Sei sulla pagina 1di 247

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UM I


films the text directly from the original or copy submitted. Thus, some
thesis and dissertation copies are in typewriter face, while others may
be from any type of computer printer.

The quality of this reproduction is dependent upon the quality o f the


copy subm itted. B roken or indistinct print, colored o r p o o r quality
illustrations and photographs, print bleedthrough, substandard margins,
and im proper alignment can adversely afreet reproduction.

In the unlikely event th a t th e author did n o t send U M I a com plete


m anuscript and there are missing pages, these will be noted. Also, if
unauthorized copyright m aterial had to be removed, a note will indicate
the deletion.

O versize m aterials (e.g., maps, drawings, charts) are rep ro d u ced by


sectioning the original, beginning at the u pper left-hand corn er and
continuing from left to right in equal sections with small overlaps. Each
original is also p h o to g ra p h e d in one exposure a n d is in clu d ed in
reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. H igher quality 6" x 9" black and white
photographic prints are available for any photographs o r illustrations
appearing in this copy for an additional charge. Contact U M I directly
to order.

U n iversity M icrofilm s International


A B ell & H ow ell Inform ation C o m p a n y
3 0 0 N orth Z e e b R o a d , A nn Arbor, Ml 4 8 1 0 6 - 1 3 4 6 U S A
3 1 3 /7 6 1 - 4 7 0 0 8 0 0 /5 2 1 - 0 6 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
O rd e r N u m b e r 9409744

Vapor bubble dynam ics in m icrogravity

Lee, Ho Sung, Ph.D .


The University of Michigan, 1993

UMI
300 N. Zeeb Rd.
Ann Arbor, MI 48106

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
VAPOR BUBBLE DYNAMICS IN MICROGRAVITY

by

Ho Sung Lee

A dissertation submitted in partial fulfillment


of the requirements for the degree of
Doctor of Philosophy
(Mechanical Engineering)
in The University of Michigan
1993

Doctoral Committee:
Professor Herman Merte,Jr., Chairman
Professor Luis P. Bernal
Professor Ross L. Judd, McMaster University
Professor Robert B. Keller
Professor Gretar Tryggvason

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
To God

ii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
ACKNOWLEDGMENTS

It is indeed a pleasure to express my deep appreciation to Professor Herman

Merte, Jr. who, as the Chairman of my committee, provided his profound guidance

and encouragement during my study at The University of Michigan. Thanks are

extended to Professors R. B. Keller, R. L. Judd, L. P. Bernal and G. Tryggvason for

serving on the committee, and for their constructive comments and suggestions.

I would like to thank the fellow doctoral candidates and friends while at the

University of Michigan, who were all a source of great encouragement: Kevin Kirk,

Jamie Ervin, Matthew Brusstar, Jeff Hood, Andrey Vinarov, and Longhu Li.

Significant help and encouragement of Kent Pruss in the design and assembly of the

experimental apparatus is gratefully acknowledged. Especial recognition is also

given to Dr. Youngil Kim for his faithful support during the qualifying exam.

I would like to extend my special thanks to the National Aeronautics and

Space Administration (NASA), including the paticular group at the Lewis Research

Center in Cleveland, Ohio for funding the research and providing facilities for the

experiments. I also wish to thank the University of Michigan library staff for their

help in obtaining literature, which made it possible to complete my work efficiendy.

Thanks are extended to the Mechanical Engineering Department of the University of

Michigan for considering my industrial experience in admitting me to the program.

Finally, I sincerely thank my beloved wife, Young-Ae for her extreme self-

sacrifice and support, and our beautiful daughter, Yu-Jin, for her patience, support

and the joy she brought me during my research.

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
TABLE OF CONTENTS

DEDICATION................................................................................................ii
ACKNOWLEDGEMENTS...........................................................................iii
LIST OF FIGURES........................................................................................ vii
LIST OF TABLES......................................................................................... xvi
LIST OF APPENDICES............................................... xvii
LIST OF SYMBOLS................................................. xviii
CHAPTER
1. INTRODUCTION................................................................................................1
2. SPHERICAL BUBBLE GROWTH IN INITIAL UNIFORM SUPERHEAT
................................................................................................................................... 5

2.1 Literature Survey............................................................................................. 6

2.2 Previous Analytical Solutions..........................................................................9

2.3 Formulation of Governing Equations.............................................................. 16

2.4 Numerical Solution............................................................................................19

2.5 Comparisons with Previous W orks................................................................. 24

2.5.1 Sodium.............................................................................................. 25

2.5.2 W ater.................................................................................................26

2.5.3 R-113.................................................................................................29

2.6 Discussion......................................................................................................... 30

iv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.6.1 Water 31

2.6.2 R-113................................................................................................ 35

2.6.3 Sodium.............................................................................................. 38

2.6.4 Comparisons between Water, R-113 and Sodium.......................... 39

3. VAPOR BUBBLE DYNAMICS IN NON-UNIFORM TEMPERATURE


FIELDS..................................................................................................................90
3.1 Literature Survey............................................................................................. 91

3.1.1 Inertia Controlled Bubble Growth and Collapse............................ 95

3.1.2 Heat Transfer Controlled Bubble Growth and Collapse................ 96

3.1.3 Vapor Bubble Collapse.................................................................... 98

3.1.3.1 Liquid inertia controlled collapse.....................................99

3.1.3.2 Heat transfer controlled collapse...................................... 100

3.1.4 Vapor Bubble Growth on a Heater Surface.................................... 101

3.2 Modeling........................................................................................................... 103

3.2.1 Initial Uniform Superheat Model for Spherical Bubble Growth 104

3.2.2 Initial Non-uniform Superheat Model..............................................104

3.2.2.1 Spherical Bubble Model................................................... 104

32.2.2 Hemispherical Bubble Model............................................105

3.3 Governing Equations........................................................................................ 106

3.4 Numerical Solution........................................................................................... 109

3.5 Vapor Bubble Dynamics in Subcooled Liquids...............................................109

3.5.1 Bubble Collapse.................................................................................109

3.5.2 Bubble Growth and Collapse............................................................ 110

3.5.3 Bubble departure mechanism............................................................ 112

3.6 Experiments in Microgravity with Saturated and Subcooled R-l 13..............112

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.6.1 Experimental Apparatus................................................................... 112

3.6.2 Experimental Results and Discussions............................................ 114

3.6.2.1 Energy Input to the Liquid................................................114

3.6.2.2 Initial Temperature Distribution....................................... 115

3.6.2.3 Bubble Dynamics...............................................................117

4. ROLE OF INSTABILITIES IN POOL BOILING BUBBLE GROWTH IN


MICROGRAVITY...............................................................................................140
4.1 Literature Survey............................................................................................141

4.2 Experimental Results...................................................................................... 144

4.3 Instability Considerations................................................................................151

4.3.1 Rayleigh-Taylor Instability...............................................................152

4.3.2 Landau Instability............................................................................. 154

4.3.3 Instability Growth in Terms of Heat Conduction........................... 156

4.4 Increased Evaporation Surface Area...............................................................158

5. CONCLUSIONS...................................................................................................178
5.1 Initial Uniform Superheat M odel................................................................... 178

5.2 Initial Non-uniform Superheat M odel........................................................... 179

5.3 Role of Instability in Pool Boiling..................................................................180

APPENDICES............................................................................................................... 182
BIBLIOGRAPHY.........................................................................................................216

vi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF FIGURES

Figures

2.1. Photographs by Bohrer (1973) for vapor bubble growth in R-l 13, demonstrating the

sphericity of the bubble............................................................................................. 42

2.2. (a) Vapor bubble at the critical size radius, (b) Bubble growth in a uniformly

superheated liquid...................................................................................................... 43

2.3. Schematic diagram of coordinate transformation...................................................... 44

2.4. Grid point and control volume notation.....................................................................44

2.5. Comparison of present results with the numerical work of Dalle Donne and Ferranti

(1975). Radius versus time for sodium.....................................................................45

2.6. Comparison of present results with the numerical work of Dalle Donne and Ferranti

(1975). Liquid-vapor interface velocity versus time for sodium............................46

2.7. Comparison of present results with the numerical work of Dalle Donne and Ferranti

(1975). Liquid-vapor interface temperature versus time for sodium..................... 47

2.8. Comparison of present results with the numerical work of Dalle Donne and Ferranti

(1975). Effect of magnitude of temperature disturbance on bubble growth for

sodium....................................................................................................................... 48

2.9. Comparison of computations with the measurements of Dergarabedian (1953).

Bubble radius versus time for water......................................................................... 49

2.10. Comparison of computations with the measurements of Kosky (1968). Bubble

radius versus time for water......................................................................................50

2.11. Comparison of computations with the measurements of Dergarabedian (1960).

Bubble radius versus time for water with low initial liquid superheats................... 51

vii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.12. Comparison of computations with the measurements of Florschuetz (1969).

Bubble radius versus time for water. An arbitrary shift of 6.5 ms is made in the time

scale............................................................................................................................ 52

2.13. Comparison of present numerical calculations with measurements of Lien (1969)

over a wide range of pressure. Bubble radius versus time for water...................... 53

2.14. Comparison of computations with measurements of Bohrer (1973) for R -l 13 at

low pressures and large liquid superheats................................................................. 54

2.15. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for water at atmospheric pressure. Radius and liquid-vapor interface

velocity.......................................................................................................................55

2.16. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for water at atmospheric pressure. Radius and liquid-vapor interface

temperature................................................................................................................ 56

2.17. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for water at atmospheric pressure. Liquid-vapor interface acceleration, for a

low initial superheat of 3.1 C....................................................................................57

2.18. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for water at atmospheric pressure. Liquid-vapor interface acceleration, for a

high initial superheat of 36 C................................................................................... 58

2.19. Example of characterization of the three domains of vapor bubble growth. Bubble

radius and liquid-vapor interface acceleration vs. time, for water at atmospheric

pressure and uniform initial superheat of 3.1 C ...................................................... 59

2.20. Comparison of present numerical calculations with various analytical solutions.

Water at atmospheric pressure uniformly superheated to 36 C, Ja=l 10.21..........60

2.21. Comparison of present numerical calculations with various analytical solutions.

Water at atmospheric pressure uniformly superheated to 36 C, Ja=l 10.21. Vapor


*
bubble radius versus time........................................................................................... 61

VIII

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.22. Comparison of the present numerical calculations with various analytical solutions

and the measurements of Lien (1969) for very low system pressure (P=1.26 kPa)

with water. Vapor bubble radius and liquid-vapor interface temperature.............. 62

2.23. Liquid-vapor interface velocity and acceleration from the present numerical

calculations for conditions of Figure 2.22................................................................. 63

2.24. Comparison of the present numerical calculations with various analytical solutions

for relatively high pressure (P=l atm.=101.325 kPa) with water. Vapor bubble

radius and liquid-vapor interface temperature.......................................................... 64

2.25. Liquid-vapor interface velocity and acceleration from the present numerical

calculations for conditions of Figure 2.24.................................................................65

2.26. Comparison of present numerical calculations with various analytical solutions.

Water at atmospheric pressure uniformly superheated to 0.8 C, Ja=2.38. Vapor

bubble radius versus time...........................................................................................66

2.27. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for R -l 13 at atmospheric pressure. Radius and liquid-vapor interface

velocity.......................................................................................................................67

2.28 (a) & (b). a) Amplification of the embryo bubble radius and velocity in Figure 2.27

for the early period of growth, b) Comparison of the time varying surfacetension

and pressure terms in Equation (2.22) with the liquid-vapor interface velocity..... 68

2.29. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for R -l 13 at atmospheric pressure. Radius and liquid-vapor interface

temperature................................................................................................................ 69

2.30. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for R-l 13 at atmospheric pressure. Radius and liquid-vapor interface

acceleration for a low initial superheat of 3.1 C ....................................................70

IX

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.31. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for R-l 13 at atmospheric pressure. Radius and liquid-vapor interface

acceleration for a high initial superheat of 36 C .....................................................71

2.32. Sensitivity of early R-113 vapor bubble growth rates to bulk liquid superheat at low

levels of superheat..................................................................................................... 72

2.33. Comparison of present numerical calculations with various analytical solutions and

one set of measurements. R-l 13 at low pressure of 8.5 kPa (1.23 psia) with initially

uniform superheat of 34 C. Ja = 396....................................................................... 73

2.34. Comparison of present numerical calculations with various analytical solutions. R-

113 at atmospheric pressure with initially uniform superheat of 36 C .................. 74

2.35. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for sodium at atmospheric pressure. Radius and liquid-vapor interface

velocity...................................................................................................................... 75

2.36. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for sodium at atmospheric pressure. Radius and liquid-vapor interface

temperature................................................................................................................76

2.37. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for sodium at atmospheric pressure. Radius and liquid-vapor interface

acceleration for a low initial superheat of 3.1 C .....................................................77

2.38. Influence of magnitude of initial temperature disturbance on the early vapor bubble

growth for sodium at atmospheric pressure. Radius and liquid-vapor interface

acceleration for a high initial superheat of 36 C .....................................................78

2.39. Comparison of present numerical calculations with various analytical solutions.

Sodium at atmospheric pressure uniformly superheated to 36 C, Ja=32.28..........79

2.40. Comparison of present numerical caculation with analytical solution of Mikic at al

(1970). Vapor bubble radius versus time for sodium............................................ 80

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.41. Comparisons of bubble radius and liquid-vapor interface velocity vs. time for water,

R-113 and sodium all at atmospheric pressure and a low superheat of 3.1 C 81

2.42. Comparisons of bubble radius and liquid-vapor interface velocity vs. time for water,

R-113 and sodium all at atmospheric pressure and a high superheat of 36 C 82

2.43. Comparisons of bubble radius and liquid-vapor interface acceleration vs. time for

water, R-113 and sodium all at atmospheric pressure and a low superheat of 3.1 C

....................................................................................................................................83

2.44. Comparisons of bubble radius and liquid-vapor interface acceleration vs. time for

water, R-113 and sodium all at atmospheric pressure and a high superheat of 36 C

...................................................................................................................................84
2.45. Critical radius versus superheat (Thulk - Tsat) for sodium, R-l 13 and water,

computed by the present model, as a function of pressure..................................... 85


2.46. Delay time versus superheat (Tbulk - Tsat) for sodium, R-113 and water, computed

by the present model, as a function of pressure...................................................... 86


2.47. Maximum liquid-vapor interface velocity versus superheat (Tbulk - Tsal) for sodium,

R-113 and water, computed by the present model, as a function of pressure 87


2.48. Maximum liquid-vapor interface acceleration versus superheat (Thulk - Tsal) for

sodium, R-113 and water, computed by the present model, as a function of pressure

88
2.49. Bubble radius at an arbitrary time of 0.01 seconds versus superheat (Tbulk - Tsat) for

sodium, R-l 13 and water, computed by the present model, as a function of pressure

...................................................................................................................................89

3.1. Illustration of variability of bubble dynamics in pool boiling during a given


experiment. Water, ATsub =42.8C , ATwsap =32.7C , Ellion(1953)............... 118

3.2. Illustration of variability of bubble dynamics in pool boiling during a given


experiment. Water, ATsub =74.5C , AJwsup =34.4C , Ellion(1953)............... 119

xi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.3. Comparison of Equations (3.4) and (3.9) with normalized data of Figure 3.1.
ATsub = 42.8C, ATwsup = 32.7C, Ellion (1953)................................................... 120

3.4. Comparison of Equations (3.4) and (3.9) with normalized data of Figure 3.2.
ATsub = 74.5 C, Arwsup = 34.4 C, Ellion (1953)................................................... 120

3.5. Numerical solutions of Florschuetz and Chao (1965) for vapor bubble collapse given

by Equations (3.11) and (3.12), for three different values of the dimensionless

parameter B, together with the inertia controlled collapse.......................................121

3.6. Numerical solutions of Florschuetz and Chao (1965) for vapor bubble collapse given

by Equations (3.13) and (3.14)..................................................................................122

3.7. Initial uniform superheat model................................................................................... 123

3.8. Initial non-uniform superheat model........................................................................... 123

3.9. Schematic diagram of a realistic hemispherical vapor bubble growing on a heater

surface..........................................................................................................................124

3.10. Schematic diagram of an ideal hemispherical bubble growing on a heater surface for

the initial non-uniform superheat m odel....................................................................124

3.11. Schematic diagram of an ideal hemispherical bubble growing on a heater surface for

the initial uniform superheat m odel........................................................................... 125

3.12. Temperature field around a downward-facing R-113 vapor bubble by means of

holographic interferometery (Straub et al. (1990)).................................................. 125

3.13. Comparison of experimental data of Florschuetz and Chao (1965) for water with

computations using a heat transfer controlled collapse and the present numerical

method........................................................................................................................ 126

3.14. Vapor bubble growth and collapse with water at atmospheric pressure. Comparison

between measurements and numerical computations carried out using estimated

parameters................................................................................................................... 127

3.15. Computed temperature distributions around a growing bubble at various times, for

the estimated parameters of Figure 3.14....................................................................128

xii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.16. Early bubble growth behavior under microgravity with different levels of heat flux

and subcooling........................................................................................................... 129

3.17. Measured mean heater surface temperature and system pressure for Test

PBMT1129.820, Figure 3.16 (b ).............................................................................. 130

3.18. Comparisons of several parameters influencing growth dynamics for the

experiments in Figure 3.16......................................................................................... 131

3.19. Transient liquid temperature distribution surrounding a spherical vapor bubble

growing from the critical size, with an initial non-uniform liquid temperature

distribution..................................................................................................................132

3.20. Transient liquid temperature distribution surrounding a spherical vapor bubble

growing from the critical size with an initial uniform liquid temperature distribution

.................................................................................................................................... 133

3.21. Comparisons of measured bubble growth data with the initially uniform and non-

uniform superheat spherical vapor bubble growth models (Figure 3.16 (a)).......... 134

3.22. Comparisons of measured bubble growth data with the initially uniform and non-

uniform superheat spherical vapor bubble growth models (Figure 3.16 (b)).......... 135

3.23. Comparisons of measured bubble growth data with the initially uniform and non-

uniform superheat spherical vapor bubble growth models (Figure 3.16 (c)).......... 136

3.24. Comparison of the liquid-vapor interface velocity between the initially uniform and

nonuniform superheat models....................................................................................137

3.25. Comparison of the liquid-vapor interface acceleration between the initially uniform

and nonuniform superheat models............................................................................. 138

3.26. Comparison of the liquid-vapor interface temperature between the initially uniform

and nonuniform superheat models............................................................................. 139

4.1. Dynamic vapor bubble growth of R -113 in microgravity at NASALewis Research

Center 5 second drop tower. Experimental Data Code PBMT1102.800................162

4.2. Enlargement of Figure 4.1 (E) for Experimental Data Code PBMT1102.800........ 164

xiii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4.3. Measured mean heater surface temperature and system pressure for Figure 4.1 (A)-

(H).............................................................................................................................. 165

4.4. Comparisons of early vapor bubble growth behavior under microgravity with

different levels of heat flux and subcooling............................................................... 166

4.5. Comparisons of experimental bubble growth data with several spherical vapor

bubble growth models. Experimental conditions correspond to Figure 4.4 (a).... 167

4.6. Comparisons of experimental bubble growth data with several spherical vapor

bubble growth models. Experimental conditions correspond to Figure 4.4 (c).... 168

4.7. Transient liquid temperature distribution surrounding a spherical vapor bubble

growing from the critical size with an initially non-uniform liquid temperature

distribution..................................................................................................................169

4.8. Photographs of n-butane nucleating at the superheat limit. (Shepherd and

Sturtevant, 1982)........................................................................................................170

4.9. Comparisons between extremely dynamic bubble growth experiments, a momentum

controlled model, and approximate and numerical solutions of the energy and

momentum equations for n-butane............................................................................171

4.10. Liquid-vapor interface velocity and acceleration vs. time for n-butane. Numerical

computations. Physical conditions identical to Figure 4.9....................................... 172

4.11. Effect of system pressure on Landau instability criteria for n-butane. The density

ratio ( a ) is assumed constant at bulk temperature of 105 C.................................. 173

4.12. Velocity potential simulation of a temperature field in the liquid at the liquid-vapor

interface. The velocity potential was superimposed on x-y ax is............................ 174

4.13. Modeling of a perturbed liquid-vapor interface...................................................... 174

4.14. Ratio of perturbed to smooth surface area of a spherical bubble as a function of the

perturbation amplitude to wave length ratio............................................................. 175

xiv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4.15. Comparison of experimental data for n-butane with the modified prediction of

Mikic et al. (1970), Equation (4.23), which includes the effect of the increased area

due to the perturbation.............................................................................................. 176

4.16. Comparison of experimental data for R-113 with the modified prediction of Mikic

et al.(1970), Equation (4.23).....................................................................................177


B -l Control Volume at liquid-vapor interface and velocity at a distance (r).................. 187

B-2 Force balance at Liquid-vapor interface.....................................................................187

xv

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF TABLES

Table

2.1 Parameters corresponding to Figure 2.14...............................................

2.2 Thermophysical properties of saturated liquids at atmospheric pressure

xvi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF APPENDICES

Appendix

A Diffusion Controlled Vapor Bubble Growth..........................................................183

B Momentum Equation for Spherical Bubble G row th............................................. 185

C Sonic Velocity of Water Vapor..............................................................................188

D Homogeneous Nucleation Temperature................................................................. 189

E A Fortran Program for Spherical Vapor Bubble Growth and Collapse...............190

xvii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
LIST OF SYMBOLS

Symbol

c specific heat
g gravity
hfs latent heat

c
Ja Jakob number Ja =
V P vtyg j

Jac condensing Jakob number (= -J a above), Ja - IjslL

k thermal conductivity

n spherical harmonic index


P., system pressure
Pv vapor pressure

PR liquid pressure adjacent the bubble wall


q, heat flux to liquid

qT total heat flux imposed

r radial coordinate
R(t) bubble radius
Rc critical radius

t time

t* time at nucleation following onset of heating


Tt liquid-vapor interface temperature

xvui

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
T' wall temperature at nucleation

Tv vapor temperature

T fluid temperature far from the liquid-vapor interface


Tbulk fluid temperature far from the liquid-vapor interface

Tsal saturation temperature corresponding to system pressure

A T !ub ( Tsa, ~ T~ )> bulk liquid subcooling

ATsuP (Tm- Tsat), bulk liquid superheat

A7;sup (7; - Tsal), heater surface sperheat

A r;sup ( r j - Tsat), heater surface superheat at nucleation

AT {Tx - Tsat), bulk liquid superheat

AP Pv -P

Subscripts

i liquid-vapor interface

I liquid inertia controlled

H heat transfer controlled

I liquid

m maximum

w wall

V vapor
00 infinite or far field

sub subcooling

sup superheat

xix

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Superscripts

at nucleation

Greek letters

a thermal diffusivity

a density ratio =
Pi
8 thermal boundary layer

O velocity potential

r ratio of perturbed to smooth area


v'V
r| a dimensionless coordinate

X wavelength
(X dynamic viscosity

p density

c surface tension

v kinematic viscosity of liquid

x dimensionless time

vapor density ratio ( = -


P
Eu
^ p,(R ef)J

Q, nondimensional growth rate

C, amplitude of perturbation

xx

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 1
INTRODUCTION

Vapor bubble dynamics plays an important role in the design of nuclear reactor

coolers and future space power plants, and also in various fields such as cavitation, vapor

explosions, and ultrasonics, and thermal and cryogenic propellant management in space

applications. Bubble dynamics incorporates the bubble growth and/or collapse in initial

uniform or non-uniform temperature fields. The driving force for the bubble growth or

collapse results not only from the phase change, vaporization or condensation, taking

place at liquid-vapor interface, but also from the effect of liquid inertia surrounding the

bubble. The mechanisms of the bubble dynamics are considerably complex such that no

appropriate prediction covering the entire range of conditions to be encountered has yet

been found. In order to study the problem of bubble growth and collapse in some detail, a

Fortran program was written to model the spherical vapor bubble dynamics with a

minimum of assumptions.

The present work deals with three basic related phenomena: bubble growth in

initially uniformly superheated liquids, bubble growth and/or collapse in initially non-

uniform temperature fields, and the role of instabilities in pool boiling bubble growth as

became manifested in microgravity. These three problems are discussed in detail in

Chapters 2, 3 and 4 respectively. The first problem is relatively simple to solve

analytically and to deal with experimentally, since it considers the bubble growth only in

initially uniformly superheated liquids. However, the second problem is more complicated

due to the non-uniform temperature fields surrounding the growing bubble, in which not

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2

only bubble growth but collapse can take place. The third problem explores the origins of

the explosive bubble growth observed during pool boiling experiments in microgravity,

ascribed to interfacial instabilities whose origins lie in the very early period of growth.

Vapor bubble growth in an initially uniformly superheated liquid is dealt with in

Chapter 2, and can be considered to be a foundation for Chapters 3 and 4. A vapor

bubble begins its growth from the critical size, whose formation is based on homogeneous

nucleation theory. The nucleus is highly unstable and fluctuates. Mechanically, the vapor

pressure in the nucleus is much higher than the adjacent liquid pressure and is balanced by

the surface tension. A small disturbance is required to initiate the growth from the critical

size when making numerical computations. The effects of the disturbance on the

computations will be discussed below. It is to be noted that the critical radius, frequently

considered to be a theoretical calculation, was demonstrated to be an experimental reality

by Dergarabedian (1960). In order to assess the validity of the computational model

developed here, it is evaluated in light of both previous experiments and numerical works,

and then tested by comparison with available experimental data for two fluids, water and

R-113, and by varying properties over a wide range of liquid superheat and system

pressures. The testing was extended to the previous numerical work of Dalle Donne and

Ferranti (1975) for sodium. As a result, the good agreement between the present and

previous works confirms the validity of the present numerical model. The present initially

uniformly superheated liquid model will be referred to here as the uniform model.

In earth gravity, measurements of vapor bubble growth and collapse as a result of

heat transfer are quite non-reproducible. One of the reasons for this non-reproducibility is

the randomness of the temperature distribution in the vicinity of the growing vapor bubble.

The randomness results from the turbulence associated with the natural and/or forced

convection, and from the liquid agitation produced by adjacent or prior bubbles. In order

to eliminate the randomness arising from natural convection, transient pool boiling

experiments were conducted under microgravity conditions as part of the present research

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3

using the 132 m drop tower at the NASA Lewis Research Center, with transient

measurements of the mean heater surface temperature and far-field pressure along with

photographs of the vapor bubbles taken at 400 fps. Results and associated analysis from

these experiments are presented here in Chapters 3 and 4. The absence of natural

convection and agitation from adjacent bubbles permits the determination of the precise

initial temperature distribution at nucleation. The present experiments reveal two

categories of bubble growth: one with bubbles having a smooth surface, which produces

moderate growth rates, is discussed in Chapter 3; the other with bubbles having a

roughened surface, which produces extremely large growth rates which might be classified

as violent, is discussed in Chapter 4.

Vapor bubble growth and/or collapse in initially non-uniform liquid temperature

fields is dealt with in Chapter 3. The model is identical to the uniform model except for

the initial condition, in which the initial superheat surrounding the initial vapor bubble is

non-uniform, and will be referred to as the non-uniform model here. The non-uniform

model not only predicts the bubble growth but also the transient temperature distributions

surrounding the growing bubble, such that the bubble rebound after the first collapse and

the bubble spring-off phenomena observed near a heater surface can be interpreted

quantitatively. Bubble dynamics in pool boiling taking place on a flat heater surface is

modeled using both the uniform and non-uniform models by mathematically considering

half of the spherical bubble model, and is called a hemispherical bubble model. As

presently posed, this model is not yet able to provide an accurate prediction of the bubble

growth for transient pool boiling on a flat heater surface, but rather gives upper and lower

bounds on the growth. The model is tested by comparing computed results with transient

bubble growth measurements obtained from the pool boiling experiments conducted in

microgravity, which provides a known initial temperature distribution at nucleation, in the

absence of any liquid disturbances. It was demonstrated that the concept of upper and

lower bounds for the bubble growth has merit.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4

The role of instabilities observed in pool boiling bubble growth taking place in

microgravity is dealt with in Chapter 4. Extremely large vapor bubble growths at

superheats far below the homogeneous nucleation superheat limit were measured in pool

boiling experiments taking place in microgravity. Photographic evidence revealed a

roughening of the bubble liquid-vapor interface as growth took place, which is believed to

be a consequence of an interface instability. The growth of the instability is explained in

terms of heat conduction, where the heat transfer to the liquid-vapor interface increases at

the trough of the interface perturbation while decreasing at the crest. This can give rise to

a differential vapor recoil at the interface, which further increases the liquid-vapor

interfacial surface area, increasing the evaporating surface area and the bubble growth

rate. The rapid bubble growth is modeled by introducing a simple expression in terms of a

circular function to account for the increase in surface evaporating area. The resulting

factor is found to be a function of the ratio of the amplitude to the wave length. The onset

of the interface roughening can be considered to originate from either a Rayleigh-Taylor

or Landau instability; however, neither of these current instability models provide the

correct quantitative physical descriptions.

The conclusions obtained in Chapters 2, 3 and 4 will be summarized in Chapter 5.

These three Chapters are, of course, closely related to one another, but each also has its

own unique characteristic independence. The present study can be described as having

two basic objectives: the dynamic behavior of vapor bubble growth during the very early

stages, when such bubbles usually are beyond the visible or measurable size, and the

dynamics of vapor bubbles in initially non-uniform temperature fields. The numerical

study of the bubble dynamics in the early stages is particularly helpful in interpreting the

instability taking place on the vapor bubble surface in pool boiling in microgravity, while

the transient temperature distributions around a growing bubble provided by the numerical

model is essential for describing the bubble dynamics in initially non-uniform temperature

fields.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 2
SPHERICAL BUBBLE GROWTH IN INITIAL UNIFORM
SUPERHEAT

The problem of spherical vapor bubble growth in an initially uniformly superheated

liquid is considered. A review of earlier analytical and numerical models is made, followed

by the development of a model whose growth is initiated from a critical size vapor bubble

by providing a small temperature disturbance at the liquid-vapor interface. Momentum,

energy, and surface tension are incorporated simultaneously. The momentum equation is

solved by the Runge-Kutta method to determine the locus of the bubble interface, which

provides the boundary condition for the energy equation. The Landau coordinate

transformation is adopted to immobilize the moving boundary. After the transformation,

the energy equation is solved using the Thomas algorithm (TDMA).

The solution with the present method is compared with various experimental,

numerical, and analytical results. Excellent comparison is demonstrated with the one

existing experimental work with water in which the measurements were initiated from the

thermodynamic critical size. Good comparisons also exist with measurements made

during the growth phase for R-l 13, and for water over a wide range of system pressure.

The results of the present model conform well with physical reality, and demonstrate clear

distinctions between the vapor-bubble growth domains governed by surface tension, liquid

momentum, and thermal diffusion. Although the computed liquid-vapor interface

acceleration appears to be of significance for only a very short period of time, it serves to

clearly delineate these three domains.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
6

2.1 Literature Survey

Bubble growth in superheated liquids has been investigated since early in the

twentieth century because it plays an important role in submarine explosions, the design of

nuclear reactor coolers, rocket engines, and space power plants, and in understanding

vapor explosions associated with the transportation of liquefied natural gas. Lord

Rayleigh (1917) first formulated the equation of motion for a growing spherical bubble,

later referred to as the inertia controlled bubble growth. In contrast to this, Bosnjakovic

(1930) proposed a model for thermally controlled bubble growth, which found

experimental support in the work of Jakob (1932). Plesset and Zwick (1952) and (1954)

suggested an asymptotic analytic solution, with an assumption of a thin thermal boundary

layer, which was in good agreement with the data of Dergarabedian (1953) for moderate

superheats, but compared poorly with the data of Dergarabedian (1960) for low

superheats. The results presented in this latter work demonstrated that the vapor bubble

growth was initiated from an observed critical size, which is established by the initial liquid

superheat. The asymptotic solution of Plesset and Zwick (1952) and (1954) considers that

momentum no longer plays an important role after the early stage of bubble growth, and

that thermal diffusion dominates the later stage of growth.

The analyses of Forster and Zuber (1954), Birkhoff et al. (1958), and Scriven

(1959) were basically in agreement with that of Plesset and Zwick (1954), and with the

experiments of Dergarabedian (1953). Florschuetz et al. (1969) reported experiments for

moderate superheats that were in good agreement with Scriven (1959). Hooper and

Abdelmessih (1966) were the first to conduct experiments with high superheats, and their

data did not correspond to the asymptotic solution of Plesset and Zwick (1954). Kosky

(1968), citing the possibility for misinterpretation in the measurement of the system

pressure in the work of Hooper and Abdelmessih (1966), subsequently performed a similar

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
7

experiment with high superheats. These results were in good agreement with the analysis

of Plesset and Zwick (1954). Lien (1969) performed experiments at low pressures and

demonstrated that liquid inertia dominates the bubble growth process at low pressures, on

the order of 1.26kPa (0.18 psia) for water. Mikic et al. (1970) formulated a single

approximate equation for the bubble growth dynamics, covering both inertia and diffusion

controlled limits, which provided good agreement with the measurements o f Lien and

Griffith (1969).

Dalle Donne and Ferranti (1975) were the first to solve both the momentum and

the energy equations numerically for the spherical bubble growth problem with sodium .

Their computations with water are in good agreement with the experiments o f Kosky

(1968), but not with those of Hooper and Abdelmessih (1966). Prosperetti and Plesset

(1978) solved the same type of bubble growth problem exactly, but the solution could not

be expressed in closed form. Their results compared well with those of Dalle Donne and

Ferranti (1975), except at low levels of liquid superheat. Shepherd and Sturtevant (1982)

conducted experiments at high superheats for Butane near the computed homogeneous

nucleation superheat, referred to as the superheat limit, and compared the results with the

analysis of Mikic et al. (1970). A discrepancy existed in this case between the measured

growth rates and the prediction of the classical diffusion limited theory, and was attributed

to the observed surface instability which produces significant increases in the liquid-vapor

surface area of the bubble.

The literature cited above covers liquid superheat levels from low to the

homogeneous nucleation superheat limit. It should be noted that the majority of the

asymptotic analytic solutions predict neither the early stages of bubble growth for any

superheat, nor the complete stages of bubble growth for low superheats. In order to

examine the early stages of vapor bubble growth, along with the complete vapor bubble

growth for all levels of liquid superheat, the governing equations for spherical vapor

bubble growth in initially uniformly superheated liquids were solved, using a computer

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
8

program written with highly accurate numerical schemes. This will be described below.

Numerical computations for sodium are compared here with the works of Dalle Donne

and Ferranti (1975), with good agreement between them. Further comparisons are made

with available experimental data through the entire range of liquid superheats. This work

is in good agreement with the measurements of Dergarabedian (1960) for low superheats

of 0.8 ~ 1.0 C; in fair agreement with that of Florschuertz (1969) and Dergarabedian

(1953) for moderate superheats of 2.0 ~ 5.0 C; in good agreement with the

measurements of Kosky (1968) for a high superheat of 36 C; and in good agreement with

the measurements of Lien (1969) for low pressures with moderate and high superheats.

As a result, it is believed that the model to be described below follows physical reality to

an extent not attained heretofore.

Comparisons will also be presented here between the numerical results obtained

here and the asymptotic analytic solutions cited above. As will be demonstrated, the

analyses of Mikic et al. (1970), Scriven (1959) and Plesset and Zwick (1954) provide

good agreement for moderate and high superheats, while the results of Prosperetti and

Plesset (1978) are in disagreement with the model here. None of the analytical solutions

available are in agreement with the numerical solutions for the cases of both low

superheats and early stages of bubble growth.

The model here considers that a vapor bubble begins its growth from the critical

size, whose formation is based on homogeneous nucleation theory. A vapor nucleus

forms in a superheated liquid as a result of molecular interactions. The size and time

required for its formation are functions of the liquid temperature and other properties,

Kwak and Lee (1991). The nucleus is highly unstable and fluctuates. Mechanically, the

vapor pressure in the nucleus is much higher than the adjacent liquid pressure and is

balanced by the surface tension. For analytical purposes the nucleus is assumed to be in a

state of metastable thermodynamic equilibrium, and a small disturbance is required to

initiate the growth from the critical size when making numerical computations. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
9

disturbance may be interpreted as a fluctuation in the state of the nucleus similar to that

required for its formation. The effects of the disturbance on the computations will be

discussed below. The nonequilibrium at the boundary of a growing vapor bubble has

been matter of considerable disagreement. Thermodynamic equilibrium at the liquid-vapor

interface is assumed in the present model. The resulting good predictions obtained

demonstrate that the effect of nonequilibrium is negligible, for the range of liquid

superheat and system pressure used here.

The bubble growth dynamics can be characterized by three domains: surface

tension dominated, momentum dominated, and thermal diffusion dominated growth. It

was predicted that the interface acceleration in the early stages of growth can reach

extremely high values, up to 4 x 107 m/s2 for the case of a high superheat of 36 C with

water, which suggests the possibility that surface instabilities can arise, which in turn can

lead to vapor explosions. This happens so rapidly, within on the order of one millisecond,

that it has been considered to this point that the interface acceleration does not

significantly affect the subsequent growth for moderate superheats. However, this may be

an important parameter for describing the transition in the pattern of growth from

moderate growth rates to the vapor explosions observed with very high liquid superheats.

2.2 Previous Analytical Solutions

Lord Rayleigh (1917) first formulated the equation of motion in terms of the

bubble radius and pressure difference as:

R R + -R 2= (2.1)
2 p,

where AP = PV-P

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
10

The effects of surface tension and viscosity were neglected here. Equation (2.1)

can be solved directly for R(t) for the special case where AP is constant. With the initial

condition R(0)=0, the solution is:

t ( 2 .2)

Equation (2..2) is referred to as the inertia controlled bubble growth, since the

growth is governed solely by the momentum equation, and does not consider the source of

the AP. It is to be noted that R is a linear function of t for this case.

Bosnjakovic (1930) was the first to propose what is now referred to as the

thermally controlled bubble growth, stating that the vaporization process at the bubble

interface is maintained by an energy transfer from the superheated liquid to the interface.

Accordingly, equating the heat transfer in the liquid at the liquid-vapor interface to the

work and internal energy change in a spherical bubble, from the first law of

thermodynamics:

(2.3)

By assuming a one dimensional semi-infinite solid transient conduction process for


the LHS of Equation (2.3), and taking Tv = Tinleiface(= Tt) = Tsat , Equation (2.3) is solved

for R(t) as:

* (0 = 4 aMo *),/2 (2.4)


K

where Ja = ''iZkzZil
Pv*/
Jakob (1932) demonstrated the validity of Equation (2.4) by experiment under

certain conditions, indicating that the inertia of the liquid at a distance far from the bubble

wall is asymptotically negligible.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
11

Plesset and Zwick (1952) solved the one-dimensional energy equation in spherical
coordinates with the initial condition T(r, 0) = T and the boundary condition ,

and suggested a formulation for the temperature at a bubble wall by assuming a thin

thermal boundary layer near the bubble wall with moderate or high liquid superheats. The

energy equation used is:

dT dT f d2T 2dT
+ = a |^ r r + I (2-5)
d t + r 3 r " a dr2 + rdr

where ur is the radial component of the local liquid velocity.

An approximate expression for the temperature at the moving spherical boundary

was given as:

3: = 3' - - ( f ] 1 ,, (2-6)

The temperature at the bubble wall (7]) is thus given in terms of R and R. Plesset

and Zwick (1954) later addeda surface tension term to the Rayleigh Equation (2.1) as:

R R + -R 2 (2.7)
2 p, p,R

Equations (2.6) and (2.7) were related by assuming the equilibrium vapor pressure
to be a linear function of the interface temperature (7J), and an asymptotic solution for

R(t) was obtained as:

R(0 = ( ^ ) Ja{at)in (2.8)

This equation is in good agreement with the experiments of Dergarabedian (1953).

However, it should be emphasized here that since Equation (2.8) is an asymptotic

solution, it is valid only for times large enough for the growth velocity to be much smaller

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
12

than the inertia-controlled velocity and for moderate or high superheats. In order to
ascertain how the temperature (7]) at the moving boundary varies as the bubble grows, the

asymptotic solution Equation (2.8) was substituted into Equation (2.6). The result

showed that the temperature at the moving boundary is independent of time and exactly
equals the saturation temperature (Tsal) corresponding to the system pressure. This is not

unexpected, since Equation (2.8) is the same form as Equation (2.4), where inertia effects

are neglected. The detailed derivation is given in Appendix A. If the solution given by

Equation (2.8) is used for early times as well as for the asymptotic times, it would imply

that the momentum in the early stage of bubble growth is negligible, so that only diffusion

controlled growth takes place, although both the momentum and diffusion equations were

initially considered by Plesset & Zwick (1954). As a result of this, the asymptotic solution

Equation (2.8) is hereafter regarded as a solution for the thermal diffusion controlled

growth.

Birkhoff et al. (1958) produced a result similar to Equation (2.8) by solving the

energy Equation (2.5) without assuming the thin thermal boundary layer, and concluded

that in ordinary boiling water, inertial effects are negligible whenever the bubble is large

enough that the effect of surface tension is negligible. Forster and Zuber (1954) solved

the problem using a method similar to that of Plesset and Zwick (1954). They combined

Equations (2.5) and (2.7) by using the thermodynamic equilibrium vapor pressure

relationship, the Clausius-Clapeyron relation as:

AP = p i T (2.9)
r ra,(vv- v ;)

where vv and v, are the specific volumes of vapor and liquid, respectively.

Forster and Zuber (1954) demonstrated that the coefficients of the inertia terms in

the momentum equation are relatively small (of the order of lO9) and therefore negligible.

Under these assumptions, an asymptotic approximation for R(t) resulted as:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
13

/?(r)=|Xj ja(aif'2
( 2 . 10)

This is also in fair agreement with Plesset and Zwick (1954) and with the

experiments of Dergarabedian (1953). Scriven (1959) solved the bubble growth problem

exactly without assuming a thin thermal boundary layer, but neglected inertia, surface

tension, and viscous terms. For bubbles over the entire range of superheats (AT), the

solution was of the form:


R(t) = 2$Ja(at)in ( 2 . 11)

where P is determined by the following relation:

<t>(e,P) = /a E=1 = 2 p 3exp(p2 + 2 ep 2)jp ;t"2ex p (-* 2- 2 e p 3x~')dx (2.11a)

where

The result of Scriven (1959) is in substantial agreement not only with Plesset and

Zwick (1954), but also with the experiments of Dergarabedian (1953), particularly in the

case for moderate or high superheats. Scriven (1959) postulated that the growth of the

vapor nucleus initially depends very strongly on the surface tension and the inertia of

surrounding fluid, but quickly becomes limited by the rate at which the latent heat of

vaporization can be supplied to the vapor bubble interface.

Mikic et al. (1970) developed a relation applicable over the entire range of the

growth, including both inertia controlled and diffusion controlled, using the Clausius-
Clapeyron relation (2.9) to evaluate dP/dT at the saturation temperature (Tsal). The result

is:

( 2 . 12)

t
where R+= ^----
B2 / A B2 / A 2

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
14

'2AThApvV * no V'2
A= ; B = { Ja2a.
3rra(p, ) V
Vn 'J

Equation (2.12) simplifies to the Rayleigh solution Equation (2.2) for t+ 1 and

to the Plesset and Zwick solution Equation (2.8) for t+ 1, and is in good agreement

with the experiments of Lien and Griffith (1969) made at conditions of low pressure,

where the vapor specific volume is large. Mikic et al. (1970) noted that Equation (2.12)

does not predict the early stage of the bubble growth well when the bubble is near the

critical radius (Rc), because the assumption that thermodynamic equilibrium persists in the

vapor is not justified near the critical size. The critical size is based on thermodynamic

equilibrium, and a simplified form may be derived by combining the Clausius-Clapeyron


Equation (2.9) with the static equilibrium relation for a spherical bubble: AP = 2 a / Rc.

The resulting expression for the radius is:

R = 2aT < (2.13)


hfspvAT

The work of Mikic et al. (1970) was examined carefully here, since it is referred to

frequently by various researchers. As a result of this, a question has arisen about a

relation used by Mikic, taken from the work of Plesset & Zwick (1954):

d R _ 1 B T -T v ^ l ( 1 2 ) m plcl{T'0- T v) ( a A U2
dt 2 tm AT 2\ k J hfgpv U J

Solving Equation (2.14) for R(TV) by integration yields, for R(0)=0:

R(T) = f I" 2 P/C/f c ~ Tv- (a,t)u2 (2.15)


\K J h/gPv

Mikic stated that Equation (2.14), or equivalently Equation (2.15), was taken from

the work of Plesset and Zwick (1954), which is Equation (2.8) here. When comparing
Equations (2.15) and (2.8), however, it is found that Tv was used in Equation (2.15)

instead of Tsat in Equation (2.8). However, it seems that the substitution does not make

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
15

the problem different, but rather more accurate, since Tv takes into account the

temperature variation in the early stage of bubble growth as a result of surface tension
effects, while Tsat is constant.

Prosperetti and Plesset (1978) re-examined the works of Plesset and Zwick (1954)

and Mikic et al. (1970). They demonstrated that the assumptions used in Plesset and

Zwick (1954), the thin thermal boundary layer with moderate or high liquid superheats,

produce valid results. Prosperetti and Plesset (1978) first solved Equations (2.6) and (2.7)

exactly for sodium by combining the momentum and the energy equation, although the

solution was not given explicitly in closed form. Their results were compared with the

numerical work of Dalle Donne (1975). Good agreement was found for moderate and

high superheats, while a substantial disagreement was found for low superheats. This was

to be expected since the assumption of a high superheat was required in deriving Equation

(2.6). Prosperetti and Plesset (1978) improved the single equation provided by Mikic et

al. (1970) slightly, now given as:

(2.16)

where R+= (i2 ; t+= Pp2f (2.17)


Rr

Equation (2.16) covers the entire range of superheats, but is valid only for bubbles

that have grown by about an order of magnitude beyond the critical radius, so that surface

tension effects have become unimportant.

Dalle Donne and Ferranti (1975) were the first to solve Equations (2.5) and (2.7)

numerically without the assumption of a thin thermal boundary layer, and considered the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
16

influence of viscosity as well as surface tension. Equation (2.16) incorporates the

influence of surface tension only, although as stated above it becomes insignificant in the

domain of validity for the solution. When the results of Dalle Donne and Ferranti (1975)

were compared with Equation (2.16), good agreement was found, and for the conditions

used demonstrate that the liquid viscosity has little effect.

The works described above consider bubble growth to take place in uniformly

superheated liquids, without considering the wall effects, and indicate that inertia does not

play an important role beyond the very early stage of bubble growth. The asymptotic

solutions discussed above can not accurately describe the early stages of bubble growth,

and some discrepancies exist in the various solutions for bubble growth with low

superheats. In view of potential applications at that time, the early stages of bubble

growth were considered to be unimportant. However, it develops that the behavior during

the early stages of growth may be quite important when transient pool boiling taking place

in microgravity is considered, particularly for the non-isothermal wall-heating cases.

2.3 Formulation of Governing Equations

A series of the photographs of the spherical vapor bubble growth in a superheated

liquid in Figure 2.1, which was taken by Bohrer (1973), demonstrates the sphericity of the

bubble. Schematic diagrams of the spherical bubble growth problem in a superheated

liquid are illustrated in Figures 2.2 (a) and (b). The bulk liquid and the critical size bubble

are initially isothermal. It is assumed that the bubble growth is initiated from the critical

size bubble. The bubble wall temperature decreases as the bubble grows at the saturation

temperature corresponding to the internal pressure of the bubble. Other assumptions are

necessary in order to solve the problem, and are summarized below.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
17

The major assumptions made in this work are:

a) Bubble remains spherical

b) A uniform pressure and temperature exist inside the vapor bubble

c) The vapor temperature inside the bubble is equal to the liquid-vapor interface
temperature, Tv = Tt

d) The liquid is incompressible

The momentum equation in spherical coordinates is given by:

3u 3u l3 P u 1 3
u = + r \ r l (2.18)
31 dr p dr p ^ r dr v dr J r

The continuity equation with assumption (d) is given by:

^ - ( r 2) = 0 (2.19)
3r
Using the continuity equation (2.19) and integrating Equation (2.18) from the

bubble surface (R) to infinity (), the momentum equation can be written in terms of the

bubble radius as (details of the derivation are in Appendix B):


r r + 1 r 2 = I jl ~
2 EP/ (2.20)

For a spherical bubble, viscosity affects only the boundary condition, so that the

balance of pressure across the liquid-vapor interface gives:

Pv( T ) - P r = ^ +4 ^ (2.21)
K K
where PR is the liquid pressure adjacent the bubble wall.

Inserting Equation (2.21) into Equation (2.20) yields

g+lp.SiZLS,.-**!
2 ep, ep,/?
1 -4 -A
ep,R
(2.22)

The initial conditions are:


7?(0) = Rc and i?(0) = 0 at t = 0 (2.23)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
18

The driving force for the bubble growth in Equation (2.22) is the pressure

difference between the vapor bubble and the liquid adjacent the bubble, which is

constrained by the surface tension. Equation (2.22) implies that static equilibrium will

exist when the first and second derivatives are zero. For the initial vapor bubble of critical
size, this gives, since Tv = initially:

Rc = 2 [ T) (2.24)
c PV(T)~P

The energy equation in spherical coordinates is given by Equation (2.5) and listed

here again as:

dT dT d 2T 2 dT
(2.25)
dt +Ur dr ~ a vv3r2 r dr

R} R
where ur = and e = ' i - * ' (2.26)
r~ P /.
The initial condition is: T(r,0) = Tm (2.27)

The boundary conditions are: jT(,0) =

< 2 -2 8 )

Equation (2.28) is the heat balance at the liquid-vapor interface of the bubble. The

term in the RHS of the equation represents the latent heat rate for evaporation of the

liquid to produce the vapor required in the bubble, while the term in the LHS represents

the conduction heat transfer rate from the liquid adjacent to the vapor bubble wall.

According to assumption (b), the conduction heat transfer from the bubble wall to the

interior of the vapor bubble is neglected. The bubble growth problem in a superheated

liquid can be solved by coupling Equation (2.22) and Equation (2.25) with the initial and

boundary conditions given in Equations (2.23) and (2.28).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
19

2.4 Numerical Solution

A Fortran program was written to solve the transient spherical bubble growth

problem in superheated liquids. The program couples the momentum and the energy

equations and marches stepwise in time, taking into account the variations of properties

with temperature. For the infinitesimal time steps, the momentum equation is solved by

means of the Runge-Kutta method to determine the locus of the bubble interface, which

provides the boundary condition for the energy equation. The complexity of the problem

is simplified by adopting the Landau coordinate transformation in order to immobilize the

moving boundary. After the transformation, the energy equation is solved by using the

Thomas algorithm, which is called the TriDiagonal Matrix Algorithm.

The solution method uses the implicit finite difference method for the energy

equation, with the explicit boundary condition given by Equation (2.28). An irregular grid

is adopted for the liquid to minimize the computation time, with the grid spacing

decreasing as the bubble interface is approached. 100 nodes are used for about 2 mm of

the thermal boundary layer in this program. The positions of the control volume faces are

determined according to the relation:

n (2.29)

where n is total number of control volume faces and p is an exponent suited to the

problem, usually taken to be between 1.5 and 2.5. If p = 1, a regular grid results; if p > 1,

an irregular grid results. It is critically important that the time step be arranged to increase

as the computational time passes in order to reduce the computation time. The time step

initially starts from on the order of 10-9 seconds and eventually increases to about 10'3

seconds, depending on the computational time limit.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
20

A small disturbance is necessary in order to initiate the bubble growth from the

critical size, which is in metastable equilibrium. In this work, a small temperature

disturbance is given once at the liquid-vapor interface, and thereafter the program runs

without further disturbances. The effect of the magnitude of the disturbances will be

discussed in section 2.5. Dalle Donne and Ferranti (1975) added a volumetric energy

source term to the energy equation to provide the disturbance necessary to make bubble

grow from its critical size, claiming that the energy source term will not affect the

subsequent growth rate of the bubble unless extremely high or unrealistic values are

assumed. Two major differences exist between the present work and that of Dalle Donne

and Ferranti (1975): (1) the former uses the Landau transformation while the latter uses

the Lagrange transformation to immobilize the moving boundary, (2) the former uses a

temperature disturbance for one time step at the liquid-vapor interface at the beginning of

the computational process, while the latter uses a constant energy source during the entire

computation in order to initiate the bubble growth from the condition of metastable

equilibrium.
To immobilize the moving boundary, a dimensionless coordinate (rj) is introduced

as shown in Figure 2.3, given by Sparrow et al. (1978) and referred to as the Landau

coordinate:

0 < r| < 1 , fl(f)<r </?(?) + 8 (2.30)

where 8 is an arbitrary distance which can be considered to correspond to a

thermal boundary layer. The conservative forms of the continuity and the energy equation

are used to begin the coordinate transformation.

(2.31)

(2.32)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
21

The transformation of the governing equations is facilitated by the following

relations, also given by Sparrow et al. (1978):

3 _ 1 8 r' 3
(2.33)
dr 8 9r| dt dt 8 3tj

, d r , s. dd dR
(2.34)
r = a7 ( v ) = ^ + *

The continuity and energy Equations (2.31) and (2.32) can be transformed to the

immobilized equations by using Equations (2.30), (2.33), and (2.34), as was done by Kim

and Kaviany (1990) based on the work of Sparrow et al. (1978). This has the significant

effect of eliminating the convection term in Equation (2.32), with the result:

df 3 ^ dF n
(2.35)
dt ^ 3 t | J 3rj

3 / .3 1 ^ 37 .
(2.36)
dt . ) 3n

dv_
where F(r|,r) = p r2u - (2.37)
dt

1( v . r2k d T
7(ti,/) = F(x\,t)h (2.38)

1
V(tl,f) = - r 3 (2.39)

r(rj, t) = tj8 + R(t) (2.40)

The term F(Tj,r) and the term J(r\,t) represent the mass flux and the energy flux,

respectively, across the face of the control volume, as shown in Figure 2.4.

Now, the transformed conservation equations (2.35) and (2.36) are discretized by

using the numerical scheme of Patankar (1980). For a typical control volume as shown in

Figure 2.4, the integration of Equation (2.36) with the aid of Equation (2.35) gives:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
22

a PTP = a eTe + a WTW+b (2.41)

where a p = a E + a w + a , (2.42)

+ max{-cFe,0} (2.43)

(2.44)

(2.45)

(2.46)

(2.47)

(2.48)

The various numerical schemes developed and used in recent years have been

formulated as different choices of the function A(IPI) in Equations (2.43) and (2.44), as

developed by Patankar(1980), which amount to different means for expressing the

derivatives between nodal points. The power law scheme suggested by Patankar(1980) is

used here:

a (|^|) = m a x {o ,(l-0 .l|n |)S} (2.49)

A(|i;,|) = nm {o,(l-0.l|P .|)s} (2.50)

The max {B, C} denotes the greater of B and C.

The previous value at time t is known and is denoted by using the superscript "o",

while the new value at time t + At is unknown and is denoted by using no superscript.

The discretized Equation (2.41) can be solved efficiently by using TDMA (TriDiagonal

Matrix Algorithm) with the two aforementioned boundary conditions Equation (2.28),

which can be discretized considering the finite volume at boundary.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
23

(2.51)

T t~ 2 F nQxk
(2.52)
1+ 2 K 1+ 2 F0

where F0 =

Equation (2.52) is used as the discretized boundary condition for the energy

equation, since dR/dt can be obtained from a time step solution of the momentum

Equation (2.22) by using the Runge-Kutta method.

Solution Methods

To begin the computation of bubble growth from the critical radius it is first
necessary to determine the critical radius ( Rc) for the given initial values of the bulk

temperature(7^,) and the system pressure(PJ, and is given by Equation(2.24)

In the actual computational process a vapor temperature slightly higher than the

initial bulk temperature is imposed for the initial time step in order to provide the

disturbance necessary for growth from the critical size, and is expressed as:
TV= T+ ATddisturb. (2.53)

The small disturbance does not affect the subsequent bubble growth except during

the very early stages of bubble growth (to be demonstrated in section 2.5). This small

disturbance could be interpreted as a small statistical fluctuation of the local liquid

temperature in the vicinity of the critical size vapor bubble. Once this disturbance is given,

the computer solves Equation (2.22) by means of the Runge-Kutta method assuming that

the pressure is constant for the short period of the time step, so that R, R, and R are

produced. TheR is used to solve Equation (2.52), which is a boundary condition at the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
24

liquid-vapor interface. The discretization Equation (2.41) then can be solved by means of
the TDMA so that a new interface temperature(TJ) is determined, which then is used to

obtain the vapor pressure corresponding to the interface temperature with assumption (c).

This process is iterated to update the temperature variation until the velocity difference

between the time steps (i - 1) and (i) converges to a certain value, here taken as 10-6 m/s

for the convergence limit. The convergence is so fast that only 2 - 3 iterations are usually

required to satisfy the convergence limit. The computation time required was

approximately 1 0 - 3 0 minutes for a personal computer with a 486/66 processor in order

to run the program up to the time limit of 1 second, using an initial time step of 10'8

seconds. The actual time required is dependent upon how the time step variation is

arranged.

2.5 Comparisons with Previous Works

Among the many studies dealing with bubble dynamics, few numerical works have

been reported to date. Most analyses used some form of an analytical approximation in

order to describe the bubble growth in superheated liquids, as discussed in Section II.

Dalle Donne and Ferranti (1975) were the first to solve the problem numerically, taking

into account the variation of properties with temperature. In order to eliminate the

complexity of the convection term in the energy Equation (2.25), a Lagrange coordinate

transformation was performed. For the finite difference method, the Babuska method

(1966) was adopted.

In the work presented here, the Landau coordinate transformation is used for the

immobilization of the moving coordinate and the Patankar (1980) procedure for the finite

difference method. Thus, it can be said that Dalle Donne and Ferrenti (1975) and the

present work use quite different solution methods, although both basically use the same

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
25

governing equations. Both works use the Runge-Kutta method to solve the equation of

motion, Equation (2.22).

2.5.1 Sodium

Computations performed here will first be compared with the results of Dalle

Donne and Ferrenti (1975) for a liquid metal. Sodium was used with three different

superheats, and the comparisons are given in Figures 2.5 ~ 2.8. The solid lines represent

the computations conducted here, while the symbols represent the discrete computations

of Dalle Donne. It is to be noted that since the data of Dalle Donne (1975) used in these

figures were taken from the original logarithmic scaled plots, an error of about 4% may

be present in the values shown. Furthermore, the property equations used for sodium in

the present work may not be identical to those used by Dalle Donne. In spite of these

differences, this work is in a good agreement with the results of Dalle Donne and Ferrenti

(1975), as shown in Figures 2.5, 2.6 and 2.7.

In Figure 2.5, the onset of the bubble growth from each corresponding critical size

radius begins earlier as the superheat increases. It seems that the surface tension

dominates the onset of bubble growth. It is noteworthy that the bubble grows to a radius

of about 1 meter in 0.1 seconds. In Figure 2.6, it is noted that for a superheat of 42 K,

the interface velocity increased from 0.001 m/s to 5 m/s, while for a high superheat of 362

K the maximum velocity encountered was on the order of 25 m/s, which can be considered

as an explosively fast growth.

Figure 2.7 represents the interface temperature vs. time. The variation of the

interface temperature during the early stages of the bubble growth has not been

determined in previous analytical works because of computational difficulties. It begins at

the initial liquid temperature and converges to the saturation temperature corresponding to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
26

the system pressure. The convergence is slow, in case of sodium, so that the saturation

temperature is not attained even in one second, as noted in Figure 2.7. This is due to its

high thermal conductivity, and is to be contrasted with the case of both water and R -l 13,

where convergence occurs within one millisecond. A discrepancy of about 10 K out of

1100 K exists between the two works for a superheat of 42 K. This is attributed to using

a different vapor-pressure curve, since the discrepancy remained consistent. Also, a slight

discrepancy is present in the computation of the saturation temperature, as shown by the

symbols at 1 second on the time scale being about 10 K above the computed values. The

radius vs. time curve in Figure 2.5 is less sensitive to this error than the interface

temperature curve in Figure 2.7.

Figure 2.8 shows, for sodium, the effect of using various small initial temperature

disturbances. The symbols are the numerical data of Dalle Donne and Ferrenti (1975) and

the solid lines represent the present work. It is interesting that the best fit corresponds to

the use of 10" C as a disturbance in this work, while Dalle Donne used the equivalent of

102 C as a heat source. Dalle Donne and Ferrenti (1975) did not use the terminology of

"disturbance" for the distributed energy or heat source in the liquid. Additionally, it is to

be recalled that the disturbance used here is characterized differently from that of Dalle

Donne. The disturbance here is applied to the vapor temperature initially at one time step

only, while the heat source term is applied continuously to the liquid surrounding the

bubble. The effect of the disturbance will be discussed in detail in section 2.6.

2.5.2 Water

Dergarabedian (1953) was among the first to conduct experiments measuring the

bubble growth in uniformly superheated water at atmospheric pressure. Superheats of up

to 6 C were obtained by heating the liquid in a glass beaker using an infrared lamp.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
27

Transient bubble sizes were measured with a 16 mm high speed camera having a framing

rate of 1000 fps. It was claimed that the growth rates were so rapid that the translational

motion due to buoyancy could be neglected. It was necessary to facilitate the

homogeneous nucleation by addition of powdered chalk or sand to the water to produce

the nucleation at the desired degree of superheat. Strictly speaking, the nucleation in these

experiments occurred at a solid surface of the particles, with bubble growth then

following. The measurement accuracy of superheat was assumed to be within 0.2 C,

and a shifting of the data along the time scale is permissible because of the uncertainty of

the appearance of the first bubble within the framing interval of 1 millisecond. Figure 2.9

presents a comparison of the data with computations performed here. The data was

arranged to provide the best fit within the uncertainty allowance. The data for superheats

of 2.1C and 3.1 C are in good agreement with the computations, while the data for

superheat 4.5 C fell about 10 % below the computed data. One possible reason may be

the variable air content in the water, although Dergarabedian boiled the distilled water for

about 1 hour to remove the air.

The measurements of Kosky (1968) for a high superheat of 36 C with water are

compared in Figure 2.10 with computations made here. Experiments with such high

superheats are quite difficult to perform because the boiling usually is initiated earlier at

some point on the wall surface of the test vessel before a high superheat can be reached.

The high superheat was reached in this case by the use of a diaphragm rupture technique

to decrease the system pressure rapidly, with a glass tube as the test vessel to eliminate

premature nucleation at the wall, using a 10,000 fps high speed camera for photography.

The data of Kosky (1968) are in good agreement with the computations.

Figure 2.11 shows additional measurements of Dergarabedian (1960) with low

bulk liquid superheats of 0.8 C and 1.0 C for water. This data is unique in that it

included measurements of vapor bubble sizes in the thermodynamic critical size range.

Bubble growth rates measured with such low bulk liquid superheats were not in good

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
28

agreement with certain earlier analytical solutions because the assumption of moderate or

high bulk liquid superheats was necessary in the analytical development of the asymptotic

solutions. Thus, solutions for low bulk liquid temperatures were not possible until

numerical computations to the fine degree necessary became practical. Although Dalle

Donne and Ferrenti (1975) solved the coupled equations for spherical growth, they

apparently were unaware of the earlier measurements of Dergarabedian (1960).

Comparisons of these are made with present computations in Figure 2.11, where the vapor

bubble growth is initiated from the thermodynamic critical size corresponding to the initial

bulk liquid superheat, given by Equation (2.24). The temperature disturbances here were

selected as 10-4 C and 10'9 C to provide the best fit along the time axis. For the low

superheats used here the critical vapor bubble radius of 0.04 mm was measurable with

appropriate optical magnification, whereas this is not possible for high levels of initial

superheat. It is to be noted that the critical radius, frequently considered to be a

theoretical calculation, was demonstrated to be an experimental reality.

Florschuetz et al. (1969) used a rapid depressurization technique to obtain a

moderate superheat 3.61 C of water, considerably lower than that used by Kosky (1968).

The specification of the system pressure was complicated by its variation during the

experiments being on the order of 5% of the mean value. The liquid was boiled for at

least 30 minutes at atmospheric pressure to remove the air. A high speed camera with a

framing rate of 1200 fps was used, and the measurement accuracy of liquid temperature

was within 0.2 C. It was noted that the initial bubble grew very slowly at first and then

grew rapidly, and compared well with the analytical work of Scriven (1959) when the data

were shifted 6.5 milliseconds to obtain the best fit. The data of Florschuetz et al. (1969)

are compared in Figure 2.12 with computations made here, with an arbitrary shift of 6.5

ms in the time origin. As may be noted in Figure 2.12, the bubble has a radius of 0.12

mm at the very early times. This was originally thought to be the critical size, but actually

is an order of magnitude larger than the critical radius of about 0.01 mm corresponding to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
29

the superheat of 3.61 C. The possibility exists that the initial bubble formed is much

larger than the computed critical radius because of the presence of air.

Figure 2.13 shows comparisons between the present numerical calculations and

the measurements of Lien (1969) for water over a pressure range of 1.26 to 38.66 kPa,

with good agreement. Figure 2.13 demonstrates the applicability of the computational

procedure for a wide range of reduced pressures. Lien (1969) basically adopted the

method of Derarabedian (1953), with the additional control of the bubble formation, by

energizing a chromel electrode of 0.00127 mm (0.005 inch) in diameter with a 0.00254

mm ( 0.001 inch) gap between the electrodes. A Pyrex glass beaker was used as a test

container, and superheats as high as 16 C were attained with two infrared lamps. A layer

of silicon oil 3.175 mm (1/8 inch) in thickness was used to prevent the evaporation of

water, and the low system pressure was maintained by means of vacuum pump. A streak

camera with strobe flash provided the high framing speed of 2000 fps. The uncertainty in

temperature is estimated to be 0.5 C and the uncertainty in pressure to be 0.2 mmHg.

2.5.3 R-113

Measurements of spherical bubble growth in initially uniformly superheated R-113

were made by Bohrer (1973). A Pyrex flask was used as a test vessel with a layer of

glycerin floating on the fluid to prevent surface evaporation. A touching probe technique,

in which the desired nucleation is achieved by quickly touching and separating two wires

suspended in the superheated liquid, was used to initiate the vapor bubble growth. Two

sizes of probe, 0.01 and 0.02 inch in diameter, were used in the experiments of Bohrer

(1973). Figure 2.14 presents data obtained with the small size probe of 0.01 inch, using a

camera framing rate of 5000 fps, together with computations obtained with the model

described here.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
30

Table 2.1 below shows the parameters for Figure 2.14. The initial temperature

was uniform at 22.78 C, and the system pressure was reduced carefully to 8.46 kPa (1.23

psia) such that the liquid became uniformly superheated at 34.12 C, and to 3.66 kPa

(0.53 psia) such that the liquid superheat was 48.21 C. The higher uniform initial

superheat, resulting from dropping the initial pressure to a lower value, produces a more

rapid growth rate, as anticipated. The experiments correspond to saturation temperatures

of -11.33 C and -25.99 C, respectively. It is to be noted that large bubbles are

associated with such low pressures, and the growth rates are not unusually high in spite of

the large initial liquid superheats. It is expected that circumstances will be quite different

with large superheats at higher pressures.

Table 2.1 Parameters corresponding to Figure 2.14.

Superheat P(system) Tsat Jacob No.

kPa/psia

34.12 C 8.462/1.227 -11.33 C 418

48.21 C 3.66/0.53 -25.99 C 1321

2.6 Discussion

It has been demonstrated that fair agreement exists between measurements and

computations over a wide range of superheat and pressure, although some difficulties have

been experienced in the conduct of experiments. The effect of varying the disturbance

parameter imposed to begin the bubble growth will now be examined, particularly during

the early stages of growth, since the magnitude of the disturbance used will be quite

arbitrary. The effect of this parameter on the transient vapor bubble radius, interface

temperature, velocity and acceleration will be presented, for H 2 O, R-113, and sodium. In

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
31

addition, the computations with the model presented here will be compared with the

various analytical models described earlier, also for H2 O, R-113, and sodium, for high/low

superheats and high/low pressures. The effect of superheat and pressure on the maximum

bubble growth rate, the delay time, and the maximum interface acceleration also will be

presented.

2.6.1 Water

Figure 2.15 shows the bubble radius and liquid-vapor interface velocity versus time

curves on the log-log scale axis for high and low superheats of 36 C and 3.1 C, from the

numerical computations, with time beginning when the disturbance is imposed. It is noted

that the bubble initially grows very slowly. After a certain time lag, the bubble then

suddenly grows rapidly due to the reduction in surface tension and the accompanying

increase in the internal vapor pressure difference. This rapid growth is referred to as the

onset of bubble growth. The time lag increases as the superheat decreases. Three

interface temperature disturbances are introduced at t=0: lCH C, lO 5 C and 10'9 C.

The onset of the bubble growth is delayed as the magnitude of the disturbance decreases.

It is noted that the bubble growth converges asymptotically no matter what the magnitude

of the disturbance, indicating that the disturbance does not affect the subsequent growth.

The critical radius for a superheat of 3.1 C is 0.01 mm, to be compared with a radius of

0.0005 mm for a superheat of 36 C. The radius of 0.01 mm is difficult to measure, while

the radius of 0.0005 mm is almost impossible to measure even with modem technology. It

was noted in connection with Figure 2.11 that the corresponding measurements of the

critical size vapor bubble indeed have been made, but this was possible only because of

low level of bulk liquid superheat used. It is to be noted in Figure 2.15 that the onset of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
32

growth does not change significantly as the magnitude of the disturbance changes,

although it appears markedly different due to the log scale axis used here.

Figure 2.15 also shows the variation of the interface velocity of the bubble for both

superheats. The maximum interface velocities are seen to be about 10 m/s for a superheat

of 36 C and 0.5 m/s for 3 C, and are independent of the magnitude of the disturbances.

This implies that the magnitude of the disturbance should have a negligible effect on the

vapor bubble size once the initial acceleration is completed. The interface velocity of 10

m/s for the high superheat is much smaller than both the sound speed of 500 m/s

calculated for water vapor and the root mean square molecular speed o f 752 m/s for water

vapor (refer to Appendix C), so that it may reasonably be assumed to be incompressible

flow. It is noted that a small dip appears on each of the interface velocity curves for the

superheat of 3.1 C at about 10'5 sec. Although its effect appears to be negligible, this dip

becomes greater in case of R113, and is related to the relative effects of surface tension

and pressure difference between the bubble vapor and bulk liquid. A more elaborate

discussion of this phenomena will be presented in section 2 below. It is noted that the

seemingly large differences in the initial interface velocities due to the different initial

temperature disturbances have relatively little effect on the bubble radii in the measurable

domains.

Figure 2.16 presents the variation of the liquid-vapor interface temperature along

with the duplication of the radius-time curves from Figure 2.15, for both superheats of 3.1

C and 36 C for water. The temperature begins to decrease rapidly from the initial liquid

temperature at the onset of the bubble growth, and eventually converges asymptotically to

the saturation temperature corresponding to the system pressure. It is noted that the rapid

descent of temperature occurs within about 1 millisecond at atmospheric pressure. This

decrease will be much slower at low pressures, as will be discussed later. It can be

deduced that the internal vapor pressure will respond in a manner similar to that of the

interface temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
33

Figures 2.17 and 2.18 show the interface acceleration vs. time for moderate and

high superheats, respectively. The radius-time curves are again duplicated from Figure

2.15 for purposes of convenience. A rapid acceleration occurs at 5x105 seconds for the

largest temperature disturbance and moderate superheat of 3.1 C, in Figure 2.17, due to

the reduction of pressure difference associated with the surface tension. It rises quickly to

10,000 m/s2, then becomes negative due to the inertia of the liquid and eventually

converges to zero, which indicates that momentum no longer plays a role. The thermal

diffusion controlled growth then becomes the primary acting mechanism. Thus, bubble

growth can be characterized by three mechanisms: surface tension dominated growth,

momentum dominated growth, and thermal diffusion dominated growth, with the divisions

between these mechanisms separated by the acceleration peaks as illustrated in Figure

2.19. It is noted that the maximum of the interface acceleration is quite sensitive to the

initial bulk liquid superheat, but not to the magnitude of the initial temperature disturbance

necessary to initiate the growth. For the high superheat of 36 C in Figure 2.18, the

enormous acceleration of 4xl07 m/s2 might provide the surface instability to produce a

roughened liquid-vapor interface if the superheat were increased further, as was observed

by Ervin et al (1992) with R-l 13. On other hand, it may be considered that even though

the magnitude of this acceleration is so large, the relative shortness of its effective action

combined with the small size of the vapor bubble at this time could produce a negligible

effect on the subsequent growth.

Computations resulting from the model of the present work will now be compared

with analytical solutions available to date, for water at atmospheric and low pressures,

with both high and low bulk liquid superheats. Five solutions are selected here for the

comparison purposes consisting of Equation (2.2) for Rayleigh (1917), Equation (2.8) for

Plesset and Zwick (1952), Equation (2.11) for Scriven (1959), Equation (2.12) for Mikic

et al. (1970), and Equation (2.16) for Prosperetti and Plesset (1978). Constant properties

are used in all the plots except in the present work.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
34

Figure 2.20 shows the comparison of the work with the five analytical solutions for

a high superheat of 36 C at atmospheric pressure, using linear scale. The Rayleigh curve

denotes the momentum controlled growth, with the corresponding fastest growth. Plesset

and Zwick (1954) and Scriven (1959) are in fair agreement with the present work here.

Prosperetti and Plesset (1978) is in poor agreement, while Mikic et al. (1970) is in

excellent agreement.

The very early stages of the bubble growth for the identical case of Figure 2.20 are

amplified in Figure 2.21 by the use of log-log scales. It is noted that none of the analytic

solutions follow the numerical solution of the present work. This is to be expected, since

both surface tension and momentum were generally assumed to be negligible in the

derivation of the analytical solutions.

Figure 2.22 shows the comparison between the present work, measurement, and

four analytical solutions for water for a superheat of 15.74 C at a low pressure of 1.26

kPa. Plesset and Zwick (1954) and Scriven (1959) overestimate the bubble growth

markedly , while Rayleigh (1917) and Mikic et al. (1970) are in fair agreement if the delay

periods are considered. The present work is in excellent agreement with the

measurements of Lien (1969). The comparisons between the predicted bubble growths at

the low pressure in Figure 2.22 are distinctly different from that at the higher pressure in

Figure 2.20. On examining the computed liquid-vapor interface temperature in Figure

2.22, it is noted that a temperature of 23.6 C is reached at 0.01 sec., which is still

considerably above the saturation temperature of 10.41 C and is to be contrasted with the

behavior at a high pressure in Figure 2.16 where the computed liquid-vapor interface

temperature reaches saturation in about 1(H sec. This implies that the vapor pressure is

considerably greater than the system pressure, so that liquid inertia still governs the

growth process. These comparisons between high and low pressures with water for the

same superheat are given explicitly in Figures 2.22 ~ 2.25.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
35

Figure 2.26 presents a comparison of the present numerical results with the

analytical solutions for a low liquid superheat of 0.8 C, and it is obvious that none of the

analytical solutions corresponds to the present numerical results. This is also to be

expected, since moderate or high superheats were assumed in deriving the analytical

solutions. A long time lag exists before the onset of the bubble growth for such a low

superheat. As the experimental work of Dergarabedian (1960) demonstrated, some

degree of time shift may be required when making comparisons with measurements due

both to the uncertainty in measurement of the true onset of bubble growth, associated with

a finite framing speed, and to the inherent randomness associated with homogeneous

nucleation.

2.6.2 R-113

The effect of the initial temperature disturbance necessary to initiate the vapor

bubble growth are presented below for R-113 at atmospheric pressure. Figures 2.27,

2.29, 2.30, and 2.31 for R-113 correspond to Figures 2.15-2.18 above for water,

respectively, with the same system pressure and initial liquid superheat. It is to be noted

that the influences of both the disturbance temperatures and the initial liquid superheat

levels on the vapor bubble growth parameters of radius, interface temperature, interface

velocity, and interface acceleration are quite similar for the two fluids, even though their

respective saturation temperature and other properties are quite different.

Figure 2.27 indicates that variation of the initial temperature disturbance has little

effect on the subsequent bubble growth for R-113. The discontinuity in the velocity

curves for the superheat of 3.1 C is a consequence of the use of a log-log scale, which

can not represent negative values. As discussed earlier, the same circumstance occurs

with water in Figure 2.15 except that negative values do not arise in the interface velocity.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
36

Although the radius curves for the delay time period ranging from 10'8 to 10'5 seconds

appears to be constant for the scale in Figure 2.27, microscale motions of the embryo

bubble interface are in fact present, as shown in Figure 2.28. In order to examine this

behavior in some detail, the time domain where this small dip appears is amplified in

Figure 2.28. The first, second, and third terms on the right hand side of Equation (2.22),

repeated below for convenience, are hereafter referred to as the pressure, surface tension,

and viscous terms, respectively. It is noted that the pressure is a function of the interface

temperature, while the surface tension term is a function of both the temperature and

radius. The sum of the terms on the right hand side of Equation (2.22) act as the driving

force for the bubble growth rate, which may be positive or negative.

J8gt3iP. f.m-&_2ge2 -4j ! i ( 2 .2 2 )


2 ep, ep ,R ep,R

Figure 2.28 (a) shows the radius and velocity versus time for the delay period up

to 10*5 sec., and Figure 2.28 (b) presents the pressure and surface tension terms versus

time. The surface tension term is always positive, while the viscous term may be neglected

since its magnitude is smaller than the other two terms by a fifth order. Since the bubble is

at the critical size at zero time, the right hand side of Equation (2.22) is mathematically

balanced at zero. Upon applying the initial required temperature disturbance, this balance

is disturbed, and which term will be initially greater is unknown, since it is dependent upon

the superheat level. However, it has been determined that for the case of the low

superheat of 3.1 C, the disturbance produces a large initial surface tension term, while the

opposite is the case for the high superheat of 36 C. For the latter case the sum of the two

terms is positive, and the velocity grows gradually, but for the former case the velocity

decreases, since the sum becomes negative. As the microsize bubble grows, the pressure

and surface tension terms decrease due to the decrease of the interface temperature and

increase in bubble size, respectively. As seen in Figure 2.28, the velocity decreases rapidly

but still remains positive, which means the radius growth slows down. This leads in turn

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
37

to a halt in the decrease of the liquid-vapor interface temperature. Since the interface

temperature corresponds to the vapor pressure, the pressure term begins to increase, due

to liquid momentum effects as the radius growth continues to slow down. Meanwhile, the

interface velocity becomes negative so that the radius decreases momentarily, and the

pressure term then rises above the surface tension term. This leads to a radius increase

again, and continues until the observed so-called nucleation occurs. The sequence of

events described above may be considered to be the dynamic growth process for a typical

microsize bubble during the delay period, leading to the onset of boiling. It is to be

emphasized that the negative velocity in Figure 2.28 (b) is not due to interface acceleration

effects associated with liquid momentum, as can be confirmed by examining the

acceleration plot in Figure 2.30, for the identical case with the large disturbance of 0.1 C.

Figure 2.32 presents computed vapor bubble radii for R-113 as a function of time

for two very low and slightly different superheat levels of 0.8 and 1.0 C, and illustrates

the extreme sensitivity of the bubble growth to small changes in superheat at this level.

Computation from the model in the present work are compared with several analytical

models in Figures 2.33 and 2.34 for R-113 at low and high pressures, respectively, and

demonstrate rather severe discrepancies at the low pressure levels.

Figure 2.33 shows the comparison between the present numerical calculations and

the various analytical solutions and the measurements of Bohrer (1973), for a high

superheat of 34.11 C and low system pressure of 8.46 kPa. Rayleigh (1917) and Plesset

and Zwick (1954) are in poor agreement with the measurements of Bohrer (1973), while

Mikic et al. (1970) and the present computed work are in fair agreement. This infers that

the liquid inertia and thermal diffusion are equally important. The solution of Mikic et al.

(1970) works reasonably for R-113 as well as water, within 10 % error.

Figure 2.34 presents the comparisons between the present numerical calculations

and the various analytical solutions for a high superheat of 36 C and atmospheric pressure

of 101.325 kPa. It is noted that the pressure changed from 8.46 kPa to 101.325 kPa

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
38

between Figure 2.33 and Figure 2.34, while the superheat is virtually the same. All

analytic works are in fair agreement with the present computation, which infers that the

bubble growth at higher pressures is governed not by liquid inertia, but rather by thermal

diffusion.

2.6.3 Sodium

The effect of the initial temperature disturbance necessary to initiate the vapor

bubble growth are presented below for liquid sodium at atmospheric pressure. Figures

2.35-2.38 for liquid sodium correspond respectively to Figures 2.15 - 2.18 above for

water, with the same system pressure and initial liquid superheat. The influence of the

initial temperature disturbance is similar to that with water.

Figure 2.36 presents the liquid-vapor interface temperature versus time for liquid

sodium at atmospheric pressure for superheats of 3.1 C and 36 C. It is noted that the

decrease of the interface temperature for the relatively high superheat of 36 C is quite

slow compared to water in Figure 2.16 or R-l 13 in Figure 2.29, and is a consequence of

the high thermal conductivity typical of liquid metals. This results in large growth rates,

seen in Figure 2.39, and is similar to that for R-l 13 at a low pressure, in Figure 2.33.

The analytic solution of Mikic et al. (1970) has been in good agreement thus far

with the various measurements and the present computation for three different fluids, and

it was desired to test it for sodium over a wider range of pressure and superheat. Figure

2.40 provides the comparisons between the numerical calculations and the analytic

solution of Mikic et al. (1970) for liquid sodium with superheats of 36 C and 200 C and

system pressures of 0.2 atm and 6 atm. The discrepancies between the computation and

Mikic's solution, still relatively small, become greater as the system pressure decreases or

as the superheat increases. The predictions of Mikic et al. (1970) underestimate the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
39

bubble growth with a maximum discrepancy of 33 % for the pressure of 0.2 atm and

superheat of 200 C. Nevertheless, the analytic solution of Mikic et al. (1970) can be used

with considerable accuracy for moderate superheats and pressures for a wide range of

fluids. The solution is given by Equation (2.12).

2.6.4 Comparisons between Water, R-113 and Sodium

Comparisons between the bubble growth parameters are presented directly here

for water, R-113 and sodium with the same superheats and pressures. The thermodynamic

properties for these three are given in Table 2.2 below for purposes of comparison.

Liquid properties Liquid sodium Water R-113

Saturation temperature (C) 850 100 47.5

Critical temperature (C) 2460 374 214

Superheat limit (C)* 2159 302 160

Critical pressure (MPa) 41.36 22.1 3.41

Prandtl number (Pr) 0.004 1.76 7.2

Conductivity (W/m.K) 47.43 0.68 0.07

Liquid density (kg/m3) 736.7 957.85 1507.3

Heat capacity (kJ/kg.K) 1.286 4.217 0.984

Thermal diffusivity (m2/s) 500 x 10-7 1.684 x 10-7 0.4753 x 10-7

Heat of vaporization (kJ/kg) 3874.5 2257 143.8

Surface tension (N/m) 0.113 0.059 0.016


* Refer to Appendix D.

It is noted that the Prandtl number and thermal diffusivity for sodium are distinctly

different than those for water and R-l 13, which are relatively close. These three fluids are

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
40

selected for use here since these are representative of most fluids. Figures 2.41 ~ 2.44

present computations of bubble radius, interface velocity and interface acceleration at

atmospheric pressure with superheats of 3.1 C and 36 C for the three different fluids to

facilitate comparisons between these bubble dynamic parameters. The critical radius,

delay time, maximum liquid-vapor interface velocity, maximum acceleration, and bubble

radius at a specific time are presented in Figures 2.45 - 2.49 in order to determine how

these parameters are effected by superheat and pressure. The initial temperature

disturbance was maintained constant at 10 5 C for these computations. Superheats up to

100 C are used, and system pressure varied from 0.1 to 4 atmospheres.

Figure 2.45 (a) ~ (c) present the critical radius versus superheat with the variation

of the system pressure for the three different fluids. The critical radius is determined from

Equation (2.24). Figure 2.45 (d) shows the comparison of the critical radius for the three

different fluids at atmospheric pressure. It is noted that sodium has a critical radius

considerably larger than that of water and R-113, and that the critical radius decreases as

both superheat and pressure increase. This implies that the probability for nucleation

increases as both superheat and pressure increase.

Figure 2.46 (a) ~ (d) present the delay time versus superheat with the pressure

variation for the three different fluids. The delay time here is defined as the period from

the imposition of the initial temperature disturbance to the time at which a significant

increase in the growth rate takes place, as seen, for example, with the lower curves in

Figures 2.41 ~ 2.44. The delay time decreases as both superheat and pressure increase.

The differences in delay times between water and R -l 13 are insignificant.

Figure 2.47 (a) ~ (d) show the maximum liquid-vapor interface velocity versus

superheat with variation of the system pressure for the three different fluids. The

maximum velocity increases as both the superheat and pressure increase. Water shows the

highest velocity among the three fluids in Figure 2.47 (c), with velocity of 36 m/s for a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
41

superheat of 100 C at a pressure of 4 atm, while sodium shows the lowest velocity of 15

m/s in Figure 2.47 (a).

Figure 2.48 (a) ~ (d) presents the maximum liquid-vapor interface acceleration

versus superheat for the three fluids and for the various pressure levels. The maximum

acceleration increases as both the superheat and pressure increase. Sodium shows the

lowest acceleration, while water and R-113 are close together in Figure 2.48 (d). One

cause could be that sodium has the largest critical radius as seen in Figure 2.45. This leads

to a reduction in the interface velocity as well as acceleration.

Figure 2.49 (a) ~ (d) presents the bubble radius at an arbitrary time of 0.01 sec

versus superheat, and indicates that the early bubble growth rates are very sensitive to the

system pressure. It is noted that the radius of a water vapor bubble is about 200 % larger

than a R-l 13 bubble, while the sodium bubble is about 600 % larger than a R-l 13 bubble.

The sodium vapor bubble grows to a radius of 6 cm in the short period of 0.01 seconds

with a superheat of 100 C at atmospheric pressure.

Unfortunately, the behavior of vapor bubble dynamics during the early stages of

growth has not been supported extensively by measurements because of the great

experimental difficulties. Dergarabedian (1960) uniquely obtained photographs of a

typical bubble growth from the critical size bubble, as was shown in Figure 2.11.

However, the present numerical results demonstrated here provide reasonable predictions,

and Figures 2.45 - 2.49 can be useful indicators when it becomes necessary to quickly

estimate the early vapor bubble growth stages.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
42

Bubble number: 29-G20


System pressure: 3.119 kPa (0.452 psia)
Bulk temperature: 22.22 C
Liquid superheat: 50.83 C
Saturation temperature: 28.61 C
Electrode diameter: 0.508 mm
Framing rate: 5000 fps
Picture 17: radius=approx. 1 cm and t=4 ms

Figure 2.1. Photographs by Bohrer (1973) for vapor bubble growth in R -l 13,
demonstrating the sphericity of the bubble.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
43

Tbulk

Superheated Liquid

Tsat

Critical Radius at t=0

(a)

T
Tbulk

Bubble Tsat

Thermal boundary layer

Bubble grown large enough that surface tension may b e n eglected .

(b)

Figure 2.2 (a) Vapor bubble at the critical size radius, (b) Bubble growth in a uniformly
superheated liquid.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
44

Figure 2.3 Schematic diagram of coordinate transformation.

^ m i

Figure 2.4 Grid point and control volume notation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
45

le+02
:Bubble Radius vs. Time for Sodium, Psys=50 kPa, Tsat=1066 K

le+01
Symbols : Dalle Donne's numerical work (1975)
le+00 Solid Lines : This work (computed)

le-01
Radius, meter

le-02

le-03 Superheat=362 K

le-04
Superheat=42 K
le-05

le-06 Superheat=201 K

le-07
le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01 le+00
Time, sec

Figure 2.5. Comparison of present results with the numerical work of Dalle Donne and
Ferranti (1975). Radius versus time for sodium.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
46

10000
: Vapor Bubble Growth rate for Sodium, Psys=50 kPa, Tsat=1066 K

Symbols : Dalle Donne's Numerical work (1975)


1000
Solid Lines : This work (computed)
Superheat=362 K
100
Supbrheat=201 K

1
;>
O Superheat=42 K
Iu

0.01

0.001

0.0001
le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01 le+00
Time, sec

Figure 2.6. Comparison of present results with the numerical work of Dalle Donne and
Ferranti (1975). Liquid-vapor interface velocity versus time for sodium.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
47

1600

Bubble Surface Temperature vs. Time for Sodium, Psys=50 kPa, Tsat=106(> K

1500 - Symbols : Dalle Donne's numerical work (1975)


Solid Lines : This work (computed)

1400
Superheat=362 K

2 1300

Superheat=201 K
1200

1100
Superheat=42 K

1000
le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01 le+00
Time, sec

Figure 2.7. Comparison of present results with the numerical work of Dalle Donne and
Ferranti (1975). Liquid-vapor interface temperature versus time for sodium.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
48

Bubble Radius vs. Time for Sodium


le+02
Superheat=201 K
le+01 Psys=50 kPa
Tsat=1066 K

le+00
Symbols : Dalle Donne's numerical work (1975)
le-01

le-02

le-03

Disturb.=le-0
le-04 isturb.=le-4

le-05

le-06
Disturb.=le-8

le-07
le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01 le+00
Time, sec

Figure 2.8. Comparison of present results with the numerical work of Dalle Donne and
Ferranti (1975). Effect of magnitude of temperature disturbance on bubble growth for
sodium.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
49

0.002
Radius vs. Time for water, Psys=l atm, Tsat=100 C
0.0018

0.0016
Symbols: Dergarabedian (1953)

0.0014 Solid lines: computations

0.0012
Radius, meter

4.5 C
0.001

0.0008
3.1 C
0.0006

0.0004 2.1 C

0.0002

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016


Time, sec

Figure 2.9. Comparison of computations with the measurements of Dergarabedian


(1953). Bubble radius versus time for water.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
50

1
I Bubble radius vs. Time for water, Psys=0.488 atm

Computed,
superheat=36 oC
0.1
Kosky's Exp. Data
(1968), 36 oC
Radius, meter

0.01

0.001

0.0001
0.0001 0.001 0.01
Time, sec

Figure 2.10. Comparison of computations with the measurements of Kosky (1968).


Bubble radius versus time for water.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
51

0.0003
Bubble Radius vs. Time for Water, Psys=l atm

0.00025

Computed,
superheat=0.8 C,
0.0002 Disturb.=le-4 C

Computed,
Radius, meter

superheat=1.0 C,
Disturb.=le-9 C
0.00015
Dergarabedian's Exp.
Data (1960), 0.8 C

Dergarabedian's Exp.
0.0001 Data (1960), 1.0 C

0.00005 x.'

0 0.005 0.01 0.015 0.02 0.025 0.03


Time, sec

Figure 2.11. Comparison of computations with the measurements of Dergarabedian


(1960). Bubble radius versus time for water with low initial liquid superheats.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
52

0.0018
' Bubble radius vs. time for water, Psys=l atm, to=6.5 ms
0.0016
Computed,
superheat=3.61 oC
0.0014 Florschuetz's Exp.
Data (1969)
0.0012

U
0 0.001
d3
e
GO
3

1 0.0008
os

0.0006

0.0004

0.0002

-0.005 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04


Time, sec

Figure 2.12. Comparison of computations with the measurements of Florschuetz (1969).


Bubble radius versus time for water. An arbitrary shift of 6.5 ms is made in the time scale.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
53

0.016
Bubble Radius vs. Time for Water

0.014 Computed

Lien's Exp. Data


(1969)
0.012

' Psys= 12.59 kPa


Superheat= 10.67 C
0.01 ~ Tbulk=61.06 C
Radius, meter

0.008 Psys=1.26 kPa


s Superheat= 15.74 C
Tbulk=26.15 C
0.006

0.004

Psys=38.66 kPa
0.002 Superheat=9.0 C
Tbulk=84.08 C

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016
Time, sec

Figure 2.13. Comparison of present numerical calculations with measurements of Lien


(1969) over a wide range of pressure. Bubble radius versus time for water.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
54

Bubble Radius vs. Time for R-l 13


0.025

This work (1993)

This work (1993)


0.02
Bohrer's Data
(1973)

Bohrer's Data
(1973)
0.015
U Superheat=48.21 C
u Tbulk=22.22 C, Psys=3.66 kPa
E

0.01

0.005

/Superheat=34.12 C
Tbulk=22.78 C, Psys=8.46 kl'a

0 0.001 0.002 0.003 0.004 0.005 0.006 0.007


Time, sec

Figure 2.14. Comparison of computations with measurements of Bohrer (1973) for R-


113 at low pressures and large liquid superheats.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
55

le+02
rBubble radius and velocity vs. time for water, Psys=l atm le+02

le+01
le+00

le+00 le-02
Dist. le-1
(Srheat=3.1,0C le-04
le-01

Interface velocity, m/s


le-06
le-02 Dist. le-5
le-08

le-03 le-10
Dist. le-9
le-04 Sup] it=3.1 oC le-12

le-14
le-05
Disturb.=le-1 le-16

le-06 Disturb.= le-5


le-18
Superheat=36 oC
le-07 le-20
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.15. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for water at atmospheric pressure. Radius and liquid-vapor interface
velocity.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
56

le+02 160
iBubble radius and temperature vs. time for water, Psys=l atm

le+01 Superheat=36 C 140


)ist. le-1
-Dist. le-5
le+00
' Dist. le-9 120

le-01
Superheat=3.1 C 100

Interface temperature,
le-02

le-03

le-04 SUDJ it=3.1 C

le-05
-Disturb.=le-1
le-06 Disturb.= le-5

Superheat=36 C >isturb.= le-9


le-07
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.16. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for water at atmospheric pressure. Radius and liquid-vapor interface
temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
57

le+02 2e+04
jBubble radius and acceleration vs. time for water, Psys=l atm
-Superheat=3.1 oC
le+01 r Dist. le-5

le+00

le+04
le-01 Dist. le-1 DisU.le-9
oe
QJ
4u.>
cd
E
le-02 13
CJ
o oCO
CO
OS <
OL>
le-03 <D
0e+00

le-04

Disturb.= le-
le-05

le-06 -le+04
le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.17. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for water at atmospheric pressure. Liquid-vapor interface acceleration, for
a low initial superheat of 3.1 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
58

le+02 6e+07
:Bubble radius and acceleration vs. time for water, Psys=l atm
Superheat=36 oC
le+01 5e+07

le+00 Dist. le-5 4e+07

le-01 3e+07

Interface acceleration, m/sA2


Dist. le- Djst. le-9

le-02 2e+07

le-03 le+07

le-04 0e+00

le-05 Disturb.=le-1 -le+07


-Disturb.= le-5
le-06 iisturb.= le-9 -2e+07

le-07 -3e+07
le-08 le-07 le-06 le-05 le-04 le-03
Time, sec

Figure 2.18. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for water at atmospheric pressure. Liquid-vapor interface acceleration, for
a high initial superheat of 36 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
59

le+04 2e+04
iBubble radius and acceleration vs. time for water, Psys=l atm
Superheat=3.1 C
le+03
Acceleration
le+02

le+01
le+04

le+00 co
a
03
Surface tension <lominated growth Uu*
3 le-01 U

' 3
2. Momentum dominated growth U
cd
CB <u
PS
le-02 3. Thermal diffusion dominated growth
0e+00
le-03

le-04

le-05 Radius

le-06 -le+04
le-05 le-04 le-03 le-02
Time, sec

Figure 2.19. Example of characterization of the three domains of vapor bubble growth.
Bubble radius and liquid-vapor interface acceleration vs. time, for water at atmospheric
pressure and uniform initial superheat of 3.1 C. Initial temperature disturbance of 0.1 C
used here.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
60

0.03
Bubble radius vs. time for water, Psys=l atm, superheat=36 C

0.025
Rayleigh (1917)

0.02

Prosperetti & Plesset (1978)


Radius, meter

0.015

Plesset & Zwick (1954)

0.01
Scriven (1959)

0.005
!ikicetal.(1970)
This work (1993)

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Time, sec

Figure 2.20. Comparison of present numerical calculations with various analytical


solutions. Water at atmospheric pressure uniformly superheated to 36 C, Ja=l 10.21.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
61

le+01
; Bubble radius vs. time for water, Psys=l atm, superheat=36 C

le+00

le-01 Rayreigh (1917)

le-02 Plesset & Zwick (1954)

Scriven .(1959)
le-03

le-04

Prosperetti & Plesset (1978)


le-05
ikic et al.(1970)
le-06
lis work (1993)
le-07
le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.21. Comparison of present numerical calculations with various analytical


solutions. Water at atmospheric pressure uniformly superheated to 36 C, Ja=l 10.21.
Vapor bubble radius versus time. The early period of Figure 2.20 is amplified with the use
of logarithmic axis.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
62

0.02 30
Radius and Interface Temperature vs. Time for Water
0.018 28
Superheat=15.74 C
Psys=1.26kPa
0.016 Tbulk=26.15 C 26

0.014 24
Liquid-vapor interface temperature oU
u 0.012 22 23
0> Plessetand Zwick (1954) C 3
0) dQ.>
t->

3 0.01
20 E
Rayleigh (1917) o
o
0.008 Scriven (1959) 18 1o
sH
H
0.006 Mikic et aL (1970) 16

Lien's Data (1969)


0.004 14

0.002 12
This work (computed)
10
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Time, sec

Figure 2.22. Comparison of the present numerical calculations with various analytical
solutions and the measurements of Lien (1969) for very low system pressure (P=1.26 kPa)
with water. Vapor bubble radius and liquid-vapor interface temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
63

le+04 5e+03
Interface Velocity and Acceleration vs. Time for Water

le+02 Superheat= 15.74 C, Acceleration


Psys= 1.262 kPa, 4e+03
Tbulk=26.15 C
le+00
3e+03

Interface Acceleration, m/s2


le-02

2e+03

le-04

le+03
le-06

0e+00
le-08

le-10 -le+03
le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.23. Liquid-vapor interface velocity and acceleration from the present numerical
calculations for conditions of Figure 2.22.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
64

0.006 120
Radius and Interface Temperature vs. Time for Water
Superheat=15.74 C
Rayleigh (1917) Psys=101.325 kPa 110
0.005 Tbulk=l 15.75 C
Tsat=100 C

100
0.004 Liquid-vapor interface temperature

Interface temperature,
Plesset and Zwick (1954)
s 0.003 Scriven (1959)

Mikic et al. (1970)


0.002

This work (computed)

0.001

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
Time, sec

Figure 2.24. Comparison of the present numerical calculations with various analytical
solutions for relatively high pressure (P=l atm.=101.325 kPa) with water. Vapor bubble
radius and liquid-vapor interface temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
65

le+04 5e+06
Interface Velocity and Acceleration vs. Time for Water

Superheat= 15.74 C,
le+02 Psys=101.325 kPa, 4e+06
Tbulk=l 15.75 C

le+00 3e+06

Interface Acceleration, m/s2


Interface velocity, m/s

Velocity
le-02 2e+06

le-04 Acceleration le+06

le-06 0e+00

le-08 -le+06
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.25. Liquid-vapor interface velocity and acceleration from the present numerical
calculations for conditions of Figure 2.24.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
66

0.001
Bubble radius vs. time for water, Psys=l atm, superheat=0.8 C
0.0009
x: Dergarabedian (1960)
0.0008

0.0007 ayleigh (1917)

0.0006 Prosperetti & Plesset (1978)

s 0.0005
Scriven (195!
0.0004

0.0003

0.0002 [ikic et al.(1970)


Plesset & Zwick (1954)
0.0001
x x x x x- This work (1993)
0
0 0.005 0.01 0.015 0.02 0.025 0.03
Time, sec

Figure 2.26. Comparison of present numerical calculations with various analytical


solutions. Water at atmospheric pressure uniformly superheated to 0.8 C, Ja=2.38.
Vapor bubble radius versus time.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
67

le+01 le+02
Bubble radius and velocity vs. time for R113, Psys=l atm, Tsat=47.55 oC
le+00
le+00
le-02

le-01 Disturb, le-1


le-04

le-02 le-06

Interface velocity, m/s


Disturb, le-5
le-08
le-03
le-10

le-04 - Disturb, le-9


le-12

le-14
le-05 Dist. le-1

Dist. le-5 le-16


Superheat=3.1 C list, le-9
le-06
le-18

le-07 le-20
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.27. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for R -l 13 at atmospheric pressure. Radius and liquid-vapor interface
velocity

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
68

12 2.70665E-06
Interface velocity vs. time forR-I I3,Psys=l atm,
10 superheat=3.1 C, Disturb.=le-1 C

8 Velocr
2.70655E-06

2.70645E-06
2 Radius

2 2.70635E-06
IE-08 IE-07 IE-06 IE-05
Time, sec

(a)

12 7 .2 6 6 1 S
Interface velocity and pressure and surface tension term in Equation
Psys=l atm, superheat=3.1 C, disturb.=le-l C
10
Surface tension term 7 .2 6 6 0 5

8
7 .2 6 5 9 5
Velocity
I a p*
r>> 6
i 7 .2 6 5 8 5 e e
wC
> 4 Pressure term s
o
7 .2 6 5 7 5
2
Icu
0 7 .2 6 5 6 5

2 7 .2 6 5 5 5

IE -0 7 IE -0 6

Time, sec

(b)

Figure 2.28 (a) & (b). a) Amplification of the embryo bubble radius and velocity in Figure
2.27 for the early period of growth, b) Comparison of the time varying surface tension
and pressure terms in Equation (2.22) with the liquid-vapor interface velocity.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
69

le+01 100
Bubble radius and temperature vs. time for R113, Psys=l atm, Tsat=47.55 C

le+00 Superheat=36 C
Disturb.le-1
le-01 Disturb, le-5
Disturb, le-9

Interface temperature, oC
le-02

le-03

le-04

le-05 Dist. le-1


)ist. le-5
Superheat=3.1 C ist. le-9
le-06

le-07
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.29. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for R -l 13 at atmospheric pressure. Radius and liquid-vapor interface
temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
70

le+01 2e+04
Bubble radius and acceleration vs. time for R113,
- Psys=l atm, Tsat=47.55 oC, superheat=3.1 C
le+00

le-01

le+04

Interface acceleration, m/s2


Disturb, le-1
le-02
a>
cj
S
cso le-03
co
OS
le-04
0e+00

le-05 Dist. le-1


Dist. le-5
Dist. le-9
le-06

le-07 -le+04
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.30. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for R-l 13 at atmospheric pressure. Radius and liquid-vapor interface
acceleration for a low initial superheat of 3.1 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
71

le+01 6e+07
iBubble radius and Acceleration vs. time for R113, Psys=l atm,
I Tsat=47.55 C, superheat=36 C
Disturb..le-5 5e+07
le+00

4e+07
le-01 Disturb, le-1

3e+07

Interface acceleration, m/s2


le-02

2e+07
le-03
le+07

le-04
0e+00

le-05
-le+07

le-06 Dist. le-1


-2e+07
Dist. le-5
Dist. le-9
le-07 -3e+07
le-08 le-07 le-06 le-05 le-04 le-03
Time, sec

Figure 2.31. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for R-l 13 at atmospheric pressure. Radius and liquid-vapor interface
acceleration for a high initial superheat of 36 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
72

2.0e-04 0.007
Radius and Interface velocity vs. Time for R-l 13, Psys=l atm
1.8e-04 0.006

1.6e-04 0.005

1.4e-04 0.004

Interface velocity, m/s


Ui 1.2e-04 0.003
O 1.0 C
OJ
E
'0.8 C
vsT 0.002

0.001

6.0e-05 1.0 C

4.0e-05 - 0.001
0.8 C
2.0e-05 - 0.002

0.0e+00 -0.003
0 0.005 0.01 0.015 0.02 0.025 0.03
Time, sec

Figure 2.32. Sensitivity of early R-l 13 vapor bubble growth rates to bulk liquid superheat
at low levels of superheat. An initial temperature disturbance of le-2 C was used for
both superheats of 0.8 C and 1.0 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
73

0.025
Bubble Radius vs. Time for R-l 13
- Psys=8.46 kPa, Superheat=34.11 C, Tbulk=22.78 C

0.02

Rayleigh (117)

0.015

-Plesset & Zwick (1954) Bohrer's Experimental Data (1973)


0.01

This work
0.005

Mikic et al (1970)

0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007
Time, sec

Figure 2.33. Comparison of present numerical calculations with various analytical


solutions and one set of measurements. R -l 13 at low pressure of 8.5 kPa (1.23 psia) with
initially uniform superheat of 34 C. Ja = 396.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
74

0.005

Bubble radius vs. time for R113, Psys=l atm,


0.0045 Tsat=47.55 C, superheat=36 C

0.004

0.0035

0.003
Radius, meter

Scriven,(1959)
0.0025
Plesset & Z|wick (1954)

0.002
This work (1993)

0.0015
Mikic et al. (1970)

0.001

0.0005

0 0.002 0.004 0.006 0.008 0.01


Time, sec

Figure 2.34. Comparison of present numerical calculations with various analytical


solutions. R-l 13 at atmospheric pressure with initially uniform superheat of 36 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
75

le+02
Bubble radius and velocity vs. time for sodium, le+02
Psys=l atm,Tsat=873.85 oC
le+01 le+00

le-02
le+00
le-04
Disturb, le-1
le-01

Interface velocity, m/s


le-06

le-08
le-02 Disturb, le-5

le-10
le-03
le-12
Disturb, le-9

le-04 le-14

Dist. le-1 le-16


Dist. le-5
le-05
^ ^ D is t. le-9 le-18
Superheat=36 oC
le-06 le-20
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.35. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for sodium at atmospheric pressure. Radius and liquid-vapor interface
velocity.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
76

le+02 940
Bubble radius and temperature vs. time for sodium,
Psys=l atm,Tsat=873.85 C 930
le+01 Disturb., le-1 920
Disturb, le-5
isturb. le-9 910
le+00 Superheat=36 C
900

890
le-01

Interface temperature,
880

le-02 Superheat=3.1 C 870

860
le-03
850

840
le-04
830
Dist. le-1
Dist. le-5 820
le-05
~~Dist. le-9 810

le-06 800
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.36. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for sodium at atmospheric pressure. Radius and liquid-vapor interface
temperature.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
77

le+02 5e+03
Bubble radius and acceleration vs. time for sodium,
Psys=l atm,Tsat=873.85 C
4e+03
le+01
Disturb. le.-9
Disturb, le-5 3e+03
le+00
Disturb, le-1

Interface acceleration, m/s2


2e+03
fe le-01
U
s
{/T le+03

1 le-02
0e+00

le-03 Dist. le-1


Dist. le-5 -le+03
Dist. le-9
le-04
-2e+03

le-05 -3e+03
le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.37. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for sodium at atmospheric pressure. Radius and liquid-vapor interface
acceleration for a low initial superheat of 3,1 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
78

le+02
Bubble radius and acceleration vs. time for sodium,
le+06
Psys=l atm,Tsat=873.85 C
le+01
Disturb, le-5
9e+05

le+00 Disturb., le-9


Disturb. 7e+05

Interface acceleration, m/s2


le-01
5e+05

s le-02
a 3e+05

le-03
le+05

le-04 - Superheat=36 C
-le+05
Dist. le-1
- Dist le-5
le-05 -3e+05
list, le-9
: Superheat=36 C

le-06 -5e+05
le-08 le-07 le-06 le-05 le-04 le-03
Time, sec

Figure 2.38. Influence of magnitude of initial temperature disturbance on the early vapor
bubble growth for sodium at atmospheric pressure. Radius and liquid-vapor interface
acceleration for a high initial superheat of 36 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
79

0.06
Bubble Radius vs. Time for sodium, Psys=l atm, Superheat=36 C

0.05 Rayleigh 1917)

0.04
Pleset & Zwick (1954)

Scriven (1959)

e 0.03 This work (computed)

0.02
Mikic at al. (1970)

0.01

0
0 0.002 0.004 0.006 0.008 0.01
Time, sec

Figure 2.39. Comparison of present numerical calculations with various analytical


solutions. Sodium at atmospheric pressure uniformly superheated to 36 C, Ja=32.28.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
80

- Bubble radius vs. time for sodium


0.09
o: Mikic et al. (1970)
: Present work (1993)
0.08

=0.2 atm
0.07 Superheat=200

0.06

s 0.05

0.04 \ P = 6 atm (607.95 k Pa)


yS'Superheat=200 C

0.03 P=0.2 atm (20.26 kPa)


Superheat=36 C
0.02

0.01 P=6 atm (607.95 k 3a)


Superheat=36 C

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014


Time, sec

Figure 2.40. Comparison of present numerical caculation with analytical solution of Mikic
at al (1970). Vapor bubble radius versus time for sodium.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
81

le+02
: Bubble radius and velocity vs. time, le+02
- Psys=l atm (101.325 kPa), superheat=3.1 C
le+01 = Water___ le+00

' Tsat=100 C (water) le-02


le+00 Freon 113
[ Tsat=47 C (R-l 13)
j Tsat=873 C (sodium) le-04

Interface velocity, m/s


le-01 Sodium
le-06

le-08
le-02
le-10
le-03
le-12

le-04 le-14

le-16
le-05 -Water.
le-18
Freon-14-^
le-06 le-20
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.41. Comparisons of bubble radius and liquid-vapor interface velocity vs. time for
water, R-l 13 and sodium all at atmospheric pressure and a low superheat of 3.1 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
82

le+02
Bubble radius and velocity vs. time, le+02
Psys=l atm (101.325 kPal. superheat=36 C
le+01 le+00

Freon 113
le+00 le-02

le-04
le-01

Interface velocity, m/s


Superheat=36 C le-06
le-02
le-08
Sodium
I le-03 le-10

le-12
le-04

Sodiui le-14
le-05
le-16

le-06 Water. le-18


Freon 113
le-07 le-20

le-08 le-07 le-06 le-05 le-04 le-03 le-02


Time, sec

Figure 2.42. Comparisons of bubble radius and liquid-vapor interface velocity vs. time for
water, R -l 13 and sodium all at atmospheric pressure and a high superheat of 36 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
83

le+02 2e+04
Bubble radius and acceleration vs. time,
Psys=l atm (101.325 kPa), superheat=3.1 C
le+01

le+00 le+04

Interface acceleration, m/sA2


le-01 Sodium
Freon 113 Water
0>

0e+00
a le-02
'B
cd
OS
le-03

le-04 -le+04

le-05 Water.
Freon 113

le-06 -2e+04
le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.43. Comparisons of bubble radius and liquid-vapor interface acceleration vs.
time for water, R -l 13 and sodium all at atmospheric pressure and a low superheat of 3.1
C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
84

le+02 6e+07
Bubble radius and acceleration vs. time,
Psys=l atm (101.325 kPa), superheat=36 5e+07
le+01

4e+07
le+00
Freon 113 Water 3e+07
le-01

Interface acceleration, m/sA2


2e+07
U
2B le-02 Sodium le+07
6
C fl
.2
106 le-03 0e+00

-le+07
le-04

I Sodii -2e+07
le-05
-3e+07

le-06 Water. -4e+07


Freon 113
le-07 -5e+07
le-08 le-07 le-06 le-05 le-04 le-03 le-02
Time, sec

Figure 2.44. Comparisons of bubble radius and liquid-vapor interface acceleration vs.
time for water, R-l 13 and sodium all at atmospheric pressure and a high superheat of 36
C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
85

Critical radius vs. superheat Critical radius vs. superheat


for sodium for R-l 13

le-04
3
'5
g le-05 S

c
u

0 20 40 60 80 100 120 0 20 40 60 80 100 120


Superheat, C Superheat, C

(a) (b)

Critical radius vs. superheat Critical radius vs. superheat


for water at atmospheric pressure

le-04 le-04

g le-05 1 le-05 Sodium

Water

0 20 40 60 80 100 120 0 20 40 60 80 100 120


Superheat, C Superheat, C

(C) (d)

Figure 2.45. Critical radius versus superheat {Tbulk - TSttl) for sodium, R -l 13 and water,
computed by the present model, as a function of pressure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
86

le-02 le-02
Delay time vs. superheat Delay time vs. superheat
le-03 Vs. for sodium le-03 forR -l13

O 0D le-04
0M> le-04 0.1 atm Vi

| le-05 1 le-05

J le-06 i le-06
!.1 atm
P=4 atm Q le-07
le-07
itm
le-08 le-08

le-09 le-09
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Superheat, C Superheat, C

(a) (b)

le-02 le-02
Delay time vs. superheat i Delay time vs. superheat

le-03 \ for water le-03 P\at atmospheric pressure

O le-04 o0> le-04


Vi
Sodium
tf le-05 | le-05
.s
jjj* le-06 atm i* le-06 .Water
"3
le-07 atm le-07

le-08 P=4 le-08


R-l 13
le-09 le-09
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Superheat, C Superheat, C

(C) (d)

Figure 2.46. Delay time versus superheat (Tbulk ~Tsal) for sodium, R-l 13 and water,
computed by the present model, as a function of pressure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
87

40 40
Maximum velocity vs. superheat Maximum velocity vs. superheat
35 for sodium 35 for R-l 13
Max. interface velocity, m/s


30 30
25 1"o 25
>
20 0) 20 P=4aj
iM
nu 15
15
10 10
(rfatm
5 5
0.1 atm
0 0
20 40 60 80 100 120 20 40 60 80 100 120
Superheat, C Superheat, C

(a) (b)

40
Maximum velocity vs. supedjeat Maximum velocity vs. superheat
for water 35
Max. interface velocity, m/s

1 at atmospheric pressure
P=4atm 30
1 25
T>B
D 20
<t:D 15

0.1 atm 10
iodium
5
0
20 40 60 80 100 120 0 20 40 60 80 100 120
Superheat, C Superheat, C

(c) (d)

Figure 2.47. Maximum liquid-vapor interface velocity versus superheat (Thulk - Tsat) for
sodium, R -l 13 and water, computed by the present model, as a function of pressure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
88

le+12 le+12
Acceleration vs. superheat Acceleration vs. superheat
for sodium forR-113
le+10 | le+10 P=4;
an
o
B
a le+08 P=4atrrLa le+08
E 0.1 atm
/ y.-'T'atm O
U
le+06 / X" le+06
uu
/ /P atm
ts / / / *ts
p le+04 1 le+04
Y
le+02 le+02
20 40 60 80 100 120 20 40 60 80 100 120
Superheat, C Superheat, C

(a) (b)

le+12 le+12
Acceleration vs. superheat Acceleration vs. superheat
for wat at atmospheric pressure
| le+10 3* le+10 Water
a 0c R-l 13
| le+08 1o le+08
C
UJ
C3 le+06 le+06 Sodium

* le+04 ~ le+04

le+02 le+02
20 40 60 80 100 120 20 40 60 80 100 120
Superheat, C Superheat, C

(C) (d)

Figure 2.48. Maximum liquid-vapor interface acceleration versus superheat (Tbtllk Tmt)
for sodium, R-l 13 and water, computed by the present model, as a function of pressure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
89

0.08 0.08
Radius at time of 0.01 sec vs. Radius at time of 0.01 sec vs.
superheat for sodium s superheat for R113
0.06 0.06
O o
o
0.04 0.1 atm / / 2 0-04
6
/x
CO
/ 4 a tm /^ 4 a tm CB (U alnto
0.02 1 0.02
CO
'IB3 1 atpi.
& OS
____ v at m
lEI, , , , . , _ , _ ff , -fl , .
0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Superheat, C Superheat, C

(a) (b)

0.08 Radius at time of 0.01 sec vs. 0.08


Radius at time of 0.01 sec vs
superheat for water . 0.07 superheat at atmospheric press lire
ou
0.06 cfl 0.06 yP
O o 0.05 Soditun
o
o 0.04 o i^ tn r0 "o 0.04
o U
E E
a 0.03
cs
3 0.02 V
aI 0.02 / ^ W a te r*
CO / '3
cs 0.01
/ P=4yitm OS

0 0 . . . '
20 40 60 80 100 120 20 40 60 80 100 120
Superheat, C Superheat, C

(C) (d)

Figure 2.49. Bubble radius at an arbitrary time of 0.01 seconds versus superheat
(Tbulk - Tsat) for sodium, R-l 13 and water, computed by the present model, as a function
of pressure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 3

VAPOR BUBBLE DYNAMICS IN NON-UNIFORM


TEMPERATURE FIELDS

Vapor bubble growth and collapse in non-uniform temperature fields are modeled

and solved numerically. The models are considered in two parts: a spherical bubble

model and a hemispherical model. The former is applicable to bubble dynamic processes

such as in cavitation or ultrasonic cleaning devices, while the latter is particularly

applicable to pool boiling. The experimental observation of vapor bubble rebounds

following successive collapses are predicted by the present spherical numerical model,

which also provides the transient vapor pressure in the bubble and the transient

temperature distributions surrounding the bubble. The hemispherical bubble model cannot

give a precise prediction of the bubble growth for pool boiling, but rather provides upper

and lower bounds. The model was tested by comparison of the computed results with

pool boiling experiments conducted as part of the present research in microgravity, which

provides a known initial temperature distribution at nucleation in the absence of any liquid

disturbances. It was demonstrated that the concept of upper and lower bounds for bubble

growth has merit.

90

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
91

3.1 Literature Survey

Vapor bubble dynamics in non-uniform temperature fields have extensive

applications in various diverse fields: pool boiling, forced convection boiling, cavitation,

ultrasonic cleaning devices, bubble jet printer industries and acoustics. A considerable

number of publications related to vapor bubble growth on a heater surface may be found

since the early work of Nukiyama (1934), while relatively few publications include both

vapor bubble growth and collapse. Gunther (1950) and Ellion (1953) observed vapor

bubbles growing and collapsing on a heater surface in subcooled nucleate boiling of water,

by means of high speed photography. The data of Ellion (1953) show considerable

divergence due to the inherent ramdomness of the temperature distributions in the liquid in

the vicinity of the heater surface. Bankoff and Mikesell(1958) derived a convenient

normalized formula for the bubble growth and collapse from the equation of motion

formulated by Lord Rayleigh (1917), applicable only to highly subcooled nucleate boiling.

Zuber (1961) also proposed a normalized formula for the vapor bubble growth and

collapse, considering only the energy balance at the liquid-vapor interface. The heat
transfer at the liquid-vapor interface to the bulk liquid, qb, was introduced as an unknown

quantity, and the bubble growth or collapse can not be predicted unless qb is given. Both

Bankoff and Mikesell (1958) and Zuber (1961) attempted to fit their respective formulas

to the experimental data of Ellion (1953), using the experimental maximum radius. It was

noted as a result that as the subcooling increased, the bubble growth and collapse were

governed more by the liquid inertia than by the heat transfer process at the liquid-vapor

interface.

Florschuetz and Chao (1965), considering only the vapor bubble collapse,

normalized the equation of motion derived by Lord Rayleigh (1917) and the zero order

solution for liquid-vapor interface temperature derived by Plesset and Zwick (1952), and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
92

solved these numerically by combining the two equations. A closed-form expression for

each was suggested for the inertia controlled and the heat transfer controlled collapses,

classifying the vapor bubble collapse mode by means of a dimensionless parameter.

Bubble collapse experiments were conducted with water in a vessel under free fall

conditions in order to eliminate distortions due to buoyancy. The experimental results

were compared with the approximate prediction for the heat transfer controlled collapse,

with fair agreement.

Han and Griffith (1965) derived a closed-form expression for the bubble growth,

with quasi steady pool boiling in earth gravity, considering one-dimensional heat

conduction surrounding the bubble, with the bubble beginning to grow from an arbitrary-

sized cavity. Mikic et al (1970) predicted the bubble growth in a non-uniform temperature

field by correcting an initial uniform superheat model, considering only heat diffusion. It is

assumed again that the liquid at a uniform temperature comes into contact with the heater

surface at a given temperature, and after a waiting time the vapor bubble forms and grows

in the non-uniform temperature field resulting during this waiting time.

The vapor bubble dynamics with nucleation in non-uniformly superheated liquids is

more complicated than that in uniformly superheated liquids since the bubble dynamics is

not only controlled by the initial temperature distribution, but is also governed by the

dynamics of the process itself. Bubble dynamics has been of special interest in the field of

vapor cavitation and ultrasonics, where the concern is more with dynamic effects rather

than thermal effects. Lauterbom and Hentschel (1985) used a laser beam to heat water

locally by focusing at a desired location near a solid wall. This produced homogeneous

nucleation and permitted the study of the vapor bubble collapse near a solid wall. The

photographs reveal that the laser-produced bubble grows and collapses, with a rebound

occurring during the first collapse and producing a long jet directed toward the solid wall.

This jet provides the mechanism for the so-called cavitation damage. In the field of

ultrasonic cleaning, on the other hand, the jet directed toward the solid wall is utilized to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
93

clean the surface. Shima et al. (1988) measured collapsing vapor bubble pressures in the

vicinity of atmospheric pressure by means of a pressure transducer installed on a solid

wall, with the bubble generated by means of an electric spark near the solid wall. The

impulsive pressures measured during collapse covered a wide range of 1 ~ 20 MPa,

depending on the bulk temperature and the distance of the spark from the solid wall.

In general, measurements of vapor bubble growth and collapse as a result of heat

transfer are quite non-reproducible. One of the reasons for this non-reproducibility is the

randomness of the temperature distribution in the vicinity of the growing vapor bubble.

The randomness results from the turbulence associated with the natural and/or forced

convection, and from the liquid agitation produced by adjacent or prior bubbles. In order

to eliminate the randomness arising from natural convection, transient pool boiling

experiments were conducted under microgravity conditions using the 132 m drop tower at

the NASA Lewis Research Center, with transient measurements of the mean heater

surface temperature and far-field pressure along with photographs of the vapor bubbles

taken at 400 fps. Results and associated analysis from these experiments are presented

here. The absence of natural convection and agitation from adjacent bubbles permits the

determination of the precise initial temperature distribution at nucleation. The present

experiments reveal two categories of bubble growth: one with bubbles having a smooth

surface, which produces moderate growth rates; the other with bubbles having a

roughened surface, which produces extremely large growth rates, which might be

classified as violent. The present work focuses on the dynamics of growing/collapsing

vapor bubbles with a smooth surface, arising from non-uniform initial temperatures

surrounding the bubbles.

To understand the mechanisms associated with the vapor bubble dynamics more

fully, a one-dimensional numerical model was formulated incorporating momentum,

energy, surface tension and viscosity simultaneously. The treatment for an initially

uniformly superheated bulk liquid is presented in Chapter 2. The model can also take into

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
94

account the vapor bubble growth and collapse arising from initially non-uniform

temperature fields, and predicts the dynamics following homogeneous nucleation. The

model can not predict the dynamics of vapor bubble growth with pool boiling from plane

heater surfaces since the dynamics of this configuration inherently involves two or three

dimensional considerations. However, as will be demonstrated, the combination of the

initially uniform and non-uniform superheated liquid models provide the upper and lower

bounds for predicting the vapor bubble growth and collapse dynamic processes. The

alternative available at present is to solve the three-dimensional case incorporating

momentum, energy, surface tension and viscosity simultaneously, with the three-

dimensional interface becoming a part of the solution process. Computational possibilities

for such are not yet available. Computations with the present model, on the other hand, is

fast even with currently available P.C.'s, and provides reasonable approximations to

observed behaviors.

Previous analytical solutions and experiments for vapor bubble dynamics in initially

non-uniform temperature fields have been surveyed. A considerable number of

publications for vapor bubble growth on a heater surface has been found for the period

following the work of Nukiyama (1934), but relatively few incorporated both bubble

growth and collapse. Gunther (1950) and Ellion (1953) observed bubbles growing and

collapsing on a heater surface in highly subcooled water, by means of high speed

photography. Typical experimental results of Ellion (1953) in Figures 3.1 and 3.2 show

the variability in the bubble dynamics arising from the inherent ramdomness of the

temperature distribution on the heater surface, which causes some difficulty when

attempting to model and hence understand the mechanisms involved. Nevertheless,

several representative analytical models associated with such data will be presented here.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
95

3.1.1 Inertia Controlled Bubble Growth and Collapse

Lord Rayleigh (1917) first formulated the equation of motion for a spherical

bubble in terms of the bubble radius and pressure difference as:

3 , P -P2
r r +- r 2 =^> (3.1)
2 p,

This applies to both bubble growth and collapse processes, depending on the sign
of Pv -P. A solution to Equation (3.1) was provided by Bankoff and Mikesell (1958),

with dRI dt = 0 when R = Rm and t = tm, where tm is the time at which the bubble reaches

its maximum radius (Rm), given by Equations (3.2) and (3.3). Effective values of Pv - Pm

were used for purposes of comparison with the data of Gunther(1950) and Ellion (1953).

However, it is not clear as to how these effective constant pressures were determined. It

is of course obvious that different values are necessary for the growth and collapse phases.

(3.2)

. t { 2 P -P \2 , , R
where \i/ = ------ *---- and c = (3.3)
Rm\. 3 p, o, , Rm

Zuber (1961) also gave a solution for Equation (3.1), for the collapse mode only,
with initial conditions of R(0) = Rm and R(0) = 0, as:

(3.4)

where tc is the time needed for the total collapse.

If only the collapse phase of the problem described by Equations (3.2) and (3.3) is

considered, with the same initial conditions used by Zuber (1961), it can be shown that

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
96

Equations (3.2) and (3.3) are identical to Equation (3.4), provided the lower limit of the

integral in Equation (3.2) is replaced by unity.

Equations (3.2) and (3.4) provide the descriptions of the bubble dynamics for the

so-called inertia controlled bubble collapse. As stated by Bankoff and Mikesell (1958),
Rm and tm in Equations (3.2) and (3.3) are empirical values, obtaining with highly

subcooled nucleate boiling. The solutions can give predictions only if Rm and tm are

known. Nevertheless, these indicate trends of the bubble growth and collapse as to

whether the collapse mode is controlled by heat transfer or liquid inertia.

3.1.2 Heat Transfer Controlled Bubble Growth and Collapse

Zuber (1961) proposed an approximate expression for the heat transfer controlled

growth and collapse in subcooled boiling, utilizing the energy balance at the liquid-vapor
interface in terms of qb, the heat transfer from the vapor interface to the bulk liquid, as:

( 3 ' 5 )

where b is the curvature correction factor, usually l< b < y [3 , and qh is an

unknown quantity.
Integrating Equation (3.5) assuming that qh remains constant yields:

, 2 A7c,p, , f, OlVtcoF
R =b yhzou 1---- 7^-------- r (3.6)
71 V v . 2k[Tw Tsat)

The maximum radius is reached when dR/dt=0 in Equation (3.5), resulting in:

k{Tw- T saI)
(3.7)
<lb

Substituting Equation (3.7) into Equation (3.6), the maximum radius is then:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
97

(3.8)

Equation (3.6) can now be rewritten in terms of the maximum radius and time

using Equations (3.7) and (3.8) as:

(3.9)

Equation (3.9) accounts only for the heat transfer controlled growth and collapse

because only heat transfer is considered in its derivation. Equation (3.9) itself can not give
a prediction for the growth and collapse unless qbis given. Zuber (1961) later postulated

the use of experimental values of Rm and tm in Equation (3.9). This has been somewhat

confusing for readers because Equations (3.7) and (3.8) were previously recommended for
Rm and tm. If Rm and tm as evaluated by Equations (3.7) and (3.8) are used in Equation

(3.9), it was found that the life time for the bubble growth and collapse is about an order

smaller, and the maximum radius several times smaller, than the experimental data of

Ellion (1953).

The individual vapor bubble growth/collapse data of Ellion (1953) in Figures 3.1
and 3.2 were normalized relative to Rm and tm, producing the data in Figures 3.3 and 3.4,

respectively. These are compared with the inertia controlled collapse of Equation (3.4)

and the heat transfer controlled growth and collapse of Equation (3.9), respectively. The

bubble growths in Figures 3.3 and 3.4 would seem to be controlled by the heat transfer

because of the reasonably good comparison with Equation (3.9). For the bubble collapse

phase, on the other hand, Figure 3.3 shows the measurements falling between the inertia

and the heat transfer controlled collapse models, while Figure 3.4 shows an obviously

inertia controlled collapse. It can thus be deduced that the higher the subcooling, the

greater the tendency for the collapse to be governed by liquid inertia effects.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
98

3.1.3 Vapor Bubble Collapse

The specific process of vapor bubble collapse was considered by Florschuetz and

Chao (1965). Distinctions between the inertia and heat transfer controlled modes, as well

as the intermediate case, were classified in terms of a dimensionless parameter which will

be described here. Vapor bubble collapse occurring in an initially uniformly subcooled

liquid is conceptually similar to the bubble growth in an initially uniformly superheated

liquid, so the equations derived for the uniform superheat case can be used directly by

simply changing the sign of the temperature difference or the pressure difference. Plesset

and Zwick (1952) formulated the zero order solution for the temperature at the moving

interface of a spherical bubble by assuming a thin thermal layer near the bubble, with the

result:

(3.10)

Florschuetz and Chao (1965) expressed Equations (3.1) and (3.10) in terms of a
dimensionless time (x,) and radius (y) to normalize the problem, as:

(3.11)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
99

and in terms of a dimensionless time (xH) as:

B!( n + | f )= ( f J(n,(e J - nw(T)) (3.13)

n (T \ 2 p , T2(2)Y(e)e(z) .
(3.14)

The nondimensional values are defined as:

( , 1/2

B = Jar P/ (3.15)
^ y ^sys K i^ b u tk )

at R Tl i - T bulk
*/ = xH= Jac r , Y = 0 = (3.16)
3 P/ , H K c % 1 R, T
l sat
Tbulk

j-j- _ Pys(t) Pyi^bulk) _ P/C/ ( ^ a, P/mlk) (3.17)


n Ja =
Psys ~ P v { P b u l k ) Pvi^bulk) /

Equation (3.12) becomes negligible when the dimensionless parameter B is


sufficiently large such that Qw becomes small and the vapor pressure remains constant, so

that the collapse will be essentially controlled by liquid inertia, hence the use of the symbol
x, for the dimensionless time in this case. Conversely, when the dimensionless parameter

B is sufficiently small, the left-hand side of Equation (3.13) becomes negligible such that
the heat transfer controls the collapse, hence the use of the symbol xH for the

dimensionless time in this case. The properties in the dimensionless parameter B are

evaluated at the saturation temperature.

3.1.3.1 Liquid inertia controlled collapse

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
100

With Qw in Equation (3.12) negligibly small and assuming the pressure difference

in Equation (3.11) to be constant, solving Equation (3.11) provides an equation for the

inertia controlled collapse as:

(3.18)

where x, is defined in Equation (3.16). This is identical to the result of Zuber

(1961) given by Equation (3.4).

3.1.3.2 Heat transfer controlled collapse

If a heat transfer controlled collapse is considered, the inertia term in Equation

(3.13) becomes negligible. Substituting Equation (3.14) into the right-hand side of
Equation (3.13) with 11^ = 1, following the transformation of variables, yields finally:

(3.19)

where xH and y are defined in Equation (3.16).

It is to be emphasized here that Equation (3.19) is an approximate solution and

takes account for the heat transfer controlled collapse. The complete neglect of liquid

inertia effects hence renders the solution physically inapplicable at the very beginning of

the collapse process.

Equations (3.11) and (3.12) are applicable only for bubble collapse, and thus the

bubble growth and collapse can not both be described simultaneously here, since Equation

(3.12) was derived assuming a thin thermal boundary layer. This means that non-uniform

temperature fields inherently can not be taken into account by this solution. Equations

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
101

(3.11) and (3.12) were solved numerically by Florschuetz and Chao (1965) for bubble

collapse only, and are plotted in Figure 3.5 together with the result of the inertia

controlled collapse given by Equation (3.18), for three different values of the

dimensionless parameter B. It is seen that at B=0.3 the numerical result of Equations

(3.11) and (3.12) are in good agreement with the analytical approximation of Equation

(3.18) for the inertia controlled collapse, while heat transfer effects become significant

later in the collapse period for smaller values of B. Numerical solutions of Equations

(3.13) and (3.14) are plotted in Figure 3.6 together with the analytical approximation for

the heat transfer controlled collapse given by Equation (3.19). It is noted that the bubble

collapse curves converge to the heat transfer controlled collapse approximation of

Equation (3.19) as the dimensionless parameter B decreases. At B=0.02 the bubble

collapse curve almost coincides with Equation (3.19), which supports the approximation

for the heat transfer controlled collapse. It is noted that the curve in Figure 3.6 for B=0.1

shows an oscillation as the bubble collapses. Similar oscillations were observed in the

experiments conducted by Florschuetz and Chao (1965).

3.1.4 Vapor Bubble Growth on a Heater Surface

Han and Griffith (1965) studied vapor bubble growth at a low heat flux on a heater

surface, considering one-dimensional transient heat conduction surrounding the bubble,

including a contact angle at the three phase line on the heater surface. The approximate

solution proposed appears to provide the solution for the bubble dynamics for the three

different cases: superheated, saturated and subcooled liquids. The normalized form of the

solution is:
1/2
r - r
, <P<W, T
e <Pv P A e. 4<PvPvV*

(3.20)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
102

where r = , x (3.21)
rc 8
/zv=heat transfer coefficient from heating surface to the vapor bubble.
8 = jK atw , (p=contact angle

(pc=curvature factor where 1 < (pc < 3I/2,

f _ sin2 (p^
cpx=surface factor .^ + (^0S(Pj , <pfr=base factor

2 + cos(p(2 + sin2cp)
(pv=volume factor

s ~ Tsat , 0CTbulk Tm

Thu[k)Tsalo
1 1 - (3.22)
(TW- Tsa ,f5PVhfS

Equation (3.22) provides the minimum and maximum cavity radii, and any cavity

radius between these can qualify as a potentially active nucleation site. Han and Griffith

(1965) also conducted pool boiling experiments at low levels of heat flux, and

demonstrated that Equation (3.20) approximately predicts the bubble growth. Equation

(3.20) does not include the liquid inertia effects, since the equation of motion was not

incorporated in the analysis, so that Equation (3.20) is valid only for heat transfer

controlled growth and collapse.

Mikic et al. (1970) considered the problem of bubble growth in non-uniform

temperature fields on a heater surface. Corrections were applied to the initial uniform

superheat model, which consists basically of the following two steps: (i) liquid at a
uniform temperature (Tbulk) comes into contact with a surface at wall temperature (Tw) and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
103

(ii) after a waiting time (tw) the bubble forms and grows in the non-uniform temperature

field produced during the waiting time ( tw). The result is:

T Tbulk fc
i (3.23)
T ~Tsat I M r _
) U+

where R+= -rJ-r t+= (3.24)


B2 / A B2 / A 2

, 1/2
1/2
_ Tmt)hfgPv
A= ; B = \ Ja2a,
3Tsa,Pi K

Equation (3.23) thus would appear to describe the vapor bubble growth and

collapse on a heater surface, including heat diffusion effects, although no comparisons

with experiments for the growth and collapse were given. The dimensional version of

Equation (3.23) was given earlier by Mikic and Rohsenow (1969).

3.2 Modeling

A number of analytical approximations for vapor bubble growth in uniformly

superheated liquids have been proposed, while a relatively small number of approximation

models exist for non-uniform temperature fields because of its complexity. In order to

study bubble dynamics in non-uniform temperature fields, a Fortran program was written

to model the one-dimensional spherical vapor bubble dynamics in non-uniform

temperature fields, identical to the initial uniform superheat model except for the initial

condition. The model is divided into two parts according to the initial temperature

distribution, which may be initially uniform or non-uniform,: initial uniform superheat

model and initial non-uniform superheat model, respectively.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
104

3.2.1 Initial Uniform Superheat Model for Spherical Bubble Growth

The initial uniform superheat model is relatively easier to solve than the initial non-

uniform superheat model. The analytical approximations for the initial uniform superheat

model have performed remarkably well, while this has not been the case for the analytical

approximation for the initial non-uniform superheat model. The schematic diagram for

the initial uniform superheat model is shown in Figure 3.7. The uniform model

incorporates two possible cases: Tbulk > Tsat and Tbulk < Tsat. The former is referred to

as a superheated liquid, in which the bubble grows continuously, and is discussed in

Chapter 2; the latter is referred to as a subcooled liquid, and will be discussed in section

3.5.1 below in connection with vapor bubble collapse.

3.2.2 Initial Non-uniform Superheat Model

Vapor bubble dynamics in non-uniform temperature fields may be heat transfer

controlled, liquid inertia controlled, or a combination of both. As will be demonstrated,

the mode acting is quite sensitive to the initial temperature distribution present around the

vapor bubble. Initial attempts at modeling the non-uniform superheat vapor bubble

dynamics will be divided into two parts in order to make the problem tractable: a spherical

bubble model and a hemispherical bubble model, described below.

3.2.2.1 Spherical Bubble Model

The initial non-uniform superheat model for a spherical bubble is shown

schematically in Figures 3.8 (a) and (b). Figure 3.8 (a) shows the initial temperature

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
105

distribution, where the highest temperature is the heater surface temperature (Tw), the

lowest temperature is the bulk temperature ( Thulk), and the saturation temperature (Tmt)

corresponds to the system pressure. Although the saturation temperature may take on any
value, only the case with Thulk < Tml < Tw is considered here since this range is of more

practical significance. Figure 3.8 (b) depicts the non-uniform temperature distribution

around the vapor bubble as it is in the process of growth. This is quite difficult to

estimate by any analytical approximation, since the bubble dynamics is governed not only

by heat transfer but also by liquid inertia.

3.2.2.2 Hemispherical Bubble Model

This model is particularly applicable to pool boiling circumstances, in which the

mechanisms are extremely complex. No appropriate predictive models have been found

for high subcooled pool boiling, in particular. What follows is an attempt to provide

upper and lower predictive bounds of the vapor bubble growth on a solid surface in pool

boiling. Figure 3.9 depicts a realistic hemispherical vapor bubble in contact with a heater

surface, where the top of the bubble has a non-uniform temperature distribution in the

liquid with the least thermal energy in the liquid boundary layer near the liquid-vapor

interface, while the lower part of the bubble is assumed to have a uniform temperature

distribution at the heater surface temperature, which is the highest possible temperature.

Considering half of a one-dimensional spherical model mathematically results in a

hemispherical vapor bubble model. If the initial non-uniform superheat model is taken

with radially symmetry corresponding to the temperature distribution at the top of the

bubble in Figure 3.9, as shown in Figure 3.10, the bubble growth taking place should be at

a minimum, and will be called the "lower bound" of the bubble growth here. If the initial

uniform superheat model is taken as shown in Figure 3.11, with the temperature

corresponding to the maximum possible level, the heater surface temperature at the lower

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
106

part of the bubble in Figure 3.9, the bubble growth should be at a maximum, and will be

called the "upper bound" of the bubble growth here. This type of behavior will be

demonstrated later by comparison with recent experiments. The temperature distribution

around a real bubble on a flat heater surface is shown in Figure 3.12, from Straub et al

(1990), using holographic interferometery. The heater is facing downward in order to

minimize convection due to buoyancy. It is noted that the thermal boundary layer formed

on the heater surface supports the schematic diagram given in Figure 3.9.

3.3 Governing Equations

The temperature distribution corresponding to the spherical vapor bubble growth

problem in a non-uniformly superheated liquid is shown in Figure 3.8. The bulk liquid and

the critical size bubble are initially isothermal. It is assumed that the bubble growth is

initiated from the critical size bubble. The bubble wall temperature decreases, as the

bubble grows, to the saturation temperature corresponding to the internal pressure of the

bubble. Other assumptions necessary to solve the problem are summarized below:

a) Bubble remains spherical

b) A uniform pressure and temperature exist inside the vapor bubble

c) The vapor temperature inside the bubble is equal to the liquid-vapor interface
temperature, Tv = Tt

d) The liquid is incompressible

The momentum equation in spherical coordinates is given by:

(3.25)

The continuity equation with assumption (d) is given by:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
107

(r2w) = 0 (3.26)
or

Using the continuity equation (3.26) and integrating Equation (3.25) from the

bubble surface (R) to infinity (>), the momentum equation can be written in terms of the

bubble radius as:


3 -2 Pp - P
RR+R = ^ ~
2 ep, (3.27)

For a spherical bubble, viscosity affects only the boundary condition, so that the

balance of pressure across the liquid-vapor interface gives:

PvW - P K= ^ p - + 4 ^ (3.28)
A A

where PR is the liquid pressure adjacent the bubble wall.

Inserting Equation (3.28) into Equation (3.27) yields:

RR+R2 = (3.29)
2 ep, ep ,R ep,R

The initial conditions are:


R(0) = Rc and R(0) = 0 att = 0 (3.30)

The driving force for the bubble growth in Equation (3.29) is the pressure

difference between the vapor within the bubble and the liquid adjacent the bubble, which is

constrained by the surface tension and the bulk liquid inertia. Equation (3.29) implies that

static equilibrium will exist when the first and second derivatives are zero. For the initial
vapor bubble of critical size, this gives, since TV=T initially:

/? (3.31)
PV(T)~P

The energy equation in spherical coordinates is given as:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
If the initial temperature distribution is produced by a constant imposed heat flux,

the initial condition will be:

I ' M ) = T_ + exp (-ry 4oa' ) - Z f e t f c (3.33)


/Ci AC/

If the initial temperature distribution is produced, on the other hand, by a constant

surface temperature, the initial condition will be:

(3.34)

The boundary conditions are: r(,0) = T,

(3.35)

Equation (3.35) is the energy balance at the liquid-vapor interface of the bubble.

The term in the RHS of the equation represents the latent heat rate for evaporation of the

liquid to produce the vapor required in the bubble, while the term in the LHS represents

the conduction heat transfer rate from the liquid adjacent to the vapor bubble wall.

According to assumption (b), the conduction heat transfer from the bubble wall to the

interior of the vapor bubble is neglected. The vapor bubble growth problem in a

superheated liquid can be solved by coupling Equation (3.29) and Equation (3.32) with

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
109

the initial and boundary conditions given in Equations (3.30) and (3.35), together with the

initial temperature distribution of either Equation (3.33) or (3.34).

3.4 Numerical Solution

A Fortran program was written to solve the transient one-dimensional spherical

bubble growth problem in non-uniform superheated liquids, where the initial non

uniformity is spherically symmetric around a critical size vapor bubble. The program

couples the momentum and energy equations and marches stepwise in time, taking into

account the variations of properties with temperature. For the infinitesimal time steps, the

momentum equation is solved by the Runge-Kutta method to determine the locus of the

bubble interface, which provides the boundary condition for the energy equation, which is

then solved by means of the finite difference method. The complexity of the problem is

simplified by adopting the Landau coordinate transformation to immobilize the moving

boundary. After the transformation, the energy equation is solved by using the Thomas

algorithm, also called the TriDiagonal Matrix Algorithm. Further details of the solution

are presented in Chapter 2.

3.5 Vapor Bubble Dynamics in Subcooled Liquids

3.5.1 Bubble Collapse

In connection with the analytical work described earlier, Florschuetz and Chao

(1965) conducted vapor bubble collapse experiments with water under free fall conditions,

in order to eliminate gravity effects and produce a spherical bubble. The test vessel was

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
110

dropped from a height of 1.83 m, giving a free fall time of 0.6 sec. The liquid in the vessel

was initially heated to its boiling point at subatmospheric pressure. Immediately after the

delivery of a bubble into the vessel, the vessel was released. A small electromagnet

controlled the opening of a cover on the top of the vessel to expose the system to

atmospheric pressure. The photographic data show that in spite of the free fall condition,

the bubble experienced small translational velocities of about 4 cm/s or less. The

measured bubble collapse is plotted in Figure 3.13 with both the heat transfer controlled

collapse of Equation (3.19) and computations using the present numerical method. It is

seen that the experimental curve generally follows the trend of the predictions. The

discrepancy between the experimental data and both predictions is attributed to the

residual translational motion mentioned above. The physical properties in Equation (3.19)

are evaluated at the saturation temperature corresponding to the system pressure. It is

also noted that the heat transfer controlled model of Equation (3.19) is in fair agreement

with the present numerical calculation for the present operating conditions, where the

dimensionless parameter B is 0.0087. This is below the upper limit of B=0.02 specified as

the heat transfer controlled criterion .

3.5.2 Bubble Growth and Collapse

A model for the bubble dynamics with initial non-uniform superheat for a spherical

bubble was described in section 3.2.2.1. An experiment corresponding to this

circumstance was conducted by Lauterborn and Hentschel (1985) in connection with

cavitation research, using water at atmospheric pressure and high speed photography at

75000 fps. Homogeneous nucleation was effectively produced by focusing a laser beam at

a desired location near a sold wall. However, details as to the intensity, concentration and

duration of the laser beam were not given, which would be necessary to derive the initial

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Ill

temperature field around an embryo bubble. Interest at that time was focused on bubble

collapse rather than bubble growth. It was observed in the experiments that the vapor

bubble rebounded after the first collapse and rebounded again following a second collapse.

Data from the series of photographs are plotted, together with the present numerical

computations, in Figure 3.14. It is noteworthy that the results of the computations predict

the two rebounds in an approximate fashion, in spite of the uncertainties in the initial

conditions, which are estimated. The computed vapor pressure within the bubble is

included in Figure 3.14, and it is noted that the vapor pressure falls below the system

pressure when the bubble is near its maximum radius. This may be interpreted to be a

consequence of the liquid inertia, where the liquid is still moving outward while the

thermal boundary layer no longer supplies evaporation, and followed by collapse so that

the vapor pressure increases, as expected.

Figure 3.15 provides the computed temperature distributions around the growing

bubble at various times for the estimated parameters of Figure 3.14. The abscissa

represents the distance from the moving liquid-vapor interface. The initial temperature

distribution at time zero is obtained from a one-dimensional transient conduction model

for a constant surface temperature, given by Equation (3.34), with an arbitrary waiting

time of 0.21 sec imposed in order to attain a surface temperature of 302 C, expected to

be required for homogeneous nucleation. The temperatures in region 1 and region 2 in

Figure 3.15 do not necessarily follow a chronological order for the elapsed times owing to

the use of a moving coordinate origin on the vapor bubble interface. However, it is noted

that the temperature curves at t4 and t5 are higher than at t3, which may appear to be an

anomaly. This could be interpreted as a manifestation of the liquid kinetic energy

conversion during the bubble collapse to the equivalent thermal energy, which then

possesses the potential for rebound since the liquid at the interface then becomes

superheated, as seen in Figure 3.15. The rebound of the bubble occurs only when the

bubble collapse is controlled by liquid inertia.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
112

3.5.3 Bubble departure mechanism

In normal gravitational fields the vapor bubbles found in normal pool boiling

depart as a result of buoyancy. On examining the computed internal vapor pressure of a

hemispherical vapor bubble growing and collapsing adjacent to a wall, as in Figure 3.14, it

is noted that it drops below the system pressure at times and may constitute an additional

mechanism for departure from a heating surface and may be particularly effective in

microgravity.

3.6 Experiments in Microgravity with Saturated and Subcooled R-113

One of the difficulties associated with conducting well-defined experiments of pool

boiling in earth gravity is the randomness of the initial thermal boundary conditions

occurring at nucleation. This is demonstrated in Figures 3.1 and 3.2 by the scatter of the

data. A significant advantage of the present experimental measurements with transient

pool boiling in microgravity is that the initial conditions at nucleation are known: the

surface temperature, thermal boundary layer, and system pressure.

3.6.1 Experimental Apparatus

Pool boiling experiments with R-113 were conducted in microgravity in the 132 m

evacuated drop tower facility at the NASA Lewis Research Center, which provided 5.18

seconds of free fall with effective body forces on the order of 10'5 g. Photographs of the

boiling process were obtained with a D.B. Milliken camera operating at 400 pictures per

second. The drop vehicle, vessel and associated electronics were designed to withstand

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
113

the vacuum conditions and the 50 g impact at the bottom of the drop tower. The test

vessel has internal dimensions of approximately 12.5 cm by 14 cm by 27.9 cm, and

consists of thick aluminum plates to assist in providing initially uniform internal

temperatures. Two heater surfaces are installed in one wall of the test vessel; each

consists of a 400 A thick semi-transparent gold film sputtered on a highly polished quartz

substrate, and serves simultaneously as a heater, with an uncertainty of 4 % in the

measurement of the heat flux, and a resistance thermometer, with an overall uncertainty of

1.0 C. The heater is rectangular in shape, 1.91 cm by 3.81 cm. Degassed commercial
grade R-113 (trichlorotrifluoroethane, CC12FCCIF2) was used because of its low normal

boiling point (47.6 C), which minimized problems associated with heat loss to the

surroundings, and because of its electrical nonconductivity, which made it compatible for

direct contact with the thin gold film heater. A pressure transducer was employed to

measure the system pressure, with an uncertainty of 0.172 kPa. Chromel-constantan

Teflon sheathed thermocouples with ice reference junctions were used to measure the

liquid and vapor temperatures, with a total uncertainty of 0.05 C. The desired

subcooling was obtained by increasing the system pressure above the saturation pressure

corresponding to the initial liquid temperature. In the absence of buoyancy, an initially

motionless liquid remains stagnant upon heating until the onset of boiling, and the

temperature distribution at the moment of incipient boiling can be determined from a

conduction heat transfer analysis.


The heat transfer to the liquid (q") is calculated by a transient one-dimensional

conduction analysis with the measured total heat flux (g"). The ratio of the heat flux

transferred to the liquid for the R-l 13/quartz combination is constant as:

-% = 0.175 (3.36)

Equation (3.36) is in good agreement with the short time experiments on the order

of 5 seconds, when comparing the one-dimensional transient predicted heater surface

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
114

temperatures with the measurements. The ratio of the heat flux for water corresponding

to Equation (3.36) is 0.49. Detailed descriptions of the experimental apparatus and

procedures used for obtaining present experimental results are given in Ervin et al. (1992).

3.6.2 Experimental Results and Discussions

Nucleation with pool boiling on a heated surface is considerably more complex

than homogeneous nucleation. The nucleation not only occurs on a foreign material, but

also takes place in a non-uniformly superheated liquid, while homogeneous nucleation is

generally regarded as being initiated in a uniformly superheated media. Figure 3.16

depicts three typical early bubble growths in microgravity with different levels of heat flux

and subcooling. Figure 3.17, corresponding to Figure 3.16 (b), shows the measurement of

the transient mean heater surface temperature and system pressure for the entire test. The

heater power is turned on at zero time so that a thermal boundary layer develops on the

heater until nucleation takes place at 1.55 sec. The decrease of the surface temperature

following nucleation is attributed to the boiling taking place on the heater surface. The

one-dimensional transient conduction prediction curve in Figure 3.17 coincides fairly well

with the measured surface temperature, and demonstrates the validity of the one

dimensional prediction for a short period experiment. This coincidence renders Equation

(3.36) and the corresponding temperature distribution in the liquid applicable to the

computation of the vapor bubble dynamics for the initial non-uniform superheat model.

3.6.2.1 Energy Input to the Liquid

The experiments presented in Figures 3.16 (a), (b) and (c) have respective

subcoolings of 2.63 C, 11.13 C and 0.81 C - almost saturated. Although the heat flux

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
115

itself does not directly influence the vapor bubble dynamics, the energy input (J/cm2) to

the liquid, defined as the product of the heat flux (q",) and the waiting time (t*), does

influence the growth dynamics. The energy input to the liquid up to the nucleation time,

together with the subcooling, heater surface superheat at nucleation, and bubble radius at

0.025s are plotted in Figure 3.18. The direct correspondence between the largest energy

input and growth rate in case (a) is noted, together with the largest heater surface

superheat at nucleation. Cases (b) and (c), however, are somewhat different. Although

case (b) has a larger energy input and superheat than case (c), the growth rate is smaller,

which is attributed to the fact that case (b) has a much higher subcooling than case (c).

The validity for the use of the upper and lower bounds in vapor bubble growths, as

determined from the initially uniform and non-uniform superheat models for the

computation of the vapor bubble dynamics on heated surfaces, can be determined by

comparing the numerical computations with the measurements for the three different cases

presented in Figure 3.16. The growing vapor bubble sizes were measured approximately,

assuming an oblate bubble, by an effective radius of (ab2)1/3/2 where "a" is a horizontal

length of the oblate and "b" is a height of the oblate. The numerical computations for both

the initially uniform and non-uniform superheat models provide the following data: bubble

size, liquid-vapor interface velocity, liquid-vapor interface acceleration, liquid-vapor

interface temperature, and temperature distributions in the liquid surrounding the growing

bubble, all as functions of time.

3.6.2.2 Initial Temperature Distribution

Figures 3.19 and 3.20 present the liquid temperature distributions surrounding the

growing vapor bubble for the initial non-uniform and the uniform superheat models,

respectively. Figure 3.19 shows that the initial temperature distribution collapses as time

passes, which results from both the evaporation at the liquid-vapor interface and the liquid

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
116

divergence around the growing bubble. The heat conduction into the subcooled bulk

liquid is negligible for the short period involved. It is quite noteworthy that the initial

thermal boundary layer is completely dissipated within 6 ms, which implies that the liquid-

vapor interface at the top of the vapor bubble in Figure 3.16 (b)-B is already conditioned

for condensation to take place at this location. Comparing Figure 3.19 with Figure 3.20,

up to a time of 2.6x1 O'6 sec., no difference in the temperature gradients at the liquid-vapor

interface can be detected between the models. From the time of 1.6x1 O'3 sec. onward, the

difference becomes considerable. Finally, at a time of 6xl0-3sec., the temperature gradient

in the non-uniform model of Figure 3.19 begins to change sign, and condensation at the

liquid-vapor interface can take place, accompanied by the collapse of a spherical bubble.

The gradient in the initially uniform superheat model of Figure 3.20 remains positive, with

continuous evaporation producing continuous bubble growth.

Measurements of the growing vapor bubbles from Figure 3.16 (a), (b) and (c) are

compared with computation from the initial uniform and the non-uniform superheat

models in Figures 3.21, 3.22 and 3.23, respectively. All measurements fall between the

resulting curves for the initially uniform and non-uniform models, which supports the

definition of the models as described in section 3.2.2. According to the models in section

3.2.2, the initially uniform superheat temperature provides the maximum possible bubble

growth, which is called the upper bound since the bulk liquid temperature was taken to be

uniformly at the highest temperature in the thermal boundary layer, while the initially non-

uniform superheat temperature provides the minimum bubble growth, which is called the

lower bound since the initial temperature distribution on the top of the critical bubble is

taken as the initial thermal boundary layer around a spherical bubble. Since the

photographs in the experiments were taken at 400 frame per second, the corresponding

uncertainty in the onset of nucleation is 2.5 milliseconds, which implies that the time scale

may be shifted up to 2.5 msec. The concept of the upper and lower bounds is useful in

that the first data point of the measured bubble radii can be conveniently be placed to fall

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
117

between the range of the upper and lower bounds, as may be seen in Figure 3.21 where

the data were shifted to the left by 0.002 seconds. By comparing Figures 3.21, 3.22 and

3.23, it is noted that as the subcooling increases, the bubble growth has a tendency to

follow the curve for the initially non-uniform superheat model, as is seen in Figure 3.22.

3.6.2.3 Bubble Dynamics

In order to explore the bubble dynamics during the early stages of growth for both

the initially uniform and non-uniform superheat models, the experiment corresponding to

Figure 3.16 (b), which has the highest subcooling among the three, was chosen. The

initially uniform and non-uniform superheat models with conditions given in Figure 3.16

(b) provided the numerical data of the bubble radius, interface velocity, interface

acceleration and interface temperature as shown in Figures 3.24, 3.25 and 3.26,

respectively. It is found, in general, that the differences in behavior between these two

models are negligible during the early stages, for times less than about, say, 10'5 seconds.

In particular, no differences between the interface accelerations can be detected in Figure

3.25. It may then be deduced that the bubble dynamics during the very early stages are

not affected by the thermal boundary layer, but rather by the initial bubble wall superheat

temperature. Further details of the bubble dynamics relative to the liquid-vapor interface

velocity, acceleration and temperature for the initially uniform superheated liquid case are

given in Chapter 2.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
118

0.8
- Bubble Radius vs. Time for Water

0.7 Psys=l atm, Tbulk=57.2 C, Tw=132.7 C, q"=363 W/cm2

0.6

0.5
Bubble Radius, mm

Avei ageRn ax_

0.4

0.3

0.2

0.1

0
0 100 200 300 400 500 600 700 800 900 1000
Time, (as

Figure 3.1. Illustration of variability of bubble dynamics in pool boiling during a given
experiment. Water, =42.8C , ATwsup =32.7C , Ellion (1953).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
119

0.5
Bubble Radius vs. Time for water

Psys=l atm, Tbulk=25.5 C, Tw=134.4 C,q"=490 W/cm2

0.4 . ________________
Average llm ax
Bubble Radius, mm

0.3

0.2

0.1

0
0 100 200 300 400 500 600 700
Time, us

Figure 3.2. Illustration of variability of bubble dynamics in pool boiling during a given
experiment. Water, ATsub =74.5C, ATwsap = 3 4 .4 C , Ellion (1953).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
120

q"= 363 W/cm2

Heat transfer controlled growth and


collapse, Equation (3.9)
0.8

0.6

0.4
Inertia controlled collapse.
Equation (3.4)
0.2

0 0.5 1 1.5 2 2.5 3 3.5 4


t/tmax

Figure 3.3. Comparison of Equations (3.4) and (3.9) with normalized data of Figure 3.1.
Arjufc=42.8 C , Arwsup=32.7C, Ellion (1953).

0.8 Heat transfer controlled growth and


collapse, Equation (3.9)

0.6

0.4 Inertia controlled cojlj


Equation (3.4)

0.2

0 0.5 1 1.5 2 2.5 3 3.5 4


t/tmax

Figure 3.4. Comparison of Equations (3.4) and (3.9) with normalized data of Figure 3.2.
A^6 =74.5C , ATwsup = 34.4 C, Ellion (1953).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
121

3
Dimensionless radius and liquid-vapor interface temperature for water
o
o-.. . Psys=101.325 kPa, Tsat=100 C
A I A
Qj f \ 2.5

1 - K co -cA 0 0 -0-0 0 0 -
O-o-O'
O I I\ O B=0.02, Dimensionless
/ 0 !A Temperature 2
Oo-O O B=0.1

A b =0.3

o B=0.02, Dimensionless 1.5 y


Radius

0------ B=0.1

Heat transfer controlled collapse


A B=0.3 1

-1 OOO-O-OO O -O O o O O O 0 -0
'O-,
0.5

Tnertin controlled collapse


Equation (3.18)

-2 0
0 0.5 1 1.5 2 2.5 3 3.5 4

Figure 3.5. Numerical solutions of Florschuetz and Chao (1965) for vapor bubble
collapse given by Equations (3.11) and (3.12), for three different values of the
dimensionless parameter B, together with the inertia controlled collapse.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
122

0.9
B=0.1, numerical
solution by Florschuetz
0.8 and Chao (1965)

B=0.02, numerical
0.7 solution by Florschuetz
and Chao (1965)

0.6 Heat transfer controlled


collapse, Equation
(3.19)
Y 0.5
O".
0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1.2

XH

Figure 3.6. Numerical solutions of Florschuetz and Chao (1965) for vapor bubble
collapse given by Equations (3.13) and (3.14), for two different values of the
dimensionless parameter B, together with the heat transfer controlled collapse of Equation
(3.19).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
123

T
Tbulk
Tbulk

R(t).
Tsat
Tsat

(a) (b)

Figure 3.7. Initial uniform superheat model, (a) Initial temperature profile surrounding a
critical size vapor bubble, (b) Liquid temperature profile surrounding a growing vapor
bubble.

Tsat
Tbulk
Tbulk

(a) (b)

Figure 3.8. Initial non-uniform superheat model, (a) Initial temperature profile
surrounding a critical size vapor bubble, (b) Liquid temperature profile surrounding a
growing vapor bubble.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
124

Figure 3.9. Schematic diagram of a realistic hemispherical vapor bubble growing on a


heater surface.

Figure 3.10. Schematic diagram of an ideal hemispherical bubble growing on a heater


surface for the initial non-uniform superheat model.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
125

l-c
A

Figure 3.11. Schematic diagram of an ideal hemispherical bubble growing on a heater


surface for the initial uniform superheat model.

Figure 3.12. Temperature field around a downward-facing R-113 vapor bubble by means
of holographic interferometery (Straub et al. (1990)).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
126

0.005

Bubble radius vs. time for water,


0.0045
Psys=101.325 kPa, Tbulk=87.8 C,
Tsat=100 C, Subcool=12.2 C
0.004
B=0.0087
oooo
0.0035

_ 0.003
0>
o3
E Heat transfer controlled collapse,
" 0.0025 Equation (3.19)
.2
'ca5
OS
0.002 h

0.0015 Numerical computation

0.001

Experiments by
0.0005 Florschuetz and Chao (1965)
O O o o
0 I '
_ ______ 1_______ I_ _ _ _ _ _ _ _ L . _l I 1 I I I L.

-0.02 0 0.02 0.04 0.06 0.08 0.1 0.12


Time, sec

Figure 3.13. Comparison of experimental data of Florschuetz and Chao (1965) for water
with computations using a heat transfer controlled collapse and the present numerical
method.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
127

5 0.6
Vapor bubble radius and vapor pressure vs. time for water
0.5
4.5
Conditions for computation: 0.4
Psys=0.101MPa, Tbulk=23 C,
4
Tw=302 C, t*=0.21 sec 0.3
3.5 0.2

Vapor pressure, MPa


3 0.1

2.5 Radius computed

Experiments by Lauterbom - 0.1


and Hentschel (1985)
2 - 0.2
Vapor pressure, MPa

.5 System pressure -0.3

-0.4
1
-0.5
0.5
- 0.6

0 -0.7
0 200 400 600 800 1000 1200
Time, ps

Figure 3.14. Vapor bubble growth and collapse with water at atmospheric pressure.
Comparison between measurements and numerical computations carried out using
estimated parameters. Times t3, t4 and t5 correspond to those in Figure 3.15.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
128

500 time elapsed


_Liquid temperature distribution for water tl= 0.0 ps
450
Conditions for computation: t2= 1.574 ps
Psys=0.101MPa, Tbulk=23 C,
400 Tw=302 C, t*=0.21 sec t3= 126.7 ps

t4= 400.4 ps
350
t5= 700.2 ps
U 300
Temperature,

250

200

150

Region 2
100

50 Region 1

0
le-09 le-08 le-07 le-06 le-05 le-04 le-03
Distance from the liquid-vapor interface, m

Figure 3.15. Computed temperature distributions around a growing bubble at various


times, for the estimated parameters of Figure 3.14.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
PBMT1115.605 PBMT1129.820 PBMT0207.600
q"x =5.75 W/cm2 q"T=7.87 W/cm2 q"T=6.33 W/cm2
t*=3.083 sec. t*=1.550 sec. t*=1.685 sec.
Psys=108.88 kPa Psys=142.34 kPa Psys=103.32 kPa
Tbulk=47.11 C Tbulk=47.06 C Tbulk=47.33 C
Tsat=49.74 C Tsat=58.19 C Tsat=48.14 C

r> 1 1 2 C 7^=110 C rJ=ioo.o C

A C p =62.26 C A C p=51.81 C A C p =51.86 C

&Tsub=2.63 C A7fui=11.13C A T ^O .81 C

(a) (b) (c)

Figure 3.16. Early bubble growth behavior under microgravity with different levels of
heat flux and subcooling.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
130

14U 150
Experiment, PBMT1129.820
q"T=7.87 W/cm2, 149
120 Tbulk=47.06 C, Tsat=58.19
Psys= 142.34 kPa
148
100

U 147

Pressure, kPa
80
V 1-D transient
I* 146
a conduction prediction
uu 60
Measured surface
temperature 145

40 Measured pressure
144

20 143

142
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time, sec

Figure 3.17. Measured mean heater surface temperature and system pressure for Test
PBMT1129.820, Figure 3.16 (b).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
131

PBMT1115.605 (a)

IH PBMT1129.820 (b)

M PBMT0207.600 (c)

_ l l l

q"(total) x Subcooling Heater Bubble


t* Energy (C) Surface radius
input Superheat (mm) at
(J/cm2) xlO (C) t=0.025 s

Figure 3.18. Comparisons of several parameters influencing growth dynamics for the
experiments in Figure 3.16.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
132

180
Liquid temperature distribution Time from onset of growth
- for non-uniform model
0.0 sec
160 R-l 13 (PBMT1129.820)
2.60E-06
Psys= 142.34 kPa, Tbulk=47.06 C,
Tsat=58.19 C, Tw*=l 10 C 2.00E-04
140
1.60E-03

6.00E-03
. 120

a,
100

le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02


Distance from liquid-vapor interface, meter

Figure 3.19. Transient liquid temperature distribution surrounding a spherical vapor


bubble growing from the critical size, with an initial non-uniform liquid temperature
distribution. The conditions correspond to Figure 3.16 (b). q"T=7.87 W/cm2,
ATJsup =51.8C, Artuft = 11.13C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
133

180
Liquid temperature distribution Time from onset of growth
- for uniform model
-----------0.0 sec
160 R-113 (PBMT1129.820)
-----------2.60E-06
Psys=142.34 kPa, Tbulk=l 10 C,
Tsat=58.19 C 2.00E-04
140
- 1.60E-03

U ----------- 6.00E-03
. 120
uV-l
w a
C
U
Q
m
Q J
I
E-
a.
100 /
/
/
80 /
/

60

40 J I ' m i l l : i i i l n i l ______ I I I I m i l _ _ _ _ _ _ I_ _ _ I i i m i l ______I I I I M i l l _ _ _ _ _ _ I_ _ _ I I I I I I I ! _ _ _ _ _ _ I I I I 1 1 III

le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02


Distance from liquid-vapor interface, meter

Figure 3.20. Transient liquid temperature distribution surrounding a spherical vapor


bubble growing from the critical size with an initial uniform liquid temperature
distribution. The conditions correspond to Figure 3.16 (b). A7^p = 51.8C

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
134

0.012
Radius vs. Time for R113, Psys=108.88 kPa, Tsat=49.74 C

0.01 Initial uniform superheat


model, Superheat=62.26
C, Tbulk=112.0 C

Initial non-unifoim
superheat model,
0.008 Tw*=ll2.0C,
Tbulk=47.11 C

EXPERIMENT,
PBMT1115.605

0.006

0.004

0.002

0
0 0.005 0.01 0.015 0.02 0.025
Time, sec
Figure 3.21. Comparisons of measured bubble growth data with the initially uniform and
non-uniform superheat spherical vapor bubble growth models. Computational and
experimental conditions correspond to Figure 3.16 (a).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
135

0.012
Radius vs. Time for R113, Psys=143.0 kPa, Tsat=58.19 C

0.01
Initial uniform superheat
model, Superheat=51.81
C,Tbulk=lI0.0 C

Initial non-uniform
0.008 superheat model,
Tw*=l 10.0 "C,
Tbulk=47.06 C

EXPERIMENT,
PBMT1129.820
0.006

0.004

0.002

.XL.

0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
Time, sec
Figure 3.22. Comparisons of measured bubble growth data with the initially uniform and
non-uniform superheat spherical vapor bubble growth models. Computational and
experimental conditions correspond to Figure 3.16 (b).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
136

0.012
Radius vs. Time for R113, Psys=103.32 kPa, Tsat=48.14 C

0.01
model, Superheat=51.86
C, Tbulk= 100.0 C

. . . . . . . initio non-uniform
0.008 superheat model,
Tw*=100.0 C,
Tbulk=47.33 C

0) EXPERIMENT,
E PBMT0207.600
t/j
a 0.006
'ca3
OS

\
\
I

0.004
\


..

\
"

0.002
i



3 \

a

1

;
!

1

0 ' - -1............1 ---- 1- ' ------ 1---------'--------- 1--------- 1---------

0 0.005 0.01 0.015 0.02 0.025 0.03


Time, sec
Figure 3.23. Comparisons of measured bubble growth data with the initially uniform and
non-uniform superheat spherical vapor bubble growth models. Computational and
experimental conditions correspond to Figure 3.16 (c).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
137

le+05 le+02
Bubble radius and interface velocity vs. time for R-l 13
le+04 Psys=143 kPa, Tsat=58.19 CExperimental data PBMT1129.820
le+01
le+03

le+02 le+00

le+01
le-01

Interface velocity, m/s


Uniform Model, Radius
le+00
Non-unifomi Model, Radius
le-02
S.
1A le-01 Uniform Model, Velocity
.2
le-02 le-03
Q
S Non-uniform Model,
Velocity
le-03
le-04
le-04

le-05 le-05
le-06
le-06
le-07

le-08 le-07
le-10 le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01
Time, sec

Figure 3.24. Comparison of the liquid-vapor interface velocity between the initially
uniform and nonuniform superheat models. Conditions correspond to the experiment in
Figure 3.16 (b).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
138

le+03 5e+08
Bubble radius and interface acceleration vs. time for R-l 13
le+02 Psys=143 kPa, Tsat=58.19 (Experimental data PBMT1129.820

le+01 4e+08
Unifonn Model, Radius

Non-uniform Model, Radius


le+00
Uniform Model, 3e+08
le-01 Acceleration

Non-uniform model, ao
le-02 acceleration
2e+08 O
T3
CO
le-03 uC3
Pi O
o
l_
le-04 3a
le+08
le-05

le-06 0e+00

le-07

mill t i tin
le-08 itiim l i itimtl t i i imil i i t I i i i Mi n i i i n mil _ i _i.iuiul.
-le+08
le-10 le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01
Time, sec

Figure 3.25, Comparison of the liquid-vapor interface acceleration between the initially
uniform and nonuniform superheat models. Conditions correspond to the experiment in
Figure 3.16 (b).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
139

le+03 140
Bubble radius and interface temperature vs. time for R-113
le+02 Psys=143 kPa, Tsat=58.19 C Experimental data PBMT1129.820
120
le+01
100
le+00

le-01

Interface temperature,
le-02

le-03

le-04

le-05
Uniform Model, Radius

le-06 Non-uniform Model, Radius

Uniform Model, Ti
-20
le-07
Non-uniform model, Ti

m il
le-08 -40
le-10 le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01
Time, sec

Figure 3.26. Comparison of the liquid-vapor interface temperature between the initially
uniform and nonuniform superheat models. Conditions correspond to the experiment in
Figure 3.16 (b).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTER 4

ROLE OF INSTABILITIES IN POOL BOILING BUBBLE


GROWTH IN MICROGRAVITY

Rapid vapor bubble growths at superheats far below the homogeneous nucleation

superheat limit were observed in pool boiling experiment in microgravity. The

experiments, using R-113 as a fluid, were conducted in an evacuated 132 m drop tower

providing 5.18 seconds at a/g-lO-5. Photographic evidence revealed a roughening of the

bubble liquid-vapor interface as growth took place, which is believed to be a consequence

of an interface instability. The growth of the instability is explained in terms of heat

conduction, where the heat transfer to the liquid-vapor interface increases at the trough of

the interface perturbation while decreasing at the crest. This can give rise to a differential

vapor recoil at the interface, which further increases the liquid-vapor interfacial surface

area, increasing the evaporating surface area and the bubble growth rate. In addition, it is

possible that distortion of the crests can produce entrainment of liquid particles, which can

increase the evaporation rate further. The rapid bubble growth is modeled by introducing

a simple expression in terms of a circular function to account for the increase in surface

evaporating area. The resulting factor is found to be a function of the ratio of the

amplitude to the wave length. The onset of the interface roughening can be considered to

originate from either a Rayleigh-Taylor or Landau instability; however, neither of the

present models successfully provide the necessary quantitative physical descriptions.

140

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
141

4.1 Literature Survey

Rapid evaporation at superheats much lower than the homogeneous nucleation

superheat limit was observed in transient boiling experiments conducted in short term

microgravity. A number of vapor bubble growth experiments conducted at the superheat

limit have been reported recently. The rapid evaporation rates observed have been

described by various terms such as vapor explosion, explosive boiling, thermal explosion,

rapid phase transitions (RPTs), and have been discussed by Reid (1983). In general, it has

been presumed that potential high pressure water cooled nuclear reactor explosions, LNG

spill explosions and high pressure boiler accidents are related to vapor explosions, so that

answers are being sought to a number of questions: what are the mechanisms involved in

vapor explosions; what are the conditions under which such can occur; and what can be

done to prevent its occurrence? The present work is intended to assist in providing a

better understanding of the mechanisms involved.

The vapor explosion type of behavior is assumed to result from an increase in the

liquid-vapor interface area arising from the instability of this interface. It is considered, in

general, that an interface becomes unstable if the growth rate of the amplitude of

perturbations is greater than that of the surface itself. The interface instability was first

reported by Taylor (1950), deducing that when two superposed fluids of different densities

are accelerated in a direction perpendicular to a plane interface, the interface is unstable

for the acceleration directed from the lighter to the heavier fluid. This has become termed

as the Rayleigh-Taylor instability. Later, Plesset (1954), Birkhoff (1956) and Plesset and

Prosperetti (1977) developed the corresponding Rayleigh-Taylor instability analysis for the

spherical case, suggesting a closed form of instability criteria which included surface

tension, viscosity and acceleration. Shepherd and Sturtevant (1982) and Frost and

Sturtevant (1986) compared the instability criteria with measurements, using a classical

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
142

bubble growth model. The results predicted the bubble growth to be stable, while the

experiments demonstrated obviously unstable behavior.

The hydrodynamic instability with rapid evaporation at a plane interface was

approached from a different perspective by Miller (1973) and Palmer (1975). The former

reported that a vaporizing plane interface can be unstable due to large density difference

between phases, with the stability analysis showing the perturbation growth rate to be

positive depending on the surface tension, density difference and heat transfer at the

interface. The latter introduced a differential vapor recoil mechanism to explain the plane

instability, which is described as local surface depressions being produced by the resultant

force exerted on the liquid surface by the vapor.

The roughening of a growing vapor bubble surface was first observed by Shepherd

and Sturtevant (1982) with n-butane, with the measured average bubble growth rate of

approximately 14.3 m/s being much greater than the classical bubble growth prediction.

Rather than apply the Rayleigh-Taylor instability to their bubble growth problem, it was

suggested rather that the instability theory of Landau (1944) be applied, which was

originally developed for plane combustion flame propagation under the assumption of a

thin flame. Istratov and Librovich (1969) developed the Landau theory for the spherical

case, obtaining a closed form of instability criteria. Testing this criteria using the classical

bubble growth model of Prosperetti and Plesset (1978) resulted in disagreement with the

experiments of Shepherd and Sturtevant (1982). Photographs of the early bubble diameter

of 0.15 mm at 10 ps clearly shows a roughening of the liquid-vapor interface. The

calculated evaporative mass flux, according to the measurements of Shepherd and

Sturtevant (1982), is approximately 400 g/cm2-s, which is extremely higher than the value

calculated by the classical bubble growth model. McCann et al. (1989) repeated the

experiments for the same conditions but with a higher camera speed, and confirmed the

results of Shepherd and Sturtevant (1982), with an average bubble growth rate of 23.8 m/s

at 10 ps slowing down to 13.9 m/s at a later time.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
143

A linearized instability analysis of the rapid evaporation for the superheat limit with

a plane surface was carried out by Prosperetti and Plesset (1984) and Higuera (1987).

The former asserted that the equivalent to the thin flame assumption of Landau (1944)

was not appropriate for the problem of evaporation and consequently was not made. The

analysis, however, depends on the undisturbed temperature distribution, which is

unknown, and therefore could not be evaluated by comparison with experiments. Using

several different temperature distributions, it was demonstrated that the surface of a

rapidly evaporating liquid can become unstable for a range of wave numbers, and at a

mass flux sufficiently large the growth resulting from the instability can be quite violent,

with characteristic growth times of a millisecond or less. Higuera (1987) included the

effect of viscosity, which is stabilizing, and vapor motion, which reinforces the instability,

and demonstrated that interfacial instabilities can occur, in agreement with Prosperetti and

Plesset (1984).

Grolmes and Fauske (1974) observed violent free plane flashing at a liquid-vapor

interface due to rapid depressurization of water to a low pressure of 0.004 atm. The

flashing consisted of vapor and entrained liquid, and these observations contributed to an

understanding of the phenomena of rapid evaporation. The effect of pressure on the

bubble growth at the superheat limit was studied by Avedisian (1982), with the result that

the violent bubble growth is suppressed by increasing the pressure. Frost and Sturtevant

(1986) and Frost (1988) examined the relationship between the explosive bubble growth

and the interface instability by means of the sensitive measurement of the far-field

pressure, The onset of the instability during the smooth bubble growth was clearly

demonstrated photographically by varying the ambient pressure, which supports the

pressure effects observed by Avedisian (1982). Direct evidence was obtained, by Frost

(1988) that fine liquid particles are tom from the liquid-vapor interface during the violently

unstable boiling, generating a vapor mass flux orders of magnitude greater than that

characteristic of ordinary boiling.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
144

Although a considerable improvement in the understanding o f rapid evaporation

has been established in terms of interface instability, a definitive demonstration of the

theory for the spherical case, with experimental confirmation, is not yet available.

Experimental results have been obtained which reveal that rapid evaporation can occur at

superheats lower than the homogeneous nucleation superheat limit. These results are

described and presented in the works of Ervin (1991) and Ervin at el.(1992). As will be

described in some detail below, both the Rayleigh-Taylor and the Landau instabilities

criteria were tested, using a numerical model introduced in Chapter 2, with the result that

stable conditions are predicted by both criteria, regardless of the spherical harmonic index,

in disagreement with experimental observations. The mechanism for the rapid evaporation

observed is interpreted here in terms of heat conduction. Any number of sources of

perturbations or instability could be present, but once a perturbation occurs, the heat

conduction in the liquid at the interface can be a principal mechanism for the subsequent

growth of this perturbation. Finally, a model for the rapid bubble growth will be presented

using the work of Mikic et al. (1970), which includes both inertia and diffusion controlled

bubble growth effects. The effect of the increased area due to the interface roughening

will also be included.

4.2 Experimental Results

Nucleation with pool boiling on a heated surface, described as heterogeneous

nucleation, is considerably more complex than homogeneous nucleation. The nucleation

not only occurs on an alien material, but takes place in a non-uniformly superheated liquid,

while homogeneous nucleation is generally regarded as being initiated in a uniformly

superheated media. Furthermore, only the latter case produces subsequent spherically

symmetric vapor bubble growth. Figures 4.1 (A)-(H) are photographs showing bubble

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
145

growths of R-113 on a flat heater surface with a step change in power imposed on the

gold film deposited on the quartz substrate. The upper part of each photograph shows the

view from the side of the heater surface, while the lower part is the view from beneath the

heater surface. As will be amplified below, the former occurs in a spacewise variable

temperature field, while the latter takes place more or less in an isothermal environment.

From such photographs it is possible to obtain both vapor bubble growth rates

perpendicular to the heater surface and liquid-vapor interface motion at and parallel to the

heater surface. Figure 4.2 is an enlargement of the side view of the frame in Figure 4.1

(E), while Figure 4.3 is the plot of the measurement of the transient mean heater surface

temperature and system pressure for the entire test of Figure 4.1.

Since the pictures were taken at 400 frames per second, the corresponding

uncertainty of the onset of nucleation is 2.5 milliseconds, which implies that the time scale

may be shifted up to 2.5 msec. The first hemispherical bubble is observed in Figure 4.1

(A) at 2.264 seconds after the heater power is initiated with a heat flux of 7.7 W/cm2. The

zero time in Figure 4.3 indicates the beginning of the heater power. The heater surface

was sufficiently smooth so that nucleation did not occur until the surface temperature

reached 122 C, as shown in Figure 4.3, the bulk temperature was 47.2 C to produce a

subcooling of 1.08 C, while the heater surface superheat at nucleation was 73.7 C. It is

noted that the system pressure jumps abruptly at nucleation from 103.6 kPa to 105.5 kPa

in about 0.04 seconds. The pressure jump could be considerably larger and destructive if

the heater surface were sufficiently large. The roughening of the bubble interface is readily

seen in Figures 4.1 (C),(D) and (E), and the measured bubble growth rate is much higher

than that predicted by the classical smooth bubble growth theory. The hemispherical

bubble in the Figure 4.2 has a radius of approximately 15 mm, and the length scale of the

roughness pattern have diameters on the order of approximately 0.5 ~ 2 mm. Due to the

large refraction effects with the lighting used, it was not possible to view the bubble

interface clearly from the underside. A hemispherical bubble forms, as shown in Figures

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
146

4.1 (A)-(C), while the lower interface of the bubble begins to spread to the right, forming

a highly rippled and relatively plane interface, and finally reaches the edge of the heater so

that vapor covers the entire heater. The evaporation which produces the bubble growth

can be considered as having two sources: the hemispherical cap of the bubble and the

lower edge of the bubble at the heater surface, both originating from the original thermal

boundary layer adjacent to the heater. The latter would include the thin liquid layer on the

heater surface itself. As will be confirmed elsewhere, the former source of evaporation is

influenced partially by the quick dissipation of the top of the hemispherical cap within 6

msec. Furthermore, as a result of physical constraints of the heater size, the latter source

of evaporation is no longer available as the interface spreads over the entire heater. The

liquid still has inertia and continues to move away from the heater, producing a so-called

overshoot so that the pressure of the vapor bubble becomes lower than the system

pressure, causing the vapor bubble growth rate to reach a maximum, which eventually lifts

the bubble slightly off the heater surface, as seen in figures 4.1 (G) and 4.1 (H). After the

bubble reaches a maximum size, it begins to collapse, which then is repeated to form an

oscillation. This oscillation results in an increase and decrease in the dry spot area of the

heater, causing rewetting with the aid of surface tension. The mean heater surface

temperature thus maintains a relatively low value for a short period of time, as may be

noted in Figure 4.3.

The liquid thermocouple 0.5 mm in diameter is located 1.65 mm above the heater

surface, and can be seen in Figure 4.1 (A). The measurements were converted to the

corresponding saturation pressures in order to determine the dynamic effects associated

with the pressure difference between the system and the vapor in the bubble. The

thermocouple measures the liquid temperature at the mean location 1.65 mm from the

heater surface until nucleation occurs, and thereafter measures the vapor temperature.

According to Figure 4.1 (B), the thermocouple is in contact with the growing vapor

bubble at about 5 ms and then begins to measure the vapor temperature in the bubble.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
147

During the interval from nucleation at 2.264 seconds to 4.3 seconds, in Figure 4.3, the

system pressure is larger than the bubble vapor pressure by a maximum 1.2 kPa,

equivalent to about 0.4 C saturation temperature difference, and these two pressures

eventually converge at 4.3 seconds, which implies that the vapor bubble has reached

thermodynamic equilibrium.

It is to be noted in Figure 4.3 that the one-dimensional transient conduction

prediction for the heater surface temperature is in good agreement with the measured

temperature until the onset of boiling. The mean heater surface temperature then drops

rapidly, but not instantaneously, following nucleation as the vapor bubble front sweeps

across the heater surface. This may be interpreted as providing evidence for the formation

and evaporation of a thin liquid layer under the growing bubble. The time required for this

temperature decrease is estimated to be about 190 ms, and the time required for the abrupt

pressure rise from (A) to (F) in Figures 4.1 and 4.3 is estimated at 40 ms. However, it is

seen in Figure 4.3 that the mean heater surface temperature does not begin to drop until

the time corresponding to (F), or 40 ms after the last frame with no vapor bubble present.

The time in Figure 4.1 (A) corresponds to the nucleation delay time t*=2.264 seconds,

defined from the onset of the heating process. It can be also be seen in Figure 4.1 (F) that

the vapor bubble now virtually covers the entire heater surface, and the surface could be

presumed to be dry. If so, the temperature decrease as a result of evaporation of any

liquid film on the surface ought to have occurred prior to this time. That the temperature

decrease followed the covering of the heater surface by the bubble implies the formation

and measurable evaporation time of a thin liquid layer under the growing vapor bubble.

The mechanism of the rippled plane-like interface spreading out over the heater, is

interpreted to be a consequence of a Rayleigh-Taylor instability, where the plane liquid-

vapor interface is unstable for the case with the light fluid accelerated toward the heavy

fluid. However, the source of the roughening with a hemispherical bubble is still uncertain,

since the parameters with a spherical shape now include the density ratio of vapor to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
148

liquid, surface tension, interface mass flux, divergence of the liquid adjacent to the bubble,

as well as the interface acceleration. This will be addressed in the discussion below.

The side views of the vapor bubble growth in Figures 4.1 (A)-(F) are compared in

Figure 4.4 with two tests having a lower level of heat flux and two significantly different

levels of subcooling. Figure 4.4 (b) is repeated from Figure 4.1, while Figures 4.4 (a) and

4.4 (c) have the same heat flux but subcoolings of 1.6 C and 11.2 C, respectively. For

the same heat flux, the heater surface superheat on nucleation is lower in Figure 4.4 (c),

with the higher subcooling, than in Figure 4.4 (a). However, the emphasis here is placed

rather on the character or appearance and growth rate of the vapor bubbles. At 2.5 msec,

the bubbles in both Figures 4.4 (a)-A and 4.4 (c)-A have the same smooth equivalent size.

However, at 5 msec, Figure 4.1 (a)-B shows a roughening of the bubble surface while

Figure 4.1 (c)-B still shows a smooth surface. The onset of an apparent interface

instability occurs between Figures 4.4 (a)-A and 4.4 (a)-B. This implies that either or both

surface tension and system pressure act to inhibit the onset of the instability, especially

when the bubble is quite small. Figure 4.4 (b) depicts the most violent or explosive bubble

growth, and has the highest heater surface superheat of 73.7 C at nucleation. It may be

concluded that as the heater surface superheat at nucleation is increased, the bubble

growth rate is increased significantly.

The growing bubble radii in the direction normal to the heater surface in Figures

4.4 (a) and (c) were measured approximately and compared with several spherical vapor

bubble growth models which assume a smooth interface, and are shown in Figures 4.5 and

4.6, respectively. The initial uniform and non-uniform superheat models were introduced

in Chapter 2 and Chapter 3 and consist of numerical solutions of the combined energy and

momentum equations. For the initial uniform superheat model, the entire bulk liquid is

assumed superheated at the level corresponding to the heater surface at nucleation; for the

initial non-uniform superheat model, the radial superheat distribution surrounding the

critical size vapor bubble is taken to be that normal to the heater surface at nucleation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
149

The former ought to produce much greater vapor bubble growth rates than actually occur,

while the latter ought to be lower than actually occur. For the analytical inertia dominated

model of Rayleigh (1917), the pressure producing the bubble growth is assumed constant

at the vapor pressure corresponding to the heater surface superheat at nucleation, an

upper bound on the growth rate. The measured growth rate in Figure 4.5 appears to be

linear, with a mean velocity of approximate 1.2 m/s, and is much higher than that of both

the uniform and non-uniform model predictions owing to the enhanced evaporation rate

associated with the increased surface area of the roughened interface. Since the uniform

superheat model assumes that the liquid temperature surrounding the hemispherical bubble

is equivalent to the heater surface temperature at nucleation, which is the highest

temperature in the actual thermal boundary layer, it can be regarded as the upper limit of

the bubble growth rate, while the non-uniform model assumes that the temperature profile

surrounding the hemispherical bubble is equivalent to the initial thermal boundary layer

profile on the plane heater, which may be regarded as the actual temperature profile on top

of the actual bubble, and provides the lower limit of the bubble growth rate. Figure 4.6

corresponds to the experimental conditions of Figure 4.4 (c), and depicts that the actual

bubble growth is slightly lower than that of the uniform superheat model, since the bubble

interface is smooth. This supports the concept that the roughening of the bubble interface

as a result of a surface instability leads to an increase in the evaporation rate. Also

included in Figures 4.5 and 4.6, for reference purposes, are the maximum liquid-vapor

interface motions at and parallel to the heater surface, referred to as "spreading" here. It is

to be noted that these motions are considerably more rapid than those normal to the

heating surface, producing a velocity of about 3.0 m/s in the case of Figure 4.5. These

large values are believed to occur because of the reduced liquid inertia effects when

compared to spherical bubbles, which must move all of the liquid in the same direction as

the liquid-vapor interface itself. This phenomena is quite complex in itself, and will not be

considered further here.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
150

A recurring question of bubble growth in nucleate boiling on flat heater surfaces is

how rapidly the energy in the superheated liquid boundary layer, formed during transient

heating on the heater surface, is converted to latent heat by evaporation as the boundary

layer is displaced away from the surface by the bubble growth. This will govern the

magnitude and duration of the evaporation and/or condensation taking place at the top of

the bubble away from the heater surface, sometimes referred to as the "bubble cap", which

in turn also influences the dynamics of the growth process. The answer requires the

simultaneous solution of the energy and momentum equations for the liquid as the bubble

growth takes place. To make the problem tractable at this time, the interfacial behavior at

the hemispherical bubble cap will be assumed to be reasonably represented by a spherically

symmetric smooth bubble growth. This model was solved numerically as shown in

Chapter 3. A typical result showing the transient temperature distribution in the liquid is

given in Figure 4.7 for R-113 using the experimental parameters indicated. The initial

temperature distribution is that given by transient conduction at the imposed heat flux for

the nucleation delay time period of t*=2.264 seconds. It is seen that the thermal boundary

layer surrounding the bubble, initially about 1.5 mm in thickness, is virtually eliminated in

approximately 6 ms. This results from both evaporation at the interface and the liquid

divergence surrounding the bubble. Since the actual bubble grows faster than that

predicted by the smooth bubble calculation, it is anticipated that the real time for the

energy dissipation in the boundary layer will be considerably shorter than 6 msec. This

means that for the vapor bubbles shown in Figures 4.4 (a)-B and 4.4 (c)-B the temperature

distribution in the liquid at the liquid-vapor interface on the top of the bubble has become

nearly completely diffused. In the case of Figure 4.4 (a)-B, the evaporation at the bubble

cap ceases to provide additional vapor since the bulk liquid is almost saturated, while in

the case of Figure 4.4 (c)-B condensation takes place on the top of the bubble since the

bulk liquid is subcooled at about 11 C. This eliminates a source of evaporation on the

cap of the bubble, although a source continues to act over the lower part of the bubble.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
151

The overall evaporation rate over the entire bubble surface is reduced, which would tend

to suppress the interface instability, as observed in Figure 4.4 (c).

4.3 Instability Considerations

Rapid evaporation is not prevalent in normal pool boiling processes, but has been

observed when special steps are taken to maximize the degree of liquid superheat, which is

termed the superheat limit. As described above, rapid evaporation of R -l 13 occurred at

superheats much lower than the superheat limit, on the order of 70 C, whereas the limit

of superheat is about 160 C. Analytical studies of a particular feature of the rapid

evaporation at the superheat limit have been conducted, attempting to quantify the source

of the instability which gives rise to the intense evaporation.

Shepherd and Sturtevant (1982) conducted experiments to achieve homogeneous

nucleation by injecting a droplet of highly volatile n-butane into the bottom of a cylindrical

vessel filled with hot immiscible ethylene glycol so that the droplet temperature reaches

the superheat limit as it rises due to buoyancy. Figure 4.8 shows the roughening of a n-

butane vapor bubble at 12 ps following nucleation, which took place at a bulk temperature

of 104.5 C and atmospheric pressure. Figure 4.8 (a) depicts a n-butane droplet 1 mm in

diameter which contains a growing vapor bubble, and Figure 4.8 (b) is a photographic

enlargement of the 150 pm diameter bubble. The resulting measured effective radius of a

spherical bubble is plotted in Figure 4.9, as a function of time, together with measurements

of McCann et al. (1989), the momentum governed growth predicted by Lord Rayleigh

(1917), the approximate solution of Mikic et al. (1970), and the numerical solution

introduced in Chapter 2. It is remarkable that the measured growth rates appear to be

linear, with a mean velocity of approximate 14.3 m/s, with good agreement between the

two sets of experiments. The growth rate is less than the momentum controlled growth of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
152

Rayleigh (1917), to be expected, but larger than the predictions of Mikic et al. (1970) and

the numerical calculations. The numerical computations provide the velocity and

acceleration in the early stages of growth, shown in Figure 4.10 for n-butane under the

same conditions as in Figure 4.9. The maximum interface acceleration is of the order of

2 x l0 10 m/s2, although for an exceedingly short time. The interface velocity reaches 32 m/s

early and then decreases to 3 m/s at 1 msec. It is instructive to compare the measured

velocity of 14.3 m/s cited above from Figure 4.9 with the predicted velocity at, say, 100

ps in Figure 4.10, giving about 6 m/s. It is smaller because the model presumes a smooth

interface. Using the information in Figure 4.10, both the Rayleigh-Taylor and Landau

instability criteria can be examined on a time basis, determining the range of the index of

spherical harmonics (n) necessary for instability. As indicated earlier, two closed form

descriptions of these instability criteria are available and will be summarized below.

4.3.1 Rayleigh-Taylor Instability

The Rayleigh-Taylor instability for a spherical interface has been established by

considering whether a small amplitude disturbance on the interface grows or diminishes.

It is understandable that as soon as spherical symmetry is excluded, the analysis of the

various aspects of bubble dynamics becomes exceedingly complex, both from a theoretical

and an experimental standpoint. An obvious possibility for circumventing this complexity


is the use of an expansion in terms of spherical harmonics Yn. The equation of the bubble

surface may be expressed as:

rs = R(t) + anYn (4.1)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
153

where R(t) is the instantaneous average radius, an is the amplitude of the spherical

harmonic of degree n. Equation (4.1) may be substituted into the governing equations and

the basic equations subtracted so that only the disturbance equations remain. The detailed

derivations may be obtained by reference to Plesset (1954), Birkhoff (1956) and Plesset
and Prosperetti (1977). After some manipulations, the differential equation for an, from

which the stability conditions can be deduced, is obtained as:

+ 3- + 2(/i + 2)(2 + 1)-^- an Aan = 0 (4.2)


A A

The general stability criteria show that oscillating bubbles are unstable when:

A > 0 , or (4.3)

, _ , d(R6A) .
A < 0 and ----- - > 0 (4.4)
dt

V,R
where A = (n - 1) 3
(4.5)

The criterion in Equation (4.4) is especially interesting because it predicts the

instability for the spherical case under circumstances for which a plane interface would be

stable. In this case the surface perturbation oscillates in time with algebraically increasing

amplitude. It is to be noted that the acceleration and surface tension are dominate

parameters in Equation (4.5).

Using the numerical calculations discussed in Chapter 2 for the operating

conditions of n-butane in Figure 4.9, the Rayleigh-Taylor instability criteria of Equations

(4.3) and (4.4) predict that the vapor bubble should be stable throughout the growth

period, contrary to the observation of interface instability shown in Figure 4.8. This

comparison was extended to the recent experiments with R-113 for the operating

conditions in Figure 4.4. Unconditional stable circumstances were predicted regardless of

the spherical harmonic index, again contrary to the unstable observations in Figures 4.4 (a)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
154

and 4.4 (b). Shepherd and Sturtevant (1982) also tested the criteria in the same manner

except that the classical bubble growth model of Prosperetti and Plesset (1978) was used,

again with disagreement with the experimental observations. This implies that the

Rayleigh-Taylor instability criteria for spherical bubbles should be reconsidered.

4.3.2 Landau Instability

The instability of Landau (1944) was originally derived for a plane combustion

flame propagation under the assumption of a thin flame. Istratov and Librovich (1969)

later developed the Landau theory for a spherical flame, suggesting a closed form of the

instability criteria. The flame is considered to be unstable if the growth rate of the

disturbance amplitude is greater than that of the radius of the flame sphere. Frost and

Sturtevant (1986) extended and applied the Landau stability, and expressed the

perturbation amplitude as:


(4.6)

where Q. is the nondimensional algebraic growth rate.

The resulting dispersion relation is:

f l 2 + bl + c = 0 (4.7)

where a (4.8)
P,

b = 3 + -n-(1+2an) (4.9)
(n + l ) a + n

Nw
a(ocn + l)(n2 -1 ) - cm2(n -1 )
(n + l ) a + n [(n + l)a+n]A^,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
155

n is the spherical harmonic index. Positive values of Q mean that the perturbation

amplitude grows more rapidly than the radius, leading to instability. It is to be noted here
that the density ratio (a)in Equation (4.10) is a dominant parameter, as stated by Istratov

and Librovich (1969).

The Landau instability criteria for spherical vapor bubble growth, expressed by

Equation (4.7), was tested for the same conditions as that for the Rayleigh-Taylor

instability above, using n-butane with the parameters given in Figure 4.9, along with the

results of the numerical calculations from the solution procedure discussed in Chapter 2.

It was noted above that the results of the instability test depend on the density ratio (a);

this is presumed to be a function of the interface temperature which can now be evaluated

during the bubble growth process. The result of the computation is that the growing

bubble is unconditionally stable regardless of the index of spherical harmonics for a density

ratio of 0.004, evaluated at the saturation temperature of -0.5 C, corresponding to

atmospheric pressure, but is quite unstable for the density ratio of 0.077, evaluated at the

bulk temperature of 105 C, which corresponds to a saturation pressure of 17.3 atm. The

density ratio of 0.004 at the atmospheric pressure saturation temperature of -0.5 C, may

appear to be a more pragmatic value than that of 0.077 for the bulk temperature

corresponding to the superheat limit of 105 C since upon nucleation and early bubble

growth the vapor produced will be close to the bulk temperature and corresponding

saturation pressure. However, if an incipient instability is present early in the vapor bubble

growth period, just beyond the critical size, it is possible that the vapor pressure will be

close to the saturation value of the bulk liquid temperature because of liquid inertia, and

the associated interface roughness produced here will be retained during further growth.

The issue of temperature dependent properties on instabilities has not been addressed to

date. Nevertheless, in order to examine the characteristics of the stability in more detail

here, particularly with regard to the effect of the system pressure, the n-butane bubble

growth was numerically computed, again with the solution procedure discussed in Chapter

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
156

2, for three different pressures with the initial temperature at the superheat limit of 105 C.

The density ratio (a) was taken to be constant at the initial bulk temperature of 105 C,

neglecting whether such might be valid in the Landau instability theory. The stability

results are plotted in Figure 4.11, and indicate that the stability area reduces as the

pressure increases, which is consistent with the experimental results for R-113 presented

in Figure 4.4 (a) and 4.4 (c). It is to be noted in Figure 4.11 that the onset of instability

occurs at a time of about 2x1 O'7 sec, while the maximum acceleration in Figure 4.10

occurs at a time of 5xl0'9 sec, which would imply that this enormous acceleration ought

not to effect the instability at all. The computed liquid-vapor interface temperature is also

included in Figure 4.11. Frost and Sturtevant (1986) applied the Landau criteria at the

superheat limit with the result that a growing n-butane bubble should be stable, which was

in disagreement with their experiments. The Landau stability test was extended here to

the experiments with R-113 shown in Figure 4.4 (a), (b) and (c), predicting stability for all

three cases regardless of the density ratio. It can be concluded from the above that the

analytical basis for the Landau instability criteria should be reconsidered for liquid-vapor

phase changes, for the spherical cases.

4.3.3 Instability Growth in Terms of Heat Conduction

Thus far the source of the disturbance producing a "rough" evaporating liquid-

vapor interface has been considered to be an instability, acting as a trigger, as it were.

Other sources of the disturbance are possible; for example, a differential vapor recoil

mechanism for a plane interface was discussed by Palmer (1976), and is defined as the

depression of the liquid-vapor interface by the momentum of the vapor departing the

interface. However, this does not ensure that the disturbance will grow. Another possible

mechanism might be the fluctuation or constriction of thermal diffusion in the liquid at the

liquid-vapor interface. By considering only the heat conduction in the liquid adjacent the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
157

bubble wall, it is hypothesized that any small disturbances could make the interface more

rough or wavy, as will be demonstrated below using velocity potential theory.

Neglecting the curvature at the liquid-vapor interface of a vapor bubble relative to

that of a disturbance on the interface, the two-dimensional Cartesian Laplace equation for

an inviscid fluid with irrotational motion is taken to represent the perturbed motion in the

liquid in the immediate domain of the liquid-vapor interface:

32<e
2 + T T =0 (4.11)
dx2 dy

A small disturbance, expressed by Equation (4.12), is imposed on the liquid-vapor

interface. It is assumed that the far field fluid is not effected by the disturbance at the

interface. Since no mass transfer occurs at x=0 and x=A/2 due to symmetry, the boundary

conditions can be written as:

B.C. | ( x , 0 ) = -C/0c o s p H ) (4.12)

? r ( x ,h ) = 0 (4.13)
dy

3d>
^ (0 ,j) = 0 (4.14)

" ^ l - O (4.15)
dx

where X indicates the wavelength and h is an arbitrary depth of the liquid, taken to

be large relative to the wavelength. Using the separation of variables method, the

solution for the velocity potential is obtained as:

(x >y) = ~ (cosh(ey)-tanh(e/i)sinh(ey))cos(ex) (4.16)


tan(e/i)

where e = 2 n / X

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
158

A representation of the velocity potential is plotted in Figure 4.12. As a result of

the transient heating of a liquid, a one-dimensional temperature gradient can be considered

initially to exist normal to the originally undisturbed liquid-vapor interface immediately

following the formation of the liquid-vapor interface. Once the small disturbances in the

velocity field in the liquid resulting from Equation (4.12) have taken place, the

temperature field in the liquid can be taken as indicated in Figure 4.12 since velocity

potentials are analogous to isotherms in a static medium. It can then be deduced that the

conduction heat transfer will be increased at the so-called troughs and decreased at crests,

which would cause the disturbances to grow. The disturbance would also tend to grow

because of the vapor pressure depression at the interface troughs. Parts of the crest can

become distorted such that detachment of particles liquid from the interface take place and

become entrained with the vapor. Such a mechanism can play an important role in

increasing the interfacial evaporative flux, as was described by Grolmes and Fauske (1974)

and experimentally confirmed by Frost (1988).

The growth of any disturbance at an interface depends on the net balance between

the stabilizing and the destabilizing factors. The stabilizing factors are the surface tension

and the divergence of the streamlines as the spherical bubble grows, while the destabilizing

elements are the differential vapor recoil and the heat transfer enhancement at the troughs.

4.4 Increased Evaporation Surface Area

In order to estimate the effect of the increased evaporation surface area due to the

interface roughness, the small perturbations are modeled by two-dimensional sinusoidal

functions of amplitude and wavelength A,, as shown in Figure 4.13, by the mathematical

formulation:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
159

(4.17)

The perturbed area(Ap) and smooth area (As) on a spherical bubble of radius R

can be calculated using double integration assuming a small perturbation compared to the

radius, A, R. It is determined that the ratio of the perturbed to the smooth area is a

function of only the ratio of the amplitude to the wavelength, and is plotted in Figure 4.14.

For example, the ratio of the perturbed to the smooth area is approximately 4 for a unity

ratio of amplitude to wavelength regardless of the wavelength or amplitude.

To determine how the increased area affects the subsequent bubble growth, a

simple analysis is performed with several assumptions: The wavelength of the

perturbation is much smaller than the radius of the spherical bubble; one-dimensional

transient heat conduction only takes place at the perturbed liquid-vapor interface. The

energy balance at the interface is written as:

(4.18)

where Ap is the perturbed area on a spherical bubble and As is the smooth area for

the spherical bubble. One-dimensional transient heat conduction gives:

(4.19)

where-\/3 is the curvature correction.

Combining Equations (4.18) and (4.19), after some arrangements, yields:

(4.20)

where T = is defined as the ratio of the perturbed area to the smooth area.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
160

Integrating Equation (4.20), assuming that Tis constant, gives Equation (4.21),

which is the well known solution for the thermal diffusion controlled bubble growth.

(1 2 V/2
rt(f) = r l j Ja( a t)'12 (4.21)

where Ja = P,C,(T~ T^ (4 22)


PA

Equation (4.21), with T=l , was first formulated by Plesset and Zwick (1954).

Equation (4.20) with T=constant is now used here to extend the work of Mikic et al.

(1970), which includes both inertia and diffusion effects. The result is:

2
R+= -[(r+r 2+ i)3/2- (?-r2)3/2- i] (4.23)
3T

where R* = ; t += 1 (4.24)
B /A B2 / A 2

A=
3TsatPl J
B JU ^ a ')T (4.25)

Bubble growths for n-butane are plotted in Figure 4.15, using Equation (4.23), for

various values of T, the ratio of the perturbed to the smooth area, along with the

experiments of Shepherd and Sturtevant (1982) and McCann et al. (1989). T=1 indicates

perfect smoothness for the spherical bubble. The experimental data fall between a range
of r =3-5, which corresponds to the range of C, / A,=0.6-1 in Figure 4.14. This indicates

that the amplitude of the perturbation is 60-100 % of the wavelength, so when viewing

the photograph of Figure 4.8 (b), the sizes of the pattern appear reasonable, although it is

difficult to speculate about the magnitude of the amplitude.

Similar calculations were carried out for the R-113 experiment PBMT1026.600

and are plotted in Figure 4.16. The results provide T ~ 2 roughly, which from Figure 4.14

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
161

gives ^/A,=0.3. When comparing the present experiment with the work of Shepherd and

Sturtevant (1982), the relative wavelength is shorter in the present experiment than in that

of Shepherd and Sturtevant (1982), which can be expressed in terms of the spherical
harmonic index (n = 2nR/ X). By measuring the diameter and the patterns of the

perturbation, the spherical harmonic index for the present experiment and that of Shepherd

and Sturtevant (1982) are calculated to be in the range of approximate 20-60 and 8-12,

respectively. The latter values fall marginally on the stable/unstable boundary in Figure

4.11.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.1. Dynamic vapor bubble growth of R-l 13 in microgravity at NASA Lewis
Research Center 5 second drop tower. Experimental Data Code PBMT1102.800.
Nucleation delay time t*=2.264 seconds.

Test conditions: = 10"5; qT = 1.1W / cm2\ ATsubC00l = 1.08C; = 103.8kPa;


&

A7;*sup= 73.70C; Tbulk=A1.2C\ ria, =48.28C; rJ = 122C

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
163

(E). t=17.5 msec (F). t=40 msec

(H). t=275 msec


(G). t=100 msec

Figure 4.1. Continued.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
164

t=17.5 msec

Figure 4.2. Enlargement of Figure 4.1 (E) for Experimental Data Code PBMT1102.800

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
165

140 - Experimental Data Code PBMT1102.800


. Tbulk=47.2 C, Tsat=48.28 C

120 _ Psys=103.8 kPa


108

uO 100 107

u>
S 106

Pressure, kPa
*-*
2u
a,
E 105

ou
Measured surface 104
t!3 temperature
Vi
1-D transient conduction 103
prediction
102
Measured pressure

Photographs 101
Measured liquid
temperature converted to 100
saturation pressure

0 0.5 1.5 2 2.5 3 3.5 4 4.5 5


Time, sec

Figure 4.3. Measured mean heater surface temperature and system pressure for Figure
4.1 (A)-(H). A thermocouple employed to measure the liquid/vapor temperature is
located 1.65 mm above the heater surface.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
166

PBMT1026.600 PBMT1102.800 PBMT1108.620


?t
, - ' $S

t=2.5 msec t=2.5 msec A


' </: .... t=2.5 msec A
,y r

t=5 msec t=5 msec


t=5 msec B

t=7.5 msec t=7.5 msec


t=7.5 msec C

t=10 msec D t= 10 msec


t=15 msec

t=12.5 msec E t=37.5 msec E


t=17.5 msec p f ' -i

t=20 msec t=40 msec =65 msec

qH-p$.3 W/cm2, q"X=7.7 W/cm2, q"Ts54 W/cm2


t*=3.923 sec, t*=2.264 sec, t*=4.325 sec,
Psys=103.85 kPa, Psys=103.8 kPa, Psys=142.2 kPa,
Tbulk=46.72 C, Tbulk=47.2 C, Tbulk=47.0 C,
Tsat=48.34 C, Tsat=48.28 C, Tsat=58.2 C,

rw
*=114 C 7;*=122 C rJ=116C

Aritt6=1.62C A 7 > 1 .0 8 C ATsub=ll.2C

A C p =65.66 C A^Csup=73-72 C A C P=57.8C

(a) (b) (C)

Figure 4.4. Comparisons of early vapor bubble growth behavior under microgravity
with different levels of heat flux and subcooling.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
167

0.024
Radius vs. Time for R113, Psys=103.85 kPa, Tsat=48.34 oC

0.02

Bubble growth
0.016 PBMT1026.600
Radius, meter

0.012

Spreading, PBMT1026.600

0.008 Initial Uniform Superheat Moc el,


Superheat=65.66 vC, Tbulk=ll 1C

0.004 hfaTNon-uniform Superheat Model,


Twall=ll4 C, Tbulk=46.72 C,
y^uperheat=65.7 C

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Time, sec
Figure 4.5. Comparisons of experimental bubble growth data with several spherical vapor
bubble growth models. Experimental conditions correspond to Figure 4.4 (a).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
168

Radius vs. Time for R113, Psys=142.2 kPa, Tsat=58.21 C


0.006

0.005 Rayleigh (1917)

Initial Uniform Superheat Model


Superheat=57.79 C, Tbulk=116 C
0.004
Radius, meter

0.003 Spreading, PBMT1108.620

Bubble growth, PBMT1108.620


0.002

0.001

Initial Non-uniform Superheat Model


Twall=l 16 C, Tbulk=47 C, Superheat=57.8 C
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016
Time, sec
Figure 4.6. Comparisons of experimental bubble growth data with several spherical vapor
bubble growth models. Experimental conditions correspond to Figure 4.4 (c).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
169

180
Liquid temperature distribution Time from onset of growth
-for Non-uniform model
0.0 sec
160 R-113 (PBMT1102.800)
2.50E-06
Psys=103.75 kPa, Tbulk=47.2 C,
Twall=122 C, Tsat=48.28 C 2.00E-04
140 q"T=7.7 W/cmA2, t*=2.264 s
1.50E-03
Superheat=73.7 C, Subcooling=0.8 C
u 6.00E-03
120
3

U
a.
I 100
E-

le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02


Distance from liquid-vapor interface, meter

Figure 4.7. Transient liquid temperature distribution surrounding a spherical vapor bubble
growing from the critical size with an initially non-uniform liquid temperature distribution.
ATJsup = 73.7 C, ArjH6=0.8C

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
170

Figure 4.8. Photographs of n-butane nucleating at the superheat limit. (Shepherd and
Sturtevant, 1982) (a) depicts a n-butane droplet containing a vapor bubble, (b) an
image-enhanced photograph of a vapor bubble at t=12 ps following nucleation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
171

Radius vs. Time for n-Butane


1.4
Psys=l atm, Superheat limit=105 C, x
Tbulk=104.5 C, Tsat= -0.5 C
1.2
Rayleigh (1917) '

Shepherd and Sturtevant (1982)

The present work


0.8 / McCann et al (1989) x x (uniform model)
\ nX

0.6

0.4 Mikic et al (1970)

0.2

0 20 40 60 80 100
Time, psec

Figure 4.9. Comparisons between extremely dynamic bubble growth experiments, a


momentum controlled model, and approximate and numerical solutions of the energy and
momentum equations for n-butane.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
172

40 3.0e+10
-Interface Velocity and Acceleration vs. Time for n-Butane
35 - Psys= 1 atm, Superheat Limit=105 C
2.5e+10
- Tbulk= 104.5 CLTsat=-0.5 C
30
Velocity
2.0e+10

Interface acceleration, m/sA2


25 Acceleration
Interface Velocity, m/s

20 1.5e+10

15
1.0e+10

10
5.0e+09
5

0.0e+00
0

5 -5.0e+09
le-10 le-09 le-08 le-07 le-06 le-05 le-04 le-03
Time, sec

Figure 4.10. Liquid-vapor interface velocity and acceleration vs. time for n-butane.
Numerical computations. Physical conditions identical to Figure 4.9.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
173

500 150
Spherical Harmonic Index vs. Time for n-Butane
450
100

400 Bulk temperature

350
e
Interface temperature

Interface temperature,
300
o
'5
o
| 250 -50
Stable
X Stable
| 200
oj -100

150
atm -150
100

-200

Unstable
-250
le-10 le-09 le-08 le-07 le-06 le-05 le-04 le-03 le-02 le-01
Time following onset of bubble growth, sec

Figure 4.11. Effect of system pressure on Landau instability criteria for n-butane. The
density ratio ( a ) is assumed constant at bulk temperature of 105 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
174

1.5
Velocity Potential

1
Liquid
0.5
Trough
0 Vapor
Crest

-0.5
0 1 2 3 4
x

Figure 4.12. Velocity potential simulation of a temperature field in the liquid at the liquid-
vapor interface. The velocity potential was superimposed on x-y axis.

S
Figure 4.13. Modeling of a perturbed liquid-vapor interface.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
175

n
*
/

/
/

fr

/
/
Ap/As

/
/

T-------

0.01 0.1 1 10 100


yx

Figure 4.14. Ratio of perturbed to smooth surface area of a spherical bubble as a function
of the perturbation amplitude to wave length ratio.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
176

- Bubble radius vs. time for n-butane


1.6
=7

1.4 Shepherd and Sturtevant (1982) =5

1.2

r= 3
I 1
t/T
.2
1 0.8

qg McCann et al. (1989)

r= 1
0.4

0.2

0 10 20 30 40 50 60 70 80 90 100
Time, ps

Figure 4.15. Comparison of experimental data for n-butane with the modified prediction
of Mikic et al. (1970), Equation (4.23), which includes the effect of the increased area due
to the perturbation. T is the ratio of rough to smooth area.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
177

0.014

Radius vs. time for R-l 13

0.012

r=3
0.01
Radius, meter

0.008
r =2

0.006

0.004
r =i

0.002

Experiment PBMT1026.600
0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014
Time, sec

Figure 4.16. Comparison of experimental data for R -l 13 with the modified prediction of
Mikic et al.(1970), Equation (4.23).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
CHAPTERS
CONCLUSIONS

5.1 Initial Uniform Superheat Model

As a result of the comparison of the results of the present numerical work with

previous numerical, experimental and analytical works, it has been demonstrated that the

present work conforms well to physical reality. The initial temperature disturbance

defined in this work and used to initiate the vapor bubble growth from the initial critical

size in thermodynamic equilibrium with the superheated liquid has a significant physical

meaning, and its use enables the accurate prediction of vapor bubble growth with water

liquid superheats as low as 0.8 ~ 1.0 C, measurements for which have not heretofore

been in agreement with analytical solutions. The variation of the disturbance does not

effect the subsequent bubble growth.

The bubble dynamics from nucleation to thermal diffusion controlled growth are

described by the present numerical simulation for initially uniformly superheated liquids.

Nucleation takes place in the uniformly superheated liquid to form a critical size bubble.

The time lag or delay time for the onset of the bubble growth increases as both the

superheat and the pressure decrease. This is defined as the time at which the bubble

grows rapidly, following the surface tension dominated growth. The liquid-vapor

interface velocity then reaches a maximum, which increases as both the superheat and the

pressure increase. The bubble growth rate is then limited by the surrounding liquid inertia,

and is defined as the momentum dominated growth. Finally, the acceleration of the bubble

178

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
179

interface reduces to zero, following which the growth is defined as thermal diffusion

dominated.

For moderate or high superheats, the analytic solution of Mikic et al. (1970) is in

excellent agreement with the present work, while those of Plesset and Zwick (1954) and

Scriven (1959) are in good agreement For low superheat and low pressure none of the

analytic solutions are in agreement with present work during the early stages of the vapor

bubble growth.

5.2 Initial Non-uniform Superheat Model

Spherical vapor bubble growth and collapse in non-uniform temperature fields can

be effectively predicted by the initial non-uniform superheat model developed, which also

provides the transient vapor pressure in the vapor bubble and the surrounding liquid

temperature distribution. Vapor bubble rebound during the collapse of a bubble has been

observed and predicted, and the numerical results indicate that the conversion of kinetic

energy to thermal energy during the collapse provides the energy for the rebound to take

place.

Transient pool boiling in microgravity can take place with a known initial

temperature field, to be contrasted with the randomness of the initial temperature field in

earth gravity experiments. The amount of energy input stored, defined as the product of

the heat flux to the liquid and the waiting time, is a significant parameter for estimating the

bubble growth, along with the bulk liquid subcooling.

A hemispherical bubble model developed here provides the upper and lower

effective bounds for the bubble growth and collapse.

Upon comparison of the numerical results between the initially uniform and non-

uniform superheat models, it is deduced that the bubble dynamics in the very early stages

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
180

of growth are affected negligibly by the thermal boundary layer, but rather by the initial

bubble wall temperature level.

5.3 Role of Instability in Pool Boiling

The possibility for the occurrence of extremely rapid evaporation at a vapor bubble

interface has been demonstrated with liquid superheats far below the superheat limit. This

took place with pool boiling in a microgravity environment, and is attributed to the growth

o f an interface instability. Such a process can cause an abrupt pressure rise and could be

destructive if it occurred over a large scale of surface area.

For spherical bubble growth taking place in the presence of an initially non-uniform

superheat, arising for example from thermal energy accumulated in the boundary layer on

a flat plate heater, it was found by numerical calculation that this thermal energy can

dissipate quickly, on the order of several milliseconds. This time period is significant since

it determines whether evaporation or condensation take place at the cap of a vapor bubble

growing on a flat heater surface.

Sources of perturbation growth other than instability can arise at the liquid-vapor

interface of a bubble. One mechanism for the growth of a perturbation can result from

heat conduction at the perturbed interface, where the heat transfer increases at a trough

while it decreases at the crest, which can produce differential vapor recoil. This can lead

to distortion of the crests such that liquid particles become entrained with the vapor,

departing the interface, which can serve to increase the evaporative flux further.

Using a simple circular function in a model to describe the liquid-vapor interfacial

area with roughness present, it is determined that the ratio of the perturbed to the smooth

area (T) for a spherical bubble is a function only of the ratio of the amplitude to the
wavelength ( / X).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
181

The rapid bubble growth can be approximated by modifying existing solutions with

a parameter which takes into account the increase in the evaporative surface area. Recent

experiments provide estimates of a factor of two for the increase in the evaporating

surface area, which conforms to the ratio of amplitude to wavelength of 0.3 for the model.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
APPENDICES

182

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
183

APPENDIX A
Diffusion Controlled Vapor Bubble Growth

The bubble growth equation of Plesset & Zwick (1952) is reconsidered. In order

to see how the temperature in a moving boundary varies as a bubble grows, Equations

(2.3), (2.6), and (2.8) are rewritten here:

( 2 -3 )

f a \ m f'
T - T I I
'Mf') V r Jr=R(x) dx ( 2 .6)
' ~ U J n f', . r

\X

/?(r) = ^ j Ja(at)112 = C / ' 2 (2.8)

Equations (2.3) and (2.8) are inserted into Equation (2.6).


After some manipulation, Equation (2.6) can be written as:

3 p Ji
T, = T -C . ---- -dx (2.6a)
' 1 2k

The integral term above can be analytically simplified, substituting with x = ty and

7C
successively y3 = z2. After some manipulation, the term yields the exact value of ,

which means that Tt is independent of time. Successively, substituting C, with the defined

Equation (2.8) yields precisely that Tt = Tsat. Hence, it is demonstrated that the

temperature (7!) at the moving boundary does not vary as the bubble grows when

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
184

Equation (2.8) is used for the bubble growth problem in a uniformly superheated liquid.

Plesset & Zwick state that Equation (2.8) will be valid for time large enough to neglect

inertia, and that the growth is thus controlled by diffusion when Equation (2.8) is used.
This coincides conceptually with the result obtained that 21 = Tsat . It is noted that the

reality is that 71 drops in a very short time to the saturation temperature Tmt from the

superheated liquid temperature (7^), and thereafter maintains the saturation temperature

as the bubble grows.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
185

APPENDIX B
Momentum Equation for Spherical Bubble Growth

The momentum equation in spherical coordinates is given by:

du du 1 dP l l f 1 9 ( id u } 2u
(2.18)
r 2 dr dr

The continuity equation with assumption (d) is given by:

d ( 2\ n <*u _u
[r u) = 0 or = - 2 (2.19)
dr dr r

It is readily demonstrated that the second term in Equation (2.18) is zero by

inserting Equation (2.19) into the term. This is to be expected, since no tangential stresses

are present due to spherical symmetry. Considering the control volume at the liquid-vapor

interface as shown in Figure B -l, continuity is:


Pi{ ur - r ) = Pv( uv - r ) (B-l)

let e = 1 - A (B-2)
. P I.
Then, ur = eR + ( 1 - e ) uv (B-3)

Since uv = 0 according to assumption (b), Equation (B-3) can be written as:

ur = eR (B-4)

Using Equation (2.19), continuity is written as:

R 2R
Mr = = e 2- (B-5)
r

Taking derivative of Equation (B-5) with respect to time (t) yields:

du ( RR2 + 2RR2'
= (B-6)
dt

Substituting Equations (2.19), (B-5), and (B-6) into Equation (2.18) yields:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
186

(R R 2+2RR2} _R*R2_ __J_3P


-e - (B-7)
p dr

Integrating Equation (B-7) from the bubble surface (R) to infinity (<*}, the

momentum equation can be written in terms of the bubble radius as:

(2.20)
2 ep,

Considering the force balance at the liquid-vapor interface as shown in Figure B-2,

one obtain:

(B-8)
/>* + T + T "* = Pv +T ,, v

The normal stress at the liquid interface can be written as:

(B-9)

Since V 7 = 0 for incompressible flow, Equation (B-9) can be written as follows,

using Equation (2.19):

'E l (B-10)

and x rn = 0 because uv = Owith Assumption (b).

Therefore, Equation (B-8) becomes

p>(T)_ p=MZ2+4M (2 .21 )


R R
where PR is the liquid pressure adjacent the bubble wall.

Inserting Equation (2.21) into Equation (2.20) yields

2C(D a R
-4 - (2 .22 )
2 ep, ep ,R ep ,R

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
187

iiR Liquid

PV

Vapor bubble
Control Volume
a t liquid-vapor interface

Figure B-l Control Volume at liquid-vapor interface and velocity at a distance (r)

rrRj
Liquid

rrv.

Vapor bubble

Figure B-2 Force balance at Liquid-vapor interface

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
188

APPENDIX C
Sonic Velocity of Water Vapor

The root mean square molecular speed of water vapor is:


v 1/2
3R T V/2 _ f 3 x 8.3145 x 103 x 409
= 752 m /s (C-l)
M ) 18.015

The sonic velocity in water vapor is:


c = ( m O 1'2 = (1.327 x 461 x 409)1'2 = 500 m / s (C-2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
189

APPENDIX D
Homogeneous Nucleation Temperature

A generalized equation for estimating the superheat limit or homogeneous

nucleation temperature is given by Reid (1983) as:


rrf= r e[(0.11Pr)+0.89] (D-l)

where

Tc : critical temperature, K

Pc : critical pressure, bar

Ts, : superheat limit, K

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
190

APPENDIX E
A Fortran Program for Sperical Vapor Bubble Growth and
Collapse

c *MAINPROGRAM
c
c ON JUNE 2,1992, PROGRAMMING BY HO SUNG LEE
c
c THIS PROGRAM SOLVES A 1-D TRANSIENT SPHERICAL BUBBLE GROWTH PROBLEM
c IN INITIAL UNFORM OR NON-UNIFORM SUPERHEATS, WHICH INCLUDES THE EQN
c OF MOTION AND A ENERGY EQUATION WITH VISCOSITY AND SURFACE TENSION.
c THIS PROGRAM ALLOWS TO IMPLY THE INITIAL THERMAL BOUNDARY LAYER BY
c GIVING THE HEAT FLUX (Q) AND WAITING TIME.
c THIS PROGRAM MUST BE COMPILED TOGETHER WITH TWO MORE SUBPROGRAMS,
c A FUNCTION PROGRAM(WATFUNC.FOR) AND A SUBROUTINE PROGRAM(SBDSUB5.FOR)
c THIS PROGRAM PRODUCED TWO OUTPUT FILES: ONE FOR GROWTH DATA AND OTHER
c FOR TEMPERATURE DISTRIBUTIONS.
c THIS PROGRAM ALSO REQUIRES INPUT FILE AS FOLLOWS:
c
C***INPUT FILE (WAT.IN) Hr***********************
c
c l.D-1 3. 100 3.D-3 FALSE FALSE 10 10.DO 300
c 0 .DO 1.550 103 .5 l.D-3 500 1000 5000 8000
c 112.DO l.D-1 l.D-10 10000 3.0 l.D-4 l.D-10
c TMAX P IL DELTA PRINTl PRINT2 NSKIP AX BX
c Q TSTAR PA(KPA) EXCITE NSPl NSP2 NSP3 NSP4
c TBULK(C) ESP ESPP NTIME PT COF1 COF2
(1
c, ,,,.NOTES FOR INPUT FILE.
L.
c DT=COFl* (KOUNT/NTIME) ''PT + COF2: : A EQUATION FOR TIME STEP (DT)
C WHICH CAN BE CONTROLLED BY PARAMETERS:C O E F 1 ,C O E F 2 ,PT AND N T I M E .
c PT: EXPONENT FOR THE EQN TO DETERMINE TIME STEP.
c NTIME: A FIXED INTERGER NUNBER TO BE 10000 FOR THE EQN.
c AX: LOW BOUND TEMP.(C) FOR HALF INTERVAL METHOD FOR SAT.TEMP.
c BX: HIGH BOUND TEMP.(C) FOR HALF INTERVAL METHOD FOR SAT.TEMP.
c ESP: ERROR TEST FOR CONVERGANCE OF INTERFACE VELOCITY.
c ESPP: ERROR TEST FOR RUNGE KUTTA TIME DETERMINATION.
c EXCITE: TEMPERATURE DISTURBANCE TO MAKE A CRITICAL SIZE BUBBLE GROW
c TMAX: MAXIMUM TIME TO COMPLETE THE PROGRAM
c P: EXPONENT FOR IRREGULAR GRID SPACE
c IL: GRID NUMBER
c DELTA: ARBITRARY PENETRATING DEPTH
c (THERMAL BOUNDARY LAYER THICKNESS)
c PRINTl: TRUE/FALSE FOR PRINTING ON SCREEN DURING RUNNING PROGRAM
c PRINT2: TRUE/FALSE FOR FINDING ERROR IN OUTPUT FILE(*.OUl)
c N S K I P : A NUMBER TO SKIP ROWS OF PRINTING DATA IN OUTPUT FILE(*.OU2)
c Q: HEAT FLUX FOR INITIAL THERMAL BOUNDARY LAYER
c TSTAR: A WAITING TIME FOR INITIAL THERMAL BOUNDARY LAYER
c PA: SYSTEM PRESSURE(KPA)
c NSP1-NSP4:TIME INTERVALS FOR PRINTING TEMPERATURE DISTRIBUTIONS
c IN OUTPUT F I L E (W A T 2 .O U 1 )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
191
C TBULK: BULK TEMPERATURE(OR EQUIVALENTLY TINF)
C

C
PROGRAM MAIN
IMPLICIT REAL*8(A-H,O-Z)
REAL*8 K
LOGICAL PRINTl,PRINT2
PARAMETER (PI=3.1415926D0,Ml=500)
DIMENSION R (Ml) ,TO(Ml) ,AP(M1) ,AE(M1) ,AW(M1) ,B(M1) ,SOLN(Ml)
DIMENSION DY (2) ,Y (2) ,F (14) ,TEMPO (Ml) ,TPRT (Ml,5), RPRT (Ml) ,TIME (5)
COMMON /AA/AP,AE,AW,B,SOLN
COMMON /BB/TO,R
C
O P E N {5,F I L E = 1w a t .i n ')
O P E N (6,F I L E = 'w a t .o u l )
O P E N (7,F I L E = 'w a t .o u 2 ')
C
R E A D (5,*)T M A X ,P ,I L ,DELTA,PRINTl,PRINT2,N S K I P ,A X ,BX
R E A D (5,*)Q,TSTAR,PA,EXCITE,NSP1,NSP2 ,N S P 3 ,NSP4
READ (5, *)T B U L K ,E S P ,ESPP,NTIME,P T ,C O F 1 ,COF2
C
N=2
T=0.DO
Y(2)= 0 .DO
C
C THE SATURATION TEMPERATURE IS OBTAINED BY USING
C THE HALF INTERVAL METHOD.
C
FA=PPSAT(AX)-PA*1000.DO
FB=PPSAT(BX)-PA*1 0 0 0 .DO
C
IX=0
3 IX=IX+1
X=(AX+BX)/2.D0
FX=PPSAT(X)-PA*1000.DO

C
I F (P R INT2)W R I T E (*,*)IX,X,FX
C
IF(DABS(FX).LE.1 .D-6) GO TO 88
IF(FX*FA) 85,88,86
85 BX=X
FB=FX
GO TO 3
86 AX=X
FA=FX
GO TO 3
88 IF(PRINT2)WRITE(*,*)'A IS TRUE VALUE'
TSAT=X
C
TINF=TBULK
C
C DETERMINE THE WALL TEMPERATURE AND INITIAL TEMPERATURE PROFILE
C BY USING 1-D TRANSIENT CONDUCTION SOLUTION.
C
K=COND(TINF)
ALPHA=DIFFUL(TINF)
TWAL=TINF + 2.D0*Q*DSQRT(ALPHA*TSTAR/PI)/K
TF=(TWAL + TINF)/2.DO

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
192
TWAL=TINF + 2.D0*Q*DSQRT(DIFFUL(TF)*TSTAR/PI)/COND(TF)
TF=(TWAL + TINF) 1 2 .DO
K=COND(TF)
ALPHA=DIFFUL(TF)
n o n

DETERMINE THE CRITICAL RADIUS (RC)

SIGMA=SIGMAL(TWAL)
RC = 2 .DO *SIG M A / (PPSAT(TWAL)-PA*10 0 0.D O )

FIRST TIME STEP


o o

CALL INITIAL(P ,DEL T A ,I L ,Q ,A L P H A ,T S T A R ,K ,T I N F ,R C )


n
o n

TWAL=TO(1)
UPDATE THE CRITICAL RADIUS
SIGMA=SIGMAL(TWAL)
RC=2.DO*SIGMA/(PPSAT(TWAL)-PA*1000.D0)
Y (1)=RC
o

W R I T E (7 ,9 7 0) Q ,TST A R ,T I N F ,P A ,T W A L ,R C ,TSAT
1 ,T M A X ,P ,I L ,DELTA,N S K I P ,E S P ,EXCITE,P T ,C O F 1 ,COF2
W R I T E (6,115 0)Q ,T S T A R ,T I N F ,P A ,T W A L ,R C ,TSAT
1 ,TMAX,P,IL,DELTA,NSKIP,ESP,EXCITE
n

DO 9 1=1,IL
T P R T (1,1)= T O (I )
R P R T {I )= R (I )- R (1)
9 CONTINUE

WRITE INITIAL CONDITIONS


o n o o o o o

W R I T E (7,1100)T,Y(1),Y(2),DY(2),TO(l)

GIVE A SLIGHT HIGHER TEMP THAN TWAL TO BREAK METASTABLE NUCLEATION

TS=TWAL+EXCITE

NCOUNT=0
KOUNT=0
JCOUNT=l
RDOT=0.DO
T I M E (1)=0.DO
ooo

ITERATION BEGINS FOR A NEW INTERFACE TEMPERATURE WITH A N E W TIME STEP

1 NCOUNT=NCOUNT+1
KOUNT=KOUNT +1
C
DO 14 1=1,IL
14 T E M P O (I )= T O (I )
C
DT=COFl*(DBLE(KOUNT)/DBLE(NTIME))**PT+COF2
TLMT=T
C
RBO = Y (1)
RDOTO=Y(2)
EPSL=1.DO-ROHV(TS)/ROHL(TS)
C
ICOUNT=0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
193
2 IC0UNT=IC0UNT+1
C
RD0T1=RD0T
n n o n

RUNGE KUTTA IS EMPLOYED TO COMPUTE A NEW LOCUS


OF THE LIQUID-VAPOR INTERFACE OF THE BUBBLE

10 L=3
M=0
50 CALL RUNGE (T,DT,N,Y,DY,F,L,M,J, IG)
IF(M.EQ.1)G0 TO 10
75 GO TO (100,200,999),L
100 D Y (1)= Y (2)
ROL=ROHL(TS)
TERM1=(PPSAT(TS)-PA*1000.DO)/ (ROL*Y(l)*EPSL)
T E R M 2 = 2 .*SIGMAL(TS)/ (ROL*Y(l)**2*EPSL)
TERM 3 = 1 .5D0*Y(2)*Y(2)/Y (1)
TERM4=4.DO*VISCOL(TS)*Y(2)/ (ROL*EPSL*Y(1)*Y(1))
D Y (2)=TERM1-TERM2-TERM3-TERM4
GO TO 50
n

200 I F (PRINT2)THEN
WRITE(6,1020)T,Y(1),Y (2)
ENDIF
non

TEST FOR TERMINATION

ERR=DABS(T-(TLMT+DT))
250 IF(T.GE.TLMT+DT.OR.ERR.LE.ESPP) GO TO 999
260 GO TO 50

999 R B = Y (1)
R D O T = Y (2)

HEAT FLUX IS CALCULATED BY USING ENERGY BALANCE AT THE LIQUID-VAPOR


n o n n n

INTERFACE

QB=-ROHV(TS) *XLATENT (TS)*RDOT/COND (TS)

IF(PRINTl)THEN
W R I T E (*,*)1#, ITER.# TS, QB, DRDOT, D T ',KOUNT,ICOUNT,TS
1 ,Q B ,RDOT1-RDOT,DT
ENDIF
noon

ENERGY EQUATION IS EMPLOYED TO COMPUTE THE TEMPERATURE DISTRIBUTION


WI T H THE HEAT FLUX COMPUTED AS A BOUNDARY CONDITION.

CALL FMTDIG(P ,I L ,DEL T A ,R B ,R D O T ,R B O ,D T ,Q B ,F 1 C ,F 2 C )


C
I F (PRINT2)THEN
DO 23 1=1,IL
W R I T E (6,980)R(I),AP(I),AE(I),AW(I),B(I)
23 CONTINUE
ENDIF
C
N5=IL-2
CALL TDIG(N5)
C
DO 30 1=1,IL-2
T O (1+1)=SOLN(I)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
194
30 CONTINUE
T O (1)= T O (2)* F1C+F2C
C
C DEFINE N E W VAPOR TEMPERATURE OR INTERFACE TEMPERATURE
C
T S = T O (1)
C
C CHECK CONVERSENCE FOR INTERFACE VELOCITY
C
I F (I C O U N T .G T .10 0)THEN
W R I T E (*,965)
965 FORMAT ( I I ' ** ITERATION FOR CONVERGENCE IS OVER 100, YOU M A Y REDUCE
1 THE TIME STEP***'//)
GO TO 9999
ENDIF
C
IF (DABS(RDOT1-RDOT) .GT.ESP)THEN
Y (1)=RBO
Y (2)=RDOTO
T=TLMT
DO 17 1=1,IL
17 T O (I )=TEMPO(I )
GO TO 2
ENDIF
C
C PRINT OUT THE RESULTS
C
I F(NCOUNT.E Q .NSKIP)THEN
W R I T E (7,1100)T,RB,RDOT,DY(2),TS
NCOUNT=0
ENDIF
C
C TEST FOR TERMINATION
C
IF(T.GT.TMAX) GO TO 9999
C
I F(K O U N T .E Q .N S P 1 .O R .KOUN T .E Q .NSP2)THEN
JCOUNT=JCOUNT+l
T I M E (JCO U N T )=T
DO 40 1=1,IL
T P R T (I ,JCOUNT)= T O (I )
40 CONTINUE
ELSEIF (K O U N T .E Q .NSP3 .O R .K O U N T .E Q .N S P 4 )THEN
JCOUNT=JCOUNT+l
TIME(JCOUNT)=T
DO 60 1=1,IL
T P R T (I ,JCOUNT)= T O (I )
60 CONTINUE
ENDIF
C
G O TO 1
C
9999 W R I T E (6,1160) (TIME(J) ,J = 1 ,JCOUNT)
DO 70 1=1,IL
W R I T E (6,1200)R P R T (I ) , (TPRT(I,J),J=1,JCOUNT)
70 CONTINUE
C
970 F O R M A T (/'#',2 X , 'SPHERICAL BUBBLE G R O W T H '
1 /'#',2X,,Q=',F7.1,1X,,Twait=',
1 F 6 .4,I X , 'Tinf= 1,F 8 .3, IX, 'k P a = 1,F 7 .3,I X , 'T w a l l = ',F 8 .3

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
195
1 /'#',2X,'R c = ',
1 DIO.4,I X , 1Tsa t = ' ,F 8 .3,2X,"Tinax=',DIO.4,IX,'P=' ,
1 F 3 .1,I X , 'I L = ,14,I X , 'D E L T A = ',F 6 .4
1 / # ' ,2X,'N S K I P = ',13,2 x , 'E S P = ',d8.2,2X,'E X C I T E = ',D8.2
1 /'#',2 X , 'P T = ',F 3 .1,2 X , 'C 0 F 1 = ',D 8 .2,2 X , 'C O F 2 = ',D 8 .2
1 /'#',5X, "I" ,12X, 'RB',10X, 'RDOT',10X, 'ACCEL',10X, 'Tv')
980 FORMAT(IX,5 (DIO.5))
1000 FORMAT(/2 X , 'T I M E = ',D12.6/8X,'R',12X,'TEMP')
1010 FORMAT(2X,D12.6,2X,D 1 2 .6,2X,D12.6)
1020 FORMAT(2X,D12.6,2(2X,D12.6))
1100 FORMAT(2X,Dll.4,2X,D 1 5 .8,2(2X,D12.5),2X,F12.7)
1150 F ORM A T (/'#',2 X , 'TEMPERATURE DISTRIBUTION'
1 /'#',2X,'Q=',F7.1,1X,'Twait=',
1 F 6 .4,I X , 'Tinf=',F 8 .3,I X , 'k P a = ',F 7 .3,I X , 'T w a l l = ',F 8 .3
1 /'#',2X,'Rc=',
1 D10.4,IX,'Tsat=',F 8 .3/'#',2X,'Tmax=',DIO.4,I X , 'P = ',
1 F 3 .1,I X , 'I L = ',14,I X , 'D E L T A = ',F6.4
1 /'#',2X,'NSKIP=',13,2 x , 'E S P = ',d8.2,2X,'E X C I T E = ,D8.2)
C
1160 FORMAT(/'#',4X,'x',10x,'T l = ',D I O .4,IX,'T2=',D10.4,
1 IX,'T3=',D I O .4,l x , 'T 4 = ',DIO.4,IX,'T5=',D10.4/)
1200 FORMAT(2X,6 (2X,D12.6))
C
STOP
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
196

Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

C GENERAL SUBROUTINES (FOR ANY PROPERTIES)

C
SUBROUTINE INI T I A L (P ,DELTA,IL,Q ,ALPHA,TSTAR,K ,T I N F ,R C )
o n o o

THIS PROGRAM DETERMINES THE INITIAL TEMPERATURE


DISTRIBUTION

IMPLICIT REAL*8(A-H,O-Z)
REAL*8 K
DIMENSION TO(500),R(500)
COMMON /BB/TO,R
P I = 3 .141592645D0
C
DO 10 1=1,IL
X=DELTA*(DBLE{I 1) /DBLE(IL-l))**P
R (I )=X+RC
A l = 2 .*Q*DSQRT(ALPHA*TSTAR/PI)/K
A2=-X * X / (4.*ALPHA*TSTAR)
A3 =Q*X/K
A 4 = X / 2 ./DSQRT(ALPHA*TSTAR)
T O (I )=A1*DEXP(A 2 )-A3 *ER F C (A 4 )+TINF
10 CONTINUE
RETURN
END

SUBROUTINE RUNGE(T,DT,N,Y,DY,F,L,M,J,IG)
IMPLICIT REAL*8 (A-H,O-Z)
DIMENSION DY(2),Y(2),F(14)
GO TO (100,110,300),L
100 GO TO (101,110),IG
101 J=1
L=2
DO 106 K=1,N
K1=K+3*N
K2=K1+N
K3=N+K
F(K1)=Y(K)
F(K3)=F(K1)
106 F(K2)=DY(K)
GO TO 406
110 DO 140 K=1,N
K1=K
K2=K+5*N
K3=K2+N
K4=K+N
GO TO (111,112,113,114), J
111 F(Kl)=DY(K)*DT
Y(K)=F(K4)+.5D0*F(K1)
GO TO 140
112 F(K2)=DY(K)*DT
GO TO 124
113 F(K3)=DY(K)*DT
GO TO 134
114 Y(K)=F(K4)+ (F(Kl)+2.DO*(F(K2)+F(K3))+DY(K)*DT)/6.D0
GO TO 140
124 Y(K)=.5D0*F(K2)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
197
Y(K)=Y(K)+F(K4)
GO TO 140
134 Y(K)=F(K4)+F(K3)
140 CONTINUE
GO TO (170,180,170,130),J
170 T = T + .5D0*DT
180 J=J+1
I F (J - 4 )404,404,299
299 M=1
GO TO 406
300 IG=1
GO TO 405
404 IG=2
405 L=1
406 RETURN
END

SUBROUTINE FMTDIG(P ,I L ,D E L T A ,R B ,R D O T ,R B O ,D T ,Q B ,F 1 C ,F 2 C )
o o n n n n o o

THIS SUBROUTINE FORMS THE TRIDIAGONAL MATRIX


FOR THE CRANK NICOLSON METHOD.

A*T(I) = B * T (1+1) + C * T (1-1) + D

IT IS NOTED THAT A=AP, B=AE, C=AW, D=B.

IMPLICIT REAL*8(A-H,O-Z)
REAL*8 KP,KE,KW,KSE,KSW
DIMENSION TO(500),AP(500),AE(500),AW(500),B(500),R(500),
1 RO(500),SOLN(500),XN(500)
COMMON /AA/AP,AE,AW,B,SOLN
COMMON /BB/TO,R
C
DO 10 1=1,IL
XN(I)= (DBLE(I-l)/DBLE(IL-1))**P
R (I )= X N (I )*DELTA+RB
R O (I )= X N (I )*DELTA+RBO
10 CONTINUE
C
RVL=ROHV(TO(1))/ROHL(TO(1))
C
DO 20 1=2,IL-1
RE=(R(I)+R(I+1))/2.D0
RW=(R(I-l)+R(I))/2 .DO
R E O = ( R O (1+1)+RO(I))/2.D0
R W O = ( R O (1-1)+ R O (I ))/2.D0
V E = 1 .DO/3.D0*RE**3
VW=1.D0/3.D0*RW**3
V E O = l .D O /3.DO *REO* *3
WO=1.DO/3.DO*RWO**3
KP=COND(TO(I ) )
KE=COND(TO(1+1))
KW=COND(TO(1-1))
KSE=2.D0*KP*KE/(KP+KE)
KSW=2.D0*KP*KW/(KP+KW)
R O L P = R O H L (T O (I ))
ROLE=ROHL(TO(1+1))
R O L W = R O H L (T O (I-1) )

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
198
R0LSE=2 .DO*ROLP*ROLE/ (ROLP+ROLE)
ROLSW=2 .DO*ROLP*ROLW/ (ROLP+ROLW)
C P = H C A P L (T O (I ))
C E = H C A P L (T O (I+1))
C W = H C A P L (T O (I-1))
CSE=2.D0*CP*CE/(CP+CE)
C S W = 2 .D0*CP*CW/(CP+CW)
RRD=RB * * 2 *RDOT * (l.dO-RVL)
FE=ROLSE*(RRD-(VE-VEO) /DT)
FW=ROLSW* (RRD- (VW-VWO) /DT)
IP (X .E Q .2 )ALPHAL=KSW/ (ROLSW*CSW)
C
D X N E = X N (1+1)- X N (I )
D X N W = X N (I )- X N (1-1)
DVPO=VEO-VWO
A P O = R O H L ( T O (I ))*CP*DVPO/DT
B (1-1)=APO*TO(I)
DE=RE**2*KSE/DELTA/DXNE
DW=RW* * 2 *KSW/DELTA/DXNW
PE=FE/DE*CSE
PW=FW/DW*CSW
C F P E = D M A X 1 (0.D O , (1.DO- O .1D0*DABS(PE)) * * 5 )
C F P W = D M A X 1 (0.D O , (1.DO-O.1D0*DABS(PW)) * * 5 )
AE (1-1)=DE*CFPE+DMAX1(-CSE*FE, 0.D O )
AW(I-l)=DW*CFPW+DMAX1 (CSW*FW, 0 .DO)
A P (1-1)= A E (1-1)+ A W (1-1) +APO
20 CONTINUE
C
D R W = R (2)- R (1)
QXK=QB*DRW
F 0=ALPHAL*D T /DRW**2
F 1 C = 2 .D 0 * F 0 / (1.D O + 2 .D0*F0)
F 2 C = ( T O (1)+2.D0*F0*QXK)/ (1.D0+2.D0*F0)
A P (1)= A P (1)- A W (1)*F1C
B (1)= B (1)+ A W (1)*F2C
AW(1)= 0 -DO
B (I L - 2 )= B (IL-2)+ A E (IL-2)* T O (IL)
A E ( I L - 2 )=0.DO
C
RETURN
END

SUBROUTINE TDIG(N)
n n n n

THIS SUBROUTINE INVERTS A TRIDIAGONAL MATRIX BY THOMAS


ALGORITHM.

IMPLICIT REAL*8(A-H.O-Z)
DIMENSION AP (500),AE(500),A W (500),B(500),P(500),
1 Q (5 0 0),SOL N (500)
COMMON /AA/AP,A E ,AW, B ,SOLN

P (1)=AE (1)/A P (1)


Q ( 1 ) = B ( 1 ) / A P (1)
DO 10 1 = 2 , N
p(i)=AE(i)/ (AP(i)-AW(i)*p(i-l))

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
199
q(i)=(AW(i)*q(i-l)+B(i)) / (AP(i)-AW(i)*p(i-l))
10 continue
soln(n)=q(n)
do 20 i = n-1,1,-1
soln(i)=p(i)*soln(i+l)+q(i)
20 continue
return
end

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
200

C THIS FUNCTIONS WERE WRITTEN FOR WATER PROPERTIES RANGE 273-600K


C * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

c --------------------------------------------------------------------------
real*8 function erfc(x)
c --------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
erfc=l.-erf(x)
return
end
c --------------------------------------------------------------------------
real*8 function erf(x)
c --------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
dimension a (0:10)
c --------------------------------------------------------------------------
a(0)=0.D0
a (1)=0.5204998778130500D0
a (2)=0.8427007929497271D0
a (3)=0.9661051464753030D0
a (4)=0.9953222650189663D0
a (5)=0.9995930479825686D0
a (6)=0.9999779095030078D0
a (7)=0.9999992569016313D0
a (8)=0.9999999845827402D0
a (9)=0.9999999998033906D0
a (10)=0.999999999998460D0

erf = l .
if(x.gt.5.49) return
p i = 3 .141592645d0
n=int(x/0.5)
xa=float(n)*0.5
erf=a(n)+simpsn(l,xa,x)*2./Dsqrt(pi)
if (erf.gt.l.) erf=l.
return
end

c ---------------------------------------------------------------------------
real*8 function simpsn(nsimp,a,b)
IMPLICIT REAL*8 (A-H,O-Z)
C ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
c integration by Simpson rule
c step size is less than dx<0.001,
c number of step is greater than 100
c
c nsimp equation number for f(x)
c a lower bound
c b upper bound
c
c definition of function
simpf(nsimp,x)=Dexp(-x*x)
c
c simpson integration
simpsn=0.
if(a.eq.b)return
ni=100
1 dx=(b-a)/(2.* float(ni))
if(dx.gt.0.001)then

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
201
ni=ni+10
goto 1
endif
c
d x 2 = 2 .*dx
s=a+dx
f2=simpf(nsimp,s-dx)
do 10 i=l,ni
f0=f2
fl=simpf(nsimp,s)
f2=simpf(nsimp,s+dx)
c
simpsn=simpsn+f0+4.*fl+f2
s=s+dx2
10 continue
simpsn=simpsn*dx/3.
c
return
end
C

REAL*8 FUNCTION PPSAT(Tl)


C
IMPLICIT REAL*8(A-H,O-Z)
C INPUT TEMP=(C), PPSAT(Pa), RANGE(273-600K) FOR WATER
T=Tl + 2 7 3 .15D0
P PSAT=1.354980D-12*T**5-3.465876D-9*T**4+3.632606D-6*T**3
1 -1.984735D-3*T**2+5.883791D-l*T-6.762496D1
P PSAT=DEXP(PPSAT)*100.DO
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHV(T)


C VAPOR ROH, TEMP.(C),RANGE(273-600K) FOR W A T E R .....
IMPLICIT REAL*8(A-H, O-Z)
C
Tl=T + 2 7 3 .15d0
R O H V = l .759194d-12*Tl**5-4.293708D-9*Tl**4+4.313782D-6*T1**3
1 -2.258077D-3*Tl**2+6.383280D-l*Tl-7.099724D1
ROHV=DEXP(ROHV)/1000.D0
RETURN
END
q* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

R EA L * 8 FUNCTION ROHL(T)
C --- LIQUID ROHL, T E M P . (C),R A N G E (273-600K)FOR WATER ......
IMPLICIT REAL*8 (A-H,O-Z)
Tl=T + 2 7 3 .15D0
ROHL=-l.336940D-13*Tl**5+2.337246D-10*T1**4-1.582645D-7*T1**3
1 +4.854679D-5*Tl**2-6.496841D-3*Tl+7.188601D0
ROHL=DEXP(ROHL)
RETURN
END
* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

R EA L * 8 FUNCTION COND(T2)
IMPLICIT REAL*8(A-H, O-Z)
C T(oC) FOR LIQUID(W/m.K) , TEMP.RANGE(273-600K) FOR WATER
T=T2 + 2 7 3 .15D0
COND=9.878536D-ll*T**5-2.183427D-7*T**4+1.933439D-4*T**3
1 -9.137936D-2*T**2+2.3 65709D1*T-1.950454D3
COND=COND*l.D-3

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
202
RETURN
END
C

REAL*8 FUNCTION XLA T E N T (T 3 )


IMPLICIT REAL*8 (A-H,O-Z)
C T(oC) FOR LATENT(J/KG), T E M P .R A N G E (273-600K) FOR WATER
T = T 3 + 2 7 3 .15D0
XLATE N T = - 3 .320494D-10*T**5+6.487444D-7*T**4-5.184260D-4*T**3
1 +2.068768D-l*T**2-4.313656Dl*T+6.308721D3
XLATENT=XLATENT*1000.DO
RETURN
END
C
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION DIFFUL(T4)


IMPLICIT REAL*8(A-H,O-Z)
C TEMP.RANGE(273-500K), DIFFUSIVITY(m2/s) FOR WATER
T=T4
DIFFUL=COND(T)/ROHL(T)/HCAPL(T)
RETURN
END
C

REAL*8 FUNCTION VISCOL(T8)


IMPLICIT REAL*8(A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE(o C ) TO VISCOSITY(N. s/m/'2)
C RANGE TEMP(270-600K) FOR WATER
T = T 8 + 2 7 3 .15D0
V I S C O = - l .885948D-12*T**5+4.473683D-9*T**4-4.273123D-6*T**3
1 +2.070945D-3 *T**2-5.179370D-l*T+5.947036D1
VISCOL=DEXP(VISCO)*l.D-6
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION S I G M A L (T9)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO
C SURFACE TENSION (N/m) FOR WATER
T = T 9 + 2 7 3 .15D0
S I G M = 3 .804613D-7*T**3-6.447664D-4*T**2+1.285080D-l*T+8.093310D1
SIGMAL=SIGM*1.d-3
RETURN
END

R E A L *8 FUNCTION HCAPL(TIO)
IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE(o C ) TO HEAT CAPACITY, J/KG.K
C FOR WATER
T = T 1 0 + 2 7 3 .15D0
C P F = 6 .258269D-12*T**5-1.232528D-8*T**4+9.620796D-6*T**3
1 -3.704042D-3*T**2+7.019596D-l*T-4.816255D1
HCAPL=CPF*1000.DO
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
203

C FUNCTIONS FOR R-113


Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

C -----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
real*8 function erfc(x)
c ---------------------------------------------------------------------------
IMPLICIT REAL*8(A-H,O-Z)
erfc=l.-erf(x)
return
end
c ---------------------------------------------------------------------------
real*8 function erf(x)
c ----------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
dimension a (0:10)
c ----------------------------------------------------------------------------
a(0)=0.D0
a (1)=0.5204998778130500D0
a(2)=0.8427007929497271D0
a (3)=0.966105146475303ODO
a(4)=0.9953222650189663D0
a (5)=0.9995930479825686D0
a (6)=0.9999779095030078D0
a (7)=0.9999992569016313D0
a (8)=0.9999999845827402D0
a (9)=0.9999999998033906D0
a (10)=0.99999999999846ODO

erf=1.
if(x.gt.5.49) return
p i = 3 .141592645d0
n=int(x/0.5)
xa=float(n)*0.5
erf=a(n)+simpsn(l,xa,x)*2./Dsqrt(pi)
if (erf.gt.l.) erf=l.
return
end

c ----------------------------------------------------------------------------
real*8 function simpsn(nsimp,a,b)
IMPLICIT REAL*8(A-H,O-Z)
c ----------------------------------------------------------------------------
c integration by Simpson rule
c step size is less than dx<0.001,
c number of step is greater than 100
c
c nsimp equation number for f(x)
c a lower bound
c b upper bound
c
c definition of function
simpf(nsimp,x ) = D e x p (-x*x)
c
c simpson integration
simpsn=0.
if(a.eq.b)return
ni=100
1 dx=(b-a)/(2.*float(ni))
if(dx.gt.0.001)then

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
204
ni=ni+10
goto 1
endif
c
dx2 = 2 .*dx
s=a+dx
f2=simpf (nsimp, s-dx)
do 10 i=l,ni
f0=f2
fl=simpf(nsimp,s)
f2=simpf(nsimp,s+dx)
c
simpsn=simpsn+f0+4.*fl+f2
s=s+dx2
10 continue
simpsn=simpsn*dx/3.
c
return
end
C

REAL*8 FUNCTION PPSAT(Tl)


C
IMPLICIT REAL*8(A-H,O-Z)
C INPUT TEMP=(C), PPSAT(Pa), RANGE(240-480K)
T=Tl+273.15D0
PPSAT=2.957142318d3-5.317878720dl*T+3.716948190d-l*T**2
1 -1.235475106d-3*T**3+1.860059877d-6*T**4-8.710006423d-10*T**5
PPSAT=PPSAT*10 0 0.DO
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHV(Tl)


C --- VAPOR ROH, TEMP.(C),R A N G E (240-400 K) .....
IMPLICIT REAL*8(A-H,O-Z)
C
T=Tl+273.15d0
R OHV=-2.996373102D2+4.974715994D0*T-3.322820191D-2*T**2
1 +1.134072022D-4*T**3-2.031981824D-7*T**4+1.585972295D-10*T**5
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHL(T)


C --- LIQUID ROHL, T E M P . (C),R A N G E (-30 - 210 C) .....
IMPLICIT REAL*8(A-H,O-Z)
R OHL=16.17677365058042D0-0.021409787008367DO*T
1 +0.000003666728067DO*T**2-0.000001210948343DO*T**3
1 +0.000000011848044D0*T**4-0.000000000037748D0*T**5
ROHL=ROHL *10 0.DO
RETURN
END

REAL*8 FUNCTION COND(T2)


IMPLICIT REAL*8(A-H,O-Z)
C T(oC) FOR LIQUID(W/m.K), T E M P .R A N G E (240-480)
T=T2+273.15D0
COND=572.672566562039D0-7.220703521017116D0*T
1 +0.044478466185909D0*T**2-0.000139306769359D0*T**3
1 +0.000000215863959D0*T**4-0.000000000132549D0*T**5
COND=COND/1000.DO

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
205
RETURN
END

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *
o

REAL*8 FUNCTION XLATENT(T3)


o

IMPLICIT REAL*8 {A-H,O-Z)


o

T (oC) FOR LATENT (J/KG) , T E M P .RANGE (240-480K)


T=T3+ 2 7 3 .15D0
XLATENT=2 737.488491466112D0-39.30036427470726D0*T
1 +0.239810862940233D0*T**2-0.000727033297258D0*T**3
1 +0.000001089453393D0*T**4-0.000000000647894D0*T**5
XLATENT=XLATENT*1000 .DO
RETURN
END

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *
n

REAL*8 FUNCTION DIFFUV(T4)


o

IMPLICIT REAL*8(A-H,O-Z)
n

T E M P . R A N G E (300-43OK), DIFFUSIVITY(m2/s)
T = T 4 + 2 7 3 .15D0
DIFFUV=303923.6357589624D0 - 3904.362680547735D0*T
1 +20.11858607644752D0*T**2-0.051927435282147DO*T**3
1 +0.000067087359484D0*T**4-0.000000034688727D0*T**5
DIFFUV=DIFFUV*1.D-8
RETURN
END
oo

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION DIFFUL(T5)


IMPLICIT REAL*8 (A-H,O-Z)
C TEMP.RANGE(300-430K), DIFFUSIVITY(m2/s*E-8)
T = T 5 + 2 7 3 .15D0
D I F F U L = 4 .798119328554736d3 -4.693879076560403dl*T
1 +2.04257142145589d-l*T**2-4.36619015926d-4*T**3
1 +4.47451584d-7*T**4-l.79222d-10*T**5
DIFFUL=DIFFUL*1.D-10
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION T E N T H A (E N T H )
IMPLICIT REAL*8(A-H,O-Z)
C TO TRANSFER DATA FROM LIQUID ENTHALPY TO TEMPERATURE (oC)
C RANGE OF ENTHALPY(10 - 270 kJ/ K ) , T E M P . (240 - 470 k)
ETP=ENTH*1.D-3
T T = 2 3 4 .4244709882D0+1.1228950385D0 *ETP-0.0002653983D0*ETP**2
1 -0.0000014715DO*ETP**3
TENTHA=TT-27 3.15D0
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

**

REAI,*8 FUNCTION ENTHAL (T 7 )


IMPLICIT REAL*8 (A-H,O-Z)
C TO TRANSFER DATA FROM T EMP E R A T U R E (o C ) TO LIQUID ENTHALPY
TEMP=T7+273.15D0
E T L = - 2 7 4 .4657037001D0+1.5944746489D0*TEMP-
1 0.0025719666D0*TEMP**2+0.0000032310D0*TEMP**3
ENTHAL=ETL*1.D+3
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
206

RE A L *8 FUNCTION VISCOL(T8)
IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO VISCOSITY(N.s/mA2 )
C RANGE TEMP(300K - 480K)
T= T 8 + 2 7 3 .15D0
V I S = 9 1 4 2 .7080404291D0-50.7337760352*T+0.0689662825D0*T*T
1 +0.0000616571DO*T**3-0.0000001400*T**4
V I SCOL=VIS*l.D-6
RETURN
END

REAL*8 FUNCTION SIGMAL(T9)


IMPLICIT REAL*8(A-H,O-Z)
C TRANSFER DATA FROM PRESS(kPa) & T E M P E R A T U R E (O C ) TO
C SURFACE TENSION (N/m)
T= T 9 + 2 7 3 .15D0
P 9 = P P S A T (T9)*1.D-3
TC=487.5D0
P C = 3 .4 1 1 D 0 / .101325D0
T B = 3 2 0 .7d0
TBR=TB/TC
TR=T/TC
Q = 0 .1 3 0 7 D 0 * (1.D0+TBR*DLOG(PC)/ (1.D0-TBR))-0.281D0
SI G M = P C * * (2.DO/3 .DO)*TC**(1.D O / 3 .D O ) * Q * (1.D0-TR)**(11.D O / 9 .D O )
SIGMAL=SIGM*1.D-3
RETURN
END
* * * * * ie * * * it * * * *

REAL*8 FUNCTION HCAPL(TIO)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO HEAT CAPACITY, J / K G .K
T= T 1 0+273.15D0
HCAPL=-2687.2937711853D0+32.1760788219D0*T-0.0967652227D0*T**2
1 +0.0001001237DO*T**3
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
207

ooooo
a * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

THIS FUNCTIONS WERE WRITTEN FOR A MAIN PROGRAM S B D 2 .FOR (SODIUM)


*******************************************************************

c ----------------------------------------------
real*8 function erfc(x)
c ----------------------------------------------
IMPLICIT REAL*8(A-H,O-Z)
erfc=l.-erf(x)
return
end
c ----------------------------------------------
real*8 function erf(x)
c ----------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
dimension a (0:10)
c ----------------------------------------------
a(0)=0.D0
a (1)=0.520499877813050ODO
a (2)=0.8427007929497271D0
a (3)=0.966105146475303ODO
a (4)=0.9953222650189663D0
a (5)=0.9995930479825686D0
a (6)=0.9999779095030078D0
a (7)=0.9999992569016313D0
a (8)=0.9999999845827402D0
a (9)=0.9999999998033906D0
a (10)=0.99999999999846ODO

erf=l.
if(x.gt.5.49) return
pi = 3 .141592645d0
n=int(x/0.5)
xa=dble(n)*0.5
erf=a(n)+simpsn(l,xa,x)*2./Dsqrt(pi)
if (erf.gt.l.) erf=l.
return
end

c----------------------------------------------
real*8 function simpsn(nsimp,a,b)
IMPLICIT REAL*8(A-H,O-Z)
c ----------------------------------------------
c integration by Simpson rule
c step size is less than dx<0.001,
c number of step is greater than 100
c
c nsimp equation number for f(x)
c a lower bound
c b upper bound
c
c definition of function
simpf(nsimp,x ) = D e x p (-x*x)
c
c simpson integration
simpsn=0.
if(a.eq.b)return
ni=100
1 dx=(b-a)/(2.*dble(ni))

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
208
if(dx.gt.O.OOl)then
ni=ni+10
goto 1
endif
c
dx2=2.*dx
s=a+dx
f2=simpf(nsimp, s-dx)
do 10 i=l,ni
f 0=f 2
fl=simpf(nsimp,s)
f2=simpf(nsimp,s+dx)
c
simpsn=simpsn+f0+4.*fl+f2
s=s+dx2
10 continue
simpsn=simpsn*dx/3.
c
return
end
C

REAL*8 FUNCTION PPSAT(Tl)


C
IMPLICIT REAL*8(A-H,O-Z)
C INPUT TEMP=(C), PPSAT(Pa), RANGE(1000-1800K) FOR SODIUM
T=Tl+273.15D0
PPSAT=-9103.5297143197D0+25.7651166667D0*T
1 -0.024643661D0*T**2+O.0000080011D0*T**3
PPSAT=PPSAT*1000.DO
RETURN
END

REAL*8 FUNCTION ROHV(T)


C VAPOR ROH, TEMP. (C),R A N G E (1000-1800K) FOR S O D I U M .....
IMPLICIT REAL*8(A-H,O-Z)
C
T l=T+273.15d0
ROHV=-17202.336704949D0+48.5939579878D0*Tl-0.046665605D0*T1**2
1 +0.0000153294DO*T1**3
ROHV=ROHV/1000.DO
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHL(T)


C .... LIQUID ROHL, T E M P . (C),R A N G E (1000-1800K)FOR SODIUM ......
IMPLICIT REAL*8 (A-H,O-Z)
Tl=T+273.15D0
ROHL=l.0086d0-2.134D-4*T1-1.75D-8*T1**2
ROHL=ROHL*1000.DO
RETURN
END
C * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION COND(T2)


IMPLICIT REAL* 8 (A-H,O - Z )
C T(oC) FOR LIQUID(W/m.K), T E M P .R A N G E (1000-1800) FOR SODIUM
T=T2+273.15D0
COND = 0 .2482D0-1.16D-4*T
COND=COND*418.68D0
RETURN

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
209
END

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *
n

REAL*8 FUNCTION XLATENT(T3)


o

IMPLICIT REAL*8(A-H,O-Z)
n

T(oC) FOR LATENT(J/KG), TEMP.RANGE(1000-1800K) FOR SODIUM


T = T 3 + 2 7 3 .15D0
XLATENT=4511.6285714DO-0.1 4 9 D 0 * T - 0 .0003429D0*T**2
XLATENT=XLATENT*1000.DO
RETURN
END
o n n

REAL*8 FUNCTION DIFFUL(T4)


IMPLICIT REAL*8(A-H,O-Z)
TEMP.RANGE(1000-1800K), DIFFUSIVITY(m2/s) FOR SODIUM
T=T4
DIFFUL=COND(T)/ROHL(T)/H C A P L (T )
RETURN
END
no

A T * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REA L * 8 FUNCTION VISCOL(T8)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO VISCOSITY (N. s/m/'2)
C RANGE T E M P (800-160OK) FOR SODIUM
T = T 8 + 2 7 3 .15D0
V I S C O = 5 0 1 .706490170722DO-0.377334049007D0*T+0.00005979232DO*T**2
VISCOL=VISCO*l.D-6
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION SIG M A L (T9)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO
C SURFACE TENSION (N/ltl) FOR SODIUM
T = T 9 + 2 7 3 .15D0
S I G M = 2 2 9 .3D0-0.1D0*T
SIGMAL=SIGM*1.d-3
RETURN
END

REAL*8 FUNCTION HCAPL(T10)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO HEAT CAPACITY, J/KG.K
C FOR SODIUM
T=T1 0 + 2 7 3 .15D0
C P F = 2 .00431746032D0-1.82562830688D-3*T+1.32708333333D-6*T**2
1 -2.47685185185D-10*T**3
HCAPL=CPF*1000.DO
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
210
* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

C THIS FUNCTIONS WERE WRITTEN FOR BUTANE PROPERTIES

-------------------------------------------------------------------------------------------------------------------
real*8 function erfc(x)
-------------------------------------------------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
erf c = l .-erf(x)
return
end
-------------------------------------------------------------------------------------------------------------------
real*8 function erf(x)
c ---------------------------------------------------------------------------
IMPLICIT REAL*8(A-H,O-Z)
dimension a(0:10)
c ---------------------------------------------------------------------------
a(0)=O.DO
a (1)=0.520499877813050ODO
a (2)=0.8427007929497271D0
a (3)=0.966105146475303ODO
a (4)=0.9953222650189663DO
a (5)=0.9995930479825686D0
a (6)=0.9999779095030078D0
a(7)=0.9999992569016313D0
a (8)=0.9999999845827402D0
a (9)=0.9999999998033906D0
a (10)=0.99999999999846ODO

erf=1.
if(x.gt.5.49) return
p i = 3 .141592645d0
n=int(x/0.5)
xa=float(n)*0.5
erf=a(n)+simpsn(l,xa,x)*2./Dsqrt(pi)
if (erf.gt.l.) erf=l.
return
end

c ---------------------------------------------------------------------------
real*8 function simpsn(nsimp,a,b)
IMPLICIT REAL*8(A-H,O-Z)
C ----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
c integration by Simpson rule
c step size is less than dx<0.001,
c number of step is greater than 100
c
c nsimp equation number for f (x)
c a lower bound
c b upper bound
c
c definition of function
simpf(nsimp,x)=Dexp(-x*x)
c
c simpson integration
simpsn=0.
if(a.eq.b)return
ni=100
1 dx=(b-a)/(2.*float(ni))
if(dx.gt.O.OOl)then
ni=ni+10

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
211
goto 1
endif
c
6 x 2 = 2 .*dx
s=a+dx
f2=simpf(nsimp, s-dx)
do 10 i=l,ni
f0=2
fl=simpf(nsimp,s)
f2=simpf(nsimp,s+dx)
c
simpsn=simpsn+f0+4.*fl+f2
s=s+dx2
10 continue
simpsn=simpsn*dx/3.
c
return
end
C
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION PPSAT(Tl)


C
IMPLICIT REAL*8(A-H,O-Z)
C INPUT TEMP=(C), PPSAT(Pa), RANGE(230-400K) FOR BUTANE
T=Tl+273.15D0
PPSAT=6.95668D-9*T**4-3.822772D-6*T**3 + 6 .014244D-4*T**2
1 -7.0909lD-4*T-4.465552D0
PPSAT=PPSAT*101325.DO
RETURN
END
q* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHV(T5)


C VAPOR ROH, TEMP. (C),RANGE(230-400K) FOR B U T A N E ......
IMPLICIT REAL*8(A-H,O-Z)
C
Tl=T5+273.15d0
ROHV=6.387127D-7*Tl**3-4.073028D-4*T1**2+1.079184D-1*T1
1 -1.042319D1
ROHV=ROHV* *2
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHL ( T 6 )


C --- LIQUID ROHL, TEMP.(C),RANGE(230-400K)FOR BUTANE ......
IMPLICIT REAL*8(A-H,O-Z)
Tl=T6+273.15D0
ROHL=1.407261D-12*T1**4-1.573344D-9*T1**3+6.651187D-7*T1**2
1 -1.228186D-4*Tl+9.821806D-3
ROHL=l.DO/ROHL
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION COND(T2)


IMPLICIT REAL*8(A-H,O-Z)
C T(oC) FOR LIQUID(W/m.K), TEMP.RANGE(230-400K) FOR BUTANE
T =T2+273.15D0
COND=-l.433322D-9*T**3+1.328689D-6*T**2-7.952385D-4*T
1 +2.740857D-1
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
REAL*8 FUNCTION XLATENT(T3)
IMPLICIT REAL*8 (A-H,O-Z)
C T(oC) FOR L A T E N T (J/KG), T E M P .RANGE(230-400K) FOR BUTANE
T=T3+273.15D0
XLATENT=-1.688303D-7*T**4+1.781551D-4*T**3-7.151602D-2*T**2
1 +1.205770Dl*T-2.648612D2
XLATENT=XLATENT*1000.DO
RETURN
END
o o

REAL*8 FUNCTION DIFFUL(T4)


IMPLICIT REAL*8 (A-H,O-Z)
n

T E M P . R A N G E (230-400K) , DIFFUSIVITY(m2/s) FOR BUTANE


T=T4
DIFFUL=COND(T)/ROHL(T)/HCAPL(T)
RETURN
END
on

REAL*8 FUNCTION VISCOL(T8)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE (o C ) TO VISCOSITY (N.s/m/s2 )
C RANGE TEMP(230-400K) FOR BUTANE
T=T8+273.15D0
V I S C O = 4 .318534D-9*T**4-5.753387D-6*T**3+2.898539D-3*T**2
1 -6.664525D-l*T+6.103799D1
VISCOL=VISCO *1.D-4
RETURN
END

REAL*8 FUNCTION S I G M A L (T9)


IMPLICIT REAL*8(A-H,O-Z)
C TRANSFER DATA FROM TEMPERATURE(o C ) TO
C SURFACE TENSION (N/m) FOR BUTANE
T=T9
A = 1 4 .87D0
B = 0 .1206D0
THI=20.DO
TC=152.DO
TR=(T+273.15D0) / (TC+273.15D0)
THIR=(THI+273.15D0)/ (TC+273.1 5 D 0 )
E=B*(TC-THI)/ (A-B*THI)
D=(A-B*THI)/ (1.D0-THIR)**E
SIGMAL=D*(1.D0-TR)**E
SIGMAL=SIGMAL/1000.DO
RETURN
END

REAL*8 FUNCTION HCAPL(T7)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERA T U RE (o C ) TO HEAT CAPACITY,J/KG.K
C FOR BUTANE
T=T7+273.15D0
C P F = 1 .114838D-7*T**3-6.744840D-5*T**2+1.656721D-2*T+5.259061D-1
HCAPL=CPF*1000.DO
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
213
Q 'k'k'kit'k'k'k'kle'k'k'k'k'kii'k'kir'kicirlcie'kitieic'kieic'k'k'kirit'k'kit'k'kitielric'k'kie'k'kit'k'kit-kicic'k'k'k'kic'kir'k'k'k'k

C FUNCTIONS FOR METHANOL


C * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

c ----------------------------------------------------------------------------
real*8 function erfc(x)
c ----------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
e r f c = l .-erf(x)
return
end
c ----------------------------------------------------------------------------
real*8 function erf(x)
c ----------------------------------------------------------------------------
IMPLICIT REAL*8 (A-H,O-Z)
dimension a (0:10)
c ----------------------------------------------------------------------------
a (0)=0.DO
a (1)=0.520499877813050ODO
a (2)=0.8427007929497271D0
a (3)=0.966105146475303ODO
a (4)=0.9953222650189663D0
a (5)=0.9995930479825686D0
a (6)=0.9999779095030078D0
a (7)=0.9999992569016313D0
a ( 8 ) = 0 .9999999845827402D0
a(9)=0.9999999998033906D0
a (10)=0.999999999998460D0

erf=l.
if(x . g t .5.49) return
p i = 3 .141592645d0
n=int(x/0.5)
xa=float(n)*0.5
erf=a(n)+simpsn(l,xa,x)*2./Dsqrt(pi)
if (erf.gt.l.) erf=l.
return
end

c ----------------------------------------------------------------------------
real*8 function simpsn(nsimp,a,b)
IMPLICIT REAL*8(A-H, O-Z)
c ----------------------------------------------------------------------------
c integration by Simpson rule
c step size is less than dx<0.001,
c number of step is greater than 100
c
c nsimp equation number for f(x)
c a lower bound
c b upper bound
c
c definition of function
simpf(nsimp,x)=Dexp(-x*x)
c
c simpson integration
simpsn=0.
if(a.eq.b)return
ni=100
1 dx=(b-a)/ (2.*float(ni))
if(dx.gt.0.001)then
ni=ni+10

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
214
goto 1
endif
c
6 x 2 = 2 .*dx
s=a+dx
f2=simpf(nsimp,s-dx)
do 10 i=l,ni
f0=f2
fl=simpf(nsimp,s)
f2=simpf(nsimp,s+dx)
c
simpsn=simpsn+f0+4.*fl+f2
s=s+dx2
10 continue
simpsn=simpsn*dx/3.
c
return
end
C
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION PPSAT(Tl)


C
IMPLICIT REAL*8 (A-H,O-Z)
C INPUT TEMP=(C), PPSAT(Pa), RANGE(250-450K)
T=Tl+273.15D0
PPSAT=2.602528D-12*T**5-5.689175D-9*T**4+5.137114D-6*T**3
1 -2.442997D-3*T**2+6.389363D-l*T-6.552792D1
PPSAT=DEXP(PPSAT)*10 0.DO
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHV(Tl)


C ___ VAPOR ROH, TEMP.(C),R A N G E (250-450 K) ......
IMPLICIT REAL*8(A-H,O-Z)
C
T =Tl+273.15d0
ROHV=6.560962D-12*T**5-1.211598D-8*T**4+9.264145D-6*T**3
1 -3.741939D-3*T**2+8.351774D-l*T-7.621540D1
ROHV=DEXP(ROHV)/1000.D0
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION ROHL(Tl)


C ___ LIQUID ROHL, T E M P .(C),RANGE(250-450K) ......
IMPLICIT REAL*8(A-H,O-Z)
T =Tl+273.15D0
R OHL=-l.049737D-9*T* *5+1.820122D-6*T**4-1.258585D-3 *T* *3
1 +4.320083D-l*T**2-7.436394Dl*T+6.003278D3
RETURN
END
Q * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION COND(T2)


IMPLICIT REAL*8 (A-H,O-Z)
C T(oC) FOR LIQUID(W/m.K), TE M P .R A N G E (250-450)
T=T2+273.15D0
CON D = 0 .3225 D 0 - 4 .785D-4*T+1.168D-7*T**2
RETURN
END
C

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
REAL*8 FUNCTION XLATENT(T3)
IMPLICIT REAL*8 (A-H,O-Z)
C T(oC) FOR LATENT(J/KG), T E M P .RANGE(250-450K)
T=T3+273.15D0
XLATENT=-3.499864D-9*T**5+6.003288D-6*T**4-4.094137D-3*T**3
1 +1.38057D0*T**2-2.310179D2*T+1.664205D4
XLATENT=XLATENT*1000.DO
RETURN
END
n n

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION DIFFUL(T4)


IMPLICIT REAL*8(A-H,O-Z)
TEMP.RANGE(250-450K), DIFFUSIVITY(m2/s)
n

DIFFUL=COND(T4)/ROHL(T4)/HCAPL(T4)
RETURN
END
no

REAL*8 FUNCTION VISCOL(T8)


IMPLICIT REAL*8(A-H,O-Z)
C TRANSFER DATA FROM TEMPERAT U R E (o C ) TO VISCOSITY(N.s/m ~ 2 )
C RANGE TEMP(250K - 450K)
T=T8+273.15D0
A = - 3 .935D1
B=4.826D3
C = 1 .091D-1
D = - l .127D-4
VIS=A+B/T+C*T+D*T**2
VISCOL=DEXP(V I S )*1.D-3
RETURN
END
* * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * *

REAL*8 FUNCTION S I G M A L (T9)


IMPLICIT REAL*8(A-H,O-Z)
C TRANSFER DATA FROM PRESS(kPa) & TEMPERATURE(o C ) TO
C SURFACE TENSION (N/m)
T=T9+273.15D0
P9=PPSAT(T9)*l.D-3
TC=512.6D0
PC=80.9/1.01325D0
TB=337.7d0
TBR=TB/TC
TR=T/TC
Q = 0 .1196D0*(1.DO+TBR*DLOG(PC)/ (1.D0-TBR))-0.279D0
SIGM=PC**(2.D0/3.D0)*TC**(1.D0/3.D0)*Q*(1.D0-TR)* * ( 1 1 .DO/9.DO)
SIGMAL=SIGM*1.D-3
RETURN
END

REAL*8 FUNCTION HCAPL(T10)


IMPLICIT REAL*8 (A-H,O-Z)
C TRANSFER DATA FROM TEMPERA T U R E (o C ) TO HEAT CAPACITY,J/KG.K
T=T10+273.15D0
HCAPL=-3.13177D-ll*T**5+5.726075D-8*T**4-4.109695D-5*T**3
1 +1.452645D-2*T**2-2.528083DO*T+1.754952D2
HCAPL=HCAPL*10 0 0.DO
RETURN
END

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
BIBLIOGRAPHY

216

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
217

BIBLIOGRAPHY

Abdelmessih, A. H., Hooper, F. C. and Nangia, S., 1972, "Flow Effects on Bubble

Growth and Collapse in Surface Boiling", Int. J. Heat Transfer, Vol. 15, pp. 115-125.

Asai, A., 1991, "Bubble Dynamics in Boiling under High Flux Pulse Heating", Journal of

Heat Transfer, Vol. 113, pp.973-979.

Avedisian, C. T., 1985, "The Homogeneous Nucleation Limits of Liquids", J. Phys.

Chem. Ref. Data, Vol. 14, No.3, pp.695-729.

Babuska, I., Prager, M., and Vitasek, E., 1966, "Numerical Processes in Differential

Equations", Chapter 6.1 SNTL, Prague

Birkhoff, G., 1956, "Stability of Spherical Bubble, Q. Appl. Math, Vol.13, pp.451.

Bankoff, S. G. and Mikesell, R.D., 1958, "Bubble Growth Rates in Highly Subcooled

Nucleate Boiling", A.I.Ch.E., Preprint No.2, A.S.M.E.-A.I.Ch.E. Heat Transfer

Conference, Chicago (1958).

Birkhoff, G . , Margulies, R.S. and Horning, W. A., 1958, "Spherical Bubble Growth",

Phys. Fluid, V ol.l, 201

B oard, S. J . and Duffey, R. B., 1971, "Spherical Vapor Bubble Growth in Superheated

Liquids", Chem. Engng. Sci., Vol.26, pp.263-274

B ohrer, T. H., 1973, "Bubble Growth in Highly Superheated Liquids", M.S. Thesis in

Chemical Engineering, Purdue University

Bosnjakovic, F., 1930, Tech.Mech. Thermodynam. 1, 358,

Dalle Donne, M. and Ferranti, M.P., 1975, "The growth of Vapor Bubble in

Superheated Sodium", Int. J. Heat Mass Transfer, Vol. 18, pp.477-493

D ergarabedian, P., 1953, "The Rate of Growth of Vapor Bubbles in Superheated

Water", J. Appl. Mech., Vol.20, pp.537-545

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
218

Dergarabedian, P., 1960, "Observations on Bubble Growth in Various Superheated

Liquids", J. Fluid Mech., Vol.9, pp.39-48

Dimic, M., 1977, "Collapse of One-component Vapor Bubble with Translatory Motion",

Int. J. Heat Transfer, Vol.20, pp.1325-1332.

Ellion, M. E., 1953, "A Study of the Mechanism of Boiling Heat Transfer", Ph.D. Thesis,

California Institute of Technology, Pasadena, California.

Ervin, J. S., 1990, "Incipient Boiling in Microgravity", Ph.D. Thesis, University of

Michigan, Ann Arbor.

Ervin, J. S., M erte, H., Keller, R.B., Kirk, K., 1992, "Transient Pool Boiling in

Microgravity", Int. J. Heat Mass Transfer, Vol.35, pp.659-674

Florschuetz, L.W. and Chao, B.T., 1965, "On the Mechanics of Vapor Bubble

Collapse", Journal of Heat Transfer, Vol.87, pp.209-220.

Florschuetz, L. W ., H enry, C. L. and Khan, A. R ashid , 1969, "Growth Rates of Free

Vapor Bubbles in Liquids at Uniform Superheats under Normal and Zero Gravity

Conditions", Int. J. Heat Mass Transfer, Vol. 12, pp. 1465-1489

Forster, H. K. and Zuber, N., 1954, "Growth of a Vapor Bubble in a Superheated

Liquid", J. Appl. Phys., Vol.25,474

Frost, D. L., 1988, "Dynamics of Explosive Boiling of a Droplet", Phys. Fluids, Vol.31,

pp.2554-2561.

Frost, D. and Sturtevant, B., 1986, "Effects of Ambient Pressure on the Instability of a

Liquid Boiling Explosively at the Superheat Limit", J. Heat Transfer, Vol. 108, pp.418-

424.

Grolmes, M. A. and Fauske, H. K., 1974, "Axial Propagation of Free Surface Boiling

into Superheated Liquids in vertical Tubes", In Proceeding of the Fifth International Heat

Transfer Conference, Tokyo, Vol.4, pp.30-34.

G unther, F. C. and K reith F., 1949, "Photographic Study of Bubble Formation in Heat

Transfer to Subcooled Liquids", Heat Fluid Mech. Inst. Berkeley, pp. 113.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
219

G unther, F. C., 1950, Jet Prop. Lab. Prog. Rept. 4-75, Calif. Inst. Technol. Pasadena.

G unther, F. C., 1951, "Photographic Study of Surface Boiling Heat Transfer to Water

with Forced Convection", Trans. Amer. Soc. Mech. Engrs,vol.73, pp. 115.

H an, C. Y. and Griffith, P., 1965, "The Mechanism of Heat Transfer in Nucleate Pool

Boiling-Part I ", Int. J. Heat Transfer, Vol.8, pp.887-904.

Hewitt, H. C. and Parker, J . D., 1968, "Bubble Growth and Collapse in Liquid

Nitrogen", Journal of Heat Transfer, Vol., pp.22-26.

Higuera, F. J., 1987, "The Hydrodynamic stability of an evaporating liquid", Phys. Fluids,

Vol.30, pp.679-686.

Hooper, F. C. and Abdelmessih, A. H., 1966, "The Flashing of Liquids at Higher

Superheats", Proc. 3rd Int. Conf. on Heat Transfer, Chicago,, IV.44, Aug7-12

Jakob, Z, 1932, Tech. Mech. Thermodynam., 1, 358

Istratov, A. G. and Librovich, V. B., 1969, "On the Stability of Gasdynamic

Discontinuities Associated with Chemical Reactions. The Case of a Spherical Flame",

Astronautica Acta, Vol. 14, pp.453-467.

Kim, C harn-Jung and Kaviany, M asso u d , 1990, "A Numerical Method for Phase-

Change Problems", Int. J. Heat Mass Transfer, Vol.33, pp.2721-2734

Kosky, P. G ., 1968, "Bubble Growth Measurements in Uniformly Superheated Liquids",

Numerical Engineering Science, Vol.23, pp.695-706

Kwak, Ho-Young and Lee, S angbu m , 1991, "Homogeneous Bubble Nucleation

Predicted by a Molecular Interaction Model", J. Heat Transfer, Vol.l 13, pp.714-712

Landau, L., 1944, "On the Theory of Slow Combustion", Acta Physicochemica U.S.S.R.,

Vol. 19, pp.76-85.

Landau, L. and Lifshitz, E. M., 1959, "Fluid Mechanics", Pergamon, problem No.2,

pp.479.

Lauterborn, W. and Hentschel, W., 1985, "Cavitation Bubble Dynamics Studied by

High Speed Photography and Holography: part one", Ultrasonics, Vol.23, pp.260-268.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
220

Lien, Y. C., 1969, "Bubble Growth Rates at Reduced Pressure", D. Sci. Thesis, MIT

Lien, Y. C. and Griffith, P., 1969, "The Bubble Growth at Reduced Pressure", Int. J.

Heat Mass Transfer, Jan. 1969

L ord Rayleigh, 1917, Phil. Mag. 34, 94,

M cCann, H., Clarke, L. J. and A. P. M asters, 1989, "An experimental study of vapour

growth at the superheat limit temperature", Int. J. Heat Mass Transfer, Vol.32, pp. 1077-

1093.

M eister, G., 1979, "Vapor Bubble Growth and Recondensation in Subcooled Boiling

Flow", Nuclear Engineering and Design, vol.54, pp.97-114.

M ikic, B. B. and Rohsenow, W. M., 1969, "Bubble Growth Rates in Non-uniform

Temperature Field", Progress in Heat and Mass Transfer n , pp.283-293

M ikic, B. B ., Rohsenow, W. M. and G riffith, P . , 1970, "On Bubble Growth Rate",

Int. J. Heat Mass Transfer, Vol. 13, pp.657-666

M iller, C. A., 1973, "Stability of Moving Surface in Fluid Systems with Heat and Mass

Transport", AIChE Journal, Vol.19, No.5, pp.909-915.

Nukiyama, S., 1934, J.Soc. Mech. Engrs, Japan, Vol.37, pp.367.

Palm er, H. J., 1976, "The hydrodynamic stability of rapidly evaporating liquids at

reduced pressure", J. Fluid Mech., Vol.75, pp.487-511.

Palm er, H. J . and Maheshri, J. C., 1981, "Enhenced Interfacial Heat Transfer by

Differential vapor recoil instabilities", Int. J. Heat Transfer, Vol.4, pp. 117-123.

P atan k ar, S. V ., 1980, "Numerical Heat Transfer and Fluid Flow", Hemisphere

Washington,DC,

Plesset, M. S., 1954, "On the Stability of Fluid Flows with Spherical symmetry", J. Appl.

Phys., Vol.25, pp.96-98.

Plesset, M . S. and Prosperetti, A., 1977, "Bubble dynamics and Cavitation", Ann. Rev.

Fluid Mech., Vol.9, pp.145-185.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
221

Plesset, M.S. and Zwick, S. A ., 1952, "A Nonsteady Heat Diffusion Problem with

Spherical Symmetry", J. Appl. Phys., Vol.23,95

Plesset, M.S. and Zwick, S. A ., 1954, "The Growth of Vapor Bubble in Superheated

Liquid", J. Appl. Phys., Vol.25,493

Prosperetti, A. and Plesset, M ., 1978, "Vapor-bubble Growth in a Superheated

Liquid", J. Fluid Mech., Vol.85, pp.349-368

Prosperetti, A. and Plesset, M. S., 1984, "The Stability of an evaporating liquid surface",

Phys. Fluids, Vol.27, pp.1590-1602.

Reid, R. C., 1983, "Rapid Phase Transitions from Liquid to Vapor", Advanced in

Chemical Engineering, Vol. 12, pp. 105-208

Scriven, L. E . , 1959, "On the Dynamics of Phase Growth", Chemical Engineering

Science, vol.l, pp. 1-13

Shepherd, J . E. and Sturtevant, B., 1982, "Rapid Evaporation at the Superheat Limits",

J. Fluid Mech., Vol. 121, pp.379-402

Shima. M ., Tomita, Y. and Ohno, T., 1988, "Temperature effects on Single Bubble

Collapse and Induced Impulsive Pressure", Journal of Fluids Engineering, Vol.l 10,

pp. 194-199.

Sparrow , E. M . and Chuck, W ., 1984, "An Implicit/Explicit Numerical Solution

Scheme for Phase-change Problems", Numerical heat transfer, Vol.7, pp.1-15

Sparrow , E. M ., Ramadhyani, S. and Patanka, S .V ., 1978, "Effect of Subcooling on

Cylinderical Melting", J. Heat Transfer, Vol. 100, pp.395-402

S traub, J., Zell, M. and Vogel, B., 1990, "Pool Boiling in a Reduced Gravity Field",

Heat Transfer 1990, Proceedings of The Ninth International Heat Transfer Conference,

Jerusalem, Israel, Vol.l, pp.91-112.

S turtevant, B. and Shepherd, J. E., 1982, "Evaporative Instability at the Superheat

Limit", Appl. Sci. Res., Vol.38, pp.85-97.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
222

Taylor, G. I., 1949, "The Stability of Liquid surfaces when accelerated in a direction

perpendicular to their planes. I, Proc. R. Soc. Lond. A201, pp. 192.

Theofanous, T., Biasi, L. and Isbin, H. S., 1969, "A Theoretical Study on Bubble

Growth in Constant and Time-dependent Pressure Fields", Chem. Engng. Sci. Vol. 26,

pp.263-274

Theofanous, T. G. and Patel, P. D., 1976, "Universal Relations for Bubble Growth", Int.

J. Heat Mass Transfer", Vol. 19, pp.425-429

W ittke, D. D. and Chao, B. T., 1967, "Collapse of Vapor Bubbles with Translatory

Motion", Journal of Heat Transfer, Vol., pp. 17-24.

Zuber, N., 1961, "The Dynamics of Vapor Bubbles in Non-uniform Temperature fields",

Int. J. Heat Transfer, Vol.2, pp.83-98.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Potrebbero piacerti anche