - 52.5 - 42.2 3000 55.1 50.2 50.4 40.2 3500 - 48.1 - - 4000 50.1 - - 38.2 4500 - 44.7 - - 5000 - - 44.7 37.1 5500 - 42.5 - - 6000 - - - 37.0 6500 43.2 41.7 41.9 - 7000 - - - 37.0 7500 - 41.6 - - 8000 42.4 - - - 8500 - 41.6 39.6 - 178 AQUEOUS Ethoquad O/12 PG SOLUTIONS T = 23°C σ T = 80°C C C σ (wppm) (Equilibrium) (wppm) 70 32.87 31.41 (50ms) 58.48 67.6 98.6 52.5 65.7 53.8 69.2 96.87 62 164.3 51.3 98.6 51.8 262.9 68.9 131.58 59 262.9 49.4 197.2 50.3 460.13 68.4 219.3 52.3 460.1 48.4 295.8 48.2 657.33 66.1 282.96 49.9 558.7 47.5 394.4 45 953.13 64 379.59 46.3 657.3 47 424.4 43.3 1248.9 62.4 471.6 44.7 953.1 44.3 591.6 40.8 1544.7 61.1 526.31 43.3 1248.9 41.7 788.8 38.6 1840.5 59.9 660.24 42.8 1544.7 39.6 1320.49 37.8 2136.3 58.7 848.88 42.3 1906.27 39.5 1037.53 41.8 1162.5 41.6 1320.49 41.6 1603.45 41.6 C (wppm) 32.87 (50ms) 98.6 69.5 164.3 71.1 σ 55.9 C σ (wppm) (Equilibrium) 32.87 56.1 AQUEOUS Ethoquad 18/25 PG SOLUTIONS 35.27 71.3 35.27 63.2 35.27 54.1 35.27 47.9 105.8 70 70.53 60 105.8 52.1 70.53 45.8 179 35.27 71.3 35.27 63.2 35.27 54.1 35.27 47.9 105.8 70 70.53 60 105.8 52.1 70.53 45.8 317.4 66.9 105.8 58.6 211.6 50.6 105.8 44.6 634.81 62.1 211.6 56 317.4 49 211.6 43.3 952.21 60.1 317.4 55 634.8 47.1 423.21 42.6 1058.01 59.6 423.2 54.2 952.2 46 740.61 42.3 1163.82 59.2 634.8 53 1375.4 43.9 1058.01 42.3 1375.42 58.3 846.4 52.5 2080.76 42.7 1728.09 57.4 1058.0 51.9 3138.78 42.6 2080.76 56.8 1269.62 51.5 6665.49 42.6 2186.56 56.5 1622.29 51.1 2503.97 56 2327.63 50.3 3561.98 55 2458.8 50.2 5325.34 53.8 3032.97 50.2 8852.05 52.4 4090.99 50.2 180 A.2 AQUEOUS POLYMER SOLUTIONS AQUEOUS HEC QP-300 SOLUTIONS T = 23°C σ (mN/m) T = 80°C σ (mN/m) σ (mN/m) (50ms) (Equilibrium) 62.5 61.5 C (wppm) σ (mN/m) 50 (50ms) 72.4 (Equilibrium) 72.3 100 72.3 72 62.5 61.3 200 71.8 69.8 62.1 60.4 300 71.2 68.6 - 59.9 400 70.5 68 60.7 59.1 500 70.1 67.5 60.3 58.7 600 - 67.3 - - 800 69.1 67.2 59.3 57.5 1000 69 67.1 58.7 57.3 1200 68.9 67 58.5 57.2 1600 68.8 66.9 58.3 57.2 2000 68.8 66.8 58.4 57.1 AQUEOUS Carbopol 934 SOLUTIONS 100 72.4 71.7 62.6 62.3 200 - - - 61.8 250 72.5 71.1 62.5 - 181 300 - - - 61.1 400 - 70.7 - 61 500 72.5 70.5 62.5 60.8 700 - 70.2 - 60.5 800 - 70.1 - - 900 - 70 - 60.1 1000 72.5 69.9 62.4 - 1200 - 69.7 - - 1300 - 69.5 - 59.9 1500 72.7 69.4 62.5 - 1800 - 69.3 - - 2000 72.9 69.2 62.4 59.9 A.3 DYNAMIC SURFACE TENSION WITH TIME Aqueous SDS Solutions C = 1250wppm Surface Age σ (mN/m) (s) 0.026 69.3 C = 2500wppm Surface Age σ (mN/m) (s) 0.028 50.7 C = 5000wppm Surface Age σ (mN/m) (s) 1.2 38 0.031 62.6 0.037 45.2 0.32 38.3 0.042 58.3 0.041 43.9 0.183 38.6 182 0.026 69.3 0.028 50.7 1.2 38 0.031 62.6 0.037 45.2 0.32 38.3 0.042 58.3 0.041 43.9 0.183 38.6 0.057 55.6 0.058 42.1 0.09 39.3 0.092 54 0.088 41.1 0.047 40.5 0.166 52.6 0.330 40.0 0.037 41.9 0.214 52.3 1.350 39.8 0.027 45 0.354 51.9 1.653 51.7 Aqueous CTAB Solutions C = 200wppm 1.3 48.4 C = 400wppm 1.29 39 C = 800wppm 1.10 38.3 0.324 48.6 0.339 39.6 0.356 38.3 0.25 49.1 0.266 40 0.240 38.7 0.186 50.4 0.200 40.9 0.181 39.4 0.108 53.3 0.135 43.5 0.147 40 0.056 57.7 0.097 45.7 0.104 41.4 0.040 61.4 0.065 48.9 0.070 43.5 0.040 53.6 0.040 48.0 Aqueous TRITON X-305 Solutions C = 500wppm Surface Age σ (mN/m) (s) C = 1000wppm Surface Age σ (mN/m) (s) 183 C = 2000wppm Surface Age σ (mN/m) (s) 1.5 53 1.5 49.4 1.5 47.9 0.9 53 0.863 49.4 0.821 48 0.5 53.2 0.5 49.5 0.5 48.2 0.348 53.4 0.357 49.8 0.33 48.5 0.277 54.1 0.27 50.5 0.26 48.9 0.2 55.9 0.21 51.5 0.21 49.5 0.153 57.5 0.15 53.2 0.151 50.7 0.117 59.2 0.113 54.8 0.114 52.4 0.07 62 0.07 57.9 0.07 54.7 0.04 64.8 0.04 61.1 0.04 58.1 Aqueous TRITON X-100 Solutions C = 100wppm 1.4 35.7 C = 200wppm 1.4 34.1 C = 400wppm 1.4 33.1 0.6 35.9 0.6 34.3 0.6 33.2 0.45 36.1 0.45 34.4 0.45 33.4 0.35 36.8 0.335 34.9 0.33 33.9 0.27 37.6 0.25 35.4 0.25 34.5 0.18 39.2 0.18 36.6 0.18 35.3 0.12 41.4 0.12 38.3 0.13 36.4 0.097 42.7 0.097 39.2 0.09 37.9 0.07 44.6 0.07 41.5 0.06 40.3 0.04 48.3 0.04 45.5 0.04 42.6 184 Aqueous Ethoquad 18/25 Solutions C = 500wppm σ C = 1000wppm (mN/m) 52.5 Surface Age (s) 3.2 1.60 52.5 0.80 C = 2000wppm σ (mN/m) 51.1 Surface Age (s) 3.0 1.6 51.1 52.8 0.8 0.585 53 0.392 σ C = 4000wppm σ (mN/m) 49.3 Surface Age (s) 3.1 (mN/m) 48.8 1.5 49.5 1.53 48.8 51.3 0.80 49.6 0.8 48.9 0.558 51.6 0.562 49.9 0.55 49.2 54 0.375 52.3 0.368 50.5 0.369 49.8 0.25 55.5 0.25 53.7 0.25 51.9 0.270 50.6 0.19 56.9 0.19 55.2 0.192 52.9 0.200 51.5 0.128 59.2 0.14 56.6 0.13 54.6 0.12 53.8 0.099 60.8 0.094 59.0 0.090 56.8 0.09 55.2 0.07 63.3 0.070 61.1 0.060 59.2 0.06 57.4 0.040 67.6 0.036 65.9 0.040 62.1 0.04 59.5 Surface Age (s) 3 185 A.4 SURFACE TENSION WITH TEMPERATURE Aqueous CTAB Solutions T (°C) C = 100wppm C = 300wppm C = 1000wppm 22.7 60.1 43.1 38.3 40 57.4 42.1 37.7 60 55.1 40.8 36.5 70 53.7 - - 80 52.1 40.1 35.8 Aqueous Ethoquad 18/25 Solutions T (°C) C = 200wppm C = 400wppm C = 1000wppm 23 58.6 54.1 51.9 40 55.0 50.1 49.3 60 49.1 47 45.2 80 44.6 42.9 42.3 186 APPENDIX B CONTACT ANGLE DATA SDS CTAB Ethoquad 18/25 TRITON X-100 C θ TRITON X-305 C θ C θ C θ C θ (wppm) deg) (wppm) (deg) (wppm) (deg) (wppm) (deg) (wppm) (deg) 10 77 10 77 10 77 10 70 10 70 30 76 20 77 20 77 20 66 20 66 50 75 30 77 30 77 30 64 30 63 100 75 50 77 50 75 50 61 50 60 200 72 70 75 100 70 70 59 100 56 500 66 100 73 200 66 90 57 200 55 700 60 150 69 500 61 100 56 500 55 1000 55 200 65 700 57 130 48 700 55 1300 53 290 63 1000 55 140 45 1000 55 1500 50.5 380 59.8 2000 52 150 41 2000 55 1700 45 500 60 2500 51 180 41 3000 55 2000 44 700 60 4000 51 300 41 2300 37.8 5000 51 500 41 2400 35.3 3300 34.5 3700 35 4500 34.4 4700 34.8 5000 35.2 187 APPENDIX C POOL BOILING DATA C.1 AQUEOUS SURFACTANT SOLUTIONS PURE WATER Increasing qw" ⇑ Decreasing qw" ⇓ ∆Tw qw" ∆Tw qw" 3.289 2.473 3.289 2.513 5.443 9.562 5.443 9.2 6.795 21.819 6.795 21.766 7.786 38.212 7.786 38.317 8.539 59.634 8.539 59.630 9.283 85.685 9.283 84.850 10.024 116.122 10.024 117.816 10.853 154.279 10.853 154.992 11.467 193.169 11.467 197.569 12.3677 238.305 12.368 244.298 AQUEOUS DTAC SOLUTIONS C = 100 wppm C = 500 wppm C = 500 wppm C = 1000 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇓ 3.101 2.318 3.077 2.453 2.977 2.453 3.996 2.503 5.065 9.718 5.000 8.767 4.990 9.467 4.650 7.583 188 6.310 21.458 6.1086 20.470 6.213 22.470 5.744 17.964 7.292 37.944 7.005 37.817 6.805 39.817 6.676 35.357 7.957 60.213 7.660 59.090 8.060 69.490 7.487 58.765 8.854 85.576 8.295 83.057 8.595 89.957 7.920 85.133 9.341 117.326 8.795 117.854 9.095 127.854 8.502 116.804 9.960 154.612 9.424 144.939 9.724 168.939 9.060 153.847 10.792 194.836 10.121 187.562 10.391 197.872 9.515 198.140 11.312 240.125 10.497 231.178 10.392 241.178 10.181 243.079 C = 2000 wppm C = 3000 wppm C = 4000 wppm C = 5000 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇓ ∆Tw qw" ⇓ ∆Tw qw" ⇓ 2.766 2.333 2.730 2.393 2.541 2.326 2.172 2.740 4.191 7.692 4.009 8.740 4.138 9.709 4.393 9.661 5.744 21.417 4.9502 21.051 4.806 22.009 5.103 21.394 6.194 39.370 5.515 38.682 5.297 38.645 5.637 39.078 7.0886 60.272 6.100 60.255 5.793 59.174 6.289 61.777 7.627 86.996 6.576 87.707 6.285 85.463 6.689 86.241 8.101 116.31 7.293 112.192 6.586 117.385 7.117 118.369 8.609 156.323 7.685 156.47 6.911 156.206 7.392 159.661 9.107 197.173 7.862 196.619 7.497 196.757 7.988 197.281 9.286 237.989 8.0284 239.852 7.696 238.016 8.442 243.322 189 C = 7000 wppm C = 10000 wppm C = 10000 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 3.391 2.342 3.772 2.525 3.701 2.393 5.213 8.790 6.127 9.615 5.1556 7.739 5.762 18.640 7.115 21.790 6.507 18.051 6.938 38.991 8.072 39.114 7.838 41.682 7.554 61.032 8.177 59.533 8.324 63.255 7.946 86.419 8.800 85.750 9.334 93.707 8.573 118.355 9.348 117.505 9.647 126.192 9.406 157.038 10.366 152.850 10.302 166.470 9.939 197.738 11.135 195.711 11.029 196.619 10.673 240.382 11.2521 220.582 11.410 244.852 AQUEOUS CTAB SOLUTIONS C = 50 wppm C = 100 wppm C = 100 wppm C = 200 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇓ 2.990 2.307 2.995 2.514 2.965 2.414 3.0379 2.788 5.0743 9.987 5.0222 9.256 5.018 8.958 4.976 9.938 6.167 23.953 6.162 21.947 5.957 20.134 6.141 23.938 7.113 39.822 7.010 38.538 7.020 36.135 6.221 37.413 7.785 61.094 7.592 60.159 7.577 58.199 6.897 60.109 190 7.999 87.449 7.692 85.436 7.667 80.567 7.165 85.654 8.115 119.338 7.962 118.739 7.962 108.345 7.551 117.437 8.397 157.693 8.196 154.145 8.190 143.321 7.708 154.735 8.768 199.952 8.580 196.676 8.611 189.458 7.983 199.052 9.011 246.856 8.808 240.408 8.873 233.804 8.395 241.134 C = 300 wppm C = 400 wppm C = 500 wppm C = 600 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇓ ∆Tw qw" ⇓ ∆Tw qw" ⇓ 2.988 2.491 2.413 2.191 2.514 2.473 3.286 2.130 3.966 7.259 4.216 8.991 4.422 9.994 5.0188 9.102 5.455 21.290 5.191 21.749 5.261 23.953 6.318 22.025 6.536 41.327 5.778 37.644 5.872 41.427 6.995 38.642 6.477 58.738 6.186 58.894 6.282 63.647 7.159 56.714 6.793 86.172 6.381 85.260 6.533 89.166 7.853 85.625 7.043 118.126 6.7845 117.002 6.879 123.991 7.9642 111.376 7.205 155.390 6.979 152.259 6.979 143.781 8.275 141.557 7.597 194.557 7.183 196.050 7.228 188.188 8.613 184.609 8.092 235.465 7.602 238.469 7.722 247.439 9.131 233.527 C = 800 wppm C = 1000 wppm C = 1000 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 4.170 2.116 4.318 2.820 4.112 2.457 191 6.102 9.592 6.253 9.733 5.587 8.712 6.770 20.926 7.213 22.031 6.810 17.899 7.694 38.325 7.721 38.665 7.422 34.981 7.849 56.897 8.109 59.116 8.179 52.567 7.991 81.339 8.834 85.939 8.679 83.107 8.405 111.005 9.314 117.629 9.325 111.547 8.980 154.055 10.270 154.701 10.199 150.108 9.404 194.616 10.954 198.896 10.921 187.236 10.102 239.443 11.833 238.603 11.577 228.615 AQUEOUS Ethoquad O/12 PG SOLUTIONS C = 200 wppm C = 200 wppm C = 400 wppm C = 400 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 2.317 2.333 2.464 2.513 3.397 2.795 2.403 2.522 5.189 9.142 5.363 9.582 5.141 9.837 4.441 9.911 6.538 21.397 6.441 21.766 6.322 21.844 6.2361 20.725 7.235 39.556 7.193 38.317 6.914 33.522 6.9191 36.498 7.640 58.441 7.453 59.630 7.292 51.173 7.663 59.748 8.099 86.587 7.982 84.850 7.916 79.479 7.981 85.631 8.389 111.064 8.397 117.816 8.034 110.983 8.295 120.128 8.573 153.230 8.568 154.992 8.227 147.686 8.195 157.901 8.873 204.064 8.947 197.569 8.576 189.749 8.655 201.673 192 9.153 238.262 C = 600 wppm 9.378 243.298 C = 600 wppm 8.702 237.45 C = 800 wppm 8.793 241.518 C = 800 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 2.633 2.620 3.098 2.556 3.189 3.952 3.244 2.782 5.028 10.293 5.042 9.525 5.843 11.144 5.109 9.933 6.000 21.323 6.410 22.079 6.323 17.796 6.277 21.535 6.7302 37.674 6.971 38.113 7.162 36.410 7.008 38.304 7.1664 57.971 7.347 58.865 7.805 59.338 7.515 59.135 7.620 85.419 7.605 84.729 8.132 91.016 7.778 84.997 7.844 117.48 8.027 117.759 8.542 119.323 8.077 115.441 8.051 157.106 8.213 151.856 8.506 163.774 8.2 150.72 8.351 203.378 8.437 195.853 8.813 212.883 8.532 190.697 8.341 246.079 8.905 241.583 8.923 249.528 9.184 246.66 C = 1000 wppm C = 1000 wppm C = 1200 wppm C = 1200 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 2.775 3.041 4.228 2.733 3.407 2.616 3.420 2.521 4.990 10.340 6.818 9.626 7.115 11.142 6.610 9.996 6.281 21.456 7.522 22.132 7.212 24.973 7.0953 22.328 8.481 37.998 8.230 36.098 7.856 41.1607 7.954 36.939 9.597 59.483 8.578 59.373 9.055 61.576 8.756 59.407 9.869 86.114 9.086 85.238 9.315 86.526 9.114 87.746 8.961 121.256 9.432 118.942 9.379 119.954 9.366 116.024 193 8.834 161.144 9.559 152.078 9.505 166.254 9.542 155.099 8.349 206.648 9.960 195.831 9.926 203.959 10.092 196.28 8.740 251.480 10.674 230.956 10.690 239.923 11.038 250.617 C = 1500 wppm C = 1500 wppm C = 3000 wppm C = 3000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 4.295 2.865 5.086 2.394 6.389 2.735 8.389 2.733 7.862 10.357 7.662 9.996 8.197 10.457 9.775 10.857 8.050 21.479 8.517 21.056 9.782 22.118 10.97232 21.118 9.060 41.765 9.499 35.453 11.402 37.697 12.407 35.697 10.828 61.543 10.785 59.596 12.869 59.429 12.999 56.429 11.574 87.706 11.822 86.492 14.318 87.491 14.213 81.491 12.471 118.505 12.203 113.264 14.954 115.720 14.916 110.206 12.479 159.852 12.248 145.424 15.007 161.479 15.019 152.479 12.059 200.774 11.689 192.078 15.267 201.170 15.271 211.1699 12.575 239.036 12.037 236.478 15.118 237.866 15.213 246.866 AQUEOUS Ethoquad 18/25 SOLUTIONS C = 200 wppm C = 200 wppm C = 500 wppm C = 500 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 2.478 2.604 3.0945 2.356 3.488 2.837 3.303 2.551 4.779 10.261 5.266 9.113 5.841 10.299 5.308 9.811 7.317 22.279 6.630 20.902 6.513 21.809 6.598 22.541 194 8.155 38.112 7.729 38.710 7.799 38.799 7.381 37.036 8.945 59.542 8.477 63.431 8.649 58.806 8.0231 59.565 9.222 86.186 9.114 87.561 8.709 89.923 8.497 82.911 9.790 121.214 9.439 121.086 8.931 121.653 8.922 115.732 9.832 158.663 9.682 155.828 9.191 160.872 9.197 158.256 9.774 201.926 9.968 206.479 9.444 205.921 9.488 199.771 10.380 243.156 10.485 245.777 10.271 245.068 10.037 242.113 C = 700 wppm C = 1000 wppm C = 1000 wppm C = 1500 wppm ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ 3.120 2.351 2.185 2.703 3.228 2.367 2.575 2.776 5.665 8.755 4.515 9.939 6.317 9.880 6.143 11.134 7.198 20.856 7.648 22.8249 7.195 18.854 7.403 22.426 7.906 37.460 8.349 37.611 8.157 32.899 8.817 39.289 8.359 58.837 9.0963 62.051 8.599 55.085 9.652 59.899 8.559 78.847 9.419 84.564 9.058 81.783 9.969 89.547 9.126 105.989 9.711 120.431 9.654 117.421 10.553 116.809 9.816 145.469 9.900 160.360 10.169 160.379 10.759 160.360 10.142 183.634 10.579 207.225 10.548 196.738 11.204 204.136 10.607 231.967 10.999 242.619 11.106 237.097 11.021 237.854 195 C = 2000 wppm C = 2000 wppm C = 3000 wppm C = 3000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 3.363 2.605 3.446 2.196 4.155 2.621 4.5022 3.008 6.796 10.697 5.422 8.847 6.107 9.933 6.946 8.459 8.921 22.954 7.108 21.4739 7.515 22.589 8.6759 21.065 11.101 38.382 8.876 37.424 8.997 37.971 10.575 37.701 11.653 60.021 9.616 59.162 9.543 59.703 12.412 57.384 12.287 86.718 10.300 84.915 10.755 91.270 12.804 80.366 12.543 123.132 10.830 115.305 11.84 124.507 13.536 110.575 12.520 163.174 11.251 149.605 12.464 165.501 13.734 146.783 12.429 203.866 11.871 187.255 12.751 208.22 14.120 187.567 12.884 244.473 12.469 235.527 13.103 251.025 14.640 233.913 C = 5000 wppm C = 5000 wppm C = 5000 wppm C = 5000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 8.218 16.727 7.259 2.624 16.727 80.782 17.375 84.608 11.959 18.185 9.276 9.676 18.184 111.393 18.568 115.599 13.1098 18.993 12.001 23.544 18.994 152.120 19.439 159.109 14.780 20.239 14.578 38.750 20.239 191.914 20.536 199.4 16.231 21.344 16.131 62.849 21.344 247.135 21.439 237.806 196 C.2 AQUEOUS POLYMER SOLUTIONS AQUEOUS HEC-QP300 SOLUTIONS C = 100 wppm C = 100 wppm C = 300 wppm C = 500 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇓ ∆Tw qw" ⇑ ∆Tw qw" ⇑ 3.527 3.2423 3.1734 2.561 3.132 2.749 2.6576 2.776 5.370 9.742 5.483 9.502 5.272 10.147 4.903 8.732 7.466 21.609 7.1880 20.870 7.195 21.903 6.638 20.105 8.822 36.450 8.864 39.068 8.393 37.026 7.610 33.416 10.005 58.689 9.874 55.465 9.443 59.442 8.849 51.833 10.819 83.373 11.083 88.182 10.260 89.928 9.707 79.335 11.621 118.335 12.012 126.059 11.079 119.690 10.517 109.844 12.405 159.421 12.838 169.028 11.850 169.207 11.110 147.576 13.340 202.736 13.684 211.758 12.671 214.09 12.075 188.616 14.166 247.398 14.453 254.611 13.278 259.741 12.442 232.257 C = 600 wppm C = 700 wppm C = 1000 wppm C = 3000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇑ ∆Tw qw" ⇑ ∆Tw qw" ⇑ 2.547 2.723 3.105 2.818 3.463 2.794 3.480 2.603 4.691 9.099 5.518 10.714 5.955 9.794 6.253 9.290 6.358 21.335 7.188 22.764 7.681 22.992 8.199 21.146 7.497 37.673 8.1994 34.724 8.758 41.730 9.595 37.479 8.523 57.724 9.159 58.268 9.893 63.717 10.479 59.655 9.378 86.562 10.071 89.823 10.312 85.098 10.895 80.556 197 10.264 121.388 10.456 123.075 10.817 118.782 11.203 127.479 10.836 159.376 11.096 165.723 11.1617 159.495 11.758 156.263 11.736 203.888 12.196 215.480 12.1467 202.343 12.610 201.237 11.984 244.399 12.388 251.148 12.626 246.124 13.404 248.535 AQUEOUS Carbopol 934 SOLUTIONS C = 100 wppm C = 300 wppm C = 500 wppm C = 1000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇑ ∆Tw qw" ⇑ ∆Tw qw" ⇑ 2.810 2.687 2.718 2.889 3.265 2.949 2.493 2.1859 5.018 9.711 5.095 8.975 5.764 10.201 5.935 9.113 7.035 20.724 7.157 20.276 7.461 20.740 8.009 20.8 8.207 34.310 8.608 35.904 8.798 35.711 9.588 37.228 9.647 54.819 9.944 55.627 10.634 56.688 10.902 56.227 10.895 79.457 11.345 80.435 11.923 81.568 12.661 83.449 12.067 110.303 12.453 107.476 13.382 110.683 14.214 109.745 13.065 144.867 13.972 143.528 14.810 147.121 15.387 146.162 13.918 182.583 14.870 178.229 15.817 180.565 16.872 190.318 14.645 220.928 16.098 221.016 17.101 221.880 18.111 226.122 C = 1500 wppm C = 3000 wppm C = 3000 wppm ∆Tw qw" ⇑ ∆Tw qw" ⇑ ∆Tw qw" ⇓ 3.852 3.157 2.304 1.715 3.962 3.010 6.535 8.763 7.195 9.954 7.073 9.976 8.737 19.816 9.186 20.820 9.424 20.539 198 10.092 36.011 10.803 35.906 10.714 37.766 11.729 56.230 12.194 56.071 12.569 56.694 13.516 82.704 14.331 82.149 14.358 83.232 15.114 112.918 16.639 111.222 16.051 113.539 16.571 146.963 17.806 148.242 17.509 146.817 17.940 187.080 19.660 190.946 18.860 183.752 18.877 224.166 21.176 225.901 20.266 228.451 199 APPENDIX D TEMPORAL AND SPATIAL DISCRETIZATION The whole simulation is solved in the domain Ω = {( x , y )|0 ≤ x ≤ X ,0 ≤ y ≤ Y} D1. Projection Methodology (Fractional Step Method or Time-Split Method With Lagged Pressure) The accurate projection method or fractional-step (Brown et al., 2001) is used, which is an improved fully second-order accurate projection algorithm over the method used by Bell and Colella (1989). The first step of the projection method: semi-implicit viscous solver for the intermediate velocity u* U* −Un y GrL n Gp n −1/ 2 L* + Ln Mn +1/ 2 θ − n +1/ 2 + n +1/ 2 − n +1/ 2 = −[(U i∇)U ]n +1/ 2 + + ρ 2ρ ρ Fr Re 2 ∆t For the intermediate field u* on boundary u* = ubn +1 Next, un+1 is recovered from the projection of u* by solving ∆t∇i 1 ρ ∇q n +1 = ∇i u* in Ω n +1/ 2 ni∇q n +1 = 0 on ∂Ω And setting un +1 = u* − ∆t∇q n +1 ρ n +1/ 2 The new pressure is computed by utilizing the correct pressure update 200 ∇p n +1/ 2 = ∇p n −1/ 2 + ∇q n +1 − ∆t∇i 1 ρ n +1/ 2 ∆t 1 ∇µ n +1/ 2∇i n +1/ 2 ∇q n +1 ρ 2 ∇q n +1 = ∇i u* 1 ∇p n +1/ 2 = ∇p n −1/ 2 + ∇q n +1 − ∇µ n +1/ 2∇i u* 2 1 p n +1/ 2 = p n −1/ 2 + q n +1 − µ n +1/ 2∇i u* 2 This formula is different from Bell et al. (1989) ∇p n +1/ 2 = ∇p n −1/ 2 + ∇q n +1 Bell’s formula is not consistent with a second-order discretization of the NavierStokes equations, because the normal component of this equation imply that ni∇p n +1/ 2 = ni∇p n −1/ 2 For all n, which cannot be correct in general. Discretization of the projection ∆t∇i 1 ρ n +1/ 2 ∇q n +1 = ∇i u* in Ω ni∇q n +1 = 0 on ∂Ω u +u −u −u v +v −v −v ( DU )i +1/ 2, j +1/ 2 = i +1, j i +1, j +1 i , j i , j +1 + i , j +1 i +1, j +1 i , j i +1, j 2 ∆x 2∆y 1 D ρ Gφ i +1/ 2, j +1/ 2 = 1 1 ( 2φi−1/ 2, j −1/ 2 + φi+1/ 2, j−1/ 2 + φi−1/ 2, j+1/ 2 − 4φi+1/ 2, j+1/ 2 ) 6h 2 ρi , j + + + 1 ρi , j +1 1 ρi +1, j ( 2φ i −1/ 2, j + 3/ 2 ( 2φ 1 ρi +1, j +1 i + 3/ 2, j −1/ 2 ( 2φ + φi +1/ 2, j +3/ 2 + φi −1/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) + φi +1/ 2, j −1/ 2 + φi +3/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) i + 3/ 2, j + 3/ 2 + φi +1/ 2, j +3/ 2 + φi +3/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) 201 MPCG matrix Ax = b A is matrix of ( n * n ), i, j = 0......n A = [......] x (1 + 1/ 2,1 + 1/ 2) b(1 + 1/ 2,1 + 1/ 2) x (2 + 1/ 2,1 + 1/ 2) b(2 + 1/ 2,1 + 1/ 2) x (3 + 1/ 2,1 + 1/ 2) b(3 + 1/ 2,1 + 1/ 2) . . . . . . * x (i + 1/ 2, j + 1/ 2) = b(i + 1/ 2, j + 1/ 2) . . . . . . x ( n − 5 / 2, n − 1/ 2) b( n − 5 / 2, n − 1/ 2) x ( n − 3/ 2, n − 1/ 2) b( n − 3/ 2, n − 1/ 2) x ( n − 1/ 2, n − 1/ 2) b( n − 1/ 2, n − 1/ 2) b(2 / 3, j + 1/ 2), b(i + 1/ 2, 2 / 3), b( n − 1/ 2, j + 1/ 2), b(i + 1/ 2, n − 1/ 2) are different from others, boundary nodes are included , corner has two boundary nodes D2. Temporal Discretization When discretizing the governing equations temporally, the diffusion terms are treated by a fully implicit scheme and the convection and source terms by explicit method. Therefore: ∇ ⋅ u n +1 = ( ρc pl m n +1 ρ2 . ⋅∇ρ + V micro ) T n +1 − T n = − ρc pl [ u ⋅ ∇T ]n + [∇ ⋅ k∇T ]n +1 ∆t φ n +1 = φ n + ∆tL(φ n ) ∆t L(φ n ) + L(φ n +1 ) 4 ∆t φ n +1 = φ n + L(φ n ) + 4 L(φ n +1/ 2 ) + L(φ n +1 ) 6 φ n +1/ 2 = φ n + 202 | ∇φ n +1 | = 1 for φ ≠ 0 1 2 φ n +1/ 2 = (φ n + φ n +1 ) ρ n +1/ 2 = ρ (φ n +1/ 2 ) µ n +1/ 2 = µ (φ n +1/ 2 ) D3. The surface Tension term (1/ W )k (φ )∇H (φ ) ( M )ij = 1 ( DN )ij (GH node )ij W n = ∇φ / | φ | k = ∇in = ∇i(∇φ / | φ |) |φ =0 ni +1/ 2, j +1/ 2 = (Gφ )i +1/ 2, j +1/ 2 | (Gφ )i +1/ 2, j +1/ 2 | (Gφ )i +1/ 2, j +1/ 2 φi +1, j +1 + φi +1, j − φi , j +1 − φi , j 2 ∆x = φi +1, j +1 − φi +1, j + φi , j +1 − φi , j 2 ∆y ni1+1/ 2, j +1/ 2 + ni1+1/ 2, j −1/ 2 − ni1−1/ 2, j +1/ 2 − ni1−1/ 2, j −1/ 2 ( DN )i , j = 2 ∆x 1 1 ni +1/ 2, j +1/ 2 − ni +1/ 2, j −1/ 2 + ni1−1/ 2, j +1/ 2 − ni1−1/ 2, j −1/ 2 + 2∆y D4. Discretization of Pressure Gradient (Gp )i , j pi +1/ 2, j +1/ 2 + pi +1/ 2, j −1/ 2 − pi −1/ 2, j +1/ 2 − pi −1/ 2, j −1/ 2 2∆x = pi +1/ 2, j +1/ 2 − pi +1/ 2, j −1/ 2 + pi −1/ 2, j +1/ 2 − pi −1/ 2, j −1/ 2 2∆y D5. ENO Scheme A third-order ENO method was used for the approximation of the convective terms in level set function. For u divergence free: 203 (u ⋅∇)φ = (uφ ) x + (vφ ) y (u ⋅∇)u = f x + g y Where 2 f = ( uuv ), g = ( uv ) v2 (uφ ) x + (vφ ) y ≈ (ui +1/ 2 , j + ui −1/ 2 , j )(φ i +1/ 2 , j − φ i −1/ 2 , j ) / (2h) + (vi , j +1/ 2 + vi , j −1/ 2 )(φ i , j +1/ 2 − φ i , j −1/ 2 ) / (2h) (uT ) x + (vT ) y ≈ (ui +1/ 2 , j + ui −1/ 2 , j )(Ti +1/ 2 , j − Ti −1/ 2 , j ) / (2h) + (vi , j +1/ 2 + vi , j −1/ 2 )(Ti , j +1/ 2 − Ti , j −1/ 2 ) / (2h) ( f 1 ) x + ( g1 ) y ≈ (ui +1/ 2 , j + ui −1/ 2 , j )(ui +1/ 2 , j − ui −1/ 2 , j ) / (2h) + (vi , j +1/ 2 + vi , j −1/ 2 )(ui , j +1/ 2 − ui , j −1/ 2 ) / (2h) ( f 2 ) x + ( g2 ) y ≈ (ui +1/ 2 , j + ui −1/ 2 , j )(vi +1/ 2 , j − vi −1/ 2 , j ) / (2h) + (vi , j +1/ 2 + vi , j −1/ 2 )(vi , j +1/ 2 − vi , j −1/ 2 ) / (2h) For computing ui +1/ 2 , j , similarly for ui , j +1/ 2 ,φ i +1/ 2 , j ,Ti +1/ 2 , j , ... , The ENO scheme is defined as i. Upwind i u ≥0 +1/ 2 , j k l = {i +1 iotherwise ii. First order f i +(11)/ 2 , j = f kl , j iii. Second order 204 a= b= f kl , j − f kl −1, j ∆x f kl +1, j − f kl , j ∆x if |a |≤ |b | otherwise c={ a b if |a |≤ |b | k 2 = {kkll −1otherwise f i +( 21)/ 2 , j = f i +(11)/ 2 , j + ∆x c(1 − 2( k l − i )) 2 iv. Third order a= b= f k2 −1, j − 2 f k2 , j + f k2 +1, j f k2 , j ( ∆x ) 2 c = {ba f ( ∆x ) 2 − 2 f k2 +1, j + f k2 + 2 , j ( 3) i +1/ 2 , j if |a |≤ |b | otherwise = f (2) i +1/ 2 , j ( ∆x ) 2 + c(3(( k 2 − i ) 2 − 1) 3 D6. Approximation of Advection Terms (U i∇ ) U n+1/ 2 This discretization of the advection terms in this algorithm is based on the method used by Puckett et al. (1997) and Sussman et al. (1999). It is a predictor-corrector method evolved from unsplit Godunov method introduced by Colella (1990). For face (i+1/2,j): U n +1/ 2, L i +1/ 2, j n ∆x ui , j ∆t n ∆t 1 ∆t n =U +( − )U x ,ij − ( vU y )ij + n (Lij − Gpijn −1/ 2 − Mijn + Fijn ) 2 2 2 ρij 2 n ij Extrapolated from cell (i,j) n ∆x ui +1, j ∆t n ∆t n +1/ 2, R n Ui +1/ 2, j = Ui +1, j + ( − )U x ,i +1, j − ( vU y )i +1, j + 2 2 2 1 ∆t n (L − Gpin+−1,1/j 2 − Mi +n1, j + Fi +n1, j ) ρin+1, j 2 i +1, j Analogous formulae are used to predict values at each of the other faces of the cell: 205 Uin, +j +1/1 2,F / B , φin++1,1/j 2, L / R , φin, +j +1/1 2, F / B U xn and φ xn are evaluated using a monotonicity-limited fourth-order slope approximation (Colella,1985): 2 1 ( q j +1 − q j −1 ) − ( q j +2 − q j −2 ) 12 = 3 2 1 ∆rj ( rj +1 − rj −1 ) − ( rj + 2 − rj −2 ) 3 12 δ qj The transverse derivative terms ( vU y ) are evaluated as below, by first extrapolating from above and below to construct edge states, using normal derivatives only, and then choosing between these states using the upwinding procedure defined below. B ∆y vij ∆t n − U i , j +1/ 2 = Uijn + U y ,ij 2 2 T ∆y vi , j +1∆t n + U i , j +1/ 2 = Uin, j +1 − U y ,i , j +1 2 2 Where U y are limited slopes in the y direction. In this upwinding procedure we first define the normal velocity on the edge: v B if v B > 0, v B + v T > 0 B T B T = 0 if v ≤ 0, v ≥ 0 or v + v = 0 T T B T v if v < 0, v + v < 0 adv v i , j +1/ 2 adv Now upwind U based on v i , j +1/ 2 : U i , j +1/ 2 U B U B + U T = 2 T U adv if v i , j +1/ 2 > 0 adv if v i , j +1/ 2 = 0 adv if v i , j +1/ 2 < 0 After constructing U i , j −1/ 2 in a similar manner, these upwind values can be used to form the transverse derivative: ( vU ) y ij = ( )( adv adv 1 v i , j +1/ 2 + v i , j +1/ 2 U i , j +1/ 2 − U i , j −1/ 2 2 ∆y 206 ) A similar upwinding procedure is used to choose the appropriate states Ui +1/ 2, j given the left and right states: uin++1/1/2,2 j u L = 0 u R if u L > 0, u L + u R > 0 if u L ≤ 0, u R ≥ 0 or u L + u R = 0 if u R < 0, u L + u R < 0 Following a similar procedure to construct Ui −1/ 2, j , U i , j +1/ 2 , Ui , j −1/ 2 . In general, the normal velocities at the edges are not divergence-free; in order to make these velocities divergence-free, the MAC projection is applied before construction of the convective derivatives. The equation 1 D MAC n G MAC ∅ = D MACU n +1/ 2 ρ is solved for ∅ , where u n +1/ 2 − uin−+1/1/2,2 j uin, +j +1/1/22 − uin, +j −1/1/22 , D MACU n +1/ 2 = i +1/ 2, j ∆x ∆y and ∅ −∅ x (G MAC∅ )i+1/ 2, j = i+1,∆j x i , j ∅ −∅ y (G MAC∅ )i+1/ 2, j = i , j+∆1 y i , j The face-based advection velocities at t n +1/ 2 are then defined by n +1/ 2 uiADV +1/ 2, j = ui +1/ 2, j − n +1/ 2 viADV , j +1/ 2 = vi , j +1/ 2 − ρin+1/ 2, j = 1 ρin+1/ 2, j 1 ρ n i , j +1/ 2 (G MAC (G MAC ∅) ∅) x i +1/ 2, j y i , j +1/ 2 1 n ρij + ρin+1, j ) ( 2 207 The next step, after construction the advective velocities Uin++1/1/2,2 j U L if u ADV > 0 1 = (U L + U R ) if u ADV = 0 2 U R if u ADV < 0 The advection terms can now be defined by n +1/ 2 (U i∇ ) U i , j 1 ui +1/ 2, j + ui −1/ 2, j = (Ui+1/ 2, j − Ui−1/ 2, j ) 2 ∆x ADV ADV 1 vi , j +1/ 2 + vi , j −1/ 2 + (Ui, j+1/ 2 − Ui , j−1/ 2 ) 2 ∆y ADV ADV D7. Discretization of the divergence of the stress tensor ∇ ⋅ (2 µD) The components of the viscous stress tensor D is estimated using central differencing: ∇ ⋅ (2 µ D ) = 2( ( µux ) x +( µ ( µ v y ) y +( µ u y +vx 2 u y + vx 2 )y )x ) The first component of the viscous term ∇ ⋅ 2 µ (φ )D is discretized as 2 µi +1/ 2, j (ui +1, j − ui , j ) − 2 µi −1/ 2, j (ui , j − ui −1, j ) + + ∆x 2 µi , j +1/ 2 (ui , j +1 − ui , j ) − µi , j −1/ 2 (ui , j − ui , j −1 ) ∆y 2 µi , j +1/ 2 ( vi +1, j +1 − vi −1, j +1 + vi +1, j − vi −1, j ) − µi , j −1/ 2 (vi +1, j − vi −1, j + vi +1, j −1 − vi −1, j −1 ) 4 ∆x∆y Where 1 2 1 2 µi +1/ 2, j = ( µ (φi , j ) + µ (φi +1, j )), µi , j +1/ 2 = ( µ (φi , j ) + µ (φi , j +1 )) The second component of the viscous term is discretized in a similar manner. 208 D8. Discretization of the Laplace L * 2 µi +1/ 2, j iui +1, j + 2 µi −1/ 2, j iui −1, j ∆x + 2 + µi , j +1/ 2 iui , j +1 + µi , j −1/ 2 iui , j −1 ∆y 2 µi , j +1/ 2 ( vi +1, j +1 − vi −1, j +1 + vi +1, j − vi −1, j ) − µi , j −1/ 2 (vi +1, j − vi −1, j + vi +1, j −1 − vi −1, j −1 ) 4∆x∆y 2( µi +1/ 2, j + µi −1/ 2, j ) ( µi , j +1/ 2 + µi , j −1/ 2 ) − + ui , j ∆y 2 ∆x 2 A 9-point stencil approximating the Laplacian 1 D ρ Gφ i +1/ 2, j +1/ 2 = 1 1 ( 2φi−1/ 2, j−1/ 2 + φi +1/ 2, j−1/ 2 + φi−1/ 2, j+1/ 2 − 4φi+1/ 2, j+1/ 2 ) 6h 2 ρi , j + + + 1 ρi , j +1 1 ρi +1, j ( 2φ i −1/ 2, j + 3/ 2 + φi +1/ 2, j + 3/ 2 + φi −1/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) ( 2φ i + 3/ 2, j −1/ 2 + φi +1/ 2, j −1/ 2 + φi +3/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) 1 ρi +1, j +1 ( 2φ i + 3/ 2, j + 3/ 2 + φi +1/ 2, j +3/ 2 + φi +3/ 2, j +1/ 2 − 4φi +1/ 2, j +1/ 2 ) A 5-point stencil approximating the Laplacian 1 D ρ Gφ i +1/ 2, j +1/ 2 = 2 1 (φi+3/ 2, j+1/ 2 − φi+1/ 2, j+1/ 2 ) 2 h ρi +1, j + ρi +1, j +1 + 1 −φ (φ ) ρi , j + ρi , j +1 i −1/ 2, j +1/ 2 i +1/ 2, j +1/ 2 + 1 −φ (φ ) ρi , j +1 + ρi +1, j +1 i +1/ 2, j +3/ 2 i +1/ 2, j +1/ 2 + 1 φi +1/ 2, j −1/ 2 − φi +1/ 2, j +1/ 2 ) ( ρi , j + ρi +1, j 209