. “a: n: $3., , .. _. . v . . s. 3153!: . 35.... ‘ 0 .4 .l .n:. 1 l i 4 v... max»: ,I- o H.- 7) V 55.33.98 L .21.; 3...! L5... . ‘3‘; I... «iiiflss. I:n:a....xx? 13:0:2: t E ~ ; 1.3 gum r—kh his : 23935:: ....:}..._..v i: S... anvfzumwwéuflfimg .. : .53 awvmmfi. LL; .3» . :22... , .. .. .2! . . .11.; 4.701.: .. z. x? . hittr. . a Ifinxnl’: 1.3.0.. 2 v. if. 5 ‘ 2A.. V :Zi: .1... a #1... «24:11 an v2. {E‘s z .513... 3'1): .3?) 5 xx“... 2“? I‘ . x. t! 5:. "Pol... A! .3}! figé . \ xx .oaxou: . . 2 .x... air. I! u. ‘ , “a. 3.1: :5 fly... a- : Ma. .33....3. ‘ .. _ . 4 I... II— .fifl . 2v .3. 7 mm M. .. QR ‘1“. I n W? V I}: 1 . . 31.? p! . 5. :9 hfluuvfla it. 5.5. 1.99;...a ‘ 51.53%“. :1! . .v ’33:. 7...! bl; . . Sin??? 3. El... r. .131: Kazan}. ‘ .C. ..- 32.2.13 c in. .......:fi:. 11!??? I. .11... 7 , : is i; +¢ T l a... c. v . :b v«..w.:£..!..¢.15. - .1fi3s?4us#uq.ly {a .5. .3... z 3|..4)v5’..| .3 .qun...» ‘ V 213-52.... 3:». !OOi LIBRARY Michigan State University -W C. rnflw—o— J This is to certify that the dissertation entitled Nucleation of Gas-Supersaturated Liquids presented by Xiaobing Liu has been accepted towards fulfillment of the requirements for Ph. D. degree in Mechanical Engineering Date 6%? inf/1.4% 20“” MS U is an Affirmative Action/ Equal Opportunity Institution 0- 12771 PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIHC/DateDuep65-9. 15 NUCLEATION OF GAS-SUPERSATURATED LIQUIDS BY XIAOBING LIU A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY DEPARTMENT OF MECHANICAL ENGINEERING 2001 Nucleation of Gas-Supersaturated Liquids IBy Xiaobing Liu AN ABSTRACT OF A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2001 Professor Giles Brereton ABSTRACT NUCLEATION OF GAS-SUPERSATURATED LIQUIDS By Xiaobing Liu Classical nucleation theory relates nucleation rate to the reversible work required to form a bubble of critical size, regardless of any irreversible effects. It has been suc- cessfully applied to predict superheat nucleation thresholds of pure liquids, but fails to predict nucleation thresholds of gas-supersaturated liquids and does not account for the effect of volumetric confinement on nucleation. In this study, the equations for dissipation of energy due to viscous and diffusive effects during bubble growing processes were derived and it was shown that, for many nucleation processes, the dissipation of energy due to diffusion may not be neglected as in the classical nu- cleation theory. Therefore, the classical nucleation rate equation was modified to include a term of dissipation of energy due to diffusion. This modified theory can predict nucleation thresholds of liquids supersaturated with different gases in good agreement with experimental data. For nucleation of gas-supersaturated liquids in small capillary tubes, a dual dependence model originally proposed by Brereton et al. was refined to account for the effect of volumetric confinement on heterogeneous nucleation thresholds. The new model can predict supersaturation nucleation thresh— olds of gas-supersaturated water in capillary tubes well. It was also found that if the shape factor for nucleation in small capillary tubes is assumed inversely propor- tional to the tube diameter, a simpler heterogeneous model can predict accurately the supersaturation nucleation thresholds of water supersaturated with different gas species, as well as the superheat nucleation limits of pure water in small capillary tubes. This dissertation is dedicated to my parents Yunyao Liu & Zixiu Tang for their love and their support, encouragement, and understanding during my years in academia. iii ACKNOWLEDGEMENTS I would like to acknowledge all the help from Dr. Giles Brereton, from the initial ideas for the foundation of this thesis, to the proof reading of the final draft. Without him, this would not have been possible. Also I would like to thank Dr. Abraham Engeda, Dr. Simon Garrett, and Dr. John McGrath for being on my committee, and providing guidance to my study. iv TABLE OF CONTENTS LIST OF TABLES viii LIST OF FIGURES x 1 INTRODUCTION 1 1.1 Applications of Nucleation Theory .................... 1 1.2 Literature Review ............................. 3 1.2.1 Classical Nucleation Theory ................... 3 1.2.2 Nucleation in Gas—Liquid Solutions ............... 5 1.2.3 Nucleation in Confined Volumes ................. 10 1.2.4 Studies of Bubble Dynamics ................... 11 1.3 Objectives ................................. 12 THE THEORY OF NUCLEATION 14 2.1 The Equilibrium State .......................... 14 2.1.1 A Pure Liquid in Equilibrium with its Vapor .......... 14 2.1.2 Equilibrium in Gas-Liquid Solutions .............. 16 2.2 The Stability of a Bubble in a Liquid .................. 19 2.2.1 The Stability of a Vapor Bubble in a Pure Liquid ....... 20 2.2.2 The Stability of a Bubble in a Gas-Liquid Solution ...... 23 2.2.3 The Critical Radius and Gas Concentration Gradients 26 2.3 The Nucleation Rate Equation ...................... 28 2.4 Effects of P, T and Concentration on Nucleation Rate ........ 32 2.5 Discussions on the Applicability of Henry’s Law and the Ideal-Gas Law at High Pressures ............................. 35 2.5.1 The Applicability of Henry’s Law ................ 35 2.5.2 Effects of the Departure from Henry’s Law on Nucleation Rate 38 2.5.3 The Applicability of the Ideal-Gas Law ............. 4O BUBBLE DYNAMICS 43 3.1 Statement of the Problem ........................ 43 3.2 Governing Equations ........................... 43 3.3 Analytical Solution ............................ 46 3.3.1 Bubble Dissolution in an Undersaturated Solution ....... 49 3.3.2 Bubble Growth in a Supersaturated Solution .......... 51 3.4 Numerical Approach ........................... 53 3.4.1 Dimensionless Form ........................ 54 3.4.2 Discretization ........................... 55 3.4.3 Stability Criterion ........................ 55 3.4.4 Moving Grids ........................... 56 3.4.5 Numerical Results ........................ 57 3.4.6 Comparison of Analytical and Numerical Results ....... 60 4 TARGET DATA FOR TESTING THEORIES OF NUCLEATION IN GAS- SUPERSATURATED SOLUTIONS 62 4.1 Nucleation of Supersaturated COg—Water Solutions in Large-Scale Sys- tems .................................... 62 4.2 Experiments on Formation of Gas-Vapor Bubbles in Supersaturated Solutions of Gases in Water ....................... 64 4.2.1 Experimental Setup and Results ................. 65 4.2.2 Problems with Finkelstein and Tamir’s Equation on Nucleation of Liquids Supersaturated with Gases .............. 67 4.2.3 Inconsistencies between F inkelstein and Tamir’s Results and Classical Nucleation Theory ................... 67 4.3 Experiments on Nucleation of Gas-Supersaturated Solutions in Small Capillary Tubes ................ . .............. 71 4.3.1 Experimental Setup and Procedures ............... 71 4.3.2 Experimental Results for Nucleation of Gas-Supersaturated So- lutions in Capillary Tubes .................... 73 4.3.3 Interpretations of Experimental Data Describing Nucleation in Capillary Tubes .......................... 75 4.4 Some Shortcomings of the Classical Nucleation Theory ........ 76 5 MODIFICATIONS TO THE CLASSICAL NUCLEATION THEORY 78 5.1 Prediction of Nucleation Thresholds of Liquids Supersaturated with Different Gases .............................. 78 5.1.1 Dissipation of Energy through Diffusion and Viscous Effects . 80 5.1.2 The Significance of Diffusion and Viscous Effects to Nucleation 83 5.1.3 A New Nucleation Model for Liquids Supersaturated with Dif- ferent Gases ........................... 89 5.1.4 Calculation Results of the New Nucleation Model for Water Supersaturated with Different Gases at a Constant Temperature 92 5.1.5 Temperature Dependence of Nucleation Thresholds of Gas- Supersaturated Liquids ...................... 97 5.2 Prediction of Nucleation Thresholds of Gas—Liquid Solutions in Small Capillary Tubes .............................. 99 5.2.1 Dual Dependence Model to Account for Effects of Volumetric Confinement on Nucleation ................... 99 5.2.2 A Preliminary Model to Account for Effects of Volumetric Con- finement on Nucleation ...................... 103 6 CONCLUDING REMARKS AND RECOMMENDATIONS 108 APPENDIX 111 A THE COEFFICIENTS OF HEN RY’S LAW FOR NITROGEN, HYDROGEN, ARGON, AND HELIUM DISSOLVED IN WATER 112 vi BIBLIOGRAPHY 119 vii 2.1 2.2 3.1 4.1 5.1 5.2 5.3 5.4 5.5 5.6 A.1 A.2 LIST OF TABLES Values of shape factor for bubble formation in conical cavities Calculation results of compressibility factors using the generalized com- pressibility chart ............................. Time required for a pure gas bubble to dissolve in water ....... Experimental data for the pressure difference for bubble formation . . The values of solubility and diffusivity of several gases dissolved in water at 30°C ............................... Comparison of the experimental data of AP" (MPa) with the predicted data from the revised nucleation equation (5.43) ............ Predicted data of nucleation thresholds from Eq. (5.42) taking into account the departure from Henry’s law ................. The estimated values of diffusivity of several gases dissolved in water at 25°C .................................. The calculated values of 111m, and q for nucleation in gas-supersaturated water at 25°C ............................... The values of solubility and diffusivity of 02 dissolved in water . . . . The solubility of nitrogen in water at different pressures ........ The solubility of hydrogen in water at different pressures ....... viii 92 93 93 96 98 113 115 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 2.13 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 4.1 4.2 4.3 5.1 5.2 5.3 5.4 LIST OF FIGURES System model of bubble formation in a liquid ............. 15 The curve of Gibbs function in the neighborhood of the equilibrium radius .............. - ..................... 23 Gas concentration profiles in the liquid and in the bubble ....... 29 A vapor bubble formed at a smooth liquid-solid surface ........ 30 Definition of geometrical parameters in a conical cavity ........ 31 The effect of the liquid pressure on the nucleation rate ........ 33 The effect of the liquid temperature on the nucleation rate ...... 34 The effect of the supersaturation pressure on the nucleation rate . . . 35 Measurement and Henry’s-law prediction of the saturation concentra- tion of 02 in water ............................ 36 Variation of H with pressure for 02 dissolved in water ......... 37 Effect of the departure from Henry’s law on nucleation rate ((15 = 0.001) 39 Effect of the departure from Henry’s law on nucleation rate ((15 = 0.005) 40 Compressibility of nitrogen ........................ 41 Radius-time relation for dissolving bubbles ............... 50 Radius-time relation for growing bubbles ................ 52 Illustration of moving grids during bubble dissolution ......... 57 Dimensionless radius vs. time curve for bubble dissolution due to dif- fusion (numerical result) ......................... 58 Dimensionless radius squared vs. time curve for dissolving bubbles due to diffusion (numerical result) ...................... 59 Concentration distribution around a dissolving bubble as a function of time .................................... 59 Dimensionless radius-squared vs. time curve for growing bubbles due to diffusion (numerical result) ...................... 60 Radius vs. time curves for a dissolving bubble ............. 61 The experimental apparatus of Finkelstein & Tamir .......... 66 Schematic illustration of the experimental apparatus used for oxygen- supersaturation studies .......................... 72 Dependence of supersaturation pressure on capillary tube diameter. The line and data points are an algebraic fit to many experimental data, measured at 22°C (from Brereton et al., 1998) .......... 74 The growth of a bubble in a gas-liquid solution ............ 85 Experimental data for AP" and the predicted data from the nucleation equation revised to include the energy of diffusive dissipation ..... 94 Saturation concentration for gas dissolved in water as a function of pressure at 25°C ............................. 96 Dependence of AP” on temperature for 02 dissolved in water, calcu- lated from Eq. (5.43) ........................... 99 ix 5.5 5.6 5.7 5.8 A.l A2 A3 A.4 A5 A6 A.7 A.8 Predicted and measured dependence of 02 supersaturation pressure on capillary tube diameter using Eq. (5.56) ................ 103 Predicted and measured dependence of 02 supersaturation pressure on capillary tube diameter using Eq. (5.58) ................ 105 Predicted and measured dependence of 02 and He supersaturation pressure on capillary tube diameter using Eq. (5.58) .......... 106 Predicted and measured superheat nucleation thresholds of pure water in capillary tubes ............................. 107 Saturation concentration of N2 in water ................. 112 Variation of H with pressure for N2 dissolved in water ........ 113 Saturation concentration of H2 in water ................. 114 Variation of H with pressure for H2 dissolved in water ........ 114 Saturation concentration of He in water ................ 115 Variation of H with pressure for He dissolved in water ........ 116 Saturation concentration of Ar in water ................. 117 Variation of H with pressure for Ar dissolved in water ........ 117 CHAPTER 1 INTRODUCTION Studies of nucleation have been carried out for many years and have led to some well- established, classical theories which describe rates of homogeneous nucleation, or, with some modifications for surface effects, rates of heterogeneous nucleation. They have been used to describe formation of embryo bubbles of vapor in the parent liquid, and of micro-bubbles of dissolved gases in dissimilar liquids. As a practical matter, these theories are typically devised to provide simplistic macrosc0pic descriptions of more complex microscopic phenomena. They therefore require calibrations through the use of adjustable coefficients if they are to be used for predictive purposes. In the work described in this study, our main interest is in providing a predic- tive theory for nucleation thresholds of liquids supersaturated with different gases. A second interest is in predicting effects on heterogeneous nucleation thresholds of the confinement of gas-supersaturated liquids to small volumes, since it is well known that such confinement can significantly increase the supersaturation threshold of nu- cleation. In this chapter, we first present some applications which motivate study of nucle- ation, followed by a literature review on nucleation research. The particular problems associated with the classical nucleation theory are pointed out, and the objectives of this study are given. 1.1 Applications of Nucleation Theory There are many industrial and natural processes in which bubble formation and growth play an important role, 6.9. o In the case of boat propellers or hydraulic machines, cavitation (the forma- tion of bubbles) is a problem that engineers try to avoid. The high rotational speed of propellers leads to low pressures in liquid flow which, together with micro-bubbles of dissolved air or various particles floating around, may trigger cavitation. The resulting bubbles produce noise, vibrations, and erosion of the propellers. Recent medical studies have shown that high concentrations of dissolved oxygen in the blood system can benefit patients with lung or heart diseases. However, formation of oxygen bubbles in the blood may cause serious problems and im- proved understanding of bubble nucleation thresholds is needed to develop safe medical oxygenation procedures. Bubble production in the effervescence of carbonated beverages, such as sparkling wine, beer, and soft drinks. Bubble nucleation in boiling and condensation. Divers may succumb to the ‘bends’ if bubbles of nitrogen form in their blood while surfacing. Pressure reduction promotes the formation of these bubbles. Understanding bubble nucleation can help to avoid contact vapor explosions. Under certain conditions, a liquid held at an ambient pressure of 1 atm may be superheated to close to 90% of its critical temperature before nucleation occurs. Once boiling is initiated in a liquid superheated to this degree, the liquid will boil explosively. Such explosive boiling presents a safety hazard in industry and may be involved in the dangerous explosive phenomenon known as contact vapor explosion, observed in many metallurgical processes and in paper smelt processing (Blander and Katz, 1973). In liquid waste treatment, dissolved air is sometimes used to produce bubbles that selectively carry impurities to the surface. In the petro—chemical industry, the foaming of liquid plastics is achieved through the nucleation and growth of bubbles via diffusion processes. o Diffusional nucleation and growth play an important role in transformations in solid metals, e.g. the formation of alloys, the decarburization of steel, and the development of oxidation resistant coating, etc. Understanding the mechanism of nucleation is helpful in the manipulation of a wide range of metallurgical microstructures. 0 Vacuum degassing processes are used to remove bubbles formed in molten steel or glass. 0 Bubble production is important in electrolytic processes. 0 Cavitation is useful in ultrasonic cleaning devices. a The eruption of Lake Nyos in Cameroon in 1986, which resulted in the loss of more than one thousand lives, was caused by the sudden release of carbon dioxide gas dissolved in the deeper regions of the lake. This sudden gas release is very similar to the explosive degassing of magma encountered in volcanoes (Jones, 1999). 1 .2 Literature Review The theory of nucleation and its applications have been studied extensively by many researchers. A literature review is now presented in four sections: (1) classical nucleation theory, (2) nucleation in gas-liquid solutions, (3) nucleation in confined volumes, (4) studies of bubble dynamics. 1.2.1 Classical Nucleation Theory The initial ideas of classical nucleation theory date back to Vohner and Weber’s paper, “Nucleus Formation in Supersaturated Systems,” which was published in 1926. In that paper, the authors proposed that the probability of having a critical nucleus in a supersaturated vapor is proportional to exp(—AW,-./ (kT)) according to fluctuation theory, and hence the nucleation rate should be proportional to this term, where W... is the reversible work required to form a critical-size nucleus from the supersaturated vapor, and k is the Boltzmann constant. Becker and D6ring (1935), and Reiss (1952) rederived Volmer and Weber’s nu- cleation theory on a kinetic basis. This resulted in corrections for the rate at which supercritical embryos return to subcritical sizes and for the fact that the finite rate of reaction decreases the population of critical embryos. Frenkel (1939) developed the statistical mechanics approach to nucleation theory which focuses attention on the partition functions of nuclei and how the concentration of nuclei is distributed with nucleus size. Fisher (1948) derived an expression for the work required for reversible formation of a bubble of radius r in a liquid under negative pressure. He also considered the effect of surface irregularities on nucleation. He examined heterogeneous bubble nucleation at a smooth rigid interface, a surface having a conical cavity, and a surface having a conical cavity with a rounded apex. Based on these results, he developed an expression for the maximum reversible work required to grow a bubble of critical size, as an = 167r03¢/{3(AP)2}, where 43 is a factor describing the bubble shape, a is the surface tension, and AP is the pressure difference across the bubble surface. One assumption of the classical nucleation theory is that some macroscopic pa- rameters, such as surface tension, can be applied to nuclei of only tens or hundreds of molecules. Kirwood (1949) and Buff (1950, 1955) developed a theory describing the change in surface free energy with curvature. LaMer and Pound (1949) found that experimental agreement with classical nucleation implies that the use of the macroscopic surface tension is valid even for nuclei at the size of 10‘8m. Kantrowitz (1951), Probstein (1951), Wakeshima (1954), and Frisch (1957) devel- oped theories for the lag time for redistribution of embryo sizes after an instantaneous change of supersaturation or undercooling of the parent phase. Cole (1974) reviewed theories and experiments related to boiling nucleation. He predicted superheat limits for several pure liquids from homogeneous nucleation the- ory, and found good agreement with experimental observations. The heterogeneous nucleation from plane surfaces, spherical projections and cavities was studied and the expressions of shape factors for those geometries were obtained. Also Cole reviewed experimental evidence for nucleation from a pre-existing gas or vapor phase in surface cavities. One of his conclusions was that nucleation in most boiling systems occurs from a pre—existing gas or vapor phase. The stability of nucleation cavities under conditions of superheat was also discussed. Blander and Katz (1975) reviewed the theoretical and experimental aspects of homogeneous and heterogeneous bubble nucleation. The superheat limits of many pure substances, include some organic liquids, were calculated based on classical nucleation theory, and the predictions were in good agreement with experimental measurements. The superheat limits for some liquids may reach 80% to 90% of the critical temperature. 1.2.2 Nucleation in Gas-Liquid Solutions The classical nucleation theory originally focused on nucleation in pure superheated liquids, condensation in supersaturated vapors, or crystal growth from a melt. Since the 19703, nucleation in gas-supersaturated liquids has been studied by many re- searchers, both theoretically and experimentally. Ward and co—workers (1970) investigated the role of a dissolved gas in the ho— mogeneous nucleation of bubbles from liquids. They presented a thermodynamic analysis of homogeneous nucleation in weak gas-liquid solutions. They also derived the equations for the critical bubble radius for nucleation in weak gas-liquid solutions, considering a constant-volume, closed system of uniform and constant temperature. It was shown that for nucleation to occur in a multicomponent solution a nucleus must be created of at least a certain radius (i. e. the critical radius). A generalized Kelvin equation was derived, which relates the partial pressure of the solvent vapor inside a nucleus to the properties of the surrounding solution. To verify the equation of critical radius in Ward’s (1970) paper, Tucker and Ward (1975) carried out an ex- perimental study of bubbles near the critical state in gas-liquid solutions. Bubbles at various sizes were introduced into supersaturated solutions and their growth or decay behaviors were observed. Those at, or close to, critical size were observed to be stable for long periods of time, while larger bubbles were observed to grow and smaller ones to shrink, on a timescale of hours. This experiment confirmed the correctness of the equation of critical radius for nucleation in gas-liquid solutions developed by Ward et al. (1970). Forest and Ward (1977) conducted experiments to measure the pressure at which homogeneous nucleation of bubbles occurs in solutions of nitrogen in ethyl ether. The presence of the dissolved nitrogen in the ethyl ether was observed to raise the pressure at which nucleation occurs. Blander and Katz (1975) discussed hydrodynamic and diffusion constraints on nu- cleation rate. A modification of classical nucleation theory was proposed for diffusion- controlled nucleation, e.g. nucleation in a mixture with one volatile component. A pre—exponential coefficient (SD, which accounts for the diffusion of a volatile component to the interface, was added to the nucleation rate equation as JD = J/(1+5o) where J D is the nucleation rate taking into account diffusion. However this theory does not appear to have been tested against experimental results, and will be discussed in detail in chapter 5. Hemmingsen (1974) conducted experiments to study nucleation / cavitation of gas- supersaturated solutions in glass tubes (0.01—0.25 cm inside diameter). He observed that water equilibrated with gas at high pressures is able to sustain very high supersat- uration without any occurrence of gas bubbles during decompression to atmospheric pressure. The maximum supersaturation tensions without cavitation, which repre- sent the lowest possible limits of spontaneous bubble formation, were measured for 02, Ar, N2, and He. Finkelstein and Tamir (1985) carried out experiments on bubble formation in water supersaturated with a number of different gases. The pressure difference for bubble formation, AP” —- an expression for the degree of supersaturation which a solution can tolerate — was measured for difl'erent gases. While an empirical equa- tion was found to correlate AP” with critical properties of the gas, the theoretical explanation of this equation was questionable and the equation is inconsistent with other experimental results. It will be discussed in detail in chapter 4. Wilt ( 1985) studied mechanisms of nucleation in supersaturated COg~water so- lutions, such as carbonated beverages. He examined several possible mechanisms: homogeneous nucleation, and heterogeneous nucleation at smooth planar interfaces, and at interfaces containing conical or spherical cavities or projections. He concluded that the predominant mode of nucleation was heterogeneous and that it occurred in conical cavities, presumed to be present in some form on almost all surfaces, but only for contact angles in the range of 94 — 130°. Moreover, his analysis did not take into account the possible influence of pre—existing gases trapped in the cavities, which can provide nuclei for the formation of gas—vapor bubbles at the onset of nucleation. Carr et al. (1995) carried out experimental studies on bubble nucleation rates for the H20/C'02 / Pyrex system over the supersaturation range 1 — 7, using the pressure release bubble nucleation technique. The dependence of nucleation rate J on supersaturation ratio w was reported as a plot of an vs. (wP)‘2, but did not yield a linear relation as might be expected from Wilt’s work. Carr explained the disagreement in this way: there exists a distribution of active nucleation sites of different sizes or geometries on the Pyrex surface; each site type displays its own rate kinetics, and the overall bubble nucleation rate is the sum of the rates from each site type. A two—site model was proposed and fitted the experimental data well. Again, Carr et al. did not consider the possible effect of pre—existing gases trapped in the cavities. Rubin and Noyes (1987) measured supersaturation thresholds for homogeneous nucleation of H2, N2, 02, CO, N O, 002 dissolved in water, using a new experimen- tal approach. Instead of saturating water with gas at high pressure then decompress- ing, as Finkelstein and Tamir had done, chemical reactions were used to generate highly supersaturated solutions. Thus the cumbersome compressors and steel vessels otherwise used to generate high pressures were not required in their experiments. Supersaturation thresholds for nucleation could then be measured at atmospheric pressure and room temperature. Bowers et al. (1999) reported temperature dependences of supersaturation thresh— olds for carbon dioxide, hydrogen, and oxygen in water at 1 atm pressure. The thresh- olds were measured by generating supersaturated solutions of the gases chemically. As temperatures increased, the supersaturation limits decreased for 002 and 02, but increased for H2, and remained at almost the same value for N2. This result was of considerable interest, since according to the classical nucleation theory, all the super- saturation limits should decrease as temperature increases. Further study is needed to explain this phenomenon. Since the nucleation rate is extremely sensitive to surface tension, it is important to obtain surface tension data of different gases dissolved in water at different pressures and concentrations. Unfortunately, the metastable nature of the gas—supersaturated solution makes this kind of experiment very diflicult. Few experimental data are avail- able. Massoudi and King (1974) measured the surface tension of various gas-water solutions at pressures up to 85 atms, using the capillary-rise method. Measurements were reported for water with He, H2, 02, N2, Ar, CO, 002, N20, CH4, CgH4, C2H6, C3H8, and n-C4H10. Their results showed that pressure has a significant effect on surface tension for C02, N20, CH4, 02H4, CQHG, C3H3, and n-C4H10 dissolved in water, but has little effect for He, H2, N2, or Ar solutions with water. Lubetkin and Akhtar (1996) measured surface tension and contact angles of carbon dioxide-water solutions at pressures in the range 1-11 bars. The surface tension of COg-water solu- tions changes from 0.072 N / m to 0.057 N / m as the pressure of 0'02 increases from 1 to 11 bars. The advancing contact angle for aqueous 002 solutions on 316 stainless steel changes from 77° at P = 2 bars to 44° at P = 10 bars. The effect of the changes of surface tension and contact angle with pressure on nucleation rate was also studied. Jones et al. (1999) reviewed bubble nucleation from cavities. They argued that the nucleation observed in most instances, which is often at supersaturation ratios of 5 or less, is invariably associated with the existence of gas cavities in walls of the container. Issues concerning the formation of these gas filled cavities, and their stability was examined. As an alternative approach to classical nucleation theory, Talanquer and Oxtoby (1995) applied a density functional method to predict the nucleation rates of bub- bles in superheated, stretched, or supersaturated binary fluid mixtures. The density functional method assumes continuous changes rather than abrupt changes of fluid properties at the interface of nuclei, and thus avoids the macroscopic assumptions of classical nucleation theory. Although the density functional method is not yet able to make quantitative comparisons with experimental results, it shows some potential for making useful predictions in nucleation studies. 1.2.3 Nucleation in Confined Volumes When carrying out studies of nucleation in gas—liquid solutions, several researchers have drawn attention to the effect of volumetric confinement on raising nucleation thresholds. Hemmingsen (1974) conducted experiments to study nucleation / cavitation of gas-supersaturated solutions in glass tubes (0.01~0.25 cm inside diameter). He ob- served that water equilibrated with gas at high pressures is able to sustain very high supersaturation without any occurrence of gas bubbles during decompression to atmo— sphere pressure. The maximum supersaturation tensions without cavitation, which represent the lowest possible limit of spontaneous bubble formation, were measured for 02, Ar, N2, and He. Brereton et al. (1998) studied nucleation in small capillary tubes, of inside di- ameters of less than 100 am. It was observed in these experiments that confinement within small capillary tubes can increase the supersaturation threshold for different gases dissolved in water. A dual-dependence model was proposed to describe nucle- ation of gas—supersaturated solution within confined volumes. Although this model could. be calibrated to agree with experimental data, some of its assumptions required more careful consideration, e.g. the assumption that the changes in the radii of cur- vature of nuclei in crevices are neglected during decompression is questionable. Also calculations of the critical radii of nucleation in gas-liquid solutions did not take into account the gas species present. These shortcomings will be discussed in detail in chapter 4 and chapter 5. Some measurements of superheat thresholds of pure water in small capillary tubes were also reported. 10 1.2.4 Studies of Bubble Dynamics Bubble dynamics has been studied extensively by many researchers. The case of stationary spherical bubbles surrounded by an infinite, constant temperature liquid is of particular interest in the later stages of nucleation. Epstein and Plesset (1950) found an approximate analytical solution of the growth/ collapse rate for a gas bubble in a gas-liquid solution. The equations governing spherically symmetric growth of a bubble in an infinite medium were formulated by Scriven (1959). Bankoff ( 1966) reviewed several cases of diffusion—controlled bubble growth: (1) asymptotic growth of a vapor bubble in an initially uniformly superheated one-component liquid; (2) bubble growth under nonuniform initial conditions; and (3) bubble growth in two-component liquids. The role of bubble formation and growth in boiling heat transfer was also studied. Bretherton (1960) studied the motion of long bubbles in a tube filled with liquid at small Reynolds numbers, and Plesset and Chapman (1971) studied the collapse of an initially spherical vapor cavity in the neighborhood of a solid boundary. Ward (1984) studied the conditions for stability of bubble nuclei in solid surfaces contacting a liquid-gas solution. Worden (1998) investigated the mass transfer of micro-bubbles by developing a dynamic model of a single micro-bubble immersed in an infinite stagnant liquid which was solved numerically. The model was used to predict mass transfer coefficients in fluids with micro-bubbles. The general conclusion from these studies is that, after formation of bubbles larger than their critical radii, their subsequent growth in uniform media without interaction with other bubbles is well understood. Theoretical and computational results are in good agreement with experiment, which indicates that the continuum-level aspects of bubble growth are quite reliable for predictive purposes. It is the micro—scale aspects of nucleation where significant difficulties lie. 11 1.3 Objectives Although the subject of nucleation in supersaturated and superheated liquids has been studied extensively, understanding of the thresholds necessary for nucleation remains incomplete and theoretical predictions can not always be reconciled with experimental results, especially those involving different gas species dissolved in water. In classical nucleation theories, one typically focuses on clusters of nuclei in a liquid and considers when they are sufficiently numerous and energetic to combine to create micro—bubbles which are large enough to avoid collapse, and grow to a finite size. Some particular problems with these classical theories, which limit their usefulness in practical applications, are: 1. they typically associate the conditions required for bubble formation only with the reversible work required to create bubble surfaces of stable size, without regard for mass transfer to the bubble surface and other irreversible effects, e.g. diffusive dissipation, viscous dissipation, etc. 2. a nucleation model calibrated for one dissolved-gas species can be highly inac— curate if applied to a different gas-supersaturated solution. 3. they do not predict any effect of volumetric confinement on nucleation thresh- olds of gas-supersaturated liquid, when experiments indicate there is one. The purposes of this study are to examine carefully the bases for nucleation the- ories and to extend them to provide predictive models of nucleation of different dissolved-gas species in liquids, and models of effects of confined volumes on nu- cleation thresholds. In this dissertation, we first present the classical nucleation theory as originally developed for pure liquid-vapor mixtures, and as extended for gases dissolved in dissimilar liquids. The reasons for its failure to predict the strong dependence of the particular gas species on nucleation thresholds of gas-supersaturated liquids will be 12 discussed. Modifications are then made to the classical nucleation theory so it can predict nucleation thresholds that are in good agreement with experimental data. Finally the prediction of nucleation thresholds in small capillary tubes is described. 13 CHAPTER 2 THE THEORY OF N UCLEATION The theory of nucleation has been developed from the early ideas of Volmer and Weber (1926), Becker and D6ring (1935), and Frenkel (1955). From the classical nucleation theory, the homogeneous bubble formation rate in a pure liquid can be expressed as (see, for example, Carey, 1992) J _—_ Nzgexp{—an/(kT)} (2.1) where Wm is the maximum reversible work required to form a bubble nucleus, N1 is the number of liquid molecules per unit volume, a is the surface tension, m is the mass of one molecule, k is the Boltzmann constant, and T is the absolute temperature. Usually the maximum work occurs when the bubble nucleus is in a state of equilibrium with the liquid (Fisher, 1948). In this dissertation, we will first examine the bases of the classical nucleation the- ory and its shortcomings. Then the classical nucleation theory is modified to describe heterogeneous nucleation in gas-liquid solutions and nucleation in small capillary tubes, based on the work of Ward, Carey, Wilt, and Brereton et al. Our discussion begins with an analysis of the state of equilibrium between liquids and gases, and the stability of bubbles. 2.1 The Equilibrium State 2.1.1 A Pure Liquid in Equilibrium with its Vapor From considerations of equilibrium, it is possible to determine the radius of a gas bubble in a liquid as a function of liquid pressure, temperature and the interfacial surface tension. For a pure liquid in equilibrium with its vapor, contained inside a 14 spherical bubble (see Figure 2.1), the temperature and the chemical potential in the two phases must be equal, (Gyftopoulos, 1991) u:n:T an #1 2 ”v (23) where the subscripts l and 1) denote liquid and vapor respectively. Due to the cur- vapor liquid L—__ ___ Figure 2.1: System model of bubble formation in a liquid vature of the interface, the pressures in the vapor bubble and the liquid are related through the Young-Laplace equation, as a statement of static mechanical equilibrium, as 2 a=e+i no 8 where re is the equilibrium bubble radius, and o is the surface tension. After inte- grating the Gibbs-Duhem property relation du = —sdT + vdP (2.5) 15 at constant temperature from P 2 PM to an arbitrary pressure P, we obtain P a -— amt : / vdP (2.6) PsatlTl) For the vapor phase, substituting the ideal gas law (1) 2 RT/ P) into Eq. (2.6) yields the following relation for the chemical potential of the vapor: Pv “1‘va Tu) I ,usatmcll) + R7“! 111 (P804710) (27) Since the liquid is virtually incompressible, v is taken to be constant. Evaluation of the integral in Eq. (2.6) for P : P, then yields ”KB, Tl) : ,usatJCFl) + vllB — Psat(:ll)l (28) Substituting Eqs. (2.7)and (2.8) into Eq. (2.3), and noting that amt,v(T) = #sat,z(T), allows the vapor pressure inside the bubble at equilibrium to be expressed as (2.9) vzlpz - Psat(Tl)l} P,, = Psat(T,) exp{ RT; Substituting Eq. (2.9) into Eq. (2.4) leads to an equation for the bubble radius (re) at equilibrium: 2o Psat(7l)eXP{vllH _ Psatawl/(RTlll — Pl (2.10) re: 2.1.2 Equilibrium in Gas-Liquid Solutions For the case of an arbitrary gas, rather than the liquid’s own vapor, dissolved in a 1iquid, Ward et al. (1970) derived equations for the equilibrium state in weak gas-liquid solutions and the corresponding equilibrium radii of dissolved-gas/ vapor 16 bubbles. The weak-solution assumption (gas concentration in the liquid C’ << 1) is almost always true for gases dissolved in liquid, even at very high pressures (e.g. 10 MPa). The conditions for equilibrium described by Eqs. (2.2) and (2.4) remain valid, but in addition the chemical potentials of each component in the two phases must be equal (Gyftopoulos, 1991) [11,1 2 [11,” (2.11) #21 = H2,v (2-12) where subscripts 1 and 2 refer to the solvent and solute respectively. In the liquid phase, the chemical potential for the solvent is (Gyftopoulos, 1991), #1,: = #01(B,T)+RTln(y1,z) 1 = P T 1 M01( 1, )+RTn1+C’ z ugl(H,T)—RTC' (2.13) where #01 is the chemical potential of pure substance 1 and y” is the mole fraction of component 1 in the liquid phase. C’ is the concentration of component 2 in the liquid phase, which is defined by the quotient of mole number of the solute over that of the solvent. If the pure solvent is assumed to be incompressible, then #1,! : “01(PsataT) + v1,l(H _ Psat) — RTC’ (214) For the solute, the chemical potential can be expressed as (Landau, 1958) n.2,, : 1MP), T) + RT ln(C') (2.15) 17 where a is a function that depends only on B and T. In the case of liquid 1 saturated with gas 2 at (B,T) across a flat surface, with the saturation concentration as CO, the chemical potential of component 2 is am 2 ¢(H,T) + RT ln(C0) (2.16) Solving for 2/2 from Eqs. (2.15) and (2.16), and substituting back into Eq. (2.15), gives the chemical potential of component 2 in the liquid phase #2,: = #02001, T) + RT 1n(C’/Co) (2.17) In the gas/ vapor phase, the chemical potentials of the vapor and the gas in bubbles are H1,v = H01(P1,mT) Z 1101(1)”, T) ‘l‘ RT1D(P1,v/Pv) 1 = P,a,T RTl P, P3,, RTl (___) #01( t )+ Ill / 0+ n1+C” #24; == #02(P2,v,T) 2 ”02(1)”, T) + RT111(P2’v/R,) : 110201,?) + RT ln(Pv/Pz) + RTln (150”) (2.18) where C” is the concentration of the gas in the gas—vapor bubble. After introducing activity coefficients V1 and V2 to account for the non-ideal effects of the gas and solution, the two above equations become “L, : ”01(1),,“ T) + RT 111(Pv/P3at) + RT 1n (1 :10”) (2.19) 18 II My ___ ”02(1),,7") + RT1n(P,,/P,) + RT ln (16+: ’23) (2.20) Substituting Eq. (2.14) and Eq. (2.19) into Eq. (2.11), Pvl V1 l : Psat(T) exp {MAB ‘— Psat) "‘ RTC' } 1 + C” RT E pm, ,7 (2.21) Usually 1) is approximately equal to unity. Substituting Eq. (2.17) and Eq. (2.20) into Eq. (2.12), one finds [ C’IV’ ] = C’ (2.22) l + CH ’60 Combining Eq. (2.21) and Eq. (2.22) with Eq. (2.4), the expression for the equilibrium radius of a spherical bubble in a gas-liquid solution can be obtained as P,“ T PC’ 77 t( 0+ I Pl} — 2.2 V1 V200 ( 3) 7‘, =20/{ If the gas-vapor mixture inside the bubble behaves as an ideal gas mixture and the solution is ideal, then V1 2 V2 2 1. If V1 2 V2 2 1 and C’ = 0, Eq. (2.23) is reduced to Eq. (2.10), i.e. the equilibrium radius of vapor bubble in pure liquid. Note that the equilibrium radii for dissolved gases are always smaller than those of pure vapor bubbles. Therefore nucleation of gas-liquid solutions takes place more readily than in pure liquids. 2.2 The Stability of a Bubble in a Liquid Having established expressions for the equilibrium size of bubbles in liquids, we now consider the question of stability, and when bubbles at their equilibrium size will be stable or unstable. 19 2.2.1 The Stability of a Vapor Bubble in a Pure Liquid If an embryonic bubble has been formed through the random collision of clusters of vapor molecules in a pure liquid, we would like to determine whether such a bubble is stable or unstable, and if unstable, whether it will grow or collapse. The system model is shown in Figure (2.1). The liquid is kept in a closed system, at constant temperature and pressure, as is typically the case in most practical applications. In that case, the system reaches stable equilibrium when its Gibbs function is a minimum, or a maximum for unstable equilibrium. For convenience, we first deduce the Helmholtz free energy F of the system after a bubble of radius 7‘ is formed and, from F, determine the Gibbs function. By definition F = U —TS, where U is the system’s internal energy and S its entropy. F is extensive, or additive over the subsystems in a composite system, since U and S are additive. Therefore the Helmholtz free energy for the whole system is the sum of contributions of the liquid, the vapor and the interface. F=E+R+E QM) where F1 = Nz(uz*-Tz81) R==Mwwowg an) F,- 2 04m"2 and N), N, are the number of moles of the liquid and the vapor. Combining these equations with the equilibrium assumption that T, Z T, = T, we obtain F : N1(u, — Tsl) + N,,(uv -— Tsv) + o 47TT2 (2.26) 20 By definition, the Gibbs free energy G is given by G _—_ U—TS+PV = F + PV (2.27) Using Eq. (2.26), we can rewrite the Gibbs free energy in the form of (Carey, 1999) G : F + P1(N1'U( + vav) : N.,g.. + Mg, + 4m2a, — va,,(P,, — 19,) (2.28) where the specific Gibbs free energies for the liquid and the vapor, g, and gv, are defined as g; = U1 — T81+ PM 91) : uv ‘- T811 + vav (2.29) The Gibbs free energy of the system before bubble formation is given by Go 2 (NU + N091 (2.30) Here we assume that the formation of a tiny bubble does not cause any change of the properties of the liquid. The change in Gibbs free energy associated with bubble formation is therefore AG 2 G—Co 2 N,,(g,, - 91) + 47rr20 — va,,(P,, — B) (2.31) For a pure liquid, u = g and at equilibrium, u) = 11.9, and P, — P, = 2o/re. Also, 21 37W”, it follows that at equilibrium, Eq. (2.31) becomes since vav = 3 A0,. = -7rr§o (2.32) If we perform a Taylor-series expansion of AG(r) in Eq. (2.31) around the equilibrium point, we find that AGO) : (Nvle + ANv)l(gv — gl)le + A(gv — gl)l +47r7'2o — gnr3[(Pv — Pz)|,.3 + A(Pv — P,)] (2.33) where Af = 2;; J" — r.) + 0er and the subscript e refers to properties at the equilibrium point. The symbol ‘0’ means ‘of the order of’. Substituting (9,, — gz)|e = 0 and (P, —- P))|e : 20/1}, into Eq.(2.33) and retaining only the first order terms, we find that AG(7‘) : 47ror2(1 — $7.1) + O(AT) (2.34) A plot of AG versus 7‘ is shown in Figure (2.2); by the nature of the local expan- sion, it is only accurate near the equilibrium point. From the graph, it can be seen that the equilibrium point at re is an unstable equilibrium. If a bubble nucleus is created by fluctuations of initial size less than the equilibrium/ critical radius, it will collapse spontaneously. A nucleus of initial size greater than the equilibrium radius will grow spontaneously. It is noteworthy that this result is based on equilibrium be- ing reached when the Gibbs function is minimized or maximized, which is equivalent to the assumption that nucleation takes place in a closed system at constant T and H. 22 3 A 0 <1 0 I I 1 0 0.5 1 1 5 rlr. Figure 2.2: The curve of Gibbs function in the neighborhood of the equilibrium radius 2.2.2 The Stability of a Bubble in a Gas-Liquid Solution In the following section, the corresponding stability of a gas-vapor bubble in a gas- liquid solution is studied. We consider a closed system of uniform and constant temperature, in which gas-phase bubble nuclei are surrounded by liquid. The pressure of the liquid is kept constant and effects of gravity are ignored here. In both phases, there are 17. components. As in the previous section, it is convenient to deduce first the system Helmholtz function, from which the desired Gibbs free energy G can be found. The Helmholtz function F can be expressed in terms of the surface area of the nucleus A, the number of molecules of each component in the liquid phase Nu and vapor phase NW, the pressure of both phases Pl, Pv, temperature, entropy, etc. as F = Z -Ni,l(ui,l — T8231) + Z Ni,v(U-i,v — TSi,v) + 014 (2-35) Where subscript i denotes the i-th component. By definition, the Gibbs free energy is given by G : U—TS+PV 23 = F + PV (2.36) Using Eq. (2.35), we can rewrite the Gibbs free energy in the form G = F + P; (N12); + Nvuv) Z Z Name: + 2 New“, + 0A — (PU — 190% (237) where u,,landu,,v are defined by [1:31 = uiJ — T8131 + szu Him : ui,v _ Tsim 'l' vaim (238) The Gibbs free energy of the system before bubble formation is given by G0 = ZUVW’ + Ni,1)/li,[ (2.39) O l The change in Gibbs free energy associated with bubble formation is therefore AG : G—Go = Z Nada... — m) + a A — (P. — PW. (2.40) In these derivations, it is assumed that the formation of a tiny bubble will not change the properties of the liquid, 6. g. 131,11”, etc. At equilibrium, at, = #1312 and P.” — B = 2o/re. Since vav = gnr3, it follows that, at equilibrium, Eq. (2.40) becomes AG, 2 gargo (2.41) 24 If we carry out a Taylor-series expansion of AG(r) in Eq. (2.40) around the equilibrium point, we find that AG(1~,N,~) = 21{(Ni.vle + AN,,.,)[(11,-,,, - Hi,l)le + A01”, — Mull} +47r7‘2 o — —7rr 3,,[(P— Pl)|e + A(P,,— P1)] (2.42) where 3_f( 8r M:— e -- Ni,v,8)} + 0(AT2, ANE, AT ANz) . ’ZloN. the subscript e refers to properties at the equilibrium point, and AN, 2 Ni,” — NW8. Substituting (11,,” — Hi,l)|e = 0 and (PU - Pl)|e = 2o/re into Eq.(2.42) and retaining only the leading and first order terms, we find that 2 AG(7‘, Ni) 2 47ro'r2(1 — 37'1)+ Z {(Nwle + AN,,v)A(1i,-,v — u,,1)} 4 — §7TT3A(PU — H) : 47(07‘2 (1 — ’37:.) + 0(A7‘, AN,) (2.43) e The leading term of AG(r, N) for a bubble formed in a gas—liquid solution is not dependent on N,, and it is identical to that in a pure liquid, Eq. (2.34). But the expressions of re are different for nucleation of pure liquids and of gas-liquid solutions. A plot of AG versus r is shown in Figure 2.2 which, through the nature of this local expansion, is only accurate near the equilibrium point. From the graph, it can be seen that the equilibrium point at re is again an unstable equilibrium. If a nucleus is created by fluctuations of initial size less than the equilibrium/ critical radius, it will collapse spontaneously. A nucleus of initial size greater than the equilibrium radius will grow spontaneously. It is worth emphasizing that this result has been 25 deduced under the assumption of a closed system at constant T and Pl. Ward et al. (1970) obtained the same result for a constant volume, closed system at uniform and constant temperature. Thus embryo bubbles of both vapors and dissolved gases which are dissimilar in composition to the host liquid can only grow to finite sizes if they can be formed by random molecular collisions, at sizes greater than the equilibrium radius re. Again, because equilibrium radii for dissolved gases are always smaller than those for vapors, dissolved gases will, with all other factors being equal, always nucleate more readily than vapor bubbles. 2.2.3 The Critical Radius and Gas Concentration Gradients It was shown in the previous section that there exists a critical radius for bubble formation in a gas-liquid solution, from the viewpoint of minimization of the Gibbs free energy. Once a bubble is formed through random fluctuations, it tends either to grow or to collapse depending on whether the radius of the bubble is larger or smaller than the critical radius. Now we consider the same problem from another viewpoint: the gradient of gas concentration across the bubble surface. Assume a gas-vapor bubble of radius r is formed in a uniform gas-liquid solution. Its subsequent growth or dissolution depends on whether the net mass transfer of gas molecules is into or out from the bubble surface. From Fick’s law of mass transfer, the concentration gradient across the bubble determines the direction of mass transfer in continua. Therefore, the gas concentration at the bubble surface can be derived and compared with the gas concentration in the liquid. From the Young-Laplace equation of mechanical equilibrium, the pressure inside the bubble is 2 P. = B + 70 (2.44) The gas concentration at the bubble surface is the saturation pressure corresponding 26 to the partial pressure of the gas inside the bubble, CIR : Csat(Xga,P,,) (2.45) where X gas is the mole fraction of gas in the bubble. For nucleation of gas-supersaturated liquids, it can be shown from Eq. (2.21) that X gas 2 1. The gas concentration in the liquid far away from the bubble is Coo : sat(Pssat) (246) where P3,“, is the saturation pressure corresponding to the initial gas concentration of the solution. Thus, the bubble will grow or dissolve according to the relative magnitudes of C I R and Coo. The bubble will grow if C | R < Coo, or 2 X90, (P, + 7") < Pm. (2.47) We can see that as r approaches zero, the relation (2.47) will not be satisfied and the bubble will always dissolve. At large r, the bubble will grow. If Henry’s law is satisfied, from Eq. (2.23) we obtained 20 7'_ z ”PsatCI—l) + Pssat — [)1 (248) At equilibrium, 2o Xgas (Pl + _) it: Xgas (”PsatCPI) + Pssat) % Pasat (2.49) Therefore at the critical size re, there is no gas concentration gradient across the bubble surface. Profiles of gas concentrations for bubbles smaller than and greater 27 than the critical radius are shown in Figure 2.3. In summary, if a bubble is created by fluctuations of initial size less than the equilibrium/ critical radius, it will dissolve spontaneously. A bubble of size greater than the critical radius will grow spontaneously. For nucleation processes, extra energy is needed to overcome the concentration gradient for a bubble to grow from sub-critical size to critical size, since these growing processes are in the opposite direction to gradient-driven Fickian diffusion. 2.3 The Nucleation Rate Equation From Figure 2.2, the maximum reversible work required to form a spherical bubble nucleus is reached at the equilibrium point, 4 an 2 Ewarg (2.50) Wm can be regarded as the activation energy barrier for bubble nucleation that may be reached by thermal and other fluctuations in the system. According to kinetic theory, the probability of such an occurrence is proportional to the factor exp(— 9513;“ . More careful considerations by Volmer and Weber (1926), Becker and Doring (1935), Reiss (1952), and Barnard (1953), et al., gave the equation of homogeneous nucleation rate in a liquid as, J = N,\/%exp{—Wm/(kr)} (2.51) where N, is the number of liquid molecules per unit volume, m is the mass of one molecule, k is the Boltzmann constant, and T is the absolute temperature. The homogeneous nucleation rate J is in units of 1/ (1113s). Substituting Eq. (2.50) into Eq. (2.51), the homogeneous nucleation rate in a pure 28 VR Rfe , bubble growing. mass transfer from the surrounding liquid Into the bubble gas diffusion C’i‘ lL————: R r Figure 2.3: Gas concentration profiles in the liquid and in the bubble 29 liquid or a weak gas-liquid solution can be obtained as J 2 N,fi%exp{—§7rarg/(kT)} (2.52) where re can be calculated from Eq. (2.23) for gas-liquid solutions, or Eq. (2.10) for pure liquid. Liquid Wmm e ///// /// ///// Solid Figure 2.4: A vapor bubble formed at a smooth liquid-solid surface For heterogeneous nucleation, a shape factor d), which depends on the contact angle and the geometry of the surface, is often introduced to describe the maximum reversible work required to form a non-spherical bubble at the surface. The shape factor is defined as the ratio of the work required to form a non—spherical bubble to the work to form a spherical bubble of the critical size. It always satisfies 0<¢gl Therefore, the heterogeneous nucleation rate is expressed as J = Nl2/3\/%exp{—§qbrrar3/(kT)} (2.53) . . . 2 3 . . where J 18 1n umts of l/mgs, and N,/ represents the number of liquid molecules immediately adjacent to the solid surface per unit of surface area. For a heterogeneous 30 Figure 2.5: Definition of geometrical parameters in a conical cavity nucleation process, only molecules near the solid surface can participate in embryo bubble formation. For bubble formation on a smooth flat surface as shown in Figure 2.4, if the bubble shape is idealized as being a portion of a sphere, then the shape factor qb takes the form _ 2 +3cos€ — c0536 4 (2.54) <13 where 6 is the contact angle measured in the liquid. For bubble formation in a conical cavity as shown in Figure 2.4, the shape factor (15 takes the form (Wilt, 1985) (25(6’, 5) = {2 — 2sz’n(0 - B) + 6036 0032(6 - (3)/Sinfi}/4 (2.55) Table 2.1 lists the values of shape factor at selected 6 and [3. 31 Table 2.1: Values of shape factor for bubble formation in conical cavities 9 ¢(@fi = 5°) ¢(@fi = 10°) 60 0.562E+00 0.414E+00 62 0.480E+00 0.362E+00 64 0.405E+00 0.314E+00 66 0.337E+00 0.269E+00 68 0.276E+00 0.227E+00 70 0.222E+00 0.190E+00 72 0.175E+00 0.157E+00 74 0.135E+00 0.127E+00 76 0.101E+00 0.101E+00 78 0.728E—01 0.784E—01 80 0.504E—01 0.594E—01 82 0.330E—01 0.436E—01 84 0.201E—01 0.308E—01 86 0.111E—01 0.207E—01 88 0.5211302 0.131E—01 90 0.1901302 0.760E—02 92 0.411E—03 0.389E—02 94 0.152E—04 0.164E—02 2.4 Effects of P, T and Concentration on Nucleation Rate The expression for the equilibrium/ critical radius of a spherical bubble in a gas—liquid solution, Eq. (2.23), is restated here as Re=2U/{ 1/1 where n is approximately equal to unity, and V1 2 V2 ideal solution. According to Henry’s law (which states that the concentration of a gas in a liquid is proportional to the pressure), 51—9, 2: Pssat, therefore Eq. (2.56) simplifies to Re : 20'/{ V1 ”PsatCFI) + BC, _- Co Pi 11200 ”Paddy-l) + Pssat 32 __H} V2 1 for an ideal gas and an log(JIJo) PI’PIO Figure 2.6: The effect of the liquid pressure on the nucleation rate As the pressure of the liquid increases, the equilibrium / critical radius will increase. This leads to the decrease of nucleation rate according to Eq. (2.52) or Eq. (2.53). Figure (2.6) shows the variation of the nucleation rate with the pressure of liquid. J0 in the figure is the heterogeneous nucleation rate for an 02-water solution at H = 1 atm, T) = 20°C, and shape factor q5 = 0.01. The gas concentration in the liquid (0’) is the saturation concentration at Pam : 10 MPa. It can be seen from Figure (2.6) that the nucleation rate will decrease roughly 10 orders of magnitude, if the pressure of the liquid increases by 1 order. Because most of the temperature-dependent quantities appear in the exponential term of the nucleation rate equation (2.52), a slight change in T; can have a signif- icant effect on the rate of nucleation. As the temperature of the liquid increases, Psat(T1) increases, 0 decreases, re decreases, and the exponential term in Eq. (2.52) increases sharply. Although a in the pre-exponential term decreases a little, overall the nucleation rate will increase exponentially with temperature. Figure (2.7) shows 33 45 — 4o — 35 - 30 - 25 ~ 20 . 15 - 1o4 5 u 0 TV 1 l l l 270 290 310 330 350 370 W") log(JIJo) Figure 2.7: The effect of the liquid temperature on the nucleation rate the variation of the nucleation rate with the temperature of liquid. As the concentration of gas in the liquid C" increases, (i.e. the supersaturation pressure increases), re decreases, so the nucleation rate J will increase exponentially. Figure (2.8) shows the variation of the nucleation rate with the supersaturation pres- sure. Presumably, this is the initial nucleation rate, because prolonged nucleation will reduce C". 34 451 405 35) soi log(JIJo) o . v T . o 5 10 15 2o Pant (MPa) Figure 2.8: The effect of the supersaturation pressure on the nucleation rate 2.5 Discussions on the Applicability of Henry’s Law and the Ideal-Gas Law at High Pressures In this section, the applicability of Henry’s Law and the Ideal-Gas Law are dis- cussed together with improvements on them which account for departures from ideal gas/ solution behavior at high pressures. 2.5.1 The Applicability of Henry ’3 Law Henry’s law states that there exists a linear relation between the partial pressure of a gas P9 and its mole fraction mg when dissolved in a dilute solutions, P9 = :29 H (2.58) where H is the Henry’s law coefficient, which usually can be regarded as a constant and be evaluated from H : lim0 Pg/arg (2.59) 35 or more precisely, H = 11510 fg/zg (2.60) where fg is the fugacity of the gas in the solution. For a gas dissolved in water, the vapor pressure of the solvent is negligible and the total pressure can be treated as the partial pressure of the gas, without any appreciable error. The reliability of Henry’s law at high pressures or dissolved gas concentrations can be checked from experimental measurements of P9 and concentration. Figure 2.9 0.005 r 0.004 2 0.003 r 0.002 * Concentration 0.001 4 0 l _ T T l 0 5 10 15 20 Pressure (MPa) b— Henry’s Law + Experimental data Figure 2.9: Measurement and Henry’s-law prediction of the saturation concentration of 02 in water shows the saturation concentration (as a mole fraction) of 02 in water at 25°C, as a function of pressure. The experimental data were obtained from Brereton et al. (1998) and are converted to molar units. The Henry’s constant for 02 in water at 25°C and 1 atm is 1.26 x 10‘3mol/ (d1113at1n) (Bowers, 1996). From Figure 2.9, when the pressure is less than about 6 MPa, the error of Henry’s law is less than 5%. To provide improved levels of accuracy, corrections should be made to Henry’s constant before applying it at higher pressures. 36 A theoretical expression of the pressure dependence of the Henry’s law coefficient is given by the Krichevsky—Kasarnovsky equation (Krichevsky, 1935), ln(H/H0) = —(P — POW/(RT) (2.61) where H is defined by P/atg, l7 is the partial molar volume of the dissolved gas in water, and subscript 0 represents a reference point. Normally, V can be treated as a constant at different pressures. It follows that there is a linear relation between log(H) and P as can be seen from Figure 2.10. A least-squares fitted line relating H 5i 4.5 f 31' 41 2 ‘--3¢—¢ 3.5—fl" 3 #1 1 1 1 0 5 10 15 20 Pressure (MPa) Figure 2.10: Variation of H with pressure for 02 dissolved in water of 02 dissolved in water at 25°C to P takes the form: log(H) = 3.6022 + 1.1305 x 10-2 P (2.62) where P is in MPa and log is 10—based. Once H is calculated for a given P, the 37 saturation concentration is readily found from 2:9 2 P/ H. The same linear relation between log(H) and P exists for some other low solubility gases dissolved in water, e.g. N2, H2, He, Ar, etc. More results are given in Appendix A. According to this discussion of Henry’s law and ideal solution behavior, it appears that, for pressures higher than 7 or 8 MPa, models for dissolved gas concentration that are more general than Henry’s law are required. 2.5.2 Eflects 0f the Departure from Henry’s Law on Nucleation Rate Having shown how the dissolved gas concentration at high pressures can be described by a generalized form of Henry’s law (Eq. 2.61), we now consider the effect of depar- tures from Henry’s law on nucleation rates. For simplicity, the classical heterogeneous nucleation rate equation (2.53) 2 J : N12/3‘/§%exp{—4i::fe } (2.63) is used to calculate the nucleation rate of 02 dissolved in water. The critical radius 1",. is affected by the concentration of gas in the liquid C' as (2.64) Pm T P ’ 7,8220/{17 .t(1)+ [C —B} V1 V200 where the concentration of gas in the liquid 0’ is evaluated from c' = P/H 38 Figures 2.11 and 2.12 show the heterogeneous nucleation rates at different shape factors ((15 = 0.001 and 0.005) using the generalized Henry’s law. The nucleation rates corresponding to Henry’s law are also shown in these figures for comparison. Some conclusions can be drawn from those figures: 0 Use of the generalized form of Henry’s law always lowers the nucleation rate. The effect is more significant at high pressures. The reason for this is that the generalized form of Henry’s law gives lower gas concentration at the same pres- sure, compared with Henry’s law. Lower concentration leads to larger critical radii according to Eq. (2.64), and thus smaller nucleation rates. 0 For smaller shape factors (which are more appropriate to heterogeneous nucle- ation region), the effect of the departure from Henry’s law on nucleation rate is less significant, but can still be important. 201 log(J) 3 supersaturation pressure (MPa) + Henry's law + Generalized Henry's law Figure 2.11: Effect of the departure from Henry’s law on nucleation rate ((0 = 0.001) 39 40 1 20 1 -20 ~ log(J) ~401 ‘50 i) 10 20 30 -80 7 -100 - supersaturation pressure (MPa) + Henry's law —x— Generalized Henry's law Figure 2.12: Effect of the departure from Henry’s law on nucleation rate (<15 2: 0.005) 2.5.3 The Applicability of the Ideal-Gas Law Having considered departures from ideal solution behavior in the previous section, we now consider the applicability of ideal-gas behavior at high pressures (e. g. > 20 MPa) is discussed in this section. Figure 2.13 (Van Wylen and Sonntag, 1976) shows the compressibility of nitrogen, where the compressibility factor Z is defined as _Pv z___ RT (2.65) Its deviation from unity indicates the departure of a gas from ideal behavior. It can seen from Figure 2.13 that e at relatively low pressures and relatively high temperatures, the compressibility factor approaches unity, i.e. nitrogen behaves as an ideal gas. For other ranges of relatively high pressures or relatively low temperatures, the behavior of the gas may deviate appreciably from ideal. 40 o the deviation of the compressibility factor from unity is less than 3% for nitrogen at a temperature of 300K and pressures of up to 10 MPa. 1.6 '- 1.4 F 1.2 l E F \ 1.0 —L _ / N 200 K / 0.3 .. Set “fated "Po:- 1 I l 0.4 ’-' I state I I 0.2 1- Saturated liqnid 0 ing—1 1 0.1 1.0 3r 2 4 P(MPa) Figure 2.13: Compressibility of nitrogen From the generalized compressibility chart (e.g. Gyftopoulos, 1991, p. 355), we can obtain the compressibility factors for any gas at various temperatures and pres- sures, provided the critical properties are known. Table 2.2 shows some relevant values of the compressibility factor obtained using the generalized compressibility chart. It can be seen from Table 2.2 and the generalized compressibility chart that at 300K, the compressibility factor is in the range of 0.95—1.05 for N2 at pressures up to about 18 MP3, for 02 up to about 35 MPa, and for Ar up to about 34 MPa. Therefore corrections for compressibility rarely amount to more than 5% for the supersatura- tion pressures considered in this study. A 5% deviation in predicting supersaturation nucleation thresholds is tolerable in this study. 41 Table 2.2: Calculation results of compressibility factors using the generalized com- pressibility chart Gas T, (K) PC(MPa) T(K) Tchr/Tc P(MPa) 103:13/10c 2:}; N2 126.2 3.39 300 2.38 18.6 5.5 1.05 02 154.6 5.05 300 1.94 35.4 7.0 1.05 Ar 150.8 4.87 300 1.99 34.1 7.0 1.05 42 CHAPTER 3 BUBBLE DYNAMICS The classical nucleation theory relates nucleation rates to the work required to form bubbles of critical size. For nucleation in dilute gas—liquid solutions, the diffusion of gas molecules from the host liquid to the bubble surface undoubtably affects mass transfer at bubble surfaces and so should play some role in nucleation, though it is typically disregarded in nucleation theories. Since an important characteristic of diffusion is the diffusion coefficient of a dissolved gas in a liquid, which takes different values for each different dissolved gas, it is useful to study the dynamics of gas bubbles in gas-liquid solutions to see the extent to which diffusion—coefficient values affect bubble growth rates, and to see if they might provide a way to extend classical nucleation theories to apply to different dissolved-gas species. 3.1 Statement of the Problem In order to understand effects of diffusion characteristics on nucleation, we consider an existing, spherical bubble in an infinite body of uniform liquid. The bubble may contain a single gas species or a mixture of several species which diffuse independently and differ in both solubility and diffusivity. Initially the concentration of gas in the liquid is assumed to be uniform. Since we are interested in how nucleation of gas- supersaturated liquids depends on the particular gas species, it is particularly useful to study the extent to which bubble growth and collapse rates depend on the gas-liquid diffusion coefficient. 3.2 Governing Equations In this section, we first present the equations which describe the growth and collapse of bubbles containing a single gas species. While the treatment of pure gas bubbles is 43 really a convenient first approximation to many practical situations, it also provides a basis for extending this analysis to bubbles with multiple gas species. In our single- species analysis, it is assumed that: 1. the system is isothermal; 2. the gas inside each bubble behaves ideally, and is well mixed; 3. the properties of the liquid are independent of dissolved gas concentration; 4. the center of the bubble does not move; 5. the pressure far away from the bubble is kept constant; 6. viscous and inertial effects are negligible. The governing equation for gas diffusion in the liquid is 60 ac _ D (620 2g —+"(")a - a? a» fit ) for TB]? (3.1) where C is the gas concentration in the liquid, r the radial coordinate, and R the radius of the bubble. The gas concentration C’ is assumed to be a function of both 7' and t. D is the diffusion coefficient, which depends on both the gas and liquid species involved and v(r) is the velocity of the liquid; it can be related to R, v and r from the continuity equation applied to the (incompressible) liquid surrounding the bubble, which can be written as 12(7") 2 I v(R) (3.2) where R(t) is bubble radius and v(R) is the outward velocity of bubble surface. The boundary conditions on concentration, for growth of a single bubble in a large liquid medium, are liIn C(r, t) : Coo; (3.3) r—nc 44 C(R,t) 2 0,, t > 0 (3.4) where C, is the saturation concentration at the bubble surface, that can be calculated from Henry’s Law as, C, : Pgas/H (3.5) Here, H is the thermodynamic property which scales gas pressure to dissolved liq- uid concentration, and is a constant for weak solutions. The subscript oo denotes prOperties at locations far from the bubble. The initial condition is C(r, 0) 2 C00, 1‘ > R; (3.6) The pressure distribution in the liquid surrounding the bubble, which can change on account of bubble growth, can be found by applying the momentum equation for incompressible Newtonian fluids in the radial direction. The equation, in a spherical coordinate system, can be arranged as 2 4 _8v3(t)£ 212%??? = —l%[: (3.7) r p r 8tr2 For diffusional processes, the radial velocity at the bubble surface ’01; is usually very small (of the order of 10‘6 m/s), so 6%) is also negligible and the pressure within the liquid scarcely changes between the bubble surface and the far field. The pressure inside the bubble can then be expressed as Pgas : P00 + 2o/R (3.8) The change of mass within the bubble satisfies Fick’s Law of mass transfer at the bub- ble surface, which, after substitution for the mass of gas from the ideal gas equation, 45 can be written as dm__d Poo+2o/R 4 3 _ 0C 2 dt _ dt iM< RgT ) (37TH )] ‘ Dar iRWTR’ (3'9) where m is the mass of gas inside the bubble, C has units of kg / m3, M is the molecular weight of gas inside the bubble and R9 is the universal gas constant. On simplification of Eq. (3.9), we obtain dR 8C Ei,’ : Dakar/M (3.10) where dR/dt can be interpreted as the bubble growth rate. 3.3 Analytical Solution For diffusion processes between a micro bubble and a liquid, v(r), which is usually of the order of 10‘6 m / s, can often be neglected. This assumption is reasonable, and has been verified by numerical solutions of the complete diffusion equation for bubbles of this size. If the convection term is neglected, the diffusion equation simplifies to BC 620 2 0C a — D (a: $5) ‘3'“) with initial condition C(7',0) 2 C00, 7‘ > R; (3.12) and boundary conditions r1390 C(r,t) 2 C00; (3.13) C(R,t) : C3, t > 0 (3.14) 46 where C, is the saturation concentration at the bubble surface, and can be calculated from Henry’s Law as as 20 C, = Pym/H = (P00 + E) /H (3.15) C, is assumed to be constant, which is reasonable if the bubble radius is not too small (above 1am) , so any surface tension effects can be neglected. Epstein and Plesset (Epstein et al. 1950) solved this equation by introducing a dependent variable, a = r(C — C3) (3.16) which allows the diffusion equation to be transformed to Bu 08221 with initial and boundary conditions, u(r,t = 0) 2 7'6, 1“ > R; u(R,t) = 0, (3.18) where 6 : C00 — C3. If one makes the second variable change, 5 : r — R (3.19) and also assumes the bubble boundary moves sufficiently slowly that R is effectively a constant, Eq. (3.17) becomes analogous to a familiar problem in heat conduction. 0'11. 82 ”a 47 The analytical solution of Eq. (3.20) can be found using the separation of variables method, in the following form, (Carslaw, 1945) u(r,t)= 2W/Ooo (-R+€) ){exp [—gzl—Bi’li — exp [—(é—IfiiJ—]}d€, (3.21) Here the quantity of interest is the concentration gradient at r = R. (g?) (R : 6{1+ R/(ert)’/2} (3-22) (96%) [R = 6{1/R + 1/(7rDt)’/2} (3.23) From considerations of the gas diffusion through the bubble surface, the mass flow rate is dm _ 2 (9C d7 —47r RD(—8—T)(R (3.24) Combining Eq. (3.23) and Eq. (3.24), the velocity of the bubble surface can be ex- pressedas (1R Dd l — —— .25 d—t p —{R+ («Dal/'2} (3 ) where p is the density of the gas. It is convenient to define f as the ratio of initial dissolved gas concentration to the gas concentration at saturation, and define d as the ratio of the saturated gas concentration to the gas density, so f : Coo/Cs (3.26) d : C's/p (3.27) To solve for R(t), two cases are considered separately: undersaturation (0 g f < 1) and supersaturation (f > 1)- 48 3.3.1 Bubble Dissolution in an Undersaturated Solution In order to solve (3.25), it is convenient to rewrite Eq. (3.25) in dimensionless form. Setting 6 = R/Ro, (3.28) x2 z (2a/Rg)t, (3.29) a = D(Cs—Coo)/p : Dd(l-—f), (3.30) Eq. (3.25) becomes de/dzi: = —a:/6 — 2) (3.31) where 7 : (gfiglffl. In general, the constant 7 is small compared to the other term 13/ 6 so that an approximate solution can be obtained by neglecting 7 which takes the form 62 = l — 2:2 (3.32) or, when recast in dimensional form, (R)2:1_ZD(C,—Coo) Ea p123 t (3.33) Figure (3.1) shows this radius-time relation for dissolving bubbles. The time required for a bubble to dissolve completely can then be found from Eq. (3.33) by setting R to zero, as , _ 933 tdzssolve — 213(C;3 _ Coo) (3.34) The time taken to dissolve is proportional to the initial area and density, and inversely proportional to the diffusion coefficient and concentration difference. Table 3.1 shows 49 (RIR0)2 \/ t Figure 3.1: Radius—time relation for dissolving bubbles the time required for bubbles of different gases to dissolve in water. The results are calculated based on Eq. (3.34). The gas properties are taken at a temperature of 20°C and the six kinds of gases studied are : air, hydrogen, nitrogen, oxygen, helium and argon. It can be seen from Table 3.1 that 1. Among the six gases, it takes the least time for an H; bubble to dissolve, and that an N2 bubble is the most difficult to dissolve. 2. Argon and oxygen have similar bubble dissolution rates. Since argon is much more inert than oxygen to chemical reaction, argon is recommended as a good substitute for oxygen in some experiments, in which effects of oxidation are troublesome. 3. It takes hours for a bubble of 1 mm size to be completely dissolved in water. Therefore it seems reasonable to wait hours to reach a saturated gas-liquid solution with the help of stirring, during the preparation of supersaturated solutions in the experiments. F inkelstein suggested 24 hours (F inkelstein et al., 1985) as the waiting time for all the dissolved gas species they prepared. 50 Table 3.1: Time required for a pure gas bubble to dissolve in water Gas T(sec.) T(sec.) T(hours) T(hours) for R0 2 1pm for R0 : 25am for R0 2 1mm for R0 = 1cm Air 0.011 6.64 2.95 295 H2 0.0057 3.53 1.57 157 He 0.0080 5.02 2.23 223 N2 0.013 7.88 3.50 350 Ar 0.0067 4.17 1.86 186 02 0.0067 4.17 1.86 186 3.3.2 Bubble Growth in a Supersaturated Solution For a supersaturated solution, the positive constant a in (3.30) is redefined as a : D(C00 — Cs)/p = Dd(f — 1). (3.35) The constant 7 is also redefined, as Coo _ Cs 1/2 7' Z (77) (3'36) The dimensionless variables 6 and :17 keep the same form and the approximate solution can be obtained by neglecting 7' as 62 : 1 + 2:2. (3.37) In dimensional form, the solution is recast as (3) =1+ (”€2.90”) . (3.... After long times, R‘2 is proportional to t, D, AC, and inversely proportional to p. 51 (R/R0)2 / . W t Figure 3.2: Radius-time relation for growing bubbles The bubble growth rate is readily derived from Eq. (3.38) as (112 _ 0(000 - C.) ”(z—t _ pH (3.39) If diffusion is the limiting factor, the bubble growth rate is proportional to the con- centration difference (AC), diffusion coefficient (D), and inversely proportional to the size of the bubble and the density of the gas. Figure (3.2) shows the radius—time relation for growing bubbles. It is useful to restate the assumptions made to reach these quasi-stationary ana- lytical solutions for bubble growth and dissolution rates. 1. In forming analytical solutions, it is assumed that the boundary of the bubble is quasi-stationary. The gas diffusion equation is not identical to the heat can- duction equation when the bubble boundary is not stationary. However, since the bubble boundary moves so slowly (of the order of 10’6 m/s) at typical dif- fusion rates, it is reasonable to assume that 9% z 0. This assumption has been verified by numerical results which are discussed later. Also from Eq. (3.25), it appears that this assumption is more accurate for larger bubbles. So the ana- lytical solution is expected to be more accurate for bubble growth than bubble dissolution. 52 2. It is assumed that the convection term in the diffusion equation can be ne— glected. Since ‘17?- z 10“6 s’1 and 6(1) = d7???- < 92% for r > R, the convection term v(r)%rcZ is usually negligibly small, except in the case of extremely large concentration gradients. 3. It is assumed that the gas concentration at the bubble surface is constant and that any effect of surface tension on the saturation concentration there is ignored. This assumption is thought to be reasonable if the bubble radius R > 10‘6m. 4. The final assumption is that the pressure is constant across the liquid field. From the radial momentum equation, it can be estimated that ~66; z —p-§§§; m 0 and that deviations from a constant liquid pressure field are negligibly small. 3.4 Numerical Approach The analytical solutions in the previous section are based on many assumptions and simplifications. In order to assess the accuracy of them, the finite difference method is adopted here to solve the complete gas diffusion equation: — + ”(Ti '5? m: 8C BC _ D 02C 20C 0t 87‘ — > for rZR (3.40) where v(r) can be found from the (incompressible) continuity equation for the sur— rounding liquid, t 2 ‘1)(7') : 127(2) ’U(R). (3.41) with boundary conditions: din; C(r,t) 2 C00; (3.42) C(R,t) = 0,, t> O (3.43) 53 Here C3 is the saturation concentration at the bubble surface, which can be calculated from Henry’s Law as 2 C, = Pgas/H = (P00 + 80) /H (3.44) The initial condition is C(7‘,0) : Coo, r > R; (3.45) Initially there is a step change in gas concentration from C3 to C00 at the bubble surface. To overcome the numerical difficulty caused by the discontinuity, the step function is approximated by a decaying exponential term C1730): Coo + (C(R0,0) - Coo) exPl_Kd(T - 130)] 7" 2 R0 (3-46) where the value of K d can be adjusted to closely approximate a step function and still preserve numerical stability. Due to the small magnitudes of D (10’9 m2/s), C and 0(7), it is better to solve these equations in dimensionless form to reduce round-off error and improve accuracy in numerical calculations. 3.4.1 Dimensionless Form The dimensionless variables of this problem are chosen as: ‘r' : r/RO D ,’: —, .47 t tiff, (3 ) C, : C/CSQ R’ = R/RO 54 where 030 is the saturation gas concentration under pressure Poo. The dimensionless form of the diffusion equation is then ac" + (d_R’ R”) 60' _ 620' 2 60' _ ___ __ ,> , . 0t' 44 w? a" are 9 aw for 7‘ — R (348) and the dimensionless form of the velocity of bubble boundary is 412' [33% 3,1297") ——,- z 40 (3.49) dt [POo + 3133.. 3.4.2 Discretization For uniform grids, the diffusion equation is discretized into the following form: 07’“ — C?’ R2 073,, — 0n 1 C." + C."- — 203 v — C." i z _ _ 1. z— : D 2+1 2—1 2 1+1 2—1 . At + MR) 7",?" 2Ar { (Ar)2 + riAr (3 50) 620 0153 For non-uniform grids, the second derivative is approximated as (Ferziger, 1996) (gr—g)” : i110? — ri_1)+ 031(7):“ " Ti) — C}’(ri+1 — 7””) (3.51) 1 0-5 (“+1 — 7"z'--1)(7“i+1 - Ti)(7‘i - “Pi—1) For non-uniform grids, the diffusion equation is discretized into the following form, CY‘H—Cn R20." —C." l 1 lil x—l tn+l_tn + U(R)7?- ri+1—ri-l __ D (73,1(Ta—TH1)+C,”_1(Ti+i—Ti)—CI’(7‘1+1—"i—1) ZCI’Ztl—Cf’.) (3 52) — 0.5 (7‘1+1—7'i—1)(1‘i+1—7’1)(7‘i—7‘i—1) 7‘1‘("'i+l-Ti—1) ' 3.4.3 Stability Criterion In order to simplify considerations of stability criteria, we consider only the stability of the equations when discretized on a uniform grid, and disregard the convection 55 term. Rearranging Eq. (3.50), we find that 33-33 (Ag—,3.) (__Dm M).3_.(_D A De) (Ar)? riAr (Ar)2 rAr (3.53) For numerical stability, the following three terms must be positive: D 1— ———At > 0 (3.54) 2W )2 DAt At D— 3. (Ar)2 + riAr >0 ( 55’ D At ——At — D— 0 3.56 (Ar)2 rAr > ( ) The stability criterion is therefore M > ZD (3 57) At ' ' or, in its corresponding dimensionless form (4702 2. ." At’ > (3 08) 3.4.4 Moving Grids One difficulty which arises in solving the gas diffusion equation numerically is the problem of how to deal with the moving bubble boundary as the bubble grows. In the approach chosen here, moving grids are used. For the case of bubble dissolution, one may consider adding one grid point for each time step as the bubble surface recedes. But owing to the slow movement of the bubble boundary and the small time step, it will lead to very non-uniform grids and the small grid ratio will cause instability. To overcome this difficulty, the following approach is adopted: initially, the grid points are uniformly spaced with a spacing of Ar. At each time step, the 56 r0 r1 r2 r3 [___ _ I ,__ I l t>0 l'o f1 f2 I3 lnse a new point here Figure 3.3: Illustration of moving grids during bubble dissolution first point is moved to the new bubble boundary. As the bubble shrinks, the space between the first and second points becomes larger and larger. Once it reaches 1.5A7‘, the first grid is split into two grids with equal spacing. A similar approach is adopted for bubble growth, as the spacing between the first two grid points becomes smaller and smaller. Once it reaches 0.5Ar, the first grid point is merged with the second grid point. 3.4 . 5 Numerical Results A typical spacing of grid points of Ar = 0.5pm and a time step size of At = 10‘5 sec- onds were used to find numerical solutions to the discretized bubble growth / dissolution equations. Figure (3.4) shows the radius change with time during a microbubble dis- solution process. It can be seen that 1. Initially, the rate of bubble shrinkage % is very large, owing to the large gas concentration gradient across the bubble surface at initiation of the calculation. 2. The rate of bubble shrinkage is also very large when the bubble is very small. The reason for this is that, for a small bubble, the pressure inside the bubble is very large on account of the surface tension. High pressure leads to a high gas concentration at the bubble surface. The large concentration gradient drives the bubble to dissolve faster. 57 1.2 - 0.8 r 0.2 “ t. Figure 3.4: Dimensionless radius vs. time curve for bubble dissolution due to diffusion (numerical result) 3. Except for the initial period, the rate of shrinkage increases as the bubble dis- solves. From Figure (3.5), it can be seen clearly that the relation between R’2 and t’ is approximately linear, in agreement with the approximate analytical result. Figure (3.6) shows gas concentration profiles in the liquid around the bubble. As time passes, more gas diffuses into the liquid. Figure (3.7) shows the change of R’2 with t’ for growing bubbles. It can be seen that the relation between R’ 2 and t’ is almost linear, which again agrees with the analytical result. In the early stage of bubble growth, the growth rate is controlled by inertia and surface tension. But these factors become unimportant as the bubble develops. Then diffusion dominates the growth of the bubble. R oc \/t is a good approximation for bubble growth, if the initial size of the bubble (R0) is small enough to be neglected. 58 1.2 _ 0.8 ~ m 0.6 4 0.4 ~ 0.2 4 Figure 3.5: Dimensionless radius squared vs. time curve for dissolving bubbles due to diffusion (numerical result) 0 2 4 6 8 10 ___—t'zo -——t'=10 ------ t’=13.5 Figure 3.6: Concentration distribution around a dissolving bubble as a function of time 59 0 4 " '” 1 I fl 0 2000 4000 6000 t0 Figure 3.7: Dimensionless radius-squared vs. time curve for growing bubbles due to diffusion (numerical result) 3.4.6 Comparison of Analytical and Numerical Results Figure (3.8) shows the analytical and numerical results for radius as a function of time for a dissolving bubble. It can be seen that the effects of the moving boundary and of surface tension offset each other. The analytical result agrees very well with the numerical result which accounts for the moving boundary, the convection term, and surface tension effects, all of which are neglected in the analytical solution. 60 RI 1 i 0.8 ~ 0.6 ~ 0.4 ~ 0.2 — o . T I 0 5 10 15 t’ —— analytical solution numerical solution Figure 3.8: Radius vs. time curves for a dissolving bubble 61 CHAPTER 4 TARGET DATA FOR TESTING THEORIES OF NUCLEATION IN GAS-SUPERSATURATED SOLUTIONS In this chapter, some experiments on nucleation of gas-supersaturated solutions are discussed. These experimental results will be used as target data against which the more general models of nucleation, developed as a part of this study, are tested in the following chapter. The significance of these results to predictive theories of nucleation is also discussed, together with theoretical interpretations of these results and prob- lems associated with their interpretations are also pointed out. The experiments of interest are those which provide target data for cases in which classical nucleation the- ory is unreliable, and which provide a challenge for the more general model developed in this study. They are experiments involving nucleation of different dissolved-gas species under otherwise identical conditions, nucleation in volumes the size of which has been systematically reduced, and nucleation in large systems in which surface geometry is considered important. 4.1 Nucleation of Supersaturated C02—Water Solutions in Large-Scale Systems This section focuses on heterogeneous nucleation of gas-liquid solutions in large-scale systems, where surface geometries play a role in the nucleation processes. Wilt (1985) studied mechanisms of nucleation in supersaturated COg—water solutions, such as carbonated beverages. He examined several possible mechanisms: homogeneous nu- cleation, and heterogeneous nucleation at smooth planar interfaces, and at interfaces containing conical or spherical cavities or projections. He concluded that the predom- inant mode of nucleation was heterogeneous and that it occurred in conical cavities (Figure 2.5, page 31), presumed to be present in some form on almost all surfaces, 62 but only for contact angles (6) in the range of 94 — 130°, over which shape factors were computed to be small. It is important to note that, in these studies, no special precautions were taken to pre—compress or pre—heat either the solutions or their con- tainer, so there were probably many active nucleation sites, already partly filled with gas, present at surfaces. Consequently, one would expect nucleation to be observed at pressures very close to the dissolved-gas saturation pressure. A unique aspect of this study was that Wilt did not approximate the work done to grow bubbles by their spherical equivalent, but calculated the mechanical work done by bubbles, initially in conical cavities, which subsequently grew first within the cavity and, second, beyond it in a shape determined by local surface-contact conditions. His study would be equally applicable to growth from cavities on container surfaces as it would to growth from cavities within impurities suspended in the liquid. Based 011 his nucleation theory analyses, in which effects of bubble shape on the work of bubble expansion were considered, Wilt proposed a model for heterogeneous nucleation as a variation on classical nucleation theory which took the form: .47“: an M} (4.1) J = N2/3f30(6’fl)\/7rmf?((0 [5) exp { 3kT where N is the number of molecules per unit volume for homogeneous nucleation, or N 2/ 3 is the number per surface area for heterogeneous nucleation, o is the surface tension, m is the mass of one molecule, 9 is the liquid—solid contact angle of the bubble with the cavity surface, 6 is the half apex angle of the cavity, and fa and f 3c are functions of 6 and 6 related to the mechanical work done in growing bubbles within and beyond surface cavities, which take the form: f1c(9, [3) : {2 — 2sin(0 — 6) + c030 c082(6 — fi)/sinfi}/4 f3c(6,/3) : l1” sin(I9 — fill/2 63 When interpreting thresholds for nucleation, it is useful to discuss them in terms of the ‘supersaturation ratio,’ w, defined as the ratio of dissolved gas concentration in the liquid to the saturation concentration across a flat surface at the same liquid pressure, minus one (so that w = 0 for saturated solutions), i.e. C — 1 Csat w: Based on the observation that a COrwater solution nucleates at a supersatura- tion ratio of about 5 at atmospheric pressure and room temperature, Wilt proposed 6 = 94°,B = 4.7° as one pair of values in the above equation which enabled J to increase substantially as the supersaturation ratio exceeded 5, and also seemed to provide a plausible description of surface features in the relevant experiments. Wilt also predicted that homogeneous nucleation and heterogeneous nucleation at either a smooth planar surface, or a surface with conical or spherical projections rather than conical cavities, will not occur at a supersaturation ratio of about 5. 4.2 Experiments on Formation of Gas-Vapor Bubbles in Supersaturated Solutions of Cases in Water Finkelstein and Tamir conducted an extensive series of experiments on bubble forma— tion in water supersaturated with a gas, for several different gas species (Finkelstein & Tamir, 1985). In this study the dissolved-gas species and the initial supersaturation pressure were systematically varied while other experimental parameters were un- changed. The experimental data from this study therefore provide a good test of the predictive capabilities of nucleation models intended to apply to different dissolved gases. 64 4.2.1 Experimental Setup and Results The experimental apparatus used by Finkelstein & Tamir is shown in Figure 4.1. Their experiments were carried out as follows. A supersaturated solution of gas in water was first prepared in a small glass beaker at a high pressure Pssat, after which the pressure of the solution was slowly and steadily reduced. During this period of pressure reduction, the solution was observed carefully for visual signs of bubble for- mation and nucleation. At the moment at which bubbles were first observed, the pressure was recorded, and noted as PI. The difference between the initial super- saturation pressure Pm“ and the pressure H, was called the pressure difference for bubble formation and was designated as AP”. After carrying out these experiments using nitrogen, helium, neon and argon at different initial saturation pressures, the following results were reported: 1. For different gases, the observed values of AP" are very different. Experimen- tally measured values of APn for different gases in water are shown in Table 4.1. The following correlation was found to describe the behavior of AP” for different gases, APR/PC = 1.3(T/TC)” (4.2) where subscript c denotes critical properties of the gas (although all the exper- iments were conducted at the same temperature!) 2. For a given gas, the pressure difference for bubble formation (APn) appears to be independent of the initial supersaturation pressure (Pssat). 65 1. Wgucylhda 2. WW(WW4) 3. slamuoolplpo 4. Gas llllor, 0.15 x 10"m 5. Addlllonal tiller, 5 X 10"m 0. Normal comm lot 138 IlN/m’ (AIIINCO) 7. Mn rolom valve 8. m gauge, 34.5 DIN/n13 0. We gauge, 201 mm»2 10. Promo col 11. We windows 12. Hanna or cooling locket 13. We 14. Glen balm Figure 4.1: The experimental apparatus of F inkelstein & Tamir 66 Table 4.1: Experimental data for the pressure difference for bubble formation Gas APn (MPa) at T =303K Helium 42.0 :t 1.4 Neon 32.4 d: 1.4 Nitrogen 13.8 d: 1.4 Argon 13.1 :l: 1.4 4.2.2 Problems with F inkelstein and Tamir’s Equation on Nucleation of Liquids Supersaturated with Gases It is important to point out a limitation of the Finkelstein & Tamir correlation —— the equation is a best fit of experimental data for experiments conducted at a single tem- perature, in which the variable T/TC is included to introduce a supposed dependence on the critical temperature (TC) of the gas, and is not necessarily applicable at arbi- trary temperatures at all. For a specific gas, as temperature increases, AP" should decrease, because nucleation always takes place more readily at higher temperature according to the classical nucleation theory, e. 9. Eq. (2.53). The incorrect inclusion of T in Eq. (4.2) leads to an intuitively incorrect result, even though it happens to fit the data at the single temperature at which they were taken. Therefore Eq. (4.2) is not as general as it might appear and seems to be misleading. 4.2.3 Inconsistencies between F inkelstein and Tamir’s Results and Classical Nucleation Theory One of the most significant findings of F inkelstein & Tamir’s experimental study was that, for solutions of different gases dissolved in water, the bubble nucleation characteristics are very different. The pressure differences for bubble formation (APn) range from 13.1 MPa for argon to 42.0 MPa for helium. Bowers’ (1996) experimental data also suggested a strong dependence of supersaturation thresholds on dissolved gas species. However, when a nucleation rate equation of the kind derived from 67 classical nucleation theory is applied to gases like argon and helium, it is found that the predicted value of AP" which corresponds to nucleation occuring (J increasing rapidly with small increases in AP”) is almost the same for all gases. It varies by less than a few percent, whereas the experimental measurements of APn can vary, from one gas to another, by an order of magnitude. Thus nucleation—theory results, with a weak dependence of AP” on the dissolved gas species, are contradictory to Finkelstein & Tamir’s experimental results. Before concluding the discussion of these target data, it is important to con- sider whether the observed discrepancy with nucleation theory is a consequence of an inadequate distinction between dissolved—gas species in the theory, or whether it might arise through high-pressure/high-concentration effects such as departures from Henry’s Law or changes in critical radii on account of the pressure of the dissolved gas. From classical nucleation theory, the heterogeneous nucleation rate can be expressed as (Eq. (2.53), page 30) . / 3o -47ro q$(6’ fi)r2 2/3 ’ e J — N —’ exp{ 3kT (4.3) where the shape factor, (Z), may take the form given by Wilt (page 63) for conical cavities, ¢3(6,[3) : {2 — 23in(0 — B) + c056 6082(6 — fi)/sinfi}/4 and the other symbols take the same meanings given previously. The equilibrium radius of bubble formation (in a weak gas-liquid solution can be calculated from Ward’s (1970) equation (Eq. (2.23), page 19), PC' re : 2o/{nPSm(T1) + C — H} (4.4) /0 68 where the symbols also take their previous meanings. The coefficient 7) is UK}?! - Psat) "’ RTC, ”29“) RT and the concentration C’ is, in this case, the saturation concentration under the supersaturation pressure Pssat. The pressure difference for bubble formation APn is therefore Pssat — Pl. Using Henry’s Law, Eq. (4.4) is simplified to re : 2o/ {nPsat(T1) + A3,} (4.5) and, since 77 z 1 and Psat(T1) << APn ( Psat=2338Pa at 20°C), the above equation is simplified further by neglecting 77pm to yield: am If we consider that Henry’s law may be inaccurate at high concentrations, a modified expression of Henry’s law can be used. CzWHW) 80 where H is a function of P, instead of a constant. More discussions of the applicability of Henry’s law can be found in Section 2.5. Thus the critical radius can be more accurately expressed as 20 re z HEP} (4.8) H(Psslat)APn For reasons of simplicity, we still assume Henry’s law is applicable in the following P Hszl "mama, when derivations, but keep in mind that APn should be replaced by A Henry’s law is no longer accurate. After substituting Eq. (4.6) into Eq. (4.3), the 69 nucleation rate for a gas-liquid solution is expressed as . 3o —167ro3 (15(9 [3) : 2/3 _ , J N exp { 3kTAP3 nm For different gas species dissolved in water, the aqueous gas solution surface ten- sion 0 is almost constant for all gases (Finkelstein & Tamir, 1985). The experimental results of Massoudi and King (1974) showed that the surface tension changes less than 8% for He, H2, 02, N2 or Ar dissolved in water, when the pressure changes from 1 atm to 80 atms. Although the supersaturation pressures Pam in Finkelstein & Tamir’s experiment are very high (13.1—42.0 MPa), the liquid pressure B at which nucleation occurs is much lower. Therefore it is reasonable to assume that the surface tension essentially depends only on the properties of the solvent for dilute solutions at H. Since the experiments were carried out in the same container, surface conditions such as the contact angle 0 and the half apex angle of conical cavities on the container surface [3 should not change much when different gases were tested. Therefore, the calculation results for values of AP" from Eq. (4.9) at which J begins to increase rapidly for all the gases are expected to be almost the same. This result is in contra- diction to the experimental result shown in Table 4.1, and indicates that the classical nucleation rate equation fails to account for effects of different gas species. In the next chapter, we will modify the classical nucleation theory to interpret Finkelstein & Tamir’s data, by considering the diffusion of dissolved gas molecules towards the bubble surface. Now, from Eq. (4.9), we can see that for a given gas, it is APn, not the initial supersaturation pressure (Pam), that affects the nucleation rate. The pres- sure difference for bubble formation APn is independent of the initial supersaturation pressure (Pam). This conclusion agrees with Finkelstein & Tamir’s experimental re- sult, and so some other feature of the nucleation—rate equation should account for the dependence of gas species on nucleation rate. 70 Based on these analyses, it follows that F inkelstein & Tamir’s experimental results provide useful target data for refined nucleation models. In particular, these data provide a good test for prediction of AP" for different gas species. 4.3 Experiments on Nucleation of Gas-Supersaturated Solutions in Small Capillary Tubes To gain a better understanding of the mechanism of nucleation in small capillary tubes, in which it has been observed that nucleation thresholds can be raised con- siderably, a series of experiments which explored nucleation thresholds as functions of dissolved-gas concentration and capillary-tube diameter were conducted by Spears and co—workers (Brereton et al., 1998). 4.3.] Experimental Setup and Procedures The setup used in capillary tube nucleation experiments is shown in Figure 4.2. In these experiments, water, in which gas was dissolved at an initial supersaturation pres- sure Pssat, (ranging from 3 to 15 MPa), was pre—compressed to even higher pressures (Pmp, a: 100 MPa). The gas-supersaturated water solution was then decompressed during flow along a silica capillary tube, from its pre—compression pressure at one end of the tube to almost atmospheric pressure at discharge at the tube’s far end. The diameters of the capillary tubes in these experiments varied from 10 to 100 um. The tube lengths were chosen to be proportional to the square of their diameters so that the cavitation number Ca 2 AP/ (pV2 / 2) remained constant from one experiment to another. Also, the maximum Reynolds numbers in any experiment were also kept be- low 2300, so that laminar flow always prevailed. These experiments were carried out with the outer surface of the capillary tube either exposed to surroundings at room temperature, or, if effects of temperature were to be studied, immersed in a water 71 Pure Water Waste Flushin Circuit «— —-—. i Discharge of Ca illa Gas-enriched . p ry Water __ E fifrr—Efl Pressure Vessel _ _ Gas-enriched H _ Water Gas Cylinder Gas lntensifier Hydraulic Cylinder Figure 4.2: Schematic illustration of the experimental apparatus used for oxygen- supersaturation studies 72 bath. In order to detect the presence of micro bubbles, fluorescein was added to the liquid prior to gas supersaturation and pre-compression. The effluent from the capil- lary tube was then illuminated with an argon-ion laser, which induced yellow-green fluorescence in bubble-free liquid and blue reflections from rising bubbles. 4.3.2 Experimental Results for Nucleation of Gas-Supersaturated Solutions in Capillary Tubes A series of experiments (Brereton et al., 1998) on oxygen-supersaturated water were carried out following the approach described above. In these experiments, the gas supersaturation pressure (or concentration) was varied systematically to determine its lowest value at which nucleation was observed in water discharged from capillary tubes of different diameters, as determined by detection of bubbles in the effluent. These experiments showed that pre—compression plays a important role in rais- ing the nucleation thresholds. Without pre-compression, nucleation was observed for supersaturation at nearly all supersaturation pressures considered. However, when the oxygen-supersaturated solution was pre—compressed to approximately 100 MPa, solutions of oxygen at relatively high concentrations could be decompressed to atmo— spheric pressure within capillary tubes, without any evidence of bubble formation. This phenomenon can be explained in the following way. It is already known that there is very likely to be some amount of gas trapped in surface cavities and that the entrapped gas can provide nuclei for the formation of gas-vapor bubbles at the onset of nucleation, thereby facilitating the initial nucleation process. At high pre- compression pressures, the solubility of gas in the liquid is also higher according to Henry’s Law. Therefore pre—compression facilitates the dissolution of trapped gas bubbles and drives liquid (rather than gas) into any cavities present at surfaces or on contaminant particles. Consequently it is more difficult to initiate nucleation as the pre—compression pressure is increased. It is also thought that pre—compression can 73 drive contaminants in the liquid into surface cavities, and eliminate some negative effects of contaminants on nucleation. Although this mechanism provides an expla— nation of this phenomenon, it has not been proved that this is what happens at the microscale. 18a 16‘ -14“ “12‘ $10— 58" n. 5* 4- 24 0 v T l T fl 0 20 40 60 80 100 d (u m) Figure 4.3: Dependence of supersaturation pressure on capillary tube diameter. The line and data points are an algebraic fit to many experimental data, measured at 22°C (from Brereton et al., 1998) Figure 4.3 shows experimental results for the maximum achievable value of P3801, (or, equivalently, dissolved-gas concentration) for which no nucleation is observed, or the nucleation threshold, as a function of tube diameter d. In the area above the data line as shown in the graph, nucleation is always observed. These experiments were all carried out in quartz glass capillaries at a delivery pressure of 100 MPa at 22°C. It is clear from the figure that smaller capillary tube diameters correspond to higher nucleation thresholds of supersaturation pressure or dissolved gas concentration, so that this elevation of the nucleation threshold appears to be a confinement effect. 74 4.3.3 Interpretations of Experimental Data Describing Nucleation in Capillary Tubes Using the results of these experiments on nucleation of gas-liquid solutions in small capillary tubes by Spears and co-workers, Brereton et al. (1998) developed a dual dependence model to interpret the experimental data. Inside capillary tubes, the preferred mode of nucleation is heterogeneous, due to the existence of surface cavities (which trap gas bubbles, and make the ‘energy barrier’ for bubble nucleation smaller). A cavity is ‘active’ when its radius exceeds the critical radius for bubble formation. Experimental data suggest that the active number of cavity sites per unit area Nactive (m’2) may be modeled as TCG’U q] Native 2 const ( ' ) (4.10) Te where raw is the average size of conical cavities at surfaces, and re is the critical radius for bubble formation. q] takes the approximate value of 0.16 for boiling of water at superheated copper surfaces (Lorenz et al., 1974.) Brereton et al. developed a dual dependence model, which predicts that nucle- ation rate in capillary tubes depends on sufficient numbers of both molecules of the vaporizing species and of active cavities, and is expressed as: 30 —47ro r2¢> J : .tANa .,.,eNs,,,.‘/—— ——£— 4.11 cons C. 1 7m eXp{ 3kT } ( l where N13,". is the number of surface molecules of the dissolved gas species per unit area, (i) is a shape factor for the bubble, varying between 1 for spheres and 0 for ‘pancakes’, and J has unit of s‘“lin‘2. Although Eq. (4.11) could be calibrated to agree with experimental data, some of the assumptions in the original paper (Brereton et al., 1998) required more careful 75 consideration, e.g. the assumption that the changes in the radii of curvature of nuclei in crevices are neglected during decompression is questionable. Also more detailed calculations of the critical radii of nucleation in gas-liquid solutions could have been made. More importantly, the effect of dissolved gas species on nucleation thresholds was not considered. The target data of Brereton et al. (1998) therefore provide a test for the capa— bilities of nucleation models in predicting effects of varying surface area (and also varying surface area to volume ratio), effects of dissolved-gas concentration, and ef- fects of precompression pressures on nucleation thresholds. In the next chapter, a nucleation model with several refinements intended to allow prediction of these ef- fects will be developed and tested. 4.4 Some Shortcomings of the Classical Nucleation Theory In classical nucleation theories, one typically focuses on clusters of nuclei in a liquid and considers when they are sufficiently numerous and energetic to combine to create micro-bubbles which are large enough to avoid collapse, and grow to a finite size. Some particular problems with these classical theories, which limit their usefulness in practical applications, are: 1. they typically associate the conditions required for bubble formation only with the reversible work required to create bubble surfaces of stable size, without regard for mass transfer to the bubble surface and other irreversible effects, e.g. diffusive dissipation, viscous dissipation, etc. It is possible that bubble forma- tion is also strongly dependent on the gas species involved, possibly through (species dependent) diffusive effects adjacent to the bubble surface, or through other species dependent effects; 76 2. a nucleation model calibrated for one dissolved-gas species can be highly inac- curate if applied to a different gas-supersaturated solution and the reason for this shortcoming will be explored in Chapter 5; 3. they do not predict any effect of volumetric confinement on nucleation thresh- olds of gas-supersaturated liquid, when experiments indicate there is one. Modifications to the classical nucleation theory, which account for these shortcom- ings and are tested against the target data described in this chapter, are presented in Chapter 5. 77 CHAPTER 5 MODIFICATIONS TO THE CLASSICAL NUCLEATION THEORY Several problems associated with the classical nucleation theory were pointed out in the previous chapter, together with target data sets with which an improved theory should provide better agreement. In this chapter, modifications to the classical theory of nucleation are described and tested. They take into account effects of diffusion of gas molecules, volumetric confinement, and departures from ideal gas and solution behavior. 5.1 Prediction of Nucleation Thresholds of Liquids Supersaturated with Different Gases In Section 4.2, F inkelstein and Tamir’s experiments on bubble formation in water supersaturated with a number of different gases were described and discrepancies be- tween the experimental results and predictions from the classical theory of nucleation were pointed out. The classical nucleation theory relates nucleation rate only with work required to create bubble surfaces of stable size without regard for mass transfer to the bubble surface and other surface effects, and this is possibly the reason it fails to predict accurately the nucleation thresholds of liquid supersaturated with different gases. From the classical nucleation theory, the predicted value of AP", an expression for the degree of supersaturation which a solution can tolerate, is almost the same for all the gases. This nucleation-theory result, with a weak dependence of AB, on the dissolved gas species, is in contradiction to F inkelstein & Tamir’s experimental data of (APn) ranging from 13.1 MPa for argon to 42.0 MPa for helium. In this section, the classical nucleation rate equation is modified to account for species-dependent dif— fusive effects in a way that allows it to reproduce Finkelstein & Tamir’s experimental 78 results for the dependence of AB, on species. Becker (1938) appears to be the first to consider effects of diffusion on nucleation and proposed the following type of expression for the rate of nucleation in a condensed system, e.g. a liquid-solid phase transformation, .12 K exp{—————Wrw +q} M, (5.1) where Wm, is the maximum reversible free energy necessary for nucleus formation, q is the energy of activation for diffusion across the phase boundary, K is a constant that accounts for the free collisions between molecules, and k is Boltzmann’s constant. Blander and Katz (1971) measured the nucleation temperatures for n-pentane and hexadecane mixtures, in which the solute, n-pentane, is relatively volatile. They proposed a modification to the nucleation rate equation to account for effects of diffusion as JDZJ/(1+(SD) (5.2) _ 20 — D\/27rka(C — C3...) 5:) where J is the nucleation rate for pure solvent at the same pressure and temperature, J D is the nucleation rate for the solution with the volatile solute, (SD accounts for the diffusion of the volatile solute to the interface, m is the mass of one molecule of the solute, C is the initial concentration of the volatile solute, and Csat is the equilibrium concentration at a vapor pressure equal to the ambient pressure. While this pre- exponential modification appeared to reconcile some of their observations with the theory, a calculation of nucleation-rate dependence on AP" using Eq. (5.2) shows that it fails to explain the large difference of AP", for different gases dissolved in water. As discussed earlier, the nucleation rate, J or JD, is insensitive to pre—exponential factors. Similar values of AR, would be obtained, even if (SD were varied by several 79 orders of magnitude. So this pre—exponential modification for diffusion effects does not seem adequate. The classical nucleation theory relates the nucleation rate to the reversible work required to create a bubble of critical size, without regard for any irreversible effects due to diffusion or internal friction. We suspect that the dissipation of energy due to viscous and diffusive effects may not be negligible during nucleation process, therefore, the equations for these dissipative terms will be derived and their order of magnitude will be estimated. 5.1.1 Dissipation of Energy through Difiusion and Viscous Efiects In this subsection, the formulas of dissipation of energy through diffusion and viscous effects during bubble growth will be derived. Their significance to nucleation of gas- supersaturated liquids will be discussed afterward. During processes such as bubble growth, the entropy of a fluid system increases as a result of the irreversible process of thermal conduction, internal friction and diffusion. The rate of increase of entropy of a binary solution in a finite volume can be expressed as (Landau and Lifshitz, 1987) - 2 % fps dV = f ——’”(gradT) dV T2 77 av,- ka 2 '01), 2 fg , 2 + fZT (317k + 393i 3611:8321) dV+ T(dZ’UV) dV -2 1 + ffidv (5.3) where s is the entropy, v is the velocity, K is the thermal conductivity, 17 is the dynamic viscosity, C is the second viscosity coefficient, a = [Dp/(Bu/aC)p,T], u is the chemical potential of the mixture fluid, D is the diffusivity, C is the concentration defined by mass fraction of the solute, p is the density of the liquid, and i is the diffusion flux, 80 i.e. the amount of the solute transported by diffusion through unit area in unit time. The first term on the right side of Eq. (5.3) is the rate of increase of entropy owing to thermal conduction, the second and third terms are due to internal friction, and the fourth corresponds to diffusion. These terms are always positive. The general form of diffusion flux is i : —pD {gradC + (kT/T)gradT + (kp/P)grad P} (5.4) For incompressible flow under constant pressure and temperature, Eq. (5.3) sim- plifies to d 01),- 0v__k_ 2 01212 — d _ —j[— ——. r dtiips V 2T (Bxk+ 0x.- 36"“ 6:1: —) dV (5") + f (ngradC)2 dV aT Further, if the flow is spherically symmetrical, and the only component of velocity is the radial one 1),, then in a spherical coordinate system W = i%(2(%i’)2+4(%)2) W + pr-—(—%C) 2(g—C)PT W (5-6) For weak solutions (C << I), the chemical potentials (in joules per molecule) for the solute and solvent are (Landau and Lifshitz, 1987, p. 234) p2 : kBTlnC + f2(P, T) (5.7) M : kBTln(1 —- C) +- f1(P, T) (5.8) The potential of the mixture (in joules per unit mass) is defined by (Landau, 1987, 81 p. 228) p.dC : p2 dng + #1 atm (5.9) where M; is Boltzmann constant, and n1,n2 are the numbers of solvent and solute molecules in unit mass of solution. The numbers n1, n2 satisfy the relation nlml + ngmg : 1 where ml (or m2) is the mass of one molecule of solvent (or solute). By definition, C = ngmg. From Eq. (5.9), it follows that H = #2/m2 — [Ll/ml (5-10) Combining Eqs. (5.7) —— (5.10), the change of [.L relative to concentration can be obtained as (6“) ~ kBT — RJ (5.11) CC PT 0mg AIC where Ru is the universal gas constant, and M is the molecular weight of the so— lute. Substituting Eq. (5.11) back into Eq. (5.6), the rate of entropy increase in an isothermal, spherically symmetrical, weak solution with spatially uniform pressure is if st — f3 2 8”” 2+4.(3’1)2 dV dt ”‘ ‘ T 07‘ 7‘ ‘ JFWD—‘M (%)2R“W (5.12) The two items in Eq. (5.12) represent the rate of viscous dissipation of energy as \Ilmm =77/f (2(80’)+ +4(3—:‘)2) (1th (5.13) 82 and the rate of dissipation of energy due to diffusive effects as Twig— )2 . WWW-on: ff ’0 0111)}?qu (5.14) 5.1.2 The Significance of Difiusion and Viscous Effects t0 Nucleation Having derived expressions for the dissipation of energy due to viscous and diffusive effects, the order of magnitude of each dissipation term will be estimated, and its significance to nucleation of gas-supersaturated liquids discussed. We consider an initially uniform gas-water solution in which a bubble of critical size (Re) forms as a result of density / thermal fluctuations. The classical nucleation theory indicates that the reversible work required to form such a bubble is, Eq. (2.50) 4 . WM, 2 gnaR: (5.15) and the homogeneous nucleation rate is determined by the reversible work through the following relation, Eq. (2.51) 30 W. J:N — — m’ 7.1 (WM m) <0 6> However, since the nucleation process may not necessarily be reversible, the dissipa— tion of energy due to viscous and diffusive effects may not necessarily be neglected, as in the classical nucleation theory. Therefore it is useful to estimate the order of mag- nitude of dissipative energy terms, so we know whether viscous and diffusive effects are significant to nucleation. Since the path of a real, irreversible nucleation process can not be specified uniquely, a simple representative path is assumed: the bubble grows from nearly 83 zero size to the critical size at a constant growth rate, i. e. (112 __ : ’Uc :: const. or R = vet dt where R is the radius of the bubble. From the continuity equation, the velocity of liquid is ———vc, 7‘ > R (5.17) The dissipation of energy due to viscous effects is then obtained as 61), 2 2),. 2 11mm, _ ffn{2( 07') ”(7) }dth (5.18) Re/vc 00 61),. 2 v? 2 2 ——/0 [R n{2(6r> +4(—r—) }47r7' drdt Re /vc : / 16777rR v3 dt 0 : 8n7rch: Comparing this viscous dissipation term with the reversible work (Eq. 5.15) required to form a bubble of critical size, the bubble growth rate should satisfy 10 l 71 x 10"3 6 1 x 10‘3 ~ 12 m/s for the two terms to be of the same order and viscous dissipation to be significant. This velocity seems implausibly large for bubble growth in physical systems. Next we will estimate the order of magnitude of the dissipation of energy due to diffusive effects. Figure 5.1 shows the growth of a bubble in a gas-liquid solution. Applying the conservation of mass principle, the mass flow rate of gas into the growing 84 bubble is dm d "(F : a 00) :000 Therefore, the concentration profile can be found by integrating Eq. (5.24), _ EOMXgasch 3 RquDr _ Ach ’I" (5.25) C(r) 2 Coo 2 C00 (5.26) Once the concentration profile is known, the dissipation of energy by diffusion can be 86 calculated as TpD( (8_C_)2Ru ‘sz'f/usion — /]{ CM dV dt (5.27) Re/vc = [0 [004 WTpDRu( LC)? 7" 2dr (1t _ Re/vc WTpDRflA 113122 ’ /o [004 7‘ (Cm—Ach/r)drdt _ Re/vc TpDE, Coo — [0 47r 1W .Avaln(C,oo 30.24116) dt _ TpDRu Coo R3 — 47f M Aln(°o — Ave) 2 8 C: 2 .— _ gwanas lIl (C——ooA—’UC) Re (028) Comparing this dissipation term with the reversible work required to form a bubble of critical size, for the two terms to be of the same order, the bubble growth rate should satisfy Coo 2Xgas 111m N 1 Coo c N 0.39—+— 5.29 v A ( ) In the above derivation, the estimation X gas ~ 1 was used. For 02 dissolved in water at 30°C, the value of A is 4 0’1”me 3 Rump 4 0.071 x 32.0 x 1.0 33314.0 x 303 x 996 x 2.8 x 10-9 : 0.43 (s/m) (5.30) A: The saturation concentration of ()2 dissolved in water at 10 MPa corresponds to a mole fraction of about 0.002 or a. mass fraction of 0.0035. If 000 takes this value, then for the dissipation of energy by diffusion to be of the same order as the reversible 87 work to form a bubble of critical size, it requires vc ~ 0.003 (m/s) At this bubble growth rate, it would take about 3.3 micro seconds to form a bubble of size at 10‘8 m, which seems plausible. In contrast, a bubble growth rate of 12 m/s, required for viscous dissipation to be significant, seems highly improbable. It is noteworthy that if we assume diffusion occurs only in a limited range, say R g r g nR, the integration of Eq. (5.27) from 'r = R to r : nR yields 8 Cm _ A C In which case, 0.0053 c N -—-—- . 2 v 1.65 — g,- (5 3 l and vc is still of the same order regardless of whether 71 is set to 10, or 100, or infinity. In real nucleation processes, bubble formation and growth may not necessarily follow spherical growth at a constant rate. However, the largest contributions to each integral in Eq. (5.12) take place as R approaches Re, and so the dissipation terms in real bubble growth are still likely to be of the same order as these estimates. If we instead assumed that the bubble grew at a constant £31,2-, the same order of magnitude for (Pm-”1,3,0” would be obtained. Therefore, for nucleation of gas-supersaturated liq- uids, it seems reasonable to add the irreversible dissipation of energy through diffusion to the reversible work in a nucleation rate equation like Eq. (2.52) or Eq. (2.53). 88 5.1.3 A New Nucleation Model for Liquids Supersaturated with Different Gases Based on the findings of the previous section, it is proposed that the energy required for nucleation to occur in a weak gas-liquid solution includes not only the reversible work of formation of a gas / vapor nucleus of critical size, but also the energy of diffusive dissipation. Therefore one term exp{—q/(kT)} is added to the classical nucleation rate equation for a weak gas-liquid solution, and the modified nucleation rate is 3 w... J ‘1 NW3 fiexp{— (7TH) $015)} (533) 2 2 NW3 %exp{(—4’;Z;— ’31,) 6,09 3)} (5.34) where q is the dissipation of energy due to diffusion which always inhibits nucleation. If Henry’s law is assumed, substituting Eq. (4.6) into the above equation yields 3 16 3 0 J z N2/3‘/ EO}? exp { (— ggflfim — g?) 42(6), 5)} (5.35) No model for q appears to have been published for nucleation in weak gas—liquid solutions, so a preliminary expression for q is developed here. While the dissipation of energy due to diffusion in nucleation processes is clearly a quantity that depends on microscopic effects, some insight into its likely dependence on measurable properties can be gained by considering the equation of diffusive dis- sipation we derived in the previous section, which is restated here T D(.— \Ildiffusion: /f p C AI) RudV dt (5.36) so, T919027.) R. , (Irv/f C 11 1 (1V dt (.37) 89 where p is the density of the liquid, D is the gas diffusion coefficient in the liquid, C is the gas concentration in the liquid, R, is the universal gas constant, and M is the molecular weight of the gas. The \Ildz- ”1,3,0" term is too complicated to use in a nucleation rate equation, so it is useful to approximate it by simplifying its functional dependence. The functional dependence of q on these variables and other coefficients is estimated as follows. 1. The larger the diffusion coefficient of a dissolved gas in a liquid, the more rapidly a uniform dissolved-vapor concentration will be restored after any perturbation and the more unlikely it will be that sufficient free collisions between vapor molecules can create a nucleus of critical size. Therefore, with J proportional to e“q, it follows that q oc D. 2. The larger the deviation of dissolved gas concentration from its equilibrium saturated concentration (AC 2 C —Csat), the more metastable the solution, and nucleation is more likely to occur. The increase of AC should lead to a smaller q, and a larger nucleation rate. Therefore q should be inversely proportional to AC, suggesting that q oc D/AC or q or D/C. For solutions approaching supersaturation nucleation limits, the following approximations can be made: C >> Csat and C m AC 3. The inversely proportionality of q to C or AC can also be verified in this way. The larger the gas concentration, the more gas molecules exist in a unit volume and the more likely that sufficient gas and vapor molecules collide together to create a nucleus of critical size. From an energy balance viewpoint, it needs more work to bring sparsely distributed gas molecules together to form a bubble. Thus a smaller C should lead to a larger q and a smaller nucleation rate. 4. A combination of din‘iensional analysis and the above reasoning suggests that 90 the energy of activation for diffusion may take the form q = f {D (3%) kT} (5.38) 343(5)} where f and F are unknown functions. A simple linear model is proposed. q could take the form: 1 q = constant (D) (KC) kT (5.40) where the constant has a unit of s/m2. If Henry’s law is assumed, it can be simplified as q : constant (D) (aAlP ) kT (5.41) In Eq. (5.40), q is proportional to DT/AC or DT/ C. The same term DT/C appears in the equation of \Ildiffusion, (5.36). The Bunsen coefficient a is adopted here to represent the solubility. It is defined as the volume of gas reduced to 20°C and 1 atm, which is absorbed by unit volume of solvent at a stated temperature under a partial pressure of gas of 1 atm. Substituting Eq. (5.40) into Eq. (5.34) gives the revised nucleation rate equation, 30 47m 7'2 1 : 2/3 _ e _ , __ F, J N 7rm exp {( 3kT constant (D) (AC)) ¢(6,fi)} (o 42) If Henry’s law is assumed, the above equation simplifies to / 3o 167ro3 1 J : N2/3 ir—m exp { (—W — constant (D) (GAP )) ¢>(6,/3)} (5.43) 91 5.1.4 Calculation Results of the New Nucleation Model for Water Supersaturated with Difierent Gases at a Constant Temperature Eq. (5.43) can be solved by setting J to a constant value (e. g. 1) and calibrating the constant term in q, and the shape factor. The values of solubility and diffusivity of different gases dissolved in water can be found from Wise and Houghton (1966), Janssen (1987), “CRC Handbook of Chemistry and Physics,” or “Solubility Data Series.” Some of them are listed in Table 5.1. Table 5.1: The values of solubility and diffusivity of several gases dissolved in water at 30°C Gas Solubility a Diffusivity Bunsen absorption coef. D (x10'9 m2/s) Helium 0.0087 8.0 Nitrogen 0.0129 3.5 Argon 0.0273 2.7 Oxygen 0.0271 2.8 Table 5.2 shows the predicted values of the pressure difference for bubble formation (APn) from the revised nucleation equation (5.43), as well as the experimental data of Finkelstein & Tamir at 30°C and of Hemmingsen at 25°C. Since Finkelstein and Hemmingsen claimed that AB, is weakly dependent on temperature in the range of 293 to 303K, we can compare AP" data at 30°C with those at 25°C. In Eq. (5.43), Henry’s law was used to estimate the dissolved gas concentration at different pressures. As discussed in Section 2.5, departures from Henry’s law should be considered for pressures higher than 7 or 8 MPa. The supersaturation thresholds in this study are in the range of 13 to 42 MPa, therefore the generalized form of Henry’s law (Eq. 2.61) is needed to calculate saturation concentration in the nucleation equa- tion (Eq. 5.42). We suspect that the large discrepancy between experimental data and calculation data of AP" for nitrogen is due to the use of Henry’s law. Table 5.3 and Figure 5.2 show the supersaturation nucleation thresholds for dif- 92 Table 5.2: Comparison of the experimental data of AR: (MPa) with the predicted data from the revised nucleation equation (5.43) Gas Experimental Result of AB, Calculation Result of AP” (MPa) (MPa) Helium 42.0 :1: 1.4 44.02 Neon 32.4 i 1.4 28.45 Nitrogen 13.8 :1: 1.4 18.66 Argon 13.1 :1: 1.4 13.44 Oxygen 13.5 - 14.0 at 25°C from Hemmingsen 13.56 * If unspecified, the experimental data are from Finkelstein at 30°C. ferent gases dissolved in water at 25°C calculated from Eq. (5.42) with (b = 0.003 and J : 1, taking into account departures from Henry’s law. The reason we choose 25°C Table 5.3: Predicted data of nucleation thresholds from Eq. ( 5.42) taking into account the departure from Henry’s law Gas Experimental Result of AP" Calculation Result of APn (MPa) (MPa) Helium 42.0 :1: 1.4 40.60 Nitrogen 13.8 :t 1.4 14.75 Argon 13.1 :t 1.4 10.50 Oxygen 13.5 - 14.0 at 25°C from Hemmingsen 14.00 * If unspecified, the experimental data are from Finkelstein at 30°C. instead of 30°C is that we only have solubility data (for different gas species dissolved in water under a wide range of pressures) at 25°C. The diffusivity data at 25°C are estimated using a model supplied by Spalding (1963), together with experimental diffusivity data at 20 and 30°C from Wise and Houghton (1966). “2) (“2) _ Z — — 5.44 (IVSC ) 20° C ( pT / pT 200 C ( l 93 A E‘ G — . l 50 x’ . E 4:: c _ . O. 40 ’I’ i <1 / “a 304 3 x’ N 20 ‘ z’ > I a a" 3;- 10 . ’,’HIH 5 a 0 ’ I T l 1 I O 10 20 30 40 50 Experimental data of A P" (MPa) A Helium I Nitrogen 2: Argon 0 Oxygen —---y=x Figure 5.2: Experimental data for AP“ and the predicted data from the nucleation equation revised to include the energy of diffusive dissipation 94 where the Schmidt number is defined as NS. = — (5.45) Table 5.4 lists the estimated diffusivity data of gases dissolved in water at 25°C. Table 5.4: The estimated values of diffusivity of several gases dissolved in water at 25°C Gas Diffusivity D (x10‘9 m2/s) Helium 7.05 Nitrogen 2.97 Argon 2.38 Oxygen 2.63 Comparing Table 5.3 with Table 5.2, it is found that the values of nucleation thresholds for nitrogen and oxygen are in much better agreement when the departure from Henry’s law is taken into account. This can be explained in this way: the departure from Henry’s law for nitrogen dissolved in water is much less than that for oxygen, as shown in Figure 5.3. As we discussed in Section 2.5, departures from Henry’s law always lower the nucleation rate, and therefore require a higher AP", for nucleation to occur. Taking into account the departures from Henry’s law leads to a greater increase in AP" for oxygen than for nitrogen, thus reducing the nucleation threshold discrepancy between oxygen and nitrogen. As we discussed in Section 2.5, accounting for the departure from Henry’s law always leads to higher nucleation thresholds. However, this trend is obscured by the calibration of the two constants in the nucleation rate equations. Due to different calibration constants used in the calculations, we can not compare the absolute values in the two tables, Table 5.2 and Table 5.3. While the use of the generalized form of Henry’s law, Eq. (5.42) produces more accurate results, the simpler model Eq. (5.43), which assumes ideal solution behavior, 95 0.004 -- g 0.003 ~ E 5 0.002 ‘ 0 5 0 0.001 - O T T l 1 0 5 10 15 20 Pressure (MPa) + Oxygen + Nitrogen Figure 5.3: Saturation concentration for gas dissolved in water as a function of pres- sure at 25°C requires much less solubility data, and can still produce useful information on the species dependence of nucleation thresholds. Table 5.5 lists the values of the reversible work required to form a bubble of critical size, and the energy of diffusive dissipation at nucleation limits at 25°C from Eq. (5.42). It can be seen that wm, and q are of the same order for 02, N2, and Ar, but not for He, for which q dominates. Those data again confirm the idea that the energy of diffusive dissipation should not be neglected for nucleation of gas-supersaturated liquids. Table 5.5: The calculated values of wm, and q for nucleation in gas-supersaturated water at 25°C 02 H8 N2 AT wm, / [CT 16020 1061 8493 15648 q / kT 8299 23258 15826 8671 Table 5.2 indicates that, after calibration, the revised nucleation equation predicts 96 the values of AP” from Finkelstein & Tamir’s experimental data quite accurately, in- troducing a strong species dependence to AP” through energy of diffusive dissipation. Therefore the addition of a diffusive dissipation term to the nucleation rate equation appears to be an important and physically realistic improvement to nucleation mod- eling. 5.1.5 Temperature Dependence of Nucleation Thresholds of Gas- S'upersaturated Liquids Several experiments have been conducted to study the temperature dependence of su- persaturation thresholds for gas dissolved in water, but the results are not conclusive. Hemmingsen (1975) concluded that “over a wide range of temperatures, from 3.5 to 40°C, there was only a relatively small change in the cavitation properties (of N2 and Ar dissolved in water). However, with decreasing temperatures over a narrow range below 35°C, there was a significant and substantial increase in the supersaturation required for the onset of any cavitation.” Finkelstein and Tamir (1985) observed that “AP,, is completely independent of temperature in the range of 293 to 303K (for all the tested gases, N2, He, Ne, and Ar).” In his technical notes, Bowers (1999) reported the temperature variation of supersaturation thresholds for oxygen, nitrogen, hydro- gen, etc. in water at 1 atm pressure. The measurements were made by generating solutions of gases chemically. Bowers reported the bubble nucleation limit for oxygen decreased from 0.15 to 0.10 mol/dm3 over the range of 283 to 298K. This corresponds to a decrease of more than 4 MPa of supersaturation pressure. Bowers’s result appears to contradict Finkelstein’s and Hemmingsen’s observations. One possible explanation is that the discrepancy may be due to the different gases studied and the different methods they adOpted to measure the supersaturation thresholds. Finkelstein and Hemmingsen used the method of saturating water with gas at high pressures then de— compressing, while Bowers used chemical reactions to generate highly supersaturated 97 solutions. The results may not be comparable. Table 5.6 shows the values of solubility and diffusivity of oxygen dissolved in water at different temperatures. Data at 20 and 30°C are taken from Wise and Houghton (1966). The diffusivity at 25°C is estimated from Spalding (1963). Solubility at 25°C is obtained from Fog (1990). Table 5.6: The values of solubility and diffusivity of 02 dissolved in water Oxygen T(°C) Solubility a Diffusivity Bunsen coef. D (x10’9 m2/s) 20 0.0323 2.3 .H 25 0.0295 2.63 30 0.0271 2.8 According to Eq. (5.43), as temperature increases from 20 to 30°C, the calculation result of AP" for 02 dissolved in water increases from 13.3 to 13.6 MPa as shown in Figure 5.4, a relatively small change. This result agrees with Finkelstein and Hemmingsen’s experimental observations, but contradicts those of Bowers. In Eq. (5.43), as temperature T increases, D increases and C or AC decreases. This leads to a decrease of wrw, but an increase in q. Thus it would be possible that J may decrease as temperature increases for some gases over some temperature ranges, as indicated by Bowers’ experimental result that “the (supersaturation nucleation) limit for hydrogen increases from 0.03 mol/dm3 at 290K to 0.08 mol/dm3 at 308K.” This phenomenon can not be explained from the classical nucleation theory. 98 14‘ ‘_———t———i A12‘ E105 '3: 6~ 4. 2_ 0 w T I I 0 10 20 3O 40 T(°C) Figure 5.4: Dependence of AP" on temperature for 02 dissolved in water, calculated from Eq. (5.43) 5.2 Prediction of Nucleation Thresholds of Gas-Liquid Solutions in Small Capillary Tubes There have been some attempts to understand the effects of confinement of gas- supersaturated solutions within small volumes and their ability to raise nucleation thresholds. Some experiments on nucleation in small capillary tubes (diameter less than 100nm) were reported by Brereton, et al. (1998), and were described in Sec- tion 4.3. 5.2.1 Dual Dependence Model to Account for Effects of Volumetric Confinement on Nucleation Brereton et al. developed a dual dependence model to predict nucleation rate, i.e. the heterogeneous nucleation in capillary tubes depended on sufficient numbers of both molecules of the vaporizing species and active cavities at the surface at which 99 nucleation took place. The equation for nucleation rate (s‘lm’2) is expressed as: / 3o —47ro 72¢ : 3 ‘t A Nae ive Nsur _ ____e_ 4 J cons z 1 7rm exp { 3kT } (5 6) More details of the dual dependence model can be found in Section 4.3. Although Eq. (5.46) could be calibrated to agree with experimental data, some of assumptions in the original paper (Brereton et al. 1998) seem questionable after closer inspection. For example, one assumption concerning the initial state of embryonic nuclei compressed within surface cavities, and another assumption that the changes in the radii of curvature of nuclei in crevices are neglected during decompression was questionable. Also more accurate estimations of the critical radii of nucleation in gas-liquid solutions could have been made. In addition, the effect of the dissolved gas species on the observed threshold for nucleation (or, maximum dissolved-gas concen- tration above which nucleation always took place) was not considered. Based on the model of nucleation of gas-supersaturated liquids proposed in Sec- tion 5.1 and Brereton’s dual dependence model, we propose an improved nucleation rate equation to describe nucleation of gas-supersaturated liquids in small capillary / 3o —47ro 73¢ qu I, J — COIlSt A Nactive leur % exp {—3k_f—— — 76?} (0.47) where Nactive and N13,), are numbers of active sites and surface molecules per unit tubes: area, and q accounts for the dependence of nucleation of gas-supersaturated liquid on diffusion, and takes the form (see page 91) q = constant (D) (23%) ktT (5.48) The critical radius (re) for bubble formation in a gas-liquid solution can be calculated from Ward’s equation, Eq. (2.23). This model (Eq. 5.47) does not make any assump— 100 tion about the initial state of embryonic nuclei compressed within surface cavities, or that the changes in the radii of curvature of nuclei in crevices are neglected during decompression. The total number of surface molecules n13“, is determined from the average spacing s between molecules and the capillary-tube surface area. The ratio of these molecules around the perimeter to those in the bulk liquid in the capillary is then nlsur 43 = — 5.49 m d ( ) For oxygen dissolved in water, nlsur 0o 4300 —’ ‘ = ‘ 5.50 NI,O2 d ( ) The concentration of oxygen in water is defined as C = —"£- (5.51) "1,1120 and can be estimated from Henry’s Law as Pssat C = ” 5.52 H < > where H is Henry’s constant for the two species and P380, is the pressure under which the saturated oxygen-water solution was prepared. Thus, 43 Pssa ”13211302 2 "(£20 d02 7t (5-53) The average spacing between oxygen molecules can be expressed as 85120 - z 5.54 802 (Pssat/H)1/3 (O ) 101 Substituting Eq. (5.54) into Eq. (5.53) yields the expression for the number of surface molecules of the dissolved gas per unit area, N lsur,02 : nlsur,02 /A 71.1,}120 43HQO Pssat 2/3 The nucleation rate J(s’1m’2) for 02-water solution in capillary tubes is / 30 —47ro r305 J : const A Nactive leur.02 g7; exP {—3-k—T—_ - %} rm.) ‘11 /3o —47ro r34) qu —— COHStlA ( T‘e ) leur,02 $exp{—3—k-T—— — Ff} (5.56) where Mama, can be calculated from Eq. (5.55), Native can be estimated from Eq. (4.10), and the critical radius (re) for bubble formation in a gas-liquid solution can be calculated from Ward’s equation, Eq. (2.23). Since the conditions for nucleation are quite insensitive to the rate at which bub— bles are formed, similar values of P330, would be obtained even if J were varied by several orders of magnitude, between, say, 1 and 1000. Eq. (5.56) can be solved by setting J to a constant value and calibrating the shape factor and constl. Figure 5.5 shows the predicted supersaturation pressure thresholds of 02 dissolved in water from Eq. (5.56), predicted over a range of different capillary tube diameters, as well as the experimental data from Brereton, et al. (1998). It can be seen from the figure that Eq. (5.56) can provide a reasonable agreement with the experimental data when suit- able calibration values have been chosen for the two constants ((15 = 8.76 x 10‘5 and J = 1). Thus Eq. (5.56) can correctly interpret the effect of volumetric confinement on nucleation thresholds for gas-liquid solutions through the use of a nucleation rate equation in which the pre—exponential scaling is jointly dependent on both the num- ber of active surface cavities and the number of molecules of dissolved gas adjacent 102 to the surface. Punt (MPa) 8 K3 0 20 4O 60 80 1 00 (NM M) +calculation I experiment Figure 5.5: Predicted and measured dependence of 02 supersaturation pressure on capillary tube diameter using Eq. (5.56) From the classical heterogeneous nucleation model, no volumetric confinement effects are predicted. Thus according to the classical model, the supersaturation thresholds for nucleation in different size of capillary tubes are almost the same, i.e. a horizontal line would be obtained in Figure 5.5. 5.2.2 A Preliminary Model to Account for Efiects of Volumetric Confinement on Nucleation Although the dual dependence model can reproduce experimental data of nucleation thresholds of oxygen dissolved in water in small capillary tubes with reasonable ac- curacy, some questions remain: 103 o The model introduces the effect of volumetric confinement on nucleation thresh- olds through pre—exponential factors in the nucleation rate equation (5.56). As we mentioned before, the nucleation rate is quite insensitive to pre-exponential factors. Therefore the predicted P330) vs. (1 curve from Eq. (5.56) usually has less curvature than the experimental curve, as shown in Figure 5.5. Also the calibrated shape factor is much smaller than expected. Those questions suggest one might instead introduce a term inside the exponential part, which depends on the diameter of the capillary tube. The following assumption was proposed: the shape factor for nucleation is inversely proportional to the diameter of capillary tube, i.e. (brxa or qbz— (5.57) where (to is a function of contact angle, and is independent on tube diameter. do is usually in the order of 10‘8 or 10‘7 meters. Some reasons which lead to this assumption include: a As capillary tube diameter decreases to the order of one micrometer, it is pos- sible that the average cavity size of the tube surface is reduced, which inhibits nucleation and corresponds to large shape factors (this hypothesis is currently under experimental investigation). 0 The maximum value of shape factor is unity, which corresponds to the occur- rence of homogeneous nucleation. Experimental data show that as capillary tube diameter tends to zero, the nucleation threshold approaches homogeneous nucleation limits. With these assumptions concerning shape factor, the heterogeneous nucleation rate equation becomes 30 47m R2 (I (150 : N2/3,/__ - e _ _. — 5’8 J nmeXp{( 3kT kT) d} ( O) 104 Eq. (5.58) was used to calculate nucleation thresholds of 02 dissolved in water in small capillary tubes. The results, shown in Figure 5.6, fit the experimental data from Spears very well, indicating a dependence of nucleation threshold on volumetric confinement. 20 - 15 r E E 10 - l 5 l 0 . . r , 0 20 40 60 80 d(u m) +calculation In experiment Figure 5.6: Predicted and measured dependence of 02 supersaturation pressure on capillary tube diameter using Eq. (5.58) Once calibration constants are obtained for oxygen dissolved in water in small capillary tubes, the nucleation thresholds for other gas species can be readily calcu- lated from Eq. (5.58) without further calibrations. The gas species dependence is introduced through the activation energy for diffusion term, q, as discussed in the previous section. Figure 5.7 shows the predicted nucleation thresholds of 02 and He dissolved water in small capillary tubes with do = 1.2 x 10‘7 and J = 1, as well as the experimental data from Spears. It can be seen from the figure that the shape factor assumption (Eq. 5.57) together with an activation energy for diffusion term 105 can produce good agreement with experimental data. 100 - § 3' 10 '1 I O. 1 I I I I I I 0 20 4o 60 80 100 120 dlu m) +He-cal. I He-exp. -o—02-cal. A OZ-exp. Figure 5.7: Predicted and measured dependence of 02 and He supersaturation pres sure on capillary tube diameter using Eq. (5.58) As another test, we attempted to predict the superheat thresholds for nucleation of pure water in small capillary tubes. The heterogeneous nucleation rate equation for pure water in capillary tubes is 30 47ro R2 d : 2/3 _ _ e _9 J N VnmeXp{( 3kT )d} The experiment was described in detail in Brereton et al. (1998) and essentially in— (5.59) volved immersing small capillary tubes filled with pro-pressurized water in oil baths. Figure 5.8 shows the comparison of predicted superheat nucleation thresholds and ex— perimental data. The reasonable agreement provides some support for the assumption that the shape factor for nucleation in small capillary tubes is inversely proportional 106 to tube diameter. G l 3' 260 ~ 2 a u a E 3 180 ~ I ¢ _ C I 3. g 140 3 O 3 z 100 1 I I I 1 O 40 80 120 160 200 Tube diameter ([1. m) +cal. I exp. Figure 5.8: Predicted and measured superheat nucleation thresholds of pure water in capillary tubes The effect of volumetric confinement is really one of reduced surface area. It is a result of heterogeneous nucleation depending on both the number of active surface nucleation sites (cavities) and the number of molecules of the dissolved-gas species at the surface, as well as on bubble shape factors and other surface-dependent ef- fects. As surface area is reduced, the nucleation rate in a dual dependence model is also reduced, possibly until homogeneous nucleation becomes more likely. This ex- planation of raising the nucleation threshold by volumetric confinement is essentially through surface area reduction, consistent with the nucleation model developed in this dissertation. 107 CHAPTER 6 CONCLUDING REMARKS AND RECOMMENDATIONS In this dissertation, we first presented the classical nucleation theory as originally developed for pure liquid-vapor mixtures, and as extended for gases dissolved in dissimilar liquids. The reason for its failure to predict the strong dependence of su- persaturation nucleation thresholds on gas species was explained primarily as one of neglecting diffusive dissipation. Finally nucleation in small capillary tubes was addressed. Some modifications were made in this study to the classical nucleation theory, by considering the irreversible effects of diffusion of gas molecules and volumet— ric confinement, so it could reproduce the experimental data of nucleation thresholds of gas-liquid solutions correctly. New contributions of this study to nucleation theory are listed below: 0 The equations for dissipation of energy due to viscous and diffusive effects dur— ing bubble growing processes were derived. The order of magnitude of these dissipative terms was estimated. The significance of the energy of dissipation due to diffusion to nucleation of gas-supersaturated liquids was pointed out. (Section 5.1) o A new model including diffusive dissipation was proposed to predict nucleation thresholds of water supersaturated with different gas species, which is superior to an earlier correlation equation proposed by Tamir & Finkelstein for the same purpose. The model predictions are in good agreement with experimental data. (Section 5.1) o Ward’s equation of critical radius and Brereton’s dual dependence model have been incorporated into nucleation theory to predict nucleation thresholds of gas-supersaturated liquids in small capillary tubes. Calculation results indicate 108 that the new model can predict the effect of volumetric confinement on rais— ing nucleation thresholds well, after an appropriate choice of two calibration constants. (Section 5.2) o The limitations of Henry’s law at high pressures were recognized. The effects of the departure from Henry’s law on nucleation were also discussed. An improved dissolved-gas concentration model was used in calculations of supersaturation thresholds. (Sections 2.5, 5.1, and 5.2) I A new model, which assumes the shape factor for nucleation in small capillary tubes to be inversely proportional to the tube diameter, was proposed. It was applied to predict the supersaturation nucleation thresholds of water supersatu- rated with different gas species, as well as the superheat nucleation limits of pure water in small capillary tubes. The predicted results were in good agreement with experimental data. (Section 5.2.2) Due to the limited time frame of this research, many questions and uncertainties still remain, and need further study. Some of them are: 0 Some experiments are recommended to study the physical chemistry of surfaces of small capillary tubes. Experimental data on how contact angle, average cav- ity size, shape factor, and other parameters vary with tube diameter and com- position would benefit the modeling of nucleation in confined volumes. Also some special treatments to surfaces or surface coating, which changes the con- tact angle and other properties, and therefore affect nucleation, are worthy of investigation . o For experiments measuring nucleation thresholds of liquids flowing through small capillary tubes, the temperature of effluent should be measured by a thermo—couple. The liquid temperature may rise by well over 20°C due to fric— 109 tion between the tube surface and the liquid, particularly when tube diameters are of micron dimensions. As an alternative approach to classical nucleation theory, the density functional method assumes continuous changes rather than abrupt changes of fluid prop— erties at the interface of nuclei, and thus avoids the macroscopic assumptions of classical nucleation theory. Although the density functional method is not yet able to make quantitative comparisons with experimental results, it shows some potential for making useful predictions in nucleation study. Further study should be conducted on this topic. In recent years, the lattice Boltzmann method (LBM) has been applied to simu- lations of single-phase and multiphase fluid flows (Chen, 1998). The fundamen- tal idea of the LBM is to construct simplified kinetic models that incorporate the essential physics of microscopic processes so that the macroscopic average properties obey the desired macroscopic equations. It is possible that the LBM may be applied to simulate the nucleation process in liquids, and improve un- derstanding of non-equilibrium phase-change problems. 110 APPENDIX 111 APPENDIX A THE COEFFICIENTS OF HENRY’S LAW FOR NITROGEN, HYDROGEN, ARGON, AND HELIUM DISSOLVED IN WATER In Section 2.5, the applicability of Henry’s Law was discussed, together with im- provements on it which accounts for departures from ideal solution behavior at high pressures. Here more values of the coefficients of Henry’s Law for Nitrogen, Hydrogen, Argon, and Helium dissolved in water, are listed. Figure A.2 shows the variation of log(H) with pressure for N2 dissolved in water at 25°C. The line which best fits these data for H of N2 dissolved in water at 25°C is log(H) = 3.9684 + 3.0061 x 10’3 P (A.1) where P is again in MPa. 1.E-02 - 8.E-03 * 6.E-03 r 4.E-03 ~ Concentration 2.E-O3 — 0.E+OO 1 . . . o 50 100 150 Pressure (MPa) Figure A.1: Saturation concentration of N2 in water 112 Figure A] shows the saturation concentration (in mole fraction) of N2 in water at 25°C, as a function of pressure. The experimental data were obtained from Krichevsky (1935), as shown in Table A1 5 - 4.5 ~ 9 / ‘5 4 _ 2 3.5 - 3 T I I 0 50 100 150 Pressure (MPa) Figure A.2: Variation of H with pressure for N2 dissolved in water Table Al: The solubility of nitrogen in water at different pressures Pressure Saturation Concentration (MPa) (mole fraction) 2.5 2.800x10’4 5 5.420x 10-4 10 1.015 x10:3 20 1.812x 10-3 30 2.455x10‘3 50 3.558x10‘3 80 4.909x10‘3 100 5.720x10'3 Figure A.3 shows the saturation concentration (in mole fraction) of H2 in water at 25°C, as a function of pressure. The experimental data were obtained from Krichevsky (1935), as shown in Table A2. Figure A.4 shows the varition of log(H) with pressure 113 2.0E-02 - 1.5E-02 A 1.0E-02 3 Concentration 5.0E-03 ~ ODE-+00 l T I 0 50 100 150 Pressure (MPa) Figure A.3: Saturation concentration of H2 in water 5 1 4.5] E ‘5 41 2 n- - ; ¢ ; ‘4. 3.5 3 3 T T I 0 50 100 150 Pressure (MPa) Figure A.4: Variation of H with pressure for H2 dissolved in water 114 for H2 dissolved in water at 25°C. The best-fit line for H of H2 dissolved in water at 25°C is log(H) : 3.7598 + 8.6111 x 10“4 P (A2) where P is again in MPa. Table A2: The solubility of hydrogen in water at different pressures Pressure Saturation Concentration ‘ (MPa) (mole fraction) _. 2.5 4.306x 10‘4 ‘ 5 8.635x10‘4 . 10 1.709><10’3 20 3.351x10-3 40 6.390x10“3 60 9.252x10—3 80 1.191x10-2 100 1.425x 10-2 5.0E-03 — 4.0E-03 ~ : 8 g 3.0E-03 3 I: 8 c 2.0E-03 _ O U 1.0E-03 ~ 1 0.0E+OO I T I I l I 0 20 40 60 80 100 Pressure (MPa) Figure A.5: Saturation concentration of He in water Figure A.5 shows the saturation concentration (in mole fraction) of He in water at 25°C, as a function of pressure. The experimental data were obtained from “Solubility 115 Data Series, Volume 1, Helium and Neon _ Gas Solubilities.” Figure A.6 shows the 1D 0 log(H) A 3 i i 0 50 1 00 Pressure (MPa) Figure A.6: Variation of H with pressure for He dissolved in water varition of log(H) with pressure for He dissolved in water at 25°C. The best-fit line for H of He dissolved in water at 25°C is log(H) = 4.2120 + 8.0044 x 10’4 P (A.3) where P is again in MPa. Figure A.7 shows the saturation concentration (in mole fraction) of Ar in water at 25°C, as a function of pressure. The experimental data were obtained from Spears. Figure A.8 shows the varition of log(H) with pressure for Ar dissolved in water at 25°C. The best-fit line for H of Ar dissolved in water at 25°C is log(H) = 3.6706 + 2.7759 x 10-3 P (A.4) where P is again in MPa. 116 1.2E-02 3 1.0E-02 3 8.0E-03 3 6.0E-03 3 4.0E-03 3 ZOE-03 3 0.0E+00 i r i i 0 20 40 60 80 Pressure (MPa) Concentration Figure A.7: Saturation concentration of Ar in water 5 I 4.5 3 EE ‘5 4 3 2 A 3 f : :4: 3.5 l 3 r i . , 0 20 40 60 80 Pressure (MPa) Figure A.8: Variation of H with pressure for Ar dissolved in water 117 BIBLIOGRAPHY 118 BIBLIOGRAPHY [1] Adamson, A., 1990, “Physical Chemistry of Surfaces,” p. 364, fifth edition, John Wiley & Sons, Inc. [2] Bankoff, S. G., 1966, “Diffusion-Controlled Bubble Growth,” Adv. Chem. Eng, 6, p. 1. [3] Becker, R., and Diiring, W., 1935, “The Kinetic Treatment of Nuclear Formation in Supersaturated Vapors,” Ann. Phys., 24, pp. 719-752. [4] Bird, R. B., Stewart, W. E., and Lightfoot, E. N., 1960, “Transport Phenomena,” John Wiley & Sons, Inc. [5] Blander, M., Hengstenberg, D., and Katz, J. L., 1971, “Bubble nucleation in n—Pentane, n—Hexane, n-Pentane + Hexadecane mixtures, and Water,” J. Phys. Chem., 75, p. 3613. [6] Blander, M., and Katz, J. L., 1973, “The Role of Bubble Nucleation in Explosive Boiling,” Fourteenth National Heat Transfer Conference, Atlanta, GA. [7] Blander, M., and Katz, J. L., 1975, “Bubble Nucleation in Liquids,” AIChE J ., 21, p. 833. [8] Bowers, P. G., Hofstetter, C., Letter, C. R., and Toomey, R., 1995, “Super- saturation Limit for Homogeneous Nucleation of Oxygen Bubbles in Water at Elevated Pressure: Superhenry’s Law,” J. Phys. Chem., 99, p. 9632. [9] Bowers, P. G., Bar—Eli, K., and Noyes, R. M., 1996, “Unstable Supersaturated Solutions of Cases in Liquids and Nucleation Theory,” J. Chem. Soc., Faraday Trans, 92, p. 2843. [10] Bowers, P. G., Hofstetter, C., Ngo, H. L., and Toomey, R. T., 1999, “Tempera- ture Dependence of Bubble Nucleation Limits for Aqueous Solutions of Carbon Dioxide, Hydrogen, and Oxygen,” J. Colloid Interf. Sci., 215, pp. 441-442. [11] Brereton, G. J., Crilly, R. J., and Spears, J. R., 1998, “Nucleation in Small Capillary Tubes,” Chemical Physics, 230, p. 253. [12] Bretherton, F. P., 1960, “The Motion of Long Bubbles in Tubes,” J. Fluid Mech., 10, p. 166. [13] Buff, F. P., and Kirkwood, J. G., 1950, “Remarks on Surface Tension of Small Droplets,” J. Chem. Phys., 18, p. 991. [14] Buff, F. P., 1955, “Spherical Interface. II. l\«Iolecular Theory,” J. Chem. Phys, 23, p. 419. 119 [15] Burmeister, L. C., 1993, “Convective heat transfer,” 2nd edition, p. 113, John Wiley & Sons, Inc. [16] Cable, M., and Frade, J. R., 1987, “Diffusion Controlled Growth of Multi- Component Gas Bubbles,” J. Materials Sci., 22, p. 919. [17] Cahn, J ., and Hulliard, J ., 1958, “Free Energy of a Nonuniform System. I. Inter- facial Free Energy,” J. Chem. Phys, 28, p. 258. [18] Cahn, J., and Hulliard, J., 1959, “Free Energy of a Nonuniform System. III. Nucleation in a Two-Component Incompressible Fluid,” J. Chem. Phys, 31, p. 688. [19] Carey, V. P., 1992, “Liquid-Vapor Phase Change Phenomena,” Hemisphere Pub- lishing Co. [20] Carey, V. P., 1999, “Statistical Thermodynamics and Microscale Thermo- physics,” Cambridge University Press. [21] Carr, M. W., Hillman, A. R., and Lubetkin, S. D., 1995, “Nucleation Rate Dispersion in Bubble Evolution Kinetics,” J. Colloid Interf. Sci., 169, pp. 135- 142. [22] Carslaw, H. S., 1945, “Introduction to the Mathematical Theory of the Conduc- tion of Heat in Solids,” p. 158, Dover Publications. [23] Chen, R, and Chen, S. H., 1985, “Diffusion of Slightly Soluble Gases in Liquids Measurement and Correlation with Implications on Liquid Structures,” Chem. Eng. Sci., 40, p. 1735. [24] Chen, S., and Doolen, G. D., 1998, “Lattice Boltzmann Method for Fluid Flows,” Annu. Rev. Fluid Mech, pp. 329-364. [25] Clever, H. L., 1979, “Solubility Data Series, Volume 1, Helium and Neon — Gas Solubilities,” Pergamon Press. [26] Cole, R., 1974, “Boiling Nucleation,” Adv. Heat Transfer, 10, p. 85. [27] Defay, R., Prigogine, I., Bellemans, A., and Everett, D. H., 1966, “Surface Ten- sion and Adsorption,” John Wiley 81. Sons, Inc. [28] Dergarabedian, P., 1953, “The Rate of Growth of Vapor Bubbles in Superheated Water,” J. Appl. Mechanics, p. 953. [29] Epstein, P. S., and Plesset, M. S., 1950, “On the Stability of Gas Bubbles in Liquid-Gas Solutions,” J. Chem. Phys, 18, p. 1505. [30] Ferziger, J. H., and Peric, M., 1996, “Computational Methods for Fluid Dynam- ics,” Springer—Verlag. 120 [31] Finkelstein, Y., and Tamir, A., 1985, “Formation of Gas Bubbles in Supersatu- raed Solutions of Gases in Water,” AIChE Journal, 31, p. 1409. [32] Fisher, J. C., 1948,“The Fracture of Liquids,” J. Appl. Phys, 19. p. 1062. [33] Fogg, P. and Gerrard, W., 1990, “Solubility of Gases in Liquids: a Critical Evaluation of Gas/ Liquid Systems in Theory and Practice,” John Wiley & Sons Ltd. [34] Forest, T. W., and Ward, C. A., 1977, “Effect of a Dissolved Gas on the Homo— geneous Nucleation Pressure of 3 Liquid,” J. Chem. Phys, 66, p. 2322. [35] Forest, T. W., and Ward, C. A., 1978, “Homogeneous Nucleation of Bubbles in Solutions at Pressures above the Vapor Pressure of the Pure Liquid,” J. Chem. Phys, 69, p. 2221. [36] Frenkel, J ., 1939, “A General Theory of Heterophase Fluctuations and Pretran- sition Phenomena,” J. Chem. Phys, 7, p. 538. [37] Frenkel, J., 1946, “Kinetic Theory of Liquids,” Chap. VII, Dover Publications, New York. [38] Frisch, J. F., 1957, “Time Lag in Nucleation,” J. Chem. Phys, 27, p. 90. [39] Geankoplis, C. J., 1983, “Transport Process: Momentum, Heat, and Mass,” Allyn and Bacon, Inc., p. 391. [40] Gyftopoulos, E. P., and Beretta, G. P., 1991, “Thermodynamics Foundations and Applications,” Macmillan Publishing Co. [41] Hemmingsen, E. A., 1970, “Supersaturation of Gases in Water: Absence of Cav- itation on Decompression from High Pressures,” Science, 167, pp. 1493-1494. [42] Hemmingsen, E. A., 1975, “Cavitation in Gas-Supersaturated Solutions,” J. Appl. Phys, 46, p. 213. [43] Hemmingsen, E. A., 1977, “Spontaneous Formation of Bubbles in Gas— Supersaturated Water,” Nature, 267, pp. 141-142. [44] Hirth, J. P., and Pound, G. M., 1963, “Condensation and Evaporation, Nucle— ation and Growth Kinetics,” Progr. Mater. Sci., 11, pp. 1-167. [45] Holman, J. P., 1990, “Heat Transfer,” 7th edition, PP 301, PP 349, McGraw-Hill. [46] Huntington, H. B., and Seitz F ., 1942, “Mechanism for Self-Diffusion in Metallic Copper,” Physical Review, 61, p. 315. [47] Huntington, H. B., and Seitz F., 1942, “Self-Consistent Treatment of the Vacancy Mechanism for Metallic Diffusion,” Physical Review, 61, p. 325. 121 [48] Jenssen, L., and Warmoeskerken, M., 1987, “Transport Phenomena Data Com- panion,” Edward Arnold Ltd. [49] Jones, S. F., Evans, G. M., and Galvin, K. P., 1999, “Bubble Nucleation from Gas Cavities .. a Review,” Adv. Colloid Interf. Sci., 80, p. 27. [50] Jost, W., 1960, “Diffusion in Solids, Liquids, Gases,” third printing with adden- dum, Academic Press. [51] Kantrowitz, A., 1951, “Nucleation in Very Rapid Vapor Expansions,” J. Chem. Phys, 19, p. 1097. [52] Katz, J. L., and Blander, M., 1973, “Condensation and Boiling: Corrections to Homogeneous Nucleation Theory for Nonideal Gases,” J. Colloid Interf. Sci., 42, p. 496. [53] Kirwood, J. G., and Buff, F. P., 1949, “The Statistical Mechanical Theory of Surface Tension,” J. Chem. Phys, 17, p. 338. [54] Krichevsky, I. R., and Kasarnovsky, J. S., 1935, “Thermodynamical Calculations of Solubilities of Nitrogen and Hydrogen in Water at High Pressures,” J. Am. Chem. Soc., 57, p. 2168. [55] Krieger, I. M., Mulholland, G. W., and Dickey, C. S., 1967, “Diffusion Coeffi- cients for Cases in Liquids from the Rates of Solution of Small Gas Bubbles,” J. Phys. Chem., 71, p. 1123. [56] Landau, L. D., and Lifshitz, E. M., 1958, “Statistical Physics,” Addison-Wesley Pub. Co., p. 275. [57] Landau, L. D., and Lifshitz, E. M., 1987, “Fluid Mechanics,” 2nd edition, Perg- amon Press. [58] LaMer, V., and Pound, G., 1949, “Surface Tension of Small Droplets from Volmer and F lood’s Nucleation Data,” J. Chem. Phys, 17, p. 1337. [59] Liebermann, L., 1957, “Air Bubbles in Water,” J. Appl. Phys, 28, p. 205. [60] Lorenz, J. J ., Mikic, B. B., and Rohsenow, W. M., 1974, Proc. Fifth Int. Heat Transfer Conf., 4, p. 35 [61] Lothe, J ., and Pound, G. M., 1962, “Reconsiderations of Nucleation Theory,” J. Chem. Phys, 36, p. 2080. [62] Lubetkin, S. D., 1988, “The Nucleation of Bubbles in Supersaturaed Solutions,” J. Colloid Interf. Sci., 26, p. 611. [63] Lubetkin, S. D., 1995, “The Fundamentals of Bubble Evolution,” Chem. Soc. Rev., 24, p. 243. 122 [64] Lubetkin, S. D., and Akhtar, M., 1996, “The Variation of Surface Tension and Contact Angle under Applied Pressure of Dissolved Gases, and the Effects of these Changes on the Rate of Bubble Nucleation,” J. Colloid Interf. Sci., 180, p. 43. [65] Massoudi, R., and King, A. D., 1974, “Effect of Pressure on the Surface Tension of Water. Adsorption of Low Molecular Weight Gases on Water at 25 °C,” J. Phys. Chem., 78, p. 2262. [66] Merte, H., 1973, “Condensation Heat Transfer,” Adv. Heat Transfer, 9, p. 181. [67] Oliver, J. F ., 1977, “Resistance to Spreading of Liquids by Sharp Edges,” J. Colloid Interf. Sci., 59, p. 568. [68] Oriani, R. A., 1962, “Emendations to Nucleation Theory and the Homogeneous Nucleation of Water from the Vapor,” J. Chem. Phys, 58, p. 2082. [69] Ozisik, N., 1980, “Heat Conduction,” John Wiley & Sons, Inc. [70] Probstein, R. F., 1951, “Time Lag in the Self-Nucleation of a Supersaturated Vapor,” J. Chem. Phys, 19, p. 619. [71] Reid, R., Prausnitz, J ., and Sherwood, T., 1977, “The Properties of Gases and Liquids,” third edition, McGraw-Hill Book Co. [72] Reiss, H., 1952, “The Statistical Mechanical Theory of Irreversible Condensation. I,” J. Chem. Phys, 20, p. 1216. [73] Rubin, M., and Noyes, R., 1987, “Measurements of Critical Supersaturation for Homogeneous Nucleation of Bubbles,” J. Phys. Chem., 91, p. 4193. [74] Rubin, M., and Noyes, R., 1992, “Thresholds for Nucleation of Bubbles of N2 in Various Solvents,” J. Phys. Chem., 96, p. 993. [75] Scriven, L. E., 1959, “On the Dynamics of Phase Growth,” Chem. Eng. Sci., 10, p. 1. [76] Sherwood, T. K., Pigford, R. L., and Wilke, C. R., 1975, “Mass Transfer,” McGraw-Hill, Inc. [77] Skripov, V. P., 1974, “Metastable Liquids,” John Wiley & Sons, Inc. [78] Smoluchowski, R., Mayer, J. E., and Weyl, W. A., 1951, “Phase Transformation in Solids”, Symposium held at Cornell University, Aug. 23—26, 1948. Sponsored by the committee on solids division of physical sciences, the national research council. John Wiley & Sons Inc. [79] Spalding, D. B., 1963, “Convective Mass Tiansfer,” h‘chraw-Hill, pp. 139-143. 123 [80] Talanquer, V., and Oxtoby, D., 1995, “Nucleation of Bubbles in Binary Fluids,” J. Chem. Phys, 102, p. 2156. [81] Tucker, A., and Ward, C. A., 1975, “Critical State of Bubbles in Liquid-Gas Solution,” J. Appl. Phys, 46, p. 4801. [82] Turnbull, D., and Fisher, J. C., 1949, “Rate of Nucleation in Condensed Sys- tems,” J. Chem. Phys, 17, p. 71. [83] Volmer, M., and Weber, A., 1926, “Nucleus Formation in Supersaturated Sys- tems,” Z. Phys. Chem., 119, p. 277. [84] Wakeshima, H., 1954, “Time Lag in the Self-Nucleation,” J. Chem. Phys, 22, p. 1614. [85] Ward, C. A., Balakrishnan, A., and Hooper, F. C., 1970, “On the Thermody- namics of Nucleation in Weak Gas-Liquid Solutions,” Trans. ASME, 92, p. 695. [86] Ward, C. A., Tikuisis, P., and Venter, R. D., 1982, “Stability of Bubbles in a Closed Volume of Liquid-Gas Solution,” J. Appl. Phys, 53, p. 6076. [87] Ward, C. A., Johnson, W. R., Venter, R. D., and Ho, S., 1983, “Heterogeneous Bubble Nucleation and Conditions for Growth in a Liquid-Gas System of Con- stant Mass and Volume,” J. Appl. Phys, 54, p. 1833. [88] Ward, C. A., and Levart, E., 1984, “Conditions for Stability of Bubble Nuclei in Solid Surfaces Contacting a Liquid-Gas Solution,” J. Appl. Phys, 56, p. 491. [89] Wilt, P. M., 1985, “Nucleation Rates and Bubble Stability in Water-Carbon Dioxide Solutions,” J. Colloid Interf. Sci., 112, p. 530. [90] Wise, D. L., and Houghton, G., 1966, “The Diffusion Coefficients of Ten Soluble Gases in Water at 10-60°C,” Chem. Eng. Sci., 21, pp. 999-1010. [91] Worden, R. M., and Bredwell, M. D., 1998, “Mass-Transfer Properties of Mi- crobubbles. 2. Analysis Using a Dynamic Model,” Biotechnol. Prog., 14, p. 39. 124