Вы находитесь на странице: 1из 226

Numerical Simulation of

the Cavitation Process


in Diesel Fuel Injectors

SERGEY MARTYNOV

A thesis submitted in partial fulfilment of the


requirements of the University of Brighton
for the degree of Doctor of Philosophy

September 2005

The University of Brighton


Abstract

Hydrodynamic cavitation is one of the phenomena which affects the nature of a


fuel spray and determines the efficiency of the combustion process in Diesel engines. A
model of cavitation is needed to predict the flow in the nozzle, and to aid in the design
of injectors and control of the fuel injection process.
The review of experimental studies provides evidence of the scaling of
macroscopic cavitation structures observed in cavitation flows running at similar flow
conditions when the flow and liquid quality scale effects are minimised. To describe
non-similarity of real cavitation flows, a number of scale effects are identified.
Models of cavitation based on the transport equation for the vapour fraction have
been found to be very efficient in predicting the main features of cavitation flows.
However, many of these models contain parameters, which have to be adjusted to
account for the liquid quality effect on the flow.
It has been shown that consideration of the spectrum of cavitation bubbles is
crucial in describing the liquid quality scale effect in a model of cavitation flow.
In the present study a scalable model of cavitation is derived from the assumption
of hydrodynamic similarity of cavitation flows. The novel feature of the cavitation
model developed is that its parameters are related to the critical number density of
cavitation nuclei, which depends on the liquid tension in the cavitation region.
Parameters of the scalable cavitation model can be adjusted using only one set of
experimental data.
The model is implemented in the PHOENICS CFD code and validated across the
experimental data on cavitation flows in small-scale nozzles. Study is completed by
analysis of the scale effects on cavitation flows in models of injection nozzles.

ii
Contents
Abstract ...................................................................................................................................................... ii
Contents ..................................................................................................................................................... iii
List of figures.............................................................................................................................................. v
List of tables ............................................................................................................................................ xiv
Acknowledgments .................................................................................................................................... xv
Declaration .............................................................................................................................................. xvi
Nomenclature ......................................................................................................................................... xvii
CHAPTER 1. INTRODUCTION ............................................................................................................. 1
1.1. CAVITATION IN DIESEL ENGINE INJECTION TECHNOLOGY ................................................................. 1
1.2. HYDRODYNAMIC CAVITATION .......................................................................................................... 3
1.3. SCOPE OF THE PRESENT STUDY.......................................................................................................... 5
CHAPTER 2. LITERATURE SURVEY ................................................................................................. 6
2.1. INTRODUCTION ................................................................................................................................. 6
2.2. THEORETICAL AND EXPERIMENTAL BACKGROUND ........................................................................... 6
2.2.1. Definitions ................................................................................................................................ 6
2.2.2. Cavitation inception and cavitation number ............................................................................ 7
2.2.3. Similarity criteria and scale effects .......................................................................................... 8
2.2.4. Cavitation bubble nuclei ........................................................................................................ 12
2.2.5. Cavitation in flowing liquid.................................................................................................... 15
2.2.6. Conclusions ............................................................................................................................ 17
2.3. EXPERIMENTAL OBSERVATIONS OF CAVITATION FLOWS IN NOZZLES .............................................. 19
2.3.1. Introduction............................................................................................................................ 19
2.3.2. Visualisations of cavitation flows in nozzles .......................................................................... 21
2.3.3. Discharge coefficient of nozzle............................................................................................... 25
2.3.4. Scale effects ............................................................................................................................ 30
2.3.5. Conclusions ............................................................................................................................ 49
2.4. MODELS OF CAVITATION FLOWS ..................................................................................................... 52
2.4.1. Introduction............................................................................................................................ 52
2.4.2. Models of two-phase flows ..................................................................................................... 52
2.4.3. Homogeneous mixture flow model ......................................................................................... 53
2.4.4. Models of cavitation ............................................................................................................... 55
2.4.5. Conclusions ............................................................................................................................ 74
2.5. NUMERICAL COMPUTATION OF CAVITATION FLOWS........................................................................ 76
2.5.1. Summary of methods of calculation of liquid-vapour flows ................................................... 76
2.5.2. Basic incompressible pressure solvers ................................................................................... 78
2.5.3. Modifications to the pressure-correction algorithm for flows with variable density ............. 81
2.5.4. Conclusions ............................................................................................................................ 84
2.6. CONCLUSIONS FOR CHAPTER 2........................................................................................................ 85
CHAPTER 3. ANALYSIS AND MODEL DEVELOPMENT ............................................................. 88
3.1. INTRODUCTION ............................................................................................................................... 88
3.2. GROWTH AND COLLAPSE OF CAVITATION BUBBLES ........................................................................ 89
3.2.1. Bubble growth ........................................................................................................................ 89
3.2.2. Bubble collapse ...................................................................................................................... 90
3.2.3. Conclusions ............................................................................................................................ 94
3.3. DEVELOPMENT OF MODEL OF CAVITATION FLOW............................................................................ 95
3.3.1. Model for liquid-vapour flow ................................................................................................. 95
3.3.2. Turbulence modelling............................................................................................................. 95
3.3.3. Model for cavitation ............................................................................................................... 97
3.4. MODEL SCALABILITY .................................................................................................................... 106
3.5. POPULATIONS OF CAVITATION BUBBLES ....................................................................................... 107
3.5.1. Measurements of nuclei populations .................................................................................... 108

iii
3.5.2. Approximations for the spectrum of nuclei........................................................................... 110
3.5.3. Conclusions .......................................................................................................................... 112
3.6. VALIDATION OF THE MODEL FOR CONCENTRATION OF CAVITATION BUBBLES .............................. 112
3.7. DISCUSSION OF THE RANGE OF THE MODEL APPLICATION ............................................................. 114
3.8. CONCLUSIONS FOR CHAPTER 3...................................................................................................... 114
CHAPTER 4. NUMERICAL METHOD AND PROCEDURE ......................................................... 116
4.1. INTRODUCTION ............................................................................................................................. 116
4.2. FINITE-VOLUME METHOD .............................................................................................................. 117
4.3. CONVECTION DISCRETISATION SCHEMES ...................................................................................... 120
4.3.1. Hybrid scheme...................................................................................................................... 120
4.3.2. SMART scheme..................................................................................................................... 122
4.3.3. Super-bee scheme................................................................................................................. 122
4.3.4. Implementation of high-order schemes ................................................................................ 122
4.4. BOUNDARY CONDITIONS ............................................................................................................... 123
4.4.1. Inlet/Outlet ........................................................................................................................... 124
4.4.2. Walls..................................................................................................................................... 124
4.4.4. Symmetry planes................................................................................................................... 125
4.5. MASS FRACTION EQUATION .......................................................................................................... 125
4.6. SOLUTION ALGORITHM ................................................................................................................. 126
4.7. CONSIDERATION OF THE MIXTURE COMPRESSIBILITY ................................................................... 127
4.8. RELAXATION AND CONVERGENCE................................................................................................. 127
4.9. CONCLUSIONS FOR CHAPTER 4...................................................................................................... 128
CHAPTER 5. APPLICATION ............................................................................................................. 130
5.1. INTRODUCTION............................................................................................................................. 130
5.2. CALCULATION OF CAVITATION FLOWS IN NOZZLES ....................................................................... 130
5.2.1. Introduction.......................................................................................................................... 130
5.2.2. Strategy of calculation.......................................................................................................... 131
5.2.3. Parameters of cavitation model............................................................................................ 131
5.2.4. Parameters of iteration process ........................................................................................... 132
5.3. APPLICATION OF THE MODEL FOR THE CRITICAL NUMBER DENSITY OF THE BUBBLE NUCLEI......... 133
5.3.1. Reference measurements ...................................................................................................... 134
5.3.2. Mesh specification................................................................................................................ 135
5.3.3. Convergence of iterations .................................................................................................... 135
5.3.4. Adjustment of the number density of cavitation bubbles ...................................................... 136
5.3.5 The effect of initial mass fraction .......................................................................................... 141
5.3.6. Study of the effect of cavitation number ............................................................................... 144
5.3.7. Demonstration of scalability of the model............................................................................ 148
5.3.8. Conclusions .......................................................................................................................... 151
5.4. VALIDATION OF THE MODEL FOR THE SHEAR STRESS INDUCED CAVITATION ................................. 153
5.4.1. Description of test cases....................................................................................................... 154
5.4.2. Mesh specification................................................................................................................ 156
5.4.3. Pattern of cavitation-free flow.............................................................................................. 157
5.4.4. Model for the critical vapour pressure................................................................................. 163
5.4.5. Results .................................................................................................................................. 165
5.4.6. Conclusions .......................................................................................................................... 173
5.5. CONCLUSIONS FOR CHAPTER 5...................................................................................................... 174
CHAPTER 6. CONCLUSIONS AND RECOMMENDATIONS....................................................... 176
6.1. CONCLUSIONS ............................................................................................................................... 176
6.2. RECOMMENDATIONS FOR FURTHER STUDIES ................................................................................. 177
REFERENCES ...................................................................................................................................... 181
APPENDICES........................................................................................................................................ 191
APPENDIX A. Q1 INPUT FILE FOR PHOENICS ..................................................................................... 191
APPENDIX B. GROUND MODULE........................................................................................................ 196
APPENDIX C. STUDY OF GRID DEPENDENCE ......................................................................................... 201
APPENDIX D. TRANSIENT HEATING OF A SEMI-TRANSPARENT SPHERICAL BODY ................................. 205

iv
List of figures
Fig. 1- 1. Two types of Diesel injectors: valve-covered orifice (VCO) injector (a) and
sac-type injector (b). ................................................................................................. 2

Fig. 1- 2. Geometry-induced cavitation inside a nozzle (a) (Schmidt, et al., 1999), and
string-type cavitation in sac-type Diesel injector (b) (Arcoumanis, et al., 2001).
Dark regions indicate liquid, light regions cavitation flow. .................................. 2

Fig. 1- 3. Jet break-up regimes (Hiroyasu, 1998, 2000). Lines a and b shows hysteresis
behaviour in transition from cavitation flow to hydraulic flip regime (a) and back
from hydraulic flip to liquid flow regime (b). .......................................................... 4

Fig. 2- 1. The state of the liquid at saturation pressure pv (point_A) and under tension

pv pl at a given temperature (superheated liquid, point_B) (Skripov, 1974). ..... 7

Fig. 2- 2. Photographs of cavitation flow in large-scale (a) and real-size (b) Diesel
injection nozzles under similar flow conditions (Arcoumanis, et al., 2000). ......... 11

Fig. 2- 3. Critical pressures in liquid with a vapour-filled bubble, comparing to critical


pressures in pure liquid (spinodal curve) and saturation vapour pressures (binodal
curve). ..................................................................................................................... 13

Fig. 2- 4. Critical radius for the stable vapour-gas bubble in liquid under tension. ....... 14

Fig. 2- 5. Couette flow between two walls. .................................................................... 17

Fig. 2- 6. Sketch of the geometry-induced cavitation flow and pressure distribution


along the nozzle in the core region and near the wall............................................. 19

Fig. 2- 7. Development of the average length of cavitation region Lcav (Fig. 2-6) with
cavitation number CN............................................................................................. 21

Fig. 2- 8. Characteristic types of spray formation while flowing through the large-scale
circular nozzle (Soteriou et al., 1995). Injection into a gas. ................................... 22

Fig. 2- 9. Density gradient field and velocity field in a small-size cavitating nozzle
(nozzle height = 0.28mm), operating at injection pressure p1 = 80 bar and

downstream (back) pressures p 2 of 21 bar (a) and 11 bar (b). Images obtained
using the shadowgraph and PIV techniques. (Experimental data of Roosen et al.
(1996) referred by Yuan and Schnerr (2001)). ....................................................... 23

v
Fig. 2- 10. Image of cavitation in a nozzle of rectangular shape obtained using a
fluorescent laser light technique (Roosen, et al., 1997, 1998). Flow direction
from left to right, working liquid decalin, upstream pressure 30 bar and
downstream pressure 1 bar. Nozzle width ~0.3 mm............................................... 24

Fig. 2- 11. Structure of the internal flow in a real-size Diesel injection nozzle captured
using the light sheet technique (Badock, et al., 1999). ........................................... 24

Fig. 2- 12. A foam of cavitation bubbles formed in the sharp entry nozzle. Series of
photographs of the water flow in a circular nozzle of 22 mm in diameter (Sato and
Saito, 2001)............................................................................................................. 24

Fig. 2- 13. Discharge coefficient and cavitation structures observed in plain nozzle at
various cavitation numbers. Dashed line shows approximation for the discharge
coefficient under the choked flow conditions (equation (2.18)):
C d 0.67 1 + 1 / CN .............................................................................................. 28

Fig. 2- 14. Discharge coefficient under cavitation flow in circular nozzles.


Experimental data by Nurick (1976) and Soteriou, et al. (1995) in comparison with
the correlation (2.18): C d 0.62 1 + 1 / CN .......................................................... 29

Fig. 2- 15. Schematic of the flow in asymmetrical nozzle. ............................................ 32

Fig. 2- 16. The flow map for cavitation flow regimes in asymmetrical two-dimensional
nozzle (Schmidt, et al., 1999). ................................................................................ 33

Fig. 2- 17. Nozzles of different configurations of the entry shape, applied in study by
Yuan et al. (2001). (a) sharp entry nozzle; (b) rounded entry nozzle; (c)
counter-bore nozzle; (d) bevelled entry nozzle.................................................... 33

Fig. 2- 18. Effect of the bevel angel on the cavitation inception ( CN inc ) and transition to

super-cavitation flow ( CN super ). Results of experiments by Laoonual, et al. (2001).

................................................................................................................................ 34

Fig. 2- 19. Effect of the nozzle shape and nozzle length on cavitation inception (a) and
transition to super-cavitation flow and hydraulic flip (b). Bergwerk (1959),
Lichtarowicz, et al. (1974), Fox and Stark (1989) plain (sharp-entry) nozzles
(Fig. 2-17, a); Laoonual, et al. (2001) plain, counter-bore and bevelled nozzles
(Fig. 2-17, a,c,d). .................................................................................................... 35

vi
Fig. 2- 20. Effect of nozzle length on variation in hydraulic resistance of a nozzle
caused by transition to the hydraulic flip. Numbers above the points indicate nozzle
length to diameter ratio L/D.................................................................................... 36

Fig. 2- 21. Time delay before establishing the total hydraulic flip in a plain nozzle
(analysis of the experimental data from Laoonual, et al., 2001). ........................... 37

Fig. 2- 22. Nozzle cap with a spray hole of conical shape and rounded inlet corner. .... 38

Fig. 2- 23. Effect of the conical shape of nozzle on the extent of cavitation region and
pressure distribution (Winklhofer, et al., 2001). For the nozzles J, U and W the K-
factors are 0 %, 5% and 10%, respectively............................................................. 38

Fig. 2- 24. Discharge coefficients of three nozzles of different shape, shown in Fig. 2-
23. ........................................................................................................................... 39

Fig. 2- 25. Effect of nozzle shape on the quasi-steady-state cavitation flow in real-scale
nozzle of diameter ~0.2 mm (Blessing and Konig, 2003). Injection pressure
p1 = 800 bar, back pressure p 2 = 1 bar. ................................................................. 40

Fig. 2- 26. Sketch of the pressure distribution along Roosens and Winklhohers nozzles
based on the numerical solutions for three-dimensional liquid flows. ................... 42

Fig. 2- 27. Cross-sectional cut of a circular nozzle in the vena contracta region (Fig. 2-
6). ............................................................................................................................ 44

Fig. 2- 28. Stream lines for the steady-state re-circulation flow (a) and transient re-
entrant jet motion (b) in the separation flow formed at the nozzle entrance. ......... 45

Fig. 2- 29. Gas entrainment in closure of the vapour cavity at the nozzle exit. ............. 46

Fig. 2- 30. Diagram of the flow regimes in an orifice L / D 2.5 (Lichtarowicz, et al.,
1974) ....................................................................................................................... 47

Fig. 2- 31. Delayed transition to the super-cavitation flow. ........................................... 48

Fig. 2- 32. The effect of void fraction on the evaporation rate according to the models of
Yuan, et al. (2001) and Alajbegovic (1999). .......................................................... 63

Fig. 2- 33. Evaporation and condensation rates versus the void fraction according to
different cavitation models (velocity scale u = 100 m/s; linear scale l = 1 mm;
|| = 1). ................................................................................................................... 66

vii
Fig. 2- 34. Spatial scales in the flow with cavitation bubbles. ....................................... 72

Fig. 2- 35. Control cell around the velocity node w for the X -component of velocity, in
one-dimensional consideration. .............................................................................. 79

Fig. 3- 1. Models of bubble growth. (Mikic, et al., 1970, completed by Sauers (2001)

12 a l 3 l
solution for the initial stage of growth). Here R* = Ja 2 is the
2 p

l c p.l (Tl Ts ( pl ))
characteristic size of a bubble and Ja = is the Jacob number.
v hlg

................................................................................................................................ 90

Fig. 3- 2. Typical behaviour of collapsing bubble at different values of criteria


B eff (Florschuetz and Chao, 1965)......................................................................... 91

Fig. 3- 3. Consideration of thermal effects on the collapse of a bubble in water ( Tl , =

90oC, pl , = 0.981 bar, the degree of vapour subcooling is -10 K) (Sauer, 2001).

R R P indicates application of the Rayleigh-Plesset equation (2.53). .................... 92

Fig. 3- 4. Effect of inertia on the collapse of a bubble in water, (Ro= 102 m, pl = 105 Pa,
Tl = 20oC) (Sauer, 2001). R Ray and R R P indicate application of the Rayleigh

model (2.27) and the Rayleigh-Plesset equation (2.53).......................................... 92

Fig. 3- 5. States of liquid (l , ) and vapour (v, o) at the beginning of bubble collapse in
P-T diagram. ........................................................................................................... 93

Fig. 3- 6. Variation in the critical size of vapour bubbles with the tension in liquid (a)
and actual (b) and cumulative spectra (c) of cavitation bubbles........................... 100

Fig. 3- 7. Discretisation of spectrum of bubbles in poly-disperse model. .................... 101

Fig. 3- 8. Measurements of the number density distribution function for cavitation


nuclei in waters of different purity (Billet, 1985; Sato and Kakutani, 1994). ...... 108

Fig. 3- 9. Cavitation nuclei distribution functions in water of different quality (Gindroz,


et al, 1997). T1 large nuclei; T2 high content of medium-size nuclei; T3 low
content of medium-size nuclei; T4 strong degassed water. .............................. 109

Fig. 3- 10. Nuclei distribution function. ....................................................................... 110

viii
Fig. 3- 11. Concentration of cavitation bubbles in liquid under tension. Approximation
of measured spectra of bubbles at different oxygen contents (ppm) and system
pressures (bar) in water (Gindroz, et al., 1997) (points) by equation (3.22) with the
variable parameter n* (lines)................................................................................ 112

Fig. 4- 1. Control volume associated with the node P in three dimensions.


Neighbouring nodes are E, W, S, N, L and H, and cell faces e, w, s, n, l and h. ... 117

Fig. 4- 2. Velocity and its normal component at the cell face f . A f is area of the face

f between the cells P and F, and VP is volume of the cell around the node P . 118

Fig. 4- 3. Nodes involved in the approximation formula for the cell-face value f in

two-point and three-point convection schemes in one dimensional case (D


downwind, U upwind, and UU upwind-upwind). .......................................... 121

Fig. 5- 1. Shape and main dimensions of planar nozzle (Yuan, et al., 2001). .............. 134

Fig. 5- 2. The structure of body-fitted mesh in the Y-Z plane of the computational
domain (Fig. 5-1). ................................................................................................. 135

Fig. 5- 3. Pattern of cavitating flow in a small-scale nozzle (Fig. 5-2) (photograph by


Roosen, et al. (1996) reported Yaun, et al., 2001) at CN=2.81 (a) in comparison
with the results of numerical predictions of the vapour field: (b) -

n = 1.3 1014 (m 3 ) ; (c) - n = 4.4 1014 (m 3 ) ; (d) - n = 2 1015 (m 3 ) . ............. 137

Fig. 5- 4. Measured velocity field and shape of the vapour pocket in cavitating nozzle
(Roosen, et al., 1996) at CN = 2.81 (a) in comparison with the results of numerical

calculations using n = 1.3 1014 (m 3 ) (b), n = 4.4 1014 (m 3 ) (c),

n = 2 1015 (m 3 ) (d). The volume of calculated cavitation region is indicated by


iso-surface where the void fraction is = 20% . ................................................. 138

Fig. 5- 5. Pressure distributions along the nozzle center-line (CL) and near the wall

predicted for the inlet cavitation (Table 5-1) using n = 4.4 1014 (m 3 ) (continuous
lines), comparing to the pressure distribution in cavitation-free flow (stroke lines).
Thick dashed line marks the vapour pressure pv ................................................. 140

ix
Fig. 5- 6. Effect of the initial mass fractions f o on the vapour field predicted for inlet

cavitation (Table 5-1), n = 4.4 1014 (m3). (a) - f o = 107 , (b) - f o = 106 , (c) -

f o = 10 5 , (d) - f o = 104 . Plots show the right half of the nozzle, flow is from top
to bottom. .............................................................................................................. 142

Fig. 5- 7. Effect of the initial mass fractions f o on the vapour field predicted for

super-cavitation flow regime (Table 5-1), n = 4.4 1014 (m3). (a) - f o = 107 , (b) -

f o = 106 , (c) - f o = 10 5 . Plots show the right half of the nozzle, flow is from top
to bottom. .............................................................................................................. 143

Fig. 5- 8. Pattern of cavitation flow in a small-scale nozzle (Fig. 5-2) (Yuan, et al.,
2001) at CN=6.27 in comparison with the results of numerical predictions of

cavitation pockets using n = 1.3 1014 (m3) (b) and n = 4.4 1014 (m3) (c). ...... 145

Fig. 5- 9. Pressure distributions along the nozzle center-line (CL) and near the wall
predicted for the super-cavitation flow (Table 5-1), using

n = 4.4 1014 (m 3 ) (continuous lines), comparing to the pressure distribution in


cavitation-free flow (stroke lines). Thick dashed line marks the vapour pressure
pv . ........................................................................................................................ 146

Fig. 5- 10. Variation in the number density of cavitation bubbles with the liquid tension.
The results of adjustments of the number density parameter n* for real-scale

nozzle n* and application of the equation (3.22) (straight line, n* = 2 1010 (m 3 ) )


to estimate the number density of bubbles for a large-scale flow. Lines shows the
limits of variation for n established for the inlet cavitation in small-scale nozzle

(dashed line for n* = 9 1010 (m 3 ) and dotted line for n* = 0.7 1010 (m 3 ) ).... 149

Fig. 5- 11. Pressure fields in small (dashed lines) and large-scale nozzles (solid lines)
predicted for the inlet cavitation at CN = 2.81, comparing to the pressure
distribution in cavitation-free flow (green lines). Solid thick curves are for pressure
in the liquid core, fine curves describe pressure distribution near the wall. Grey
dashed line marks the vapour pressure. ................................................................ 150

Fig. 5- 12. Calculated pressure distributions in the small-scale (stroke lines) and large-
scale nozzles (continuous lines) predicted for the super-cavitation flow (Table 5-1)

x
using equation (3.22) with the number density parameter n* = 2 1010 (m 3 ) . Grey
dashed line marks the vapour pressure. ................................................................ 151

Fig. 5- 13. Geometry and main sizes of planar rectangular nozzle (Winklhofer, et al.,
2001). .................................................................................................................... 154

Fig. 5- 14. Pressure field before start of cavitation (a) and under critical cavitation
conditions (b) (Table 5-3) in planar nozzle (Fig. 5-13), 5 bar steps between isolines
(Winklhofer, et al., 2001)...................................................................................... 155

Fig. 5- 15. Mesh structure in Y-Z plane inside the nozzle (shown for half of a nozzle,
Fig. 5-13). ............................................................................................................. 156

Fig. 5- 16. Velocity fields in Roosens (a) and Winklhofers (b) nozzles predicted
neglecting the cavitation process. ......................................................................... 159

Fig. 5- 17. Fields of relative pressure p p 2 predicted for Roosens (a) and
Winklhofers (b) nozzles neglecting the cavitation process. ................................ 159

Fig. 5- 18. Distributions of the kinetic energy of turbulence k , m 2 /s 2 in Roosens (a)


and Winklhofers (b) nozzles predicted neglecting the cavitation process. ......... 160

Fig. 5- 19. Distributions of the component of the rate of strain S yz , sec 1 (5.5) in

Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process. ................................................................................................................. 160

Fig. 5- 20. Distributions of the shear stresses yz = 2( + t ) S yz , Pa in Roosens (a)

and Winklhofers (b) nozzles predicted neglecting the cavitation process. ......... 161

Fig. 5- 21. Distributions of coefficient of the turbulent viscosity t , m 2 /sec , in


Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process. ................................................................................................................. 161

Fig. 5- 22. Variations in absolute pressure and the turbulent component of shear stresses
along the nozzle (schematically)........................................................................... 162

Fig. 5- 23. Void fractions predicted neglecting the effect of liquid stress on critical
vapour pressure. Inlet cavitation in Winklhofers nozzle (Table 5-3). (a) - n =

xi
1.6 1016 (m3); (b) - n = 2 1018 (m3) . (Figures show only first section of the
right half of the nozzle)......................................................................................... 164

Fig. 5- 24. Void fractions predicted neglecting the effect of liquid stress on critical
vapour pressure for the critical cavitation flow in Winklhofers nozzle (Table 5-3).

(a) n = 1.6 1016 (m3), (b) n = 2 1018 (m3) , (c) n = 1.6 1019 (m3), (d) n =

1.6 10 22 (m3). ..................................................................................................... 164

Fig. 5- 25. Distributions of the void fraction in the Roosens nozzle under inlet (a, b)
and super-cavitation (c, d) flow regimes (Table 5-1), predicted assuming pcr = pv

(a, c) and using equation (5.7) with Ct = 10 (b, d). n = 4.4 1014 (m3). (Figures
show only right half of the nozzle, the flow runs from top to bottom)................. 165

Fig. 5- 26. Distributions of the void fraction at the nozzle throat for the inlet cavitation
(Table 5-3), predicted taking it into account the effect of shear stress on critical

pressure by the equation (5.7) using n = 2 1018 (m3). (a) Ct = 10 , (b)

Ct = 20 . (Figures show only first section of the right half of the nozzle). .......... 166

Fig. 5- 27. Distributions of the void fraction at the nozzle throat for the critical
cavitation (Table 5-3), predicted taking it into account the effect of shear stress on
critical pressure by the equation (5.7) with a constant Ct = 10 and various number

densities of cavitation bubbles n (m3): (a) 1.6 1013 , (b) 2 1015 , (c)

1.6 1016 , (d) 2 1018 , (e) 1.6 1019 . ............................................................... 167

Fig. 5- 28. The shape of the vapour pocket at the nozzle entry (Table 5-3) (iso-surfaces
for the void fraction = 20% (a) and = 50% (b)) predicted for the critical

cavitation using Ct = 10 and n = 1.6 1016 (m3). The entry part of the nozzle is
shown for half of the height and half of the width. Flow from left to right, back
wall is shown in grey. ........................................................................................... 167

Fig. 5- 29. Distributions of the void fraction at the nozzle throat for the critical
cavitation (Table 5-3), predicted taking it into account the effect of shear stress on
critical pressure by the equation (5.7) with a constant Ct = 20 and various number

densities of cavitation bubbles n (m3): (a) 1.6 1013 , (b) 2 1015 , (c)

1.6 1016 , (d) 2 1018 . ........................................................................................ 168

xii
Fig. 5- 30. Iso-surfaces for the tension p p v (a, c) and local variable p p cr (b, d),

predicted using Ct = 10 and n = 2 1018 (m3). (a) p p v < 0 ; (b) p p cr < 0 ;

(c) p p v < 4 10 5 (Pa); (d) p p cr < 4 10 5 (Pa). (Flow from right to left around
the nozzle inlet corner). ........................................................................................ 170

Fig. 5- 31. Iso-surfaces for the tension p p v < 0 (a) and local variable p p cr < 0

(b), predicted using Ct =20 and n = 2 1018 (m3). (Flow from right to left around
the nozzle inlet corner). ........................................................................................ 170

Fig. C- 1. Pressure distributions along the nozzle centreline (CL) and near the wall,
predicted using different meshes (Table C-1)....................................................... 202

Fig. C- 2. Effect of mesh size on the pressure drop p1 p 2 , relative minimum pressure
in the nozzle throat p min p 2 . ............................................................................ 203

xiii
List of tables
Table 2- 1. Physical properties of liquids (20oC, 1 bar). ................................................ 31

Table 2- 2. Developing and super-cavitation flows in rectangular models of Diesel


injectors. Comparison of the measurements by Roosen, et al. (1996) and
Winklhofer, et al. (2001)......................................................................................... 41

Table 2- 3. Features of models of cavitation flows......................................................... 55

Table 5- 1. Parameters of cavitation flows of a tap water in planar nozzle (Fig. 5-1)
(Pressure measurements by Roosen, et al., 1996)................................................. 134

Table 5- 2. Parameters of the cavitation flows of tap water in a large-scale planar nozzle
(Fig. 5-1). .............................................................................................................. 148

Table 5- 3. Integral parameters of cavitation flow in a straight nozzle (Fig. 5-13)


(according to the measurements by Winklhofer, et al., 2001). ............................. 155

Table C- 1. Grids used in the study of mesh dependence............................................. 201

Table C- 2. The pressure drops, the minimum pressures at the nozzle throat, the
discharge coefficients of nozzle predicted and the CPU time spent for
computations using different meshes (Table C-1). ............................................... 202

xiv
Acknowledgments
First of all I would like to thank my supervisor Dr. David Mason and Prof.
Morgan Heikal for their advisory, continuous attention and support of this project.
I would like to acknowledge Prof. Leonid Dombrovskii, Dr. David Kennaird,
Dr. Tao Bo and Alexander Shevelko for their criticism, generic and useful
discussions, and valuable contribution to this work.
I feel deep gratitude to Dr. Keiichi Sato and Yasuhiro Saito and Dr. Celia
Soteriou for useful experimental information, and PHOENICS support team for their
consultancy.
Also, my thanks go to research students and staff from the School of
Engineering at the University of Brighton for their help during the period of my
study.
I would like express my appreciation to my parents, my sister and my friends,
who have encouraged me throughout this project.
At last, but not least, I would like to express my gratitude to my tutors and
colleagues from the Institute for High Temperatures of Russian Academy of Sciences,
and Moscow Power Engineering Institute Dr. V. Kurganov, Prof. Yu. Zeigarnik, Dr.
F. Ivanov, Dr. V. Artemov, Prof. V. Yagov, Dr. V. Yankov and Dr. N.
Medvedskaya.
I would like to thank the Ricardo U.K. Ltd for their financial support during the
initial phase of this project.

xv
Declaration

I hereby certify that this thesis is my own work. Any data and results of other
authors are supported by citations and references.

Signed: Date:

xvi
Nomenclature
Latin symbols
a coefficients in discretised convection-diffusion equation (4.5);
A cross-section area of the flow, m2;
Af area of the cell face f (Fig. 4.2) , m2;
c speed of sound, m/sec;
c constant in equation (3.5);

C = l / l cav rate constant of cavitation (dimensionless);

Ccav = 1 / l cav cavitation parameter in equation (2.44), m1;

Cc contraction coefficient in equation (2.18) (dimensionless);

Cd coefficient of discharge (dimensionless);

Cnb convection contribution to anb in equation (4.4);

C numerical constant in equation (2.73) (dimensionless);

Ct constant in equation (5.7) (dimensionless);

Cv velocity coefficient in equation (2.18) (dimensionless);


cp heat capacity, J/ (kg K);
d coefficient in equation (2.67), m/ (Pa sec);
D diameter of nozzle, m;
Dnb diffusion contribution to anb in equation (4.4);
f mass fraction of the vapour phase (dimensionless);
fo initial mass fraction in equation (5.2)
F frequency of oscillations of cavitation region, Hz;
H height of nozzle, m;
h enthalpy, J/kg;
hlg latent heat of evaporation J/kg;
i, j = 1, 2, 3 indexes of coordinates;
k kinetic energy of turbulence, m2/sec2;
K factor of conicity of a nozzle, %;
L length of nozzle, m;
Lcav length of cavitation region, m;
l cav cavitation length scale, m;

xvii
l hydrodynamic length scale of the flow, m;
l hydrodynamic length scale of the flow, m;
n number density of cavitation bubbles, m3;
no number density of cavitation bubble nuclei, m3;

n* number density parameter in equation (3.22), m3;


N (R ) number density distribution function, m4;
p1 pressure at the inlet of nozzle, Pa;
p2 pressure at the outlet of nozzle, Pa;
p1,o total pressure at the nozzle inlet, Pa;

pl , pressure in liquid far from the bubble, Pa;

pv pressure of the vapour in cavitation region, Pa;


pmin minimum absolute pressure in cavitation region, Pa;
p pressure correction, Pa ;

p* guessed pressure value, Pa;

~ p p2
p= dimensionless pressure;
p1 p 2
R radius of cavitation bubble, m;
Rcr critical radius of cavitation bubble, m;
R* characteristic scale of growing/collapsing bubble, m;
S source term in the volume fraction equation, sec1 ;
~
S source term in dimensionless volume fraction equation;

S ij strain arte tensor, sec1 ;

Sv source term in the mass fraction equation (4.13), kg m3 sec1 ;


SP source term in equation (4.5);
t time, sec;
~
t = t U / l dimensionless time for the flow motion;

t + dimensionless time for bubble growth/ collapse ;


t false- time relaxation in equation (4.17), sec;

T temperature, K;
Tl , temperature in liquid far from the bubble, K;

xviii
u local velocity, m/sec;
u velocity correction, m/sec;

u* guessed value of velocity, m/sec;

U = 2( p1 p 2 ) / l velocity scale according to the Bernoulli equation, m/sec;

u mean volumetric velocity of the flow in nozzle throat, m/sec;

u + = u / ut dimensionless velocity;

u~ = u / U dimensionless velocity;
VP volume of cell at node P;
W width of nozzle, m;

y + = y ut / dimensionless coordinate;
x, X, Y, Z coordinates, m;
~
x = x / l dimensionless coordinate;

Greek symbols
volume fraction of the vapour phase;
linear relaxation factor in equation (4.16)

ij Kroneker delta tensor;


Z size of grid cells in the straight part of a nozzle, m;
rate of dissipation of the kinetic energy of turbulence, m2/sec3;
coefficient of compressibility, Pa1;
coefficient of dynamic viscosity, Pa sec;
~ = / l dimensionless viscosity of mixture;
coefficient of kinematic viscosity, m2 /sec;
surface tension coefficient, Pa m;
density, kg /m3;
density correction, kg /m3;

* guessed density, kg /m3;


~ = / l dimensionless density of mixture;
ij deviatoric viscous shear stress tensor, Pa;

ij stress tensor, Pa;

xix
generalised flow variable;

Criteria
B eff criterion for regime of the bubble collapse defined by equation (3.1);

u
Fr = Froude number;
g l

l c p ,l (T Ts )
Ja = Jacob number;
v hlg

p1 p2
CN = cavitation number;
p2 pv
M Mach number;

Cf
Pe f = cell Peclet number;
Df

Re Reynolds number;
Str Strouhal number;
We Weber number;
pv p
= local cavitation parameter;
1
2 l u 2

Indexes
B bubble;
c contraction;
cr critical conditions for the cavitation onset;
coll bubble collapse;
cond condensation;
ev evaporation;
f, nb = e,ee,w,ww,s,n,h,l index of cell face (Fig. 4.1, Fig. 4.2);
F, NB = E,EE,W,WW,S,N,H,L index of cell (Fig. 4.1, Fig. 4.2);
g gas phase;
H based on the nozzle height;
inc inception;

xx
inlet at the nozzle inlet;
l liquid phase;
large large-scale flow;
m mixture of phases;
min minimum;
new value of variable estimated at current iteration;
old value of variable calculated at previous iteration;
outlet at the nozzle outlet;
o cavitation nuclei;
P index of discretisation cell/ node (Fig. 4.1, Fig. 4.2);
Ray referring to equation (2.27);
R-P referring to Raleigh-Plesset equation (2.53);
s saturation conditions;
small small-scale flow;
super supercavitation;
t turbulent;
v vapour phase;
1 upstream the nozzle;
2 downstream the nozzle;
mean hydrodynamic scales of the flow.

Abbreviations
CFD computational fluid dynamics;
GALA gas and liquid algorithm;
PHOENICS Parabolic Hyperbolic Or Elliptic Numerical Integration Code Series;
ppm part per million (unit of concentration, 1 ppm = 106 kg/litre);
SIMPLE Semi-Implicit Method for Pressure Linked Equations;
SIMPLEST SIMPLE ShorTend;
VCO valve covered orifice;
VOF volume of fluid.

xxi
Chapter 1. Introduction 1

Chapter 1. Introduction
1.1. Cavitation in Diesel engine injection technology
During last fifteen years, significant progress has been achieved in the
development of Diesel fuel injection technology. Its biggest advance is a result of
development of efficient high-pressure common-rail technology (Ashley, 1997), which
provides flexibility in operation of engine over the entire range of engine speeds. In
direct injection Diesel engines equipped with a common rail system high injection
pressures (up to 2000 bar) are used to enhance the spray atomisation and air-fuel
formation in the combustion chamber. The more homogeneous the air-fuel mixture is,
the shorter the ignition delay, the local temperatures are lower and CO, NOx and soot
emissions are lower. As a result, the efficiency of performance of the engine is much
affected by the fuel atomisation (Heywood, 1988).
Modern Diesel engines are compact, efficient and produce few pollutants. Thanks
to their high economical efficiency and good drivability, Diesel engines are now widely
used for automobiles. Thus, over 40% of newly registered passengers cars in Western
Europe come equipped with a Diesel engine. However, recent developments in the
major Diesel technology still leaves space for further improvements in the ecological
and economical features of the Diesel engine. One of the possible ways here is to
control atomization of the fuel spray.
The characteristics of a fuel spray depend on the fuel properties, geometry of the
injector and the flow conditions upstream (inside) and downstream of the injection
nozzle. Two main types of Diesel injectors are used in direct injection systems: valve-
covered orifice (VCO) injectors and sac-type injectors (Fig. 1-1). Particular number and
arrangement of nozzle holes, as well as their geometries may vary depending on the
desirable conditions of operation of Diesel engine. Computational fluid dynamics
provides a useful tool for the prediction of the flow in a Diesel injector and, thus, design
and optimisation of the injection process. Although numerical simulations have become
more efficient, the experimental method of study is considered more reliable for use in
the design of Diesel injectors. This is due to the uncertainties in specification of many
parameters, such as geometry imperfections of real injectors, and also difficulties in the
numerical modelling of cavitation phenomena, which always accompany the flow in
Diesel injectors (Soteriou, et al., 1995; Badock, et al., 1999).
Chapter 1. Introduction 2

(a) (b)
Fig. 1- 1. Two types of Diesel injectors: valve-covered orifice (VCO) injector (a)
and sac-type injector (b).

Here, the term cavitation is applied to describe a process of the growth and
collapse of the vapour phase in the liquid when the local pressure drops below the
saturation pressure at a given temperature. Hydrodynamic cavitation occurs when the
reduction in static pressure is caused by the hydrodynamic motion of the liquid. This
phenomenon is different to acoustic cavitation, initiated by pressure waves in the liquid.

(a) (b)
Fig. 1- 2. Geometry-induced cavitation inside a nozzle (a) (Schmidt, et al.,
1999), and string-type cavitation in sac-type Diesel injector (b) (Arcoumanis, et al.,
2001). Dark regions indicate liquid, light regions cavitation flow.

When the pressure reduction is caused by an abrupt change in the geometry of the
flow passage, this is known as geometry-induced cavitation (Fig. 1-2, a). Also
cavitation can be dynamically-induced, when the pressure reduction is caused by the
pressure waves travelling in the liquid (Soteriou, 1995). In multi-hole injectors
cavitation in one nozzle hole can affect the flow in the other. In a sac volume of
Chapter 1. Introduction 3

injectors, string or vortex cavitation is observed (Fig. 1-2 b, Arcoumanis, et al., 2001;
Schmidt, et al., 1999; Laoonual, et al., 2001).
Experiments have revealed that spray disintegration occurs mainly as a result of
the aerodynamic break-up and disturbances to the flow caused by turbulent pressure
fluctuations and cavitation events inside the fuel injection nozzle (Hiroyasu, 2000). In
modern Diesel injectors, to enhance the aerodynamic break-up of the fuel spray and to
provide atomisation of fuel into droplets of a sub-micron scale, the injection is
performed at high speeds (200 400 m/s) through the nozzles of very small diameters
(~0.2 mm) at high injection pressures (up to 2000 bar) (Ashley, 1997). Under such flow
conditions cavitation becomes an essential feature of the flow. Therefore, control of the
injection process requires understanding and reliable prediction of cavitation
phenomenon.

1.2. Hydrodynamic cavitation


The hydrodynamic cavitation is known as the process of spontaneous growth and
further collapse of bubbles in a liquid, when variations in the pressure field have a
hydrodynamic nature. Small bubbles or particles, which are present in most technical
liquids, provide a source of nuclei for cavitation bubbles. Experimental studies of flows
in standard injection nozzles have shown that cavitation typically is present for the
whole or part of the injection period at most operation conditions (Soteriou, et al.,
2001). Cavitation bubbles formed inside the injection nozzle could produce pressure
fluctuations when collapsing in the high-pressure region downstream of the nozzle, and,
thus enhance the spray disintegration.
Though cavitation is expected to be beneficial for spray break-up, it may have
undesirable effects on the fuel injection performance. For example, experiments have
revealed that under certain conditions, cavitation can result in the formation of
hydraulic flip flow, which is not beneficial for atomisation (Fig. 1-3). Also, undesirable
effects of cavitation are associated with flow instabilities, excessive noise generation
and erosion, which causes the damage to nozzle walls.
Due to the complexity of the phenomenon of hydrodynamic cavitation, its control
has not yet become a part of the technology for the control of injection of fuel in Diesel
engines. Both experimental and numerical methods have been applied to develop
models of the cavitation process in Diesel injection nozzles.
Chapter 1. Introduction 4

Smooth jet
Transition flow
Jet break-up length
Incomplete Complete
Wavy jet spray spray

Hydraulic
flip Cavitation-
a free flow

Cavitating
flow

Jet velocity
Fig. 1- 3. Jet break-up regimes (Hiroyasu, 1998, 2000). Lines a and b shows
hysteresis behaviour in transition from cavitation flow to hydraulic flip regime (a) and
back from hydraulic flip to liquid flow regime (b).

At present experimental methods provide the only reliable basis to design and
control flows with hydrodynamic cavitation. However, studies of cavitation in nozzles
of sub-millimetre diameter are difficult, because of the special equipment and
techniques required for the measurement and visualisation of the flow. Therefore,
common practice is to predict the flow in a real-scale nozzle using the theory of
hydrodynamic similarity. According to this theory, the necessary conditions for
similitude, which allow the experimental observations of cavitation flow in one scale to
be transferred to another, are geometrical similarity of the flows and identity of
similarity criteria, such as Reynolds number, cavitation number, etc. In practice,
however, these conditions are not sufficient for hydrodynamic scaling, which may be
disturbed by viscous and liquid quality effects (Lecoffre, 1999). The scale effects can
not always be matched in experiment, causing the difference in cavitation flows at the
natural- (small-) scale and large-scale conditions. Therefore, the analytical and
Chapter 1. Introduction 5

numerical methods, which are able to cope with these scale effects, have become more
attractive for the prediction of cavitation flows of arbitrary scales. For theoretical
modelling, the key issue is identification of the two-phase pattern of the flow and
establishing the scale effects, which determine specifics of cavitation flow.

1.3. Scope of the present study


The present study is focused on the development of a model of cavitation, which
is capable of predicting the hydrodynamic similarity of cavitation flows in nozzles.
First, types of cavitation flows and factors responsible for the hydrodynamic
similarity (scale effects) are identified from analysis of the experimental data. Then, an
appropriate frame-work for cavitation modelling is chosen and additional assumptions
and simplifications about the flow are made. Further, the analysis of cavitation flow
models is performed and the problem in specification of the cavitation scale, which
describes the liquid quality effect, is clarified.
To develop a scalable model of cavitation, the governing equations are rearranged
into dimensionless form and the cavitation length scale is shown to be linked to the
number density of cavitation nuclei. Assuming the distribution of active cavitation
nuclei by their size, the form of this distribution is established to satisfy the
hydrodynamic similarity of geometrically homothetical cavitation flows of a given
liquid. This results in the development of a model for the number density of critical
cavitation nuclei as a function of liquid tension in the flow.
The correlation developed is compared to direct measurements of cavitation
bubbles and applied to simulate steady-state regimes of hydrodynamic vaporous
cavitation in real-scale models of Diesel injection nozzles.
From the analysis of experimental data on the hydrodynamic cavitation of the tap
water and Diesel fuel it was made a conclusion that critical pressure for the cavitation
onset can decrease with the viscosity of flowing liquid. To explain this effect, the
hypothesis of D. Joseph (1995) about the stress-generated cavitation in high-strain
flows. This hypothesis is applied to develop the model for the stressed-induced
cavitation and describe cavitation flows of Diesel fuel at high system pressures
(Winklhofer, et al., 2001).
Chapter 2.1. Literature survey. Introduction 6

Chapter 2. Literature survey


2.1. Introduction
In order to understand the nature of cavitation in Diesel injection nozzles
numerous experimental and theoretical studies have been conducted. This chapter
reviews the range of experimental studies conducted with the aim of highlighting the
key physical phenomena that need to be modelled and identifying data sets suitable for
validation purposes. The review of the theoretical studies examines the types of
modelling strategy employed in order to determine the most suitable mathematical
description to apply in this case and identify the issues associated with the numerical
solution of this type of model.

2.2. Theoretical and experimental background


2.2.1. Definitions

Cavitation is commonly known as the process of formation of voids in a liquid


due to a sudden pressure drop, when the local tension p v p exceeds the tensile

strength of the liquid p v p cr (Brennen, 1995). The tensile strength of a liquid depends
on the presence of weak spots in the liquid, which provide the nuclei for the phase
transition process. Because the density of the vapour phase is usually much smaller than
the density of the liquid, the amount of heat consumed locally for evaporation can be
neglected, so that the cavitation can be considered as isothermal process (Fig. 2-1).
In a more general sense cavitation is a process of formation and also consequent
collapse of bubbles in a liquid under a local decrease in pressure (Knapp, Daily,
Hammitt, 1970). Under the collapse stage, due to the inertia of the liquid and
compressibility of the gas-vapour bubble content, pressures and temperatures can
become extremely high inside the bubble, producing light-emission (sono-
luminescence) and causing erosive wearing of the working surfaces of hydraulic
systems.
Chapter 2.2. Literature survey. Background 7

p Critical
Saturation curve point
(binodal)

pv A
Tensile strength
Liquid of pure liquid
tension pv p cr

pl B
Liquid
spinodal
Tl T

Fig. 2- 1. The state of the liquid at saturation pressure pv (point_A) and under

tension pv pl at a given temperature (superheated liquid, point_B) (Skripov, 1974).

Depending on the topology of the vapour structures in the flow, cavitation can run
in a form of travelling bubbles, or vapour pockets, extending over the partial length of
nozzle/ body (cloud cavitation and sheet cavitation), or super-cavitation, when the
vapour region extends over the whole length of nozzle/ body (Knapp, et al., 1970;
Rood, 1991). Detail classification for the regimes of cavitation in nozzles is given in
what follows.

2.2.2. Cavitation inception and cavitation number


The term cavitation inception is applied to identify the flow conditions when
hydrodynamic cavitation just appears in the flow, to establish the boundary between
cavitating and non-cavitating flows. The term cavitation desinence is used for the
conditions, when cavitation vanishes from the flow under small variations in parameters
of the flow.
A cavitation number can be defined, which describes the nature of the flow. It can
be defined as an integral parameter of the flow, which relates the pressure drop (or
dynamic head) to the local static pressures. A critical cavitation number can be found
Chapter 2.2. Literature survey. Background 8

which marks the inception of cavitation and allows the flow to be classified as non-
cavitating or cavitating.
Various definitions of cavitation number have been applied for external and
internal flows. The present study uses the definition for cavitation number introduced by
Bergwerk (1959) and applied by Nurick (1976) and Soteriou, et al. (1995):

p1 p2
CN = , (2.1)
p 2 pv

where p1 and p2 are pressures at the inlet and outlet of nozzle and pv is the vapour
pressure, usually associated with the saturation pressure in the liquid.
Experimental studies of cavitation flows in nozzles and orifices and also flows
around bodies of different shapes have revealed a relationship between the cavitation
number and the extent of the cavitation region. This has resulted in classification of
cavitation flow regimes depending on the cavitation number, namely: incipient,
developed (sub-cavitation and transitional cavitation), and super-cavitating (Sato and
Saito, 2001; Stinebring, 2001).
Many experiments have proved the effect of cavitation bubble nuclei on the
cavitation inception number (Billet, 1985; Rood, 1991; Gindroz, et al., 1997). Lecoffre
and Bonnin (1979) have concluded that a consistent definition for the cavitation
inception number should use the actual tensile strengths of the liquid, which accounts
for all sort of impurities present in the flow (liquid quality effects).

2.2.3. Similarity criteria and scale effects

To determine the structure of a cavitation flow in a nozzle, hydraulic


measurements are accompanied by visual observations of the flow inside the nozzle. A
difficulty in determining the structure of cavitation flow in nozzles appears due to the
problem of visualizing the flow in real-scale transparent nozzles, due to their small size.
This requires the application of special methods and equipment. To overcome this
difficulty, scaling theory is applied, so that larger models can be used. The results from
these large scale experiments have then been extrapolated to real-scale flows (Soteriou,
et al., 1995; Arcoumanis, et al., 2001; Kim, et al., 1997).
Chapter 2.2. Literature survey. Background 9

The following parameters have been found to be significant in the description of


cavitation flow: densities and coefficients of dynamic viscosity of liquid and vapour
phases l , v , l and v , coefficient of surface tension , speed of sound in liquid c ,

spatial scale of the flow l , pressures upstream and downstream the cavitation region
p1 and p 2 , and velocity scale of the flow u (Lecoffre, 1999). Applying dimensional
analysis (Batchelor, 1980) the following dimensionless groups can be identified:

Cavitation number, to describe the intensity of cavitation:


p1 p2
CN = , (2.1)
p 2 pv
Reynolds number, to describe viscous effects on the flow:
l u l
Re = , (2.2)
l
Weber number, to describe the effect of surface tension:

l u 2 l
We = , (2.3)

Froude number, to describe the effect of gravity force:
u
Fr = , (2.4)
g l

Mach number, to describe the compressibility effect:


u
M= , (2.5)
c
Discharge coefficient, to describe the effect of viscous dissipation:

l u 2
Cd = , (2.6)
2( p1,o p 2 )

and Strouhal number, to describe unsteadiness of the flow, when it can be characterised
by periodical motion at natural frequency F :

F l
Str = . (2.7)
u
Chapter 2.2. Literature survey. Background 10

These dimensionless groups completed by ratios l / v and l / v appear in


the similarity equation (scaling law) of cavitation flow:


f Re, CN, We, Fr, M, Str, C d , l , l = 0. (2.8)
v v

This formulation describes the hydrodynamic similarity and neglects thermal


scaling and any detail features of the flow, including geometry and two-phase
structures.
In high-speed cavitation flows with prevailing inertia forces, We and Fr numbers
are usually neglected in equation (2.8). Experimental studies of cavitation flows in real-
size nozzles (Bergwerk, 1959; Nurick, 1976) have confirmed that the cavitation number
CN and Reynolds number Re are the most important criteria, which describe similarity.
The transient effects and compressibility of fluid, described by Mach and Strouhal
criteria, can become important for high-speed and unsteady flow conditions in Diesel
injectors. To account for the specifics of cavitation flow, the number of dimensionless
criteria in equation (2.8) can be increased. In order to establish an explicit form of
equation (2.8), for example to express C d in terms of the other dimensionless groups,
experimental or numerical studies are needed.
Experiments have shown that real flows do not always obey the classical scaling
theory. The reason for this are scale effects caused by the liquid quality (the presence of
dissolved gases in the liquid, small particles and gas-vapour nuclei), bubble dynamics,
geometrical differences due to manufacturing and wall roughness, particular flow
regime and turbulent motion (Bergwerk, 1959; Nurick, 1976). These scale effects mean
that the similarity criteria (2.1) (2.7) are not sufficient for hydrodynamic scaling
(Lecoffre, 1999). In practice, to make two cavitation flows similar, the scale effects
associated with the liquid quality and viscous nature of the flow should be minimised.
From the point of view of accurate and reliable prediction of cavitation flows it
becomes important to identify the scale effects, which determine the particular regime
of cavitation flow.
The liquid quality effects are associated with the presence of cavitation nuclei in
the liquid. They can be described using the theory of bubble dynamics, when the nature
of the cavitation nuclei has been recognised.
Chapter 2.2. Literature survey. Background 11

The flow effects depend on the geometry of the flow domain (geometry effects),
Reynolds number and parameters of turbulent motion (viscous effects). They determine
the structure of cavitation-free flow (laminar or turbulent, location of separation point
and extent of the recirculation region) and govern the onset and development of
cavitation flow.

(a) (b)
Fig. 2- 2. Photographs of cavitation flow in large-scale (a) and real-size (b)
Diesel injection nozzles under similar flow conditions (Arcoumanis, et al., 2000).

Arcoumanis, et al. (2000) have determined from experimental studies that the
general steady-state macroscopic structure of cavitation is similar for real-size and large
scale nozzles (Fig. 2-2). The structure mainly depends on the nozzle geometry and flow
effects determined by the Reynolds and cavitation numbers. The differences in the
details of cavitation flows in the small- and large-scale nozzles can be referred to the
scale effects caused by imperfections in the nozzle geometry and cavitation nuclei, etc.
In the following, to emphasise these effects on cavitation flow, the term small-, or real-
scale nozzle is applied when speaking about nozzles of diameter less than 1mm, which
is typical for Diesel injectors.
Several experimental studies have been performed to quantify the influence of
scale effects on the cavitation inception in nozzles and orifices (Knapp, et al., 1970;
Tullis, 1973; Schmidt and Corradini, 2001; Keller and Huber, 2001). They result in
empirical correlations for the cavitation inception number, applicable for one-
dimensional predictions of the early stages of cavitation. At the same time a quantitative
description of the flow effects requires consideration of the actual three-dimensional and
transient behaviour of viscous flow. In turn, this requires the development of a local
model of cavitation, which may be incorporated in the appropriate CFD code.
Comparison of numerous cavitation and related phenomena models (such as flashing
Chapter 2.2. Literature survey. Background 12

and choking), which will be described in the next chapter, shows that the most
challenging aspect is the modelling of the liquid quality effects, associated with the
presence of cavitation nuclei in liquid.
Lecoffre (1979, 1999) assuming a bubbly type of cavitation flow, has concluded
that the necessary condition for hydrodynamic similarity of cavitation flows is scaling
for the number density of bubble nuclei no according to the hydrodynamic scale of the

flow l :

1
no ~ (2.9)
l3

Many cavitation models contain dimensional parameters for the liquid quality
effect, which require scaling according to (2.9) to describe the hydrodynamic similarity
of cavitation flows. However, this approach is rather limited, because the adjustable
parameters depend not only on the liquid quality effects, but also on the hydrodynamic
length scale l in (2.9). So, the major problem, which is a subject of the present study
is the development of a scalable model of cavitation which could match the condition
(2.9).

2.2.4. Cavitation bubble nuclei


Many observations have proved that the hydrodynamic cavitation occurs in the
bulk liquid (Brennen, 1995), where the vapour-gas bubbles provide the main
contribution to the nucleation process. Nucleation is the process of the formation of
small voids in the liquid. Two types of nucleation can be defined depending on the
nature of the weaknesses, which initiate the bubble growth, namely homogeneous and
heterogeneous.
Homogeneous nucleation is the process of macroscopic bubble development from
small voids, which appear in the liquid due to the thermal molecular motion. In classical
kinetic theory of liquids the thermal motion of molecules is considered as the only
mechanism of homogeneous nucleation. In real systems micro-bubbles filled with a gas
can also initiate homogeneous nucleation. These gas micro bubbles can be present in
Chapter 2.2. Literature survey. Background 13

crevices at solid boundaries or small suspended particles. Free small bubbles also can
exist in the liquid (Yount et al, 1977).
The heterogeneous mechanism of nucleation becomes active when the weakness
is formed on a small contaminant particle or at the junction between the liquid and solid
wall.

p Saturation curve
Critical pressure (binodal)
in liquid with a
vapour bubble

p v (Tl ) 0

pv 2 / R Tensile strength of liquid


with a vapour bubble

Liquid Tension in liquid


spinodal
pv p
Tl T

Fig. 2- 3. Critical pressures in liquid with a vapour-filled bubble, comparing to


critical pressures in pure liquid (spinodal curve) and saturation vapour pressures
(binodal curve).

The presence of a sufficient number of impurities in the liquid makes the


evaporation of the liquid more likely. The thermo-molecular mechanism of
homogeneous nucleation becomes important for hot liquids, when the degree of liquid
superheat is above 10-20% of the critical (saturation) temperature (Lecoffre and
Bonnin, 1979). In this case the free-stream nuclei become irrelevant for the vapour
formation process.
At the curvilinear inter-phase boundary of a spherical bubble, the phase
equilibrium conditions are changed to account for the surface tension phenomena. The
static equilibrium condition for pressure at the bubble interface is:

2
pl = pB , (2.10)
R
Chapter 2.2. Literature survey. Background 14

where pl is pressure in the liquid, and pB = pv + p g is the pressure inside the bubble

comprised by partial pressures of the vapour and the gas. Equation (2.10) shows that
equilibrium pressure inside the bubble is always bigger than pressure in surrounding
liquid. The bubble will start grow in radius when pressure in the liquid drops below the
equilibrium level (2.10), i.e. pl < pv 2 / R (Fig. 2-3).

Unstable bubbles

Equation
Stable bubbles (2.11)

pv pl

Fig. 2- 4. Critical radius for the stable vapour-gas bubble in liquid under tension.

Stability of the vapour-gas bubble due to a step change in the pressure of the
surrounding liquid can be analysed with a help of equation (2.10), (Brennen, 1995).
This analysis clarifies the definitions for cavitation and cavitation onset. Neglecting the
gas diffusion and assuming the mass of gas inside the bubble and its temperature remain
the same:
3
R
p g = p g ,0 0
R .

The radius of a bubble, which can stay in stable equilibrium with its surrounding
liquid, can be determined (Blakes critical radius, Brennen, 1995):

4
Rcr =
3 pv pl . (2.11)
Chapter 2.2. Literature survey. Background 15

The bubbles with a radius smaller than the Blake radius (2.11) will remain stable
being subjected under tension pv pl , changing their radius to match the equilibrium
condition (2.10). This stage corresponds to the premature cavitation (some times also
referred as gaseous cavitation, or degasation). Bubbles of radii bigger than Rcr defined
by (2.11) are unstable and explosively grow in size, producing the inception of true
vapour cavitation (Rood, 1991). The theory gives an estimate for the critical tension of
liquid containing a spherical bubble of initial radius Ro at pressure po (Knapp, 1970;

Brennen, 1995) (Fig. 2-4):

1 / 2
4 po pv
( pv pl )cr = 31 +
3 Ro 2 / Ro . (2.12)

The onset of cavitation, i.e. the local flow conditions where cavitation bubbles
start to develop, depends on the local hydrodynamics of the flow and tensile strength of
the liquid. Equations (2.10) and (2.12) show that the presence of bubble nuclei can
significantly decrease the tensile strength of the liquid, comparing to the degree of
tensile strength of pure liquid (Fig. 2-3). The tensile strength of real liquids is dependent
upon the liquid quality, i.e. the presence of impurities, which may initiate growth of
cavitation bubbles. Real liquids are characterised by a non-uniform spectrum of the
bubble nuclei. Since the critical radius decreases with the pressure drop (2.11), more
nuclei in a given distribution become active cavitation sites. Therefore, the actual
tensile strength of liquids is a function of nuclei spectrum. Experimental studies of
cavitation were undertaken to reveal the effects of cavitation nuclei, presence of
dissolved gases and other sort of impurities, on the tensile strength of liquid and
cavitation inception.

2.2.5. Cavitation in flowing liquid

The above definition for the cavitation (Brennen, 1995) is based upon the
observations on liquid rupturing at static or quasi-static conditions, when the static
pressure in major part of the liquid volume is much higher than the viscous stresses
caused by the liquid flow. The classical experiment on the tensile strength of static
Chapter 2.2. Literature survey. Background 16

liquids was established by Berthelot (1850). Though this definition is widely used in
predictions of cavitation phenomenon, the actual stresses acting on the element of
viscous liquid in motion are comprised by the shear and normal stresses caused by the
velocity gradients in the flow. This results in alternative definition for the onset of
cavitation, i.e. the state of liquid when fracture occurs (Joseph, 1995).
The equation of motion of moving liquid represents the balance of forces acting
on an infinitely small element of mass of the liquid:

u j u k u j jk
+ = Fj + ,
t x k x k

where F j is a j -component of the resultant mass force and jk is the stress tensor. In

Newtonian liquids, the stress is proportional to velocity gradient (Batchelor, 1980):

ui 1 ui
jk = p jk bulk jk + 2 S jk jk ,
x 3 x
424i 44
144 3 1444 4244i 44 3
Normal viscous stress Shear viscous stress = jk

1 u j u k
is the rate-of-
where is dynamic coefficient of shear viscosity, S jk = +
2 x k x j

strain tensor, and bulk is coefficient of the bulk viscosity. In an incompressible fluid

u i xi = 0 , so that the equation for the stress tensor is simplified to

jk = p jk + 2 S jk .

According to a hypothesis by Daniel Joseph (1995), the liquid may rupture when
the maximum of the principal component of the stress tensor overcomes the vapour
pressure in liquid pv :

max(~ii ) p v , (2.13)
Chapter 2.2. Literature survey. Background 17

where ~ii denotes the principal components of the stress tensor , with unequal

components of the normal stress ~11 ~22 ~33 and zero shear stress ~ik = 0 ( i k ).

The criterion (2.13) determines the onset of cavitation in the flowing liquid.
In flows with the hydrodynamic cavitation, regions of the flow with large pressure
gradients are accompanied by high rates of strain. Therefore, for the hydrodynamic
cavitation, the concept (2.13) may complete the conventional low-pressure
condition p < p v .

y
u( y)

Fig. 2- 5. Couette flow between two walls.

In two-dimensional shear flow (Fig. 2-5) of an incompressible liquid the condition


(2.13) can be simplified to (Joseph, 1995):

u
p < pv + . (2.14)
y

Equations (2.13) and (2.14) show that at high system pressures, the liquid may
start to cavitate even when the static pressure does not fall below the critical pressure
(here, associated with the saturation vapour pressure pv ).

2.2.6. Conclusions
Cavitation is a complex phenomenon, whose appearance depends on the physical
properties of the flowing substance and presence of initial nuclei in the liquid.
Hydrodynamic cavitation starts to develop from the bubble nuclei in the bulk
liquid. To initiate the explosive (cavitation) growth of bubbles, a critical tension is
required. The tensile strength of real liquids depends on the spectrum of all the nuclei
present in the liquid and can be reduced by the viscous shear stresses in the flow.
Chapter 2.2. Literature survey. Background 18

To match the hydrodynamic similarity of cavitation flows, the scale effects (fluid
quality and flow scale effects), should be taken into account. The flow scale effects,
caused by the viscous nature of the flow can be identified from experimental
observations of cavitation, and then described in a computational model of the flow.
Analysis of experimental observations of hydrodynamic cavitation in nozzles will be
performed to clarify the scale effects on the two-phase structure of cavitation flow. To
determine the liquid quality scale effects a theoretical approach will be applied
(Stinebring, et al., 2001), with additional analysis of experimental observations about
cavitation nuclei in liquids.
As a result of numerous investigations, the main scale effects have been identified
and several empirical equations have been suggested to take them into account when
predicting cavitation inception (Tullis, 1973; Keller and Huber, 2001; Hsiao, Chahine
and Liu, 2003). At the same time, experimental studies still do not give a clear answer
regarding the influence of scale effects on the cavitation development and particular
features of the flow, such as a pattern of two-phase flow, discharge characteristic and
turbulent fluctuations. Quantification of these effects is crucial to developing on
understanding of the physical nature of cavitation and accurate prediction of cavitation
flows.
The following review collects the results of experimental studies of cavitation
flows in single-hole nozzles to clarify the structure and regimes of cavitation flow and
influence of scale effects on the flow.
Chapter 2.3. Literature survey. Experimental observations 19

2.3. Experimental observations of cavitation flows in nozzles


2.3.1. Introduction

In order to obtain better understanding of the nature of cavitation, as a process of


formation and decay of the vapour phase in liquid, many experiments have been
performed for idealised flow conditions and simplified configurations of injectors.
These studies were conducted to clarify the patterns and main features of cavitation
flows in Diesel injectors. They clarify specific features of cavitation to be considered in
a model of cavitation flow. The main results of some of these studies are described
below.

Lcav
Liquid Vapor Nozzle wall
with
1 nuclei 2
Ac A

A1 Separation Lsep
point Reattachment
Vena contracta point L

Core of the flow


p1
Near-wall region

p2
Z
pv
pv pmin = liquid tension
p min

Fig. 2- 6. Sketch of the geometry-induced cavitation flow and pressure


distribution along the nozzle in the core region and near the wall.

Geometry-induced cavitation in nozzles occurs when the pressure drop at the


nozzle entrance is accompanied by flow reattachment and formation of a recirculation
Chapter 2.3. Literature survey. Experimental observations 20

region (Fig. 2-6). The presence of the recirculation region in the flow determines the
distribution of tensions, and therefore has a crucial effect on the formation and
development of cavitation. However, as will be shown later, flow reattachment may not
happen in flows at low Reynolds numbers and in nozzles with a smooth entry
configuration. Therefore, specific features of the liquid flow and its interaction with the
growth and collapse of cavitation structures should be carefully achieved when
developing a model of cavitation flow (see discussion of the bubble dynamics scale
effects, Stinebring, et al., 2001).
At present, a number of experimental works have been published in the literature
to determine the behaviour of cavitation flows in nozzles and orifices. These
experiments help to identify the structure of cavitation flow, i.e. the pattern of two-
phase flow and the hydraulic resistance of nozzle, which determine impacts of
cavitation on the flow and the nozzle performance.
Significant progress in clarification of the structure of cavitation flows in nozzles
has been achieved as a result of studies by Bergwerk (1959) and Spikes and Pennington
(1959), Lichtarowicz, et al. (1974), Nurick (1976), Arai, Shimizu and Hiroyasu (1985),
Fox and Stark (1989), Chaves, Knapp and Kubitzek (1995), Soteriou, Andrews and
Smith (1995), Roosen with colleagues (1996-1998), Sato and Saito (2001), Laoonual, et
al. (2001) and Winklfofer, Kull, Kelz and Morozov (2001). These studies shed a light
on the phenomena of hydrodynamic cavitation in nozzles, in particular concerning the
following aspects:
Identification of the pattern of quasi-steady-state cavitation flows in nozzles;
Clarification of the scale effects, which cause the difference in cavitation flows in
large- and small-scale nozzles;
Identification of the variety of scale effects on cavitation flow (caused by
variations micro- and macro-geometry of nozzle, flow conditions and liquid
quality) including the Diesel injector-specific scale effects (injector configuration,
needle position and unsteadiness of injection process).

In the following the main experimental observations concerning the above listed
issues are reviewed. This is completed by a discussion of the effect of cavitation on the
nozzle flow and the spray formation.
Chapter 2.3. Literature survey. Experimental observations 21

2.3.2. Visualisations of cavitation flows in nozzles

The structure of cavitation flow in nozzles is influenced by the pattern of two-


phase flow and hydraulic resistance of the nozzle.
Analysis of experimental studies of hydrodynamic cavitation shows that the
pattern of two-phase flow mainly depends on the integral parameters of the flow, such
as Reynolds and cavitation numbers and geometry of the flow domain (Rouse and
McNown, 1948; Bergwerk, 1959; Spikes and Pennington, 1959; Lichtarowicz, et al.,
1974). In sharp-entrance nozzles, the flow reattachment at the nozzle entry (Fig. 2-6)
determines formation and development of two-phase structures in the nozzle.
Early experimental observations on cavitation in nozzles have revealed that under
developed turbulent flow conditions, the cavitation flow mainly depends on the
cavitation number and has little dependence on the Reynolds number (Bergwerk, 1959).
This conclusion concerns the quasi-steady-state cavitation in axisymmetrical nozzles.
However, at small Reynolds numbers, the effect on the flow separation and discharge
becomes more important (Schmidt, et al., 1999). The effects of the flow asymmetry,
Reynolds number, and transience are discussed later in this chapter.

Lcav

L
Super-
cavitation

Transitional
cavitation
Sub-
cavitation
Cavitation
0 inception

0 1 -1
CN super CN inc CN -1

Fig. 2- 7. Development of the average length of cavitation region Lcav (Fig. 2-6)

with cavitation number CN.


Chapter 2.3. Literature survey. Experimental observations 22

Studies of cavitation flows in nozzles have revealed that the cavitation number
determines an extent of the region inside the nozzle filled with the vapour. Development
of the length of cavitation region in nozzle with cavitation number is schematically
shown in Fig. 2-7, which is based on the observations of Nurick (1976), Chaves, et al.
(1995), Roosen, et al. (1997), Hiroyasu (2000) and Sato and Saito (2001). The
relationship between the cavitation number and the extent of cavitation region have
allowed Sato and Saito (2001) to make the following classification to the regimes of
quasi-steady-state cavitation flows in nozzles:
Cavitation inception when cavitation starts;
Sub-cavitation stage when vapour fills the separation region;
Transitional cavitation when cavity extends downstream the nozzle entrance
(shedding cavities and cavitation clouds);
Supercavitation when cavity length reaches the nozzle outlet.
At a certain value of cavitation number CNinc inception of cavitation occurs in
vena contracta region (zone of recirculation flow downstream the nozzle entrance).
During the sub-cavitation stage cavitation bubbles fill the separation region, and
changes in CN have little effect on the length of cavitation region. Further increase in
CN enlarges the cavity length (transitional cavitation) and at certain point (CNsuper)
cavitation zone rapidly extends to the outlet of nozzle (supercavitation). At high
cavitation numbers CN > CNsuper (Fig. 2-7) cavitation zone exceeds the nozzle hole and
forms jet cavitation when the nozzle is submerged in a liquid (Sato and Saito, 2001).

(a) (b) (c) (d)


Fig. 2- 8. Characteristic types of spray formation while flowing through the
large-scale circular nozzle (Soteriou et al., 1995). Injection into a gas.
Chapter 2.3. Literature survey. Experimental observations 23

In experiments on the hydrodynamic cavitation in single-hole nozzles, two main


patterns of cavitation flow have been observed: bubbly flow pattern and separated flow
pattern.
Observations of cavitation in large-scale nozzles (Chaves, et al., 1995; Soteriou, et
al., 1995; Arcoumanis, et al., 2000; Sato and Saito, 2001) have revealed the presence of
cavitation bubbles and bubble foams at the incipient, sub-cavitation and transitional
stages (Soteriou, et al., 1995, Fig. 2-8 b, c), while under the developed stage of
cavitation several experiments (Arcoumanis, et al., 2000; Ganippa, et al., 2001; Sato
and Saito, 2001) shown formation of transparent glossy sheet cavitation at the nozzle
entry. This sheet-type cavitation is similar to the transparent vapour films observed at
high cavitation numbers CN (transitional and supercavitation stages) in small-scale
nozzles (Fig. 2-9, Fig. 2-10 and Fig. 2-11) (Bergwerk, 1959; Chaves, et al., 1995;
Roosen, et al., 1996-1998; Badock, et al., 1999; Schmidt, et al., 1999; Winklhofer, et
al., 2003).

(a) (b)
Fig. 2- 9. Density gradient field and velocity field in a small-size cavitating
nozzle (nozzle height = 0.28mm), operating at injection pressure p1 = 80 bar and

downstream (back) pressures p 2 of 21 bar (a) and 11 bar (b). Images obtained using
the shadowgraph and PIV techniques. (Experimental data of Roosen et al. (1996)
referred by Yuan and Schnerr (2001)).
Chapter 2.3. Literature survey. Experimental observations 24

Fig. 2- 10. Image of cavitation in a nozzle of rectangular shape obtained using a


fluorescent laser light technique (Roosen, et al., 1997, 1998). Flow direction from left
to right, working liquid decalin, upstream pressure 30 bar and downstream pressure 1
bar. Nozzle width ~0.3 mm.

Fig. 2-11. Fig. 2-12.

Fig. 2- 11. Structure of the internal flow in a real-size Diesel injection nozzle
captured using the light sheet technique (Badock, et al., 1999).
Fig. 2- 12. A foam of cavitation bubbles formed in the sharp entry nozzle. Series
of photographs of the water flow in a circular nozzle of 22 mm in diameter (Sato and
Saito, 2001).

When the outflow is into a gas, experiments (Nurick, 1976; Fox and Stark, 1989;
Soteriou, 1995; Laoonual, et al., 2001) have revealed that the flow reattachment at the
Chapter 2.3. Literature survey. Experimental observations 25

nozzle exit disappears and the jet becomes separated from the nozzle walls. This flow
regime is known as the hydraulic flip (Fig. 2-8). In relatively short symmetrical nozzles
(L/D < 10) the liquid core can become completely separated from the nozzle walls (total
hydraulic flip) causing a dramatic decrease in the discharge coefficient (~0.62 for
circular nozzles according to Nurick, 1976), and reduction in the spray penetration
(Soteriou, et al., 2001).
Long symmetrical nozzles (L/D > 10) are unlikely to become flipped (Laoonual et
al., 2001; Fox and Stark, 1989). Rounded inlet corners of the nozzle make transition to
hydraulic flip flow more unlikely (Laoonual, et al., 2001).
Soteriou, et al. (2001) have concluded from experiments on models of Diesel
injectors that hydraulic flip is less likely to occur under transient conditions and at low
needle lifts.
Many authors have pointed out shedding of cavitation region and unstable
behaviour of transitional and supercavitation flows (Lichtarowicz and Pearce, 1974;
Sato and Saito, 2001). Thus, detailed measurements of cavitation in a small-scale
rectangular nozzle performed by Roosen, et al. (1997, 1998) showed a sheet-type
cavitation formed at the nozzle entry and a row of bubbles periodically detaching from
the rear side of the cavity (Fig. 2-10). Chaves, et al. (1995) have concluded that under
high injection pressure conditions, it was difficult to identify the structure of the
cavitating flow, and to determine whether it is a rough vapour film or foam of small
bubbles. From analysis of experimental studies of cavitation in Diesel injectors,
Soteriou, et al. (2001) have concluded that different types and levels of cavitation may
exist simultaneously in adjacent flow regions within the same nozzle hole.

2.3.3. Discharge coefficient of nozzle

Visualisations of cavitation flows have been used to identify the basic two-phase
patterns for cavitation flows in nozzles. However, these studies were not able to clarify
the scale effects on cavitation that requires qualitative measurements in the flow.
The efficiency of the injection process is determined by the quality of the spray
atomisation and the hydraulic resistance of the nozzle, both of which are affected by
cavitation.
The measure of the hydraulic resistance of a nozzle is the discharge coefficient:
Chapter 2.3. Literature survey. Experimental observations 26

m& actual 1
lu 2
Cd = = 2
(2.15)
m& ideal p1,o p 2

where m& actual is the actual mass flow rate through the nozzle, m& ideal = A 2 l ( p1,o p 2 ) , A

is the cross-sectional area of the nozzle, p1,o is the total pressure at the nozzle inlet and

u is the mass average velocity of the flow in a nozzle.

Imaging of cavitation flows in nozzles has revealed that cavitation happens in the
low-pressure region formed at the nozzle entry (Fig. 2-6) (Dumont, et al., 2000). Half-
empirical correlations for the discharge coefficient of a cavitating nozzles are based on
this observation and take into account the local reduction in the liquid flow area in the
vena contracta region.
Before the discussion of the scale effects on cavitation flow (two-phase structure
and parameters of the flow), some basic scale effects (geometry and viscous effects) on
the liquid (cavitation-free) flow are reviewed.

2.3.3.1. Cavitation-free flow


Experimental studies of liquid flows in axisymmetrical nozzles and orifices have
shown that the discharge coefficient is mainly affected by the following factors (see, for
example, Lefebvre, 1989):
- the Reynolds number and inlet turbulence, which determine formation and extent
of the flow separation region;
- the length of a nozzle, L/D (due to the increase in friction losses with L/D);
- the ratio of nozzle diameter and supply tube diameter D/Ds;
- the shape of inlet corners, which also affects the flow separation.

The above geometry-scale effects can be taken into consideration when


constructing the model for the discharge coefficient of a nozzle. To achieve separately
changes in the momentum of the flow caused by contraction in the flow area and
irreversible friction losses, the discharge coefficient of orifices (i.e. short nozzles, which
length is less than the extent of separation region) is presented in a form:

Cd = CvCc ,
Chapter 2.3. Literature survey. Experimental observations 27

where C c = Ac / A is coefficient of contraction in the vena contracta region (Fig. 2-6)

and C v = u / u ideal is the velocity coefficient, which accounts for the friction losses and

u ideal = 2 l ( p1,o p 2 ) is the average velocity in the nozzle according to the Bernoulli

equation. The velocity coefficient for the developed turbulent flow is close to unity
C v 1 . The contraction coefficient depends on the shape of the nozzle. Potential flow

theory gives the contraction coefficient for the free jet emerging from a plane orifice
C c = /( + 2) 0.61 (Batchelor, 1980). For finite area of the inlet plenum, the

contraction coefficient C c is related to the flow area ratio = A / A1 (Nurick, 1976):

C c 0.62 + 0.38 3 . (2.16)

The shape of the nozzle entry determines formation of vena contracta, which can
disappear in nozzles with radii of inlet corners Rinlet bigger than 0.14 D (Nurick,

1976). To describe the effect of rounded inlet corners on the discharge coefficient of
nozzle, Nurick (1976) has suggested the following equation:

[
C c = C c,2sharp 11.4 Rinlet / D ]
0.5
, (2.17)

which is valid in the range Rinlet / D < 0.14 .


To describe both the effects of the nozzle length and Reynolds number on the
discharge coefficient of cavitation-free turbulent flows, Lichtarovizc, et al. (1965) have
proposed the correlation:

1 1 20
= + (1 + 2.25 L / D ) ,
C d 0.827 0.0085 L / D Re

which is valid for Re < 2 10 4 and nozzle lengths L / D from 2 to 10.


Chapter 2.3. Literature survey. Experimental observations 28

2.3.3.2. Cavitation flow


Visualisations of cavitation flows have revealed the major effect of cavitation
number on the pattern of steady-state cavitation flow in nozzles. (Fig. 2-7, 2-8).

Fig. 2- 13. Discharge coefficient and cavitation structures observed in plain


nozzle at various cavitation numbers. Dashed line shows approximation for the
discharge coefficient under the choked flow conditions (equation (2.18)):
C d 0.67 1 + 1 / CN .

Recently, Winklhofer et al (2001) have published results of extensive studies of


the structure of cavitation flow in a real-size model of Diesel injection nozzle
(transparent nozzle of rectangular cross-section). These experiments were performed for
injection into liquid, when no hydraulic flip could happen. Results of hydraulic
measurements by Winklhofer, et al. (2001) are plotted in Fig. 2-13 as the discharge
coefficient versus the cavitation number. This figure shows that at the early stages of
cavitation have only minor effect on the hydraulic resistance of the nozzle; the
discharge coefficient is significantly affected by cavitation number at super-cavitation
flow regime, when the flow become choked and mass flow rate become independent on
the downstream pressure (Winklhofer, et al., 2001). Fig. 2-13 also emphasise the
relationship between the structure of cavitation flow (presented by the flow pattern and
discharge coefficient) and cavitation number:
Chapter 2.3. Literature survey. Experimental observations 29

Fig. 2- 14. Discharge coefficient under cavitation flow in circular nozzles.


Experimental data by Nurick (1976) and Soteriou, et al. (1995) in comparison with the

correlation (2.18): C d 0.62 1 + 1 / CN .

Under the liquid flow and cavitation onset the mass flow rate is dependent
on the pressure drop across the nozzle.
At a critical point ( CN sup er ) cavitation flow changes from the transitional

regime (in terms of Sato and Saito, 2001) to choked flow.


Under the choking flow conditions (or super-cavitation) the mass flow rate
becomes unaffected by the downstream pressure.

The last observation allows an estimate for the discharge coefficient of a


cavitating nozzle to be made. Actually, the pressure in the cavitation region attains the
vapour pressure p v . When the vapour region occupies the whole length of the nozzle,
the mass flow rate becomes virtually independent on the back pressure (the flow is
choked):

m& chocked = A C c C v 2 l ( p1,o p v ) ,

while the discharge coefficient becomes only a function of cavitation number:


Chapter 2.3. Literature survey. Experimental observations 30

1
Cd = Cc Cv 1 + . (2.18)
CN
where the contraction coefficient C c can be calculated using the correlations for

cavitation-free flow (2.16) and (2.17).


This correlation has been proposed by Bragg and Crossby in 1959 (discussion to
the paper of Spikes and Pennington, 1959) and became a useful basis for quantification
of the discharge measurements of cavitation flows (Fig. 2-13, 2-14).
When the jet emerges into a gas, the flow can become separated from the walls at
the nozzle exit. This may cause a sudden decrease in the hydraulic resistance of the
nozzle and corresponding reduction in the discharge coefficient down to the limit at
CN in equation (2.18) (Nurick, 1976).

2.3.4. Scale effects


Depending on their nature, scale effects can be classified in two types: viscous
effects and bubble-dynamics effects (Stinebring, et al., 2001; Keller and Huber, 2001):
Viscous effects. These are the flow effects described by Reynolds and cavitation
numbers and effects of turbulent pressure fluctuations; and also micro-geometry effects,
caused by the wall roughness and any imperfections in nozzle geometry caused by
manufacturing, and macro-geometry effects, associated with the nozzle shape.
Bubble-dynamics effects. These are the liquid quality effects, associated with the
cavitation nuclei and presence of dissolved gases in the liquid, and also effects caused
by specific features of cavitation vapour structures (inertial effects, viscous effects,
thermal effects, surface tension effects, etc.).
Quantification of the scale effects is important for the development of a model
of cavitation flow. The purpose of the following review is to clarify the particular scale
effects on cavitation flows in nozzles.

2.3.4.1. Flow scale effects


Two groups of the flow scale effects can be recognised. First group includes
integral scale effects, and second group describes the local scale effects.
The integral flow scale effects can be described using the dimensionless criteria,
such as Reynolds and cavitation number, based on the integral geometrical and
Chapter 2.3. Literature survey. Experimental observations 31

hydrodynamic scales of the flow. These dimensionless groups are present in equation
(2.8). The effects of Reynolds and cavitation numbers on the structure of cavitation flow
were already described in the previous section of this chapter.
The local flow scale effects are determined by the local mechanism and specific
structure of cavitation flow. Description of these effects requires consideration of three-
dimensional and transient nature of cavitation. The idea of CFD models of cavitation is
to consider the basic phenomena, which describe locally the cavitation flow. Then the
whole pattern of cavitation flow and its integral characteristics, such as the extent of the
cavitation region, amount of vapour produced, hydraulic resistance, etc., can be
predicted for an arbitrary geometry.
Experimental studies of the geometry scale effects have helped to reveal their
effect on the structure of cavitation flow. In the following section these effects are
discussed first, followed by analysis of the flow scale effects, which determine the local
mechanism of cavitation. These local scale effects are caused by specifics in thermo-
physical properties of cavitating liquid (viscosity, surface tension, and compressibility)
and turbulent motion in the flow.
The equilibrium thermo-physical properties of a liquid depend on its content,
temperature and pressure conditions. The working liquid also can make a solution with
other liquids and gases, which affects the saturation pressure. Diesel fuel is a multi-
component mixture of liquids, with different boiling points. This results in evaporation
starting from the more volatile components present in the mixture. Density, viscosity,
thermal conductivity, surface tension and latent heat of evaporation are varied for
different type of Diesel fuels. Most experiments on cavitation are performed using the
light Diesel fuel, which properties are listed in the Table. 2-1 (Heywood, 1988;
Pulkrabek, 1997).

Table 2- 1. Physical properties of liquids (20oC, 1 bar).


Liquid Dynamic Thermal Surface Latent heat of Vapour
Heat capacity,
Liquid density, viscosity, conductivity, tension, evaporation, pressure,
3 . .
J/(kg.K)
kg/m Pa sec W/(m K) N/m J/kg Pa
Water
998 0.001 0.06 4180 0.074 2.54.106 2340
20oC
Diesel
840 0.003 0.11 1800 0.028 3.8.106 300
fuel
Chapter 2.3. Literature survey. Experimental observations 32

2.3.4.1.1. Geometry scale effects


The following review concerns the effects of nozzle shape on the pattern of
cavitation flow, discharge coefficient of nozzle, and the spray formation according to
experimental observations. In particular, the effects of the nozzle length, shape of the
nozzle entrance, conical shape of nozzle (K-factor) and imperfections/ roughness of the
walls, are described.

Flow asymmetry
In real VCO injectors, the flow in each nozzle hole is substantially
asymmetrical. This results in the formation of flow separation and a cavitation region
over a part of total perimeter of the nozzle.

Inlet flow separation

Exit flow separation

Fig. 2- 15. Schematic of the flow in asymmetrical nozzle.

Schmidt, et al. (1999) have visualised cavitation in as asymmetrical two-


dimensional model of a real-scale injection nozzle (Fig. 2-15 and Fig. 1-2, a). They
have observed cavitation onset at the inlet corner of the nozzle. In contrast to
axisymmetrical nozzles, cavitation flow in asymmetrical nozzle was accompanied by
separation of the flow at nozzle exit (Fig. 2-15). This phenomenon has been found to be
dependent on the cavitation and Reynolds numbers (Fig. 2-16).
Experimental studies of sprays developed from VCO multi-hole injectors
(Soteriou, et al., 1995) have concluded that at high cavitation numbers total hydraulic
flip, which is accompanied by the jet contraction, never occurs. The term partial
hydraulic flip has been introduced by Soteriou, et al. (1995) for flows in asymmetric
nozzles when the jet surface become smooth on one side due to the flow detachment
inside the nozzle.
Chapter 2.3. Literature survey. Experimental observations 33

1 / CN
Fig. 2- 16. The flow map for cavitation flow regimes in asymmetrical two-
dimensional nozzle (Schmidt, et al., 1999).

Shape of the nozzle entry


D Rinlet Dbore bev
Dinlet
Lbore

a) b) c) d)
Fig. 2- 17. Nozzles of different configurations of the entry shape, applied in
study by Yuan et al. (2001). (a) sharp entry nozzle; (b) rounded entry nozzle; (c)
counter-bore nozzle; (d) bevelled entry nozzle.

Roundness of the nozzle entry (Fig. 2-17b) affects the recirculation flow formed at
the nozzle throat and the discharge coefficient of nozzle (2.17). Rounded nozzle entry
reduces the length of separation region and suppresses the turbulent motion in the flow.
Many experiments on cavitation flows in nozzles (Bergwerk, 1959; Spikes and
Pennington, 1959; Nurick, 1976; Hiroyasu, 2000; Laoonual, et al., 2001; Winklhofer, et
al., 2001; Konig and Blessing, 2003; Benajes, et al, 2004) have resulted in conclusion
about the delay (increase) in cavitation numbers for the inception and transition to
Chapter 2.3. Literature survey. Experimental observations 34

super-cavitation and hydraulic flip flows ( CN inc , CN sup er and CN flip ) caused by the

rounded shape of the nozzle entry. This is similar to the effect of bevelled entry on
cavitation flow. The results of experimental studies by Laoonual, et al. (2001) (Fig. 2-
18) revealed that sharper bevel angle require higher cavitation numbers to form super-
cavitation in a nozzle.

10
L/D = 10
9
8 bev
7
6
Dinlet
CN

5
4
3 CN_inc

2 CN_super

1
0
30 60 90 120 150 180
Angle, deg.

Fig. 2- 18. Effect of the bevel angel on the cavitation inception ( CN inc ) and

transition to super-cavitation flow ( CN super ). Results of experiments by Laoonual, et al.

(2001).

Nozzle length
The nozzle length is a geometrical parameter, whose effects on cavitation has been
studied for a long time. The nozzle length determines the frictional losses in the region
downstream of the flow reattachment (Fig. 2-6) and therefore affects the discharge
coefficient of a nozzle. For developed turbulent cavitation-free flow, longer nozzles
have higher hydraulic resistance (Eq. 2.17). Under turbulent flow in sharp-entry circular
nozzles of length L / D < 10 (which are typical for Diesel injectors), the nozzle length
has virtually no effect on the cavitation inception number (Lichtarowicz and Pearce,
1974 (Fig. 2-19a).
Chapter 2.3. Literature survey. Experimental observations 35

a) b)
Fig. 2- 19. Effect of the nozzle shape and nozzle length on cavitation inception
(a) and transition to super-cavitation flow and hydraulic flip (b). Bergwerk (1959),
Lichtarowicz, et al. (1974), Fox and Stark (1989) plain (sharp-entry) nozzles (Fig. 2-
17, a); Laoonual, et al. (2001) plain, counter-bore and bevelled nozzles (Fig. 2-17,
a,c,d).

In circular plain nozzles (Fig. 2-17, a) Laoonual, et al. (2001) and Sato and Saito
(2001) have observed no effect of nozzle length on transition to supercavitation. At the
same time, this effect arises for the nozzles with bevelled entry (Fig. 2-17, d) (Laoonual,
et al., 2001).
Bergwerk (1959), Nurick (1976) and later Fox and Stark (1989) have concluded
that longer nozzles require higher cavitation numbers to initiate the hydraulic flip (Fig.
2-19, b). The difference in particular numbers CN flip may be explained by specifics in

the flow geometry (and the shape of an upstream manifold), viscous effects and
imperfections in the nozzle shape (inlet contour and wall roughness).
Fig. 2-20 illustrates variations in the discharge coefficient of a nozzle under
transition to the hydraulic flip regime caused by small increase in the cavitation number
CN (Fox and Stark, 1989). Open dots shows that under the pre-flip flow conditions the
hydraulic resistance of nozzles with the length L / D > 6 , is higher than resistance of
relatively short nozzles ( L / D 6 ). This can be explained by changes in the flow
structure when friction losses make a significant contribution to the pressure drop. For
all the lengths, experiments revealed an increase in the hydraulic resistance as a result of
Chapter 2.3. Literature survey. Experimental observations 36

transition from super-cavitation flow (pre-flip) to flipped flow conditions (post-flip).


The strongest effect of transition on the discharge coefficient is observed for nozzles of
L/D 47.

Fig. 2- 20. Effect of nozzle length on variation in hydraulic resistance of a nozzle


caused by transition to the hydraulic flip. Numbers above the points indicate nozzle
length to diameter ratio L/D.

As a result of their studies Fox and Stark (1989) have proposed design
recommendations, when cavitation in a nozzle is not desirable due to its effect on the
hydraulic resistance and unstable performance of the flow. Thus, at moderate pressure
drops when the cavitation number does not exceed the critical value CN flip , high values

of the discharge coefficient can be obtained for relatively short nozzles of length
L / D < 6 . At high pressure drops when transition to flipped flow may happen, long
nozzles with L / D > 14 should be used to reduce the effect of flow instabilities on the
nozzle performance. In real Diesel injectors the length of the nozzle holes L / D is
around 5. This is considered as an optimum to achieve moderate hydraulic resistance to
the flow and avoid hydraulic flip, which is likely to happen in short nozzles.
Chapter 2.3. Literature survey. Experimental observations 37

250
transition from
super-cavitation
developing to the total imperfect
200 cavitation hydraulic flip hydraulic flip
Time delay (ms)

150

100

50

0
1 1.5 2 2.5 3 3.5
CN

L/D=5 L / D = 10

Fig. 2- 21. Time delay before establishing the total hydraulic flip in a plain
nozzle (analysis of the experimental data from Laoonual, et al., 2001).

An important observation about the effect of the nozzle length on transient


behaviour and establishment of steady-state cavitation flow pattern was made by
Laoonual, et al. (2001). They have found that total hydraulic flip occurs at moderate
cavitation numbers and requires a certain establishing time. This time delay is bigger for
longer nozzles and decreases with the cavitation number (Fig. 2-21). At high cavitation
numbers (CN>3 at Fig. 2-21) total hydraulic flip becomes unstable and imperfect
hydraulic flip happens.
Chapter 2.3. Literature survey. Experimental observations 38

Conical shape of nozzle (K-factor)

Dinlet

3
Rinlet
Doutlet

Fig. 2- 22. Nozzle cap with a spray hole of conical shape and rounded inlet
corner.

The effect of the conical shape of nozzle on cavitation can be described in terms
of the K-factor (Blessing and Konig, 2003; Benajes, et al., 2004). The conical shape
factor (K-factor) can be defined as (Fig. 2-22):

Dinlet Doutlet
K= 100% .
Dinlet

Fig. 2- 23. Effect of the conical shape of nozzle on the extent of cavitation region
and pressure distribution (Winklhofer, et al., 2001). For the nozzles J, U and W the K-
factors are 0 %, 5% and 10%, respectively.
Chapter 2.3. Literature survey. Experimental observations 39

0.90
Choking (2.18):
C d 0.67 1 + 1 / CN
0.85
Nozzle J
0.80
Nozzle U
Cd

0.75

0.70 Nozzle W

Cavitation
0.65
inception
boundary
0.60
0.00 0.50 1.00 1.50 2.00
1 / CN

Fig. 2- 24. Discharge coefficients of three nozzles of different shape, shown in


Fig. 2-23.

The nozzle contraction (K>0) affects the discharge coefficient of cavitation-free


flow in a nozzle. Stronger nozzle contraction tends to reduce the thickness of boundary
layer and thus provides higher hydraulic resistance and lower discharge coefficients
(Fig. 2-24). Winklhofer, et al. (2001) have observed shortening of the length of
separation region by the nozzle contraction (Fig. 2-23). This has damping effect on
cavitation inception (Winklhofer, et al., 2001; Benajes, et al., 2004) and delays the
transition to super-cavitation (Fig. 2-24).
Benajes, et al. (2004) have reported an increase in the discharge coefficient of a
nozzle due to nozzle contraction (K>0). This can be explained by the damping of the
turbulent motion in the flow with favourable pressure gradients that may cause the
delay in development of the turbulent boundary layer.
According to the observations by Winklhofer, et al. (2001), the K-factor has no
apparent effect on the variation of the discharge coefficient with the cavitation number
when the flow is choked. The nozzle shape only affects the hydraulic resistance of
nozzle under liquid flow and determines the inception and critical flow conditions (Fig.
2-24).
Chapter 2.3. Literature survey. Experimental observations 40

K < 0 ; sharp entry K = 0 ; sharp entry K > 0 ; rounded entry

Fig. 2- 25. Effect of nozzle shape on the quasi-steady-state cavitation flow in


real-scale nozzle of diameter ~0.2 mm (Blessing and Konig, 2003). Injection pressure
p1 = 800 bar, back pressure p 2 = 1 bar.

Blessing and Konig (2003) have observed in their experiments that the diffuser-
type nozzle ( K < 0 ) produced more cavitation bubbles than the straight nozzle ( K = 0 ),
which resulted in increase in the spray cone angle from 10o to 12.5o (Fig. 2-25). In
comparison to the sharp-entry nozzles, the nozzle contraction accompanied with the
rounded inlet corners ( K > 0 ) completely suppressed cavitation and caused decrease in
the spray angle down to 5.5o.

Wall roughness
Wall roughness results in higher shear stresses in the liquid near the wall and
produces additional disturbance to the velocity. Higher velocities are accompanied by
lower pressures. Therefore, the wall roughness arranged locally in the region of pressure
reduction makes easier the cavitation onset (Lecoffre, 1999). This has been
experimentally observed by Winklhofer, et al. (2001), who found that rough surface of
the nozzle wall decreases the cavitation inception number CN inc . Similar conclusion

regarding imperfections in the nozzle shape has been made by Chaves, et al. (1995).
Soteriou, et al. (1995) have introduced the term imperfect hydraulic flip for the nozzles
with geometry imperfections (rough walls) to distinguish it from the total hydraulic flip
in nozzles with perfectly smooth walls.

2.3.4.1.2. Viscous stress scale effects


The tensile strength of a flowing liquid pv p cr depends on the liquid viscosity
and the rate of strain in the flow (2.13).
To clarify the effect of viscous stress on the onset of cavitation, it would be
helpful to compare the cavitation formation for two flows characterised by the same
Chapter 2.3. Literature survey. Experimental observations 41

strain rates, but different viscosity. Analysis of the available experimental data on
cavitation flows reveals that the data collected by Roosen, et al. (1996) and Winklhofer,
et al. (2001) could be suitable for this purpose. These authors have reported
measurements for nozzles of nearly identical shapes and have shown different threshold
pressures for similar cavitation flows with different working liquids (water and Diesel
fuel). The aim of the following is to clarify the effect of viscous stress on cavitation
flow.

Table 2- 2. Developing and super-cavitation flows in rectangular models of


Diesel injectors. Comparison of the measurements by Roosen, et al. (1996) and
Winklhofer, et al. (2001).

p1 p2 p1 p2
Flow Authors Liquid Images of the flow CN Re
(bar) (bar) (bar)

Roosen, et
Water 80 21 59 2.81 ~30000
Inlet cavitation

al, (1996)

Winklhofer, Diesel
100 43 57 1.35 ~8000
et al. (2001) fuel

Roosen, et
Water 80 11 69 6.27 ~30000
Super-cavitation

al, (1996)

Winklhofer, Diesel
100 34 66 1.95 ~8000
et al. (2001) fuel

Firstly, consideration the images of cavitation flows reported by Roosen, et al.


(1996) and Winklhofer, et al. (2001) (Table 2-2) reveals very similar cavitation vapour
structures. Thus, under pressure difference around 60 bar both authors have observed a
vapour pocket at the nozzle entry (inlet cavitation). When the pressure drop is about 69
bar, Roosen, et al. (1996) have observed super-cavitation flow, and Winklhofer, et al.
(2001) have reported critical cavitation with the transition to super-cavitation flow at 66
bar. However, the system pressures (downstream of the nozzle exit) were completely
different in these studies (11 and 21 bar in study by Roosen, et al., and for 34 to 43 bar
by Winklhofer, et al.). Because the flow conditions in experiments performed by
Chapter 2.3. Literature survey. Experimental observations 42

Roosen, et al. (1996) and Winklhofer, et al. (2001) were not identical in these two sets
of experiments, it is interesting to consider them in details.
Both studies used rectangular shape nozzles, formed by steel lamellas placed in
between the two pieces of glass. Though the nozzle manifolds had different shapes (Fig.
5-1 and Fig. 5-13), their heights, which determine the hydraulic resistance to the flow,
which is effectively two-dimensional, were very similar (0.28 mm for Roosen, et al.
(1996) and 0.30 mm for Winklhofer, et al. (2001)), and the nozzle lengths were exactly
the same (1mm). Also, the roundness of the nozzle entry, which results from the
imperfections in manufacturing of nozzle, was estimated in both cases of the same order
of magnitude (30 m in Roosens and 20 m in Winklhofers study).
Because the nozzle heights and lengths were very similar and the flow is turbulent
in both cases, one may assume similar hydrodynamic flow fields. The difference in the
flow may result from inequality in the width of nozzles (0.200 mm for Roosens nozzle,
and 0.3 mm for Winklhofer) and properties of the working liquid, which was Diesel
fuel in Winklhofers study and tap water in Roosens case.

p (bar)

100 Winklhofers nozzle


100 bar : 43 bar
80
Roosens nozzle
60
zW 80 bar : 21 bar
40

20

0
Z
-20

-40 z R

Fig. 2- 26. Sketch of the pressure distribution along Roosens and Winklhohers
nozzles based on the numerical solutions for three-dimensional liquid flows.

Assuming similarity of the flow patterns in Roosens and Winklhofers nozzles


(the same pressure distributions, velocity fields and strain rates), the maximum pressure
drop at the nozzle entry becomes a function of the outlet pressure (under given pressure
Chapter 2.3. Literature survey. Experimental observations 43

difference across the nozzle) (Fig. 2-26). Because the system pressure (at the nozzle
outlet) is higher for Winklhofers nozzle, the volume of liquid subjected to negative
pressures is smaller than the Roosens nozzle zW < z R (Fig. 2-26). From this one
would expect a lower rate of cavitation in Winklhofer experiments. The fact that the
cavitation structures observed solely depend upon the pressure difference p1 p 2
(Table 2-2) demonstrates that the cavitation number CN is not a sufficient criterion to
describe the development of cavitation. Assuming the liquid quality and the pressure
and velocity fields are similar for both flows, the difference in pressure levels required
to produce similar cavitation structures can be attributed to the viscous effects.
While the density of Diesel fuel is of the order of magnitude of the density of
water, the dynamic viscosity of Diesel fuel is nearly three times bigger than the
viscosity of water (Table 2-1). This causes stresses that are three times higher in the
flow of Diesel fuel compared to water flow at the same strain rates. Thus, the similarity
of cavitation structures observed for these flows can be explained by the effect of the
stress in the liquid on the critical pressure threshold. According to the hypothesis of
Joseph (1995), the cavitation threshold is defined for the principal component of viscous
shear stress in the flowing liquid (2.13). In flows of high-viscosity liquids, the shear
stress can decrease the tensile strength of the liquid (increase the critical pressure
threshold), and therefore cavitation inception may start develop at higher system
pressures.

2.3.4.1.3. Turbulence scale effects


From experiments it is known that the flow behaviour downstream of the nozzle
entrance is strongly affected by the inlet turbulence (Soteriou, et al, 1995; Badock, et al,
1999; Hiroyasu, 2000). Here, experimental observations of the effect of inlet turbulence
on the performance of cavitation flow in nozzle, are described. The issue of the
interaction of cavitation bubbles with the flow turbulence will also be considered later
in analysis of the bubble dynamics.
Many authors anticipate that collapsing cavitation bubbles enhances the
turbulent motion inside the jet improving the spray break-up (Hiroyasu, 2000;
Alajbegovic, et al., 1999). Hiroyasu (2000) has reported the effect of the inlet
turbulence generated by a wire net placed upstream the nozzle on the spray
Chapter 2.3. Literature survey. Experimental observations 44

disintegration. Experiments have revealed that the inlet turbulence enhanced


disintegration of the liquid spray and prevented transition to hydraulic flip flow.
To examine the effect of inlet turbulence on cavitation flow, Sato and Saito
(2001) have installed a trip wire at the plenum wall around the nozzle entrance. Using
high-speed photography, the authors have detected a higher frequency of shedding in
the cavitation region at the transition cavitation stage in a nozzle equipped with the
wire. This was explained by a decrease in the length of separation region and shortening
of the vortex formation time, as a result of augmentation of inlet turbulence. The shorter
separation region resulted in higher cavitation numbers, for the transition to super-
cavitation.
Soteriou, et al. (2001) have speculated that at lower levels of turbulence, the
larger cavitation bubbles may form, while higher turbulent pulsations disrupts the
cavitation pockets into small bubbles.

D
Liquid

Vapour
Nozzle
wall

Fig. 2- 27. Cross-sectional cut of a circular nozzle in the vena contracta region
(Fig. 2-6).

2.3.4.1.4. Surface tension effects


Several experiments on cavitation in nozzles (Tullis, 1979; Lichtarovicz, et al,
1974; Sato and Saito, 1995) have have been conducted to study the effect of diameter of
nozzle and surface tension of liquid on cavitation flow. These effects on the structure of
cavitation flow have not been clearly understood.
Analysis of dimensions gives the criterion to account for the surface tension
effect on the liquid-vapour flow, which is known as the Weber number. One may
introduce the turbulent Weber number:
Chapter 2.3. Literature survey. Experimental observations 45

l kD
We t =

The higher the Weber number, the less stable is the vapour-liquid interface. Thus,
higher surface tension stabilises the jet flow formed in the vena contracta region
(Fig. 2-27). At the same time a decrease in the turbulence intensity and nozzle diameter
reduces We t , and may cause disintegration of the vapour structures formed at the

nozzle entry. The turbulent kinetic energy of the cavitating liquid k in definition of
We t depends on the turbulence intensity upstream the nozzle entry, and can be

suppressed by the flow acceleration (the effect of favourable pressure gradient) in the
vena contracta.

2.3.4.1.5. Transient effects


Cavitation as the process of growth and collapse of bubbles in a liquid is a
transient phenomenon. Experimental studies of cavitation flows in single-hole straight
nozzles (Sato and Saito, 2001; Schmidt, et al., 1999; Roosen, 1997; Laoonual, et al.,
2001) and Venturi nozzle (Stutz and Reboud, 1997; Sato and Shimojo, 2003) have
emphasized the transient nature of the cavitation phenomenon. Due to this transient
nature, cavitation flows generate noise and vibration. Understanding of cavitation
instabilities would help to design an adequate model of the phenomenon and control the
cavitation flow.

(a) (b)
Fig. 2- 28. Stream lines for the steady-state re-circulation flow (a) and transient
re-entrant jet motion (b) in the separation flow formed at the nozzle entrance.

The intrinsic instability of partial cavitation flows originates from the instability
of re-entrant jet motion in the flow separation region (Franc, 2001), Fig. 2-28b. This
instability causes periodical oscillations of the flow at a certain frequency, which may
Chapter 2.3. Literature survey. Experimental observations 46

be described in terms of Strouhal number as the ratio of the time needed for the re-
entrant jet to reach the cavity leading edge, and time period of cavity oscillations:
Str = F Lcav / u reentr (Franc, 2001). Studies of cavitation-free flow have confirmed
pulsations of the flow in the separation region, and periodical variations in the discharge
coefficient of a nozzle (Su and Schmidt (1995) and also discussion by Hall to the papers
by Bergwerk (1959) and Spikes and Pennington (1959)). Under cavitation flow, the re-
entrant flow instability produces quasi-periodical shedding of the cavitation region
(Sato and Saito, 2001, Fig. 2-12).

pv
p2

Fig. 2- 29. Gas entrainment in closure of the vapour cavity at the nozzle exit.

Instability of super-cavitation flow


When the cavitation region reaches the nozzle exit (super-cavitation flow), the
downstream pressure may propagate upstream the flow through the cavitation pocket,
causing its rapid collapse. High-speed photography of super-cavitation flows reveals the
stochastic character of these oscillations (Sato and Saito, 2001; Arai, et al., 1985).

2.3.4.1.6. Compressibility effects


The speed of sound in the liquid-vapour fluid depends on the local structure of
the flow. Under cavitation conditions, the system pressure (downstream the nozzle) is
usually much higher than the vapour pressure. In this sense the cavitation is different
from flashing, or flash boiling, which happens when the liquid is superheated at the
nozzle exit (Chisholm, 1983). Similar phenomena can occur in smooth Venturi nozzles
when the bulk liquid in the whole nozzle cross-section is subject to a pressure below the
saturation pressure, which results in a nearly homogeneous bubbly structure of the flow.
In the case of cavitation, the liquid core is usually present in the flow because of the
local reduction in pressure. High pressure and density gradients occur in cavitation
flow, thus the flow locally achieves the properties of a compressible medium.
Chapter 2.3. Literature survey. Experimental observations 47

Measurements have revealed a decrease in the speed of sound with the void
fraction in bubbly mixtures (Brennen, 1995). However, the pressure propagation in real
cavitation flows differs from the case of a static bubbly mixture because of the effect of
the phase transition (growth and collapse of bubbles) on the void fraction and, thus the
speed of sound in the mixture (Biesheuvel, and Wijngaarden, 1984).

2.3.4.2. Liquid quality effects


In practice the cavitation susceptibility of a liquid depends on its molecular
thermo-physical properties and the presence of cavitation nuclei (small gas-filled
bubbles and contaminating particles). The following section reviews the influence of
liquid quality effects on cavitation flow associated with the presence of cavitation
nuclei.

CN

3
delayed
cavitation
delayed
aeration
2 Super-cavitation
cavitation inception
cavitation
cavitation flow
desinence
1 aeration inception:
aeration
aeration 5% air
desinence
liquid flow 100% air
0
p 2 (bar )
0 10 20

Fig. 2- 30. Diagram of the flow regimes in an orifice L / D 2.5 (Lichtarowicz,


et al., 1974)

Lichtarowizc, et al. (1974) have studied cavitation flows in liquids of different


air content. Because of the state of experimental techniques at that time, only the total
air content has been measured and no difference between the dissolved and non-
dissolved air was made. In experiments the total air concentration has been varied from
Chapter 2.3. Literature survey. Experimental observations 48

5% to 200% of normal saturation content. In the following, the main results of this
study (Lichtarowicz, et al., 1974) are described.
Experiments by Lichtarowicz, et al. (1974) have revealed that due to the
presence of gases in liquid, aeration, or gaseous cavitation happens before the true
vapour cavitation.
The aeration inception number varies with the pressure, which affects the
cavitation nuclei. When a liquid, which contains gas-filled bubble nuclei, is subjected to
a high pressure, the gas will dissolve in the liquid according to Henry law, so that the
scale and number of nucleation bubbles will be reduced. Consequently the probability
of cavitation will be reduced as well.
When aeration precedes cavitation, cavitation inception happens at unique
cavitation number, which depends on the nozzle geometry. It was anticipated that
higher pressures and lower gas content would affect the cavitation nuclei and make the
cavitation inception more regular.
At high outlet pressures ( p 2 > 20 bar ) inception and desinence coincide. The
authors suggested that the range where inception and desinence coincide could be
affected by the pressure.

Cd

delayed
CN sup er CN inc CN

Fig. 2- 31. Delayed transition to the super-cavitation flow.

At lower pressures ( p 2 < 20 bar ), cavitation and aeration inception may start at
higher CN with a time delay (Fig. 2-30). However, there is no delay in cavitation when
aeration inception occurs because the gas-filled bubbles perform as active cavitation
nuclei.
delayed
Without aeration the cavitation inception can start at CN inc > CN sup er

accompanied by a sudden decrease in C d (Fig. 2-31). The authors concluded that upper
Chapter 2.3. Literature survey. Experimental observations 49

delayed
and lower limits for the cavitation inception ( CN inc and CN des ) depend on the orifice

geometry and sizes.


The pressure effect on cavitation inception and development has been found to
be similar for all the air contents studied. Aeration almost disappears at an air content of
5% (Fig. 2-30). The cavitation inception number CN inc decreases with the increase in

the air content, which can be explained by the effect of dissolved gas on the saturation
pressure.
At high Reynolds numbers ( Re > 2 10 4 ) desinence has been found to be
unaffected by the pressure level, velocity of the flow, size of nozzle, air content and
fluid properties. Therefore for the desinence no scale effects have been established.
Both cavitation and aeration desinence are regular, and depend on the nozzle length
L / D for cavitation and also on the air content for aeration.
The delay in cavitation and aeration may be explained by the effect of pressure
on the number of active bubble nuclei. At higher pressure there are more bubbles which
perform as active cavitation nuclei and make cavitation inception occur earlier. At high
pressures the limit for the inception number may be explained by the decrease in the
number density of bubble nuclei.

2.3.5. Conclusions

The reviewed experimental data on the hydrodynamic cavitation in nozzles and


Diesel injectors shows that cavitation is a transient, three-dimensional phenomenon,
which affects the jet atomization and therefore should be considered when modelling
the flow in Diesel injector.

The studies have been performed using the natural-scale and large-scale models of
Diesel injectors, and their simplified models.
Various experimental methods and techniques have been applied in studies of
cavitation flows. They have resulted in:
visualisations of the two-phase instantaneous and time-averaged patterns
of cavitation flow;
measurements of the discharge coefficients of cavitating nozzles;
Chapter 2.3. Literature survey. Experimental observations 50

measurements of the pressure fields and also the local averaged and
turbulent components of the velocity field.
The local fields of the volume fraction of the vapour and amount of vapour
generated in cavitation flow have not been studied for true vapour cavitation in high-
speed flows. The only void-fraction measurements published come from the study of
cloud cavitation in a large-scale transparent Venturi nozzle by a French group (Stutz
and Reboud, 1997).
Visual observations of cavitation flows have revealed the flow patterns of
cavitating flows in single-hole injection nozzles. The bubbly structure of cavitation flow
has been revealed in experiments on cavitation inception. At the same time, developed
cavitation flows may consist of opaque bubble clouds (bubbly mixture) and clearly-
defined transparent vapour pockets (separated flow pattern). The transient nature of
cavitation flows has been observed for the transitional cavitation with re-entrant jet
instability and super-cavitation flows with stochastic oscillations of cavitation at the
nozzle exit. Experimental studies of cavitation in real-size and large-scale Diesel
injectors (of multi-hole configuration and, possibly, unsteady flow conditions), have
revealed that the flow structure could be rather complex, when several types and stages
of cavitation can be found in the same flow.
Arcoumanis, et al. (2000, 2001) have matched the similarity conditions to study
the hydrodynamic scaling of cavitation flows in Diesel injectors of different scales.
Visualisations of the flows in real-size and magnified-scale injectors have revealed
similar macroscopic steady-state cavitation structures (Arcoumanis, et al., 2000). At the
same time, the structure of two-phase flow in nozzles of different scales has been found
different in detail (Arcoumanis, et al., 2000). Similarity of the macro-scale cavitation
structures can be explained by the dominant effect of the hydrodynamics of mean flow
on the development of cavitation, while the dissimilarity of micro-scale structures can
be referred to the difference in the residence and life time of bubbles in large and small-
scale nozzles (Dumont, et al., 2000).
Generally, deviations from the scaling laws can be explained by the presence of
scale effects (Stinebring, et al., 2001; Keller and Huber, 2001), which make difficult to
match all the similarity conditions in a physical experiment. Within the numerical
model of cavitation flow, which permits the hydrodynamic scaling, the scale effects can
accurately be taken into account, but sub-models for these scale effects require
validation with experimental data.
Chapter 2.3. Literature survey. Experimental observations 51

Depending on their effect on the flow and behaviour of cavitation structures, the
following scale effects may be identified: flow effects (viscous and geometry effects)
and bubble-dynamics effects (including the liquid quality, or nuclei effects).
The flow effects, associated with dimensionless parameters of the flow (Reynolds
number, Strouhal number, cavitation number, etc.) geometry of nozzle, and turbulent
motion, determine the flow pattern of cavitation-free flow (pressure field, flow
separation, etc.), and also define the cavitation onset and its development. Deviation
from the hydrodynamic scaling of cavitation flows in large- and small scale injection
nozzles is caused by the difference in the micro-geometry of nozzles as a result of
imperfections in manufacturing of small-scale injectors. The actual shape of nozzle
entry and roughness of nozzle walls may become crucial for reliable prediction of the
liquid flow and cavitation behaviour in small-scale nozzles.
Comparison of the experimental data on high-speed cavitation flows of a tap water
(Roosen, et al., 1996) and Diesel fuel (Winklhofer, et al., 2001) reveals that Diesel fuel
starts cavitating at higher system pressures. This observation can be described by the
effect of viscous stress on the critical tension in liquid. This hypothesis, originally
suggested by Daniel Joseph (1995), has never been applied in computer simulations of
cavitation flows, and its validation using the measurements by Roosen, et al. (1996) and
Winklhofer, et al. (2001) is of the subjects of the present study.
Other scale effects on cavitation flows are associated with the liquid quality,
determined by the presence of cavitation nuclei in the liquid. Specification of
parameters of bubble nuclei requires special measurements, which usually do not
accompany cavitation experiments. Consideration of cavitation nuclei present in real
liquids is crucial for the accurate prediction of cavitation inception. Reliable prediction
of cavitation flows also requires modelling of the thermo-physical properties of
cavitation fluid and consideration of the bubble-dynamics scale effects, which determine
the specifics of growth and collapse of the vapour pockets and interaction of cavitation
with the turbulent motion.
Chapter 2.4. Literature survey. Models of cavitation 52

2.4. Models of cavitation flows


2.4.1. Introduction

Measurements of cavitation flows in diesel injectors are difficult because of the


small diameters of the nozzle holes and very short duration of the injection process.
Numerical modelling offers an alternative means to investigate this phenomenon, and
also permits spray break-up and atomization to be predicted.
Experimental studies have revealed the scale effects on the formation and
development of cavitation flows. The flow scale effects and the liquid quality scale
effects were identified from the experimental observations of cavitation flows, as
described in the previous chapter. These scale effects have to be considered when
developing the numerical model of the flow.
Recently, progress has been made in the development of numerical models for
calculation of cavitation flows. Though the models may differ in terms of realisation
(using the single-fluid or multi-fluid frame-work, the Eulerian- Eulerian or Eulerian-
Lagrangian approaches), all of them are empirical to a certain level and contain scalable
parameters. The aim of the present review is to demonstrate that these parameters
characterise the cavitation nuclei, i.e. they are responsible for the liquid quality effects.
Correct description of the liquid quality effects is essential when describing the
hydrodynamic scaling of cavitation flows.
This chapter reviews the physical aspects in the modelling of the cavitation
phenomenon. Issues of turbulence modelling in cavitating flows and numerical aspects
of simulation of cavitation are discussed in the subsequent chapters.

2.4.2. Models of two-phase flows


There are two main approaches used to model multi-phase flows: the concept of
interpenetrating continua and non-interpenetrating continua.
To describe flows with a clearly-defined interface between vapour and liquid, the
concept of non-interpenetrating continua have been developed and several techniques
designed for numerical calculations (interface tracking; front-capturing; level-set;
bubble-tracking). However, these methods have not been widely adopted for calculation
of three-dimensional transient cavitation structures in flows of arbitrary geometry.
Chapter 2.4. Literature survey. Models of cavitation 53

Many cavitation models developed in literature are based on the concept of


interpenetrating continua. This concept employs the concept of the average volume
fraction of phases present at a spatial-time location in the flow. Most of the cavitation
models assume thermal and dynamic equilibrium between the liquid and vapour phases.
To achieve the specifics of cavitation flow, caused by the presence of the liquid,
vapour and gas phases, models of cavitation have been developed employing the single-
fluid and multi-fluid (Alajbegovic, et al., 1997, 1999) continuum approaches.
Multi-fluid models implies that the phases, present at the same spatial-time
location in the flow, can be characterised by different velocities and temperatures, i.e.
they are not in equilibrium. This concept is usually applied for simulation of separated
flows and non-equilibrium dispersed flows with a clearly-defined flow structure.
However, validation of the closure relationships for the multi-fluid models of cavitation
is difficult because of the lack of experimental information about the phenomena.
An alternative concept in two-phase modelling is the single-fluid approach, when
the set of conservation laws is applied to only one fluid (primary phase or mixture
of phases). The homogeneous mixture models assume both phases are uniformly mixed
together and no clear two-phase structures (flow patterns) and inter-phase boundaries
can be identified in the flow. Calculations with these models can be performed using the
Eulerian approach. Then the essence of a model is the closure relationship for the
volume fraction of the vapour phase. The models, which trace the vapour pockets
within the Lagrangian framework do not use the concept of volume fraction, but require
assumptions to be made about the pattern of the flow.

2.4.3. Homogeneous mixture flow model

Significant progress has been achieved recently in the development of


homogeneous-mixture models for the simulation of three-dimensional transient
cavitating flows (Chen and Heister, 1995, 1996; Kunz, et al., 2001; Ahuja, et al., 2000;
Yuan et al., 1999; Singhal, et al., 2002; Kubota, et al., 1990). These models allow
single-fluid solvers to be applied to the conservation equations for the mixture, without
increase in computational cost due to the increase in the number of conservation
equations when applying the multi-fluid flow concept.
In the present study, the homogeneous mixture modelling approach was chosen as
the most robust and well developed in the literature for numerical simulation of
Chapter 2.4. Literature survey. Models of cavitation 54

cavitation flows. Liquid and vapour phases are assumed to be in thermal equilibrium
and treated as interpenetrating continua, whose properties are strongly affected by the
pressure field. Effects of surface tension and inter-phase velocity slip are neglected. The
rate of evaporation/ condensation is governed by the pressure difference between liquid
and vapour phases. In the vapour phase, the pressure is assumed to be equal to the
saturation pressure at given temperature.
Physical properties of the mixture are functions of the properties of the pure
phases and volume fraction of phases. The majority of authors apply linear
approximations:

= v + (1 ) l , (2.19)
= v + (1 ) l , (2.20)

where the properties of the pure phases are kept constant (incompressible formulation).
The set of the Reynolds-averaged continuity and momentum equations is as
follows:

u j
+ =0
t x j
, (2.21)
u i u i u j p ij
+ = +
t x j xi x j
, (2.22)
u i u j 2 u k
ij = ( + t ) + ij , (2.23)
x j xi 3 x k

Modelling of the turbulent flow regimes requires application of closure


relationships for the turbulent viscosity t . In the present study turbulent viscosity was
modelled for a mixture as a whole. The effects of cavitation on the turbulent motion are
discussed in the next chapter.
The set of equations (2.21) (2.23) of the continuum single-fluid model of
vapour-liquid flow should be completed by the constitutive relationships of the
cavitation model.
Chapter 2.4. Literature survey. Models of cavitation 55

2.4.4. Models of cavitation

In the last decade a significant progress has been achieved in the development of
single-fluid models of cavitation flows (Dumont, et al, 2000). Most of the cavitation
models imply the phase transition is driven by local pressure. Classification of the
models of cavitation can be made depending on the type of equations they use to
calculate the phase content in the flow, namely algebraic or differential.
Algebraic models of cavitation assume an instantaneous effect of local pressure on
the density of the homogeneous mixture. Therefore they are also known as the
barotropic equilibrium models, or equation of state (EOS) models. These models have
been applied by Dellanoy and Kueney (1990); Avva, et al. (1995); Schmidt, et al.
(1997); Dumont, Simonin and Habchi (2001), and others.

Table 2- 3. Features of models of cavitation flows.

Models based on Models based on


Equation of state
Phenomenon the transport the Rayleigh-
(EOS) models
equation Plesset equation

Equilibrium phase Non-equilibrium, at Non-equilibrium,


Phase transition
content steady-state rates transient rates
Pressure-density
coupling Barotropic model Baroclinic model Baroclinic model
(compressibility)
Nuclei effects
Not described Can be described Can be described
on cavitation
Neglected, or Described for the Described for the
Turbulent motion described for the mixture + damping mixture, or for the
mixture effect of bubbles liquid

In comparison to the algebraic models, the differential models of cavitation enable


the baroclinic nature of cavitation flows to be described. The most simple approach is
based on the evaluation of the void fraction from the transport equation with the source
term governed by the difference between the local pressure in the mixture and the
vapour pressure (Kunz, et al., 2001; Ahuja, et al., 2000; Yuan et al., 2002). This
Chapter 2.4. Literature survey. Models of cavitation 56

approach has been found to be robust and efficient for the prediction of cavitation flows,
and not as computationally expensive as the models, which utilize the second-order
differential equation based on the Rayleigh-Plesset equation of bubble dynamics theory
(Kubota, Kato and Yamaguchi, 1990; Schnerr and Delale, 2001; Heister, et al., 1996).
Table 2-3 compares the features of three widely applied types of cavitation models
(EOS model, transport equation based models, and Rayleigh-Plesset equation based
models).

2.4.4.1. EOS models of cavitation


To close the set of equations (2.19) (2.23), an additional equation for the volume
fraction is needed. Simplifying the energy equation, Schmidt, et al. (1997) obtained
the following barotropic equation:

D Dp
c m2 =
Dt Dt

where c m is the acoustic speed of sound in the mixture of liquid and vapour.
For the homogeneous mixture of two phases, the isothermal speed of sound can be
calculated from the modelling equation (Minnaert, 1933; Wallis, 1969):

1 m 1
= = +
c 2 c 2 , (2.24)
p
m
c m2 v v l l

where m = v + (1 ) l is the density of mixture (2.19).

Integration of this equation produces the pressure-density relationship, which


originates the algebraic models of cavitation (Schmidt et al. 1999; Schmidt and
Corradini, 2001; Qin, et al., 2001; Dumont et al. 2001):

l 1 c
2

p( m ) = pl ,s + v log m m (2.25)
v l ( l cl ) + ( v cv )
2 2
l cl

where pl , s is the pressure level at which the liquid starts turn into the vapour. Densities

and sound velocities of pure liquid and vapour-gas phases are assumed to be constants.
Chapter 2.4. Literature survey. Models of cavitation 57

Under given pressure, equations (2.25) determine the mixture density, and thus, the
volume fraction of the vapour phase .
Despite the advantage of the explicit algebraic pressure-density relationship (2.25)
the EOS models have been developed for simulation of the flows with homogeneous
bubbly structure, while they can not describe the liquid quality effects on cavitation
flow. Studies of cavitation using the EOS models have revealed that they are unable to
describe the baroclinic nature of cavitation flow, when the density variation in the
cavitation region depends on the pressure history upstream the flow. To correctly
describe the baroclinic nature of hydrodynamic cavitation, differential models of
cavitation have been developed .

2.4.4.2. Rates of the phase transition in cavitation flows


A simple differential equation model of cavitation, that describes the effect of the
pressure history on the density variations in a cavitation flow, has been developed by
Chen and Heister (1995):

D
= C ( p pv ) , (2.26)
Dt

where C is an empirical dimensional parameter, which here is referred as the rate


parameter of the cavitation model. The function present on the right-hand side of this
equation determines the rate of the phase transition in a cavitation flow.
While the relaxation-type equation (2.26) is purely heuristic, several models of
cavitation have been developed from theory, resulting in differential equations for the
density of the fluid.
One popular approach uses the Rayleigh equation (Rayleigh, 1917) for the rate of
bubble growth (Yuan, et al. 2001; Singhal et al., 2002):

dR 2 p s (Tl ) pl
=
dt 3 l
, (2.27)

where R is the radius of cavitation bubble, and p s (Tl ) is saturation pressure in the

liquid at given temperature.


Chapter 2.4. Literature survey. Models of cavitation 58

Other models define the rate of phase transition from the kinetic theory of
condensation (Nigmatulin, 1991):

dR v p v p s (Tl )
=
dt l 2Bv T T
v l , (2.28)

where < 1 is the coefficient of condensation, v is correction factor (Akhatov, et al.,

2001) and Bv is the gas constant.

2.4.4.3. Transport-equation based models of cavitation


The transport equation for the volume fraction of the vapour phase can be
presented in a conservative form:

v v u j
+ = S
t x j
. (2.29)

The source term in the void fraction equation (2.29) is a function of the fluid
properties and the phase transition process (evaporation or condensation).
Comparison of transport-equation based cavitation models, can be performed
using the dimensionless form of the transport equation (2.29). Applying the
hydrodynamic spatial, time and velocity scales of the flow l , t and u = l / t ,
equation (2.29) can be non-dimensionalised:

u~k ~
+ ~ = S , (2.30)
~
t x k

~ ~ ~ S
where t = t / t , l = l / l , u~k = u k t / l and S = t .
v
The dimensionless volume fraction transport equation (2.30) is not commonly
applied in calculations. This happens because models of cavitation do not usually permit
hydrodynamic scaling of cavitation flows. Validation of this feature of cavitation
models by comparison with measurements is difficult due to the actual dissimilarity of
cavitation flows caused by the presence of scale effects. However, there is experimental
Chapter 2.4. Literature survey. Models of cavitation 59

evidence of similarity of macroscopic vapour structures in cavitation flows of different


scales (Arcoumanis, et al., 2000). Therefore the potential ability to achieve
hydrodynamic similarity within the transport equation-based models is of interest for
this study.

Chen and Heister (1995)


The cavitation model suggested by Chen and Heister (1995) (2.26) can be
rearranged in the dimensionless form (2.30) with a source term described by the
equation:

~ l2 p p
S = C v 2 , (2.31)
m ( l v ) l u

where pv is the vapour pressure, and C is the empirical dimensionless rate constant, and
u is the mean flow velocity scale.
Chen and Heister (1995) emphasised that a large C results in a rapid variation of
the mixture density with small variations in pressure, i.e. the differential equations
(2.26) and (2.30) become stiff. Adjustment of the rate constant with the experimental
data of Rouse and McNown (1948) on cavitation flow around the cylindrical bodies,
have revealed the dependence of the rate constant C on the Reynolds number of the
mean flow. Application of the model to the simulation of cavitation flow in an orifice
has revealed that the predicted flow becomes transient at high values of parameter C. In
their work, Chen and Heister (1995) conducted transient flow simulations. They found
that for small values of C the flow behaves as steady-state.

Kunz, et al. (2000)


Kunz, et al., (2000) have employed different equations for the evaporation and
condensation rates:

~ p p
S = C ev (1 ) 1 v 2 , when p < p v , (2.32)
2 l u

~
S = C cond (1 ) 2 , when p > p v . (2.33)
Chapter 2.4. Literature survey. Models of cavitation 60

where Cev and Ccond are the rate constants, and u is the free stream (or mean flow)
velocity scale.
The above equations (2.32) and (2.33) have been applied to the simulation of sheet
and super- cavitation in flows around cylindrical bodies (experiments by Rouse and
McNown, 1948), blades of centrifugal pumps, and cavitation in a Venturi section (Stutz
and Reboud, 1997). To match the experimental data, the cavitation parameters have
been adjusted to different numbers. Thus, in the original version of this model, Kunz,
Boger, et al. (2000) have used Cev = Ccond =100 in the calculations of cavitation flows
around cylindrical bodies (Rouse and McNown, 1948). Later, for the same type of flows
Lindau, Venkateswaran, Kunz, and Merkle (2001) have applied Cev = 105 and Ccond =1.
In studies of sheet cavitation in the impeller section of a centrifugal pump Medvitz,
Kunz, et al. (2002) have used Cev = 100 and Ccond =1000. In studies of cavitation flows
around cylindrical bodies and cavitation in a Venturi nozzle, Lindau, Kunz, et al. (2002)
have adjusted the rate constants to Cev =105 and Ccond =200. Thus, one may see that the
model contains rate constants, which are flow-dependent.
The standard high-Reynolds number form of k- turbulence model has been
applied for the turbulence closure.

Ahuja, et al. (2001)


The rates of evaporation and condensation were both approximated as linear
functions of pressure:

~ p p
S = C ev l (1 ) 1 v 2 , when p < p v , (2.34)
v 2 l u

~ p p
S = C cond 1 v 2 , when p > p v , (2.35)
2 l u

where and Cev = Ccond = 103 are empirical rate constants and and u is the velocity scale
of mean flow. The rate constants have been tuned to best fit the experimental data on
steady-state sheet cavitation for flow around cylindrical bodies (Rouse and McNown,
1948) and NACA 66 hydrofoil (Shen and Dimotakis, 1989). The study has not been
finished and the rate parameters were roughly adjusted to achieve only qualitative
comparison with the experiments.
Chapter 2.4. Literature survey. Models of cavitation 61

Turbulence was modelled using the low-Reynolds number version of the k-


turbulence model (So, et al., 1975), applied to the mixture.

Singhal, et al. (2002)


Expressions for the calculation of the source term in the void fraction equation in a
model suggested by Singhal, et al. (2002) can be represented in a form:

k l2 2 pv p
(1 ) , when p < pv,
~
S = u ev t (2.36)
m 3 l

~ k l2 2 p pv
S = u cond t , when p > pv, (2.37)
m 3 l

where k is the turbulence kinetic energy and is surface tension of liquid. The
recommended values for the rate parameters are uev = 0.02 (m/s) and ucond = 0.01 (m/s).
The vapour pressure in (2.36) and (2.37) accounts for the effect of the turbulent
pressure:

0.39
pv = p s + k . (2.38)
2

The model of Singhal, et al. (2002) can be simplified to examine the effect of the
flow time/ velocity scale on the evaporation and condensation rates. Assuming the
linear relationship between the mean flow velocity and turbulence velocity scales:

k 3
2 u Tu ,

the phase transition rates can be expressed in a form:

l2 2 pv p
(1 ) , when p < pv,
~
S = C ev u t
2
(2.39)
m 3 l u 2

~ 2 2 p pv
S = C cond u 2 t l , when p > pv, (2.40)
m 3 l u 2
Chapter 2.4. Literature survey. Models of cavitation 62

where the rate parameters can be estimated as C ev = 0.02 (sec.m2/kg) and

C cond = 0.01 (sec.m2/kg).

The model of Singhal, et al. (2002) contains dimensional parameters and provides
non-zero rates of the phase transition only under turbulent flow in the cavitation region.
The authors anticipated that the rate parameters could be applied to predict a variety of
steady-state regimes of cavitation for both external flows and flows in nozzles.

Yuan, et al. (2001)


In the model of Yuan, et al. (2001) the void fraction equation has been derived
assuming that evaporation and condensation occur in spherical bubbles present in
liquid. The following expression has been obtained for the source term in (2.29):

(1 ) l v n d 4 3
S = R sign( pv p ) , (2.41)
m 1 + n 43 R 3 dt 3

where n is the number density of cavitation bubbles per unit volume of liquid and R is
the bubble radius.
Taking into account equation (2.27) and applying the definition for the volume
fraction of vapour :

n 43 R3
= . (2.42)
1 + n 43 R3

the source term may be rewritten in a form:

l v pv p
S = C cav f ( ) sign( p v p ) (2.43)
m 1
2 l

where C cav is the dimensional cavitation parameter:

1
C cav = = 6 3 43 no (2.44)
l cav
Chapter 2.4. Literature survey. Models of cavitation 63

and f ( ) = (1 ) 4 / 3 2 / 3 is a function of the void fraction (Fig. 2-32).

Fig. 2- 32. The effect of void fraction on the evaporation rate according to the
models of Yuan, et al. (2001) and Alajbegovic (1999).

Applying this model to the calculation of steady cavitation in a small-scale model


of injection nozzle (experimental data by Roosen, et al., 1996) Yuan, et al. (2001) have

adjusted the number density of cavitation bubbles to n = 1.5 1014 (m 3 ) . This number

gives an estimate for the cavitation parameter (2.44) C cav = 2.1 10 5 (m -1 ) .


The dimensionless form of the source term (2.43) is:

~ C pv p
S = l cav l f ( ) sign( p v p ) . (2.45)
2 m 1
2 l u 2

The model assumes the number density of cavitation bubbles is an integral


parameter of the flow n = const . The initial size of cavitation bubbles Ro also is
considered as a parameter in the model. These two parameters are able to describe the
nuclei (liquid quality) effect on cavitation.
Because the liquid quality may vary from one liquid to another, the numbers for
the parameters n and Ro , are not unique. Thus, to fit the experimental data of Roosen
et al. (1996) on cavitation of water in a small-size nozzle, Yuan et al. (2001) have
Chapter 2.4. Literature survey. Models of cavitation 64

applied n = 1.5 1014 (m 3 ) and Ro = 0.3m ( o = 0.0017 % ); in simulations of water

flow over a hydrofoil, Basuki, Schnerr and Yuan (2002) have used n = 108(m3) and
Ro = 3m , while Frobenius et al. (2003) have applied Ro = 30 m .
For simulation of the turbulent flows Yuan et al. (2001) have applied the
modification of the Wilcox k- turbulence model with the wall functions.

Alajbegovic et al. (1999)


To account for the interactions between phases in cavitation flow, the multi-fluid
approach has been applied. The source term for the void fraction equation (2.29) was
calculated from the correlation:

dR
S ,ev = v n B 4R 2 , (2.46)
dt

where n B is the volumetric number density of cavitation bubbles in liquid, and R is the
bubble radius related to the volume fraction of the vapour phase:

n B 43 R 3 . (2.47)

The rate of bubble growth/ collapse was calculated according to the Rayleigh
model (2.27). The number density of cavitation bubbles has been modelled taking into
account the coalescence of bubbles at large void fractions:

no < 0.5
nB = (2.48)
1 + 2 (no 1)(1 ) > 0.5

where no is the number density of cavitation nuclei initially present in the liquid (no =
1012 nuclei/m3, according to Fujimoto et al., 1994).
Substituting (2.47) and (2.48) into (2.46) the source term may be expressed in
terms of the volume fraction:
Chapter 2.4. Literature survey. Models of cavitation 65

~ pv p
S ,ev = l C ev f A ( ) , (2.49)
l u 2 / 2

where f A ( ) = (n B / no ) 2 / 3 is shown in Fig. 2-32 and Cev = 2.8.104 (1/m) is the


1/ 3

cavitation parameter, defined as:

C ev = 3 3 43 no . (2.50)

By comparison with the photographs of steady-state cavitation structures in a


small-scale asymmetrical planar nozzle (under steady and transient injection pressure
conditions), Alajbegovic, et al. (1999) demonstrated the ability of the model to predict
the cavitation development.

2.4.4.4. Comparison of the transport-equation models


To compare several models of cavitation based on the transport equation for the
fraction of the vapour phase, they have been reformulated in a similar dimensionless
form (VF). Comparison of the source terms offered in the models of cavitation
suggested by Kunz, et al (2000), Ahuja, et al. (2002), Singhal, et al. (2001),
Alajbegovic, et al. (1999) and Yuan, et al. (2001) reveals the following:
All of them, except the model for condensation employed by Kunz, et al. (2000),
consider the driving effect of the local pressure on the formation and destruction of the
vapour. In dimensionless formulation this effect is described through the local
cavitation parameter:

pv p
= , (2.51)
l u 2 / 2

which can be also can be expressed in terms of the cavitation number CN, commonly
applied in studies of cavitation in nozzles and orifices:

pv p 1
= = ~
p , (2.52)
l u / 2
2
CN
Chapter 2.4. Literature survey. Models of cavitation 66

p p2 2( p1 p 2 )
where ~
p= and u = .
p1 p 2 l

All the models describe the effect of the void fraction on the rate of the phase
transition. The rates of evaporation in a pure vapour ( 1 ) and condensation in pure
liquid ( 0 ) reasonably turn to zero. The models of Alajbegovic, et al. (1999) and
Yuan, et al. (2001) also provide zero rates of condensation in pure vapour and
evaporation in pure liquid. Similar behaviour is seen in the condensation model adopted
by Kunz, et al (2000). This implies that a certain amount of new phase is needed to
initiate the phase transition.

Fig. 2- 33. Evaporation and condensation rates versus the void fraction according
to different cavitation models (velocity scale u = 100 m/s; linear scale l = 1 mm;
|| = 1).

Fig. 2-33 shows the rates of evaporation and condensation calculated from
different models for a set of parameters, which model the flow conditions in a Diesel
injector. The models developed by Kunz, et al. (2000), Alajbegovic, et al. (1999) and
Yuan, et al. (2001) can predict similar rates of cavitation only over a limited range of
the flow parameters. At the same time, the models by Ahuja, et al. (2001) and Singhal,
et al. (2002) produce the source terms several orders magnitude greater compared to
Chapter 2.4. Literature survey. Models of cavitation 67

each other and the models of Kunz, et al. (2000), Alajbegovic, et al. (1999) and Yuan, et
al. (2001) (Fig. 2-33). These deviations in the rates of evaporation and condensation can
be explained by the variety of cavitation flows, for which these models have been
originally designed and tuned. All the reviewed transport-equation based models of
cavitation are empirical to a certain degree, and require justification of the laws of
condensation and evaporation they use.

2.4.4.5. Conclusions about transport-equation models


Chen and Heister (1995), Kunz, et al. (2000) and Ahuja, et al. (2001) have applied
a simplistic linear approximation for the source term as a function of the local cavitation
parameter . This resulted in a scalable form of the void-fraction equation, which is
essential to describe the hydrodynamic similarity of cavitation flows. However,
calculations have revealed the rate constants in Kunzs model are not unique and
require tuning for particular flow conditions. The reason for this could be the limited
range of application for the laws for the phase transition and dependency of these
parameters on the liquid quality. Chen, et al. (1995) have found the solution to be very
sensitive to the cavitation rate constant, with high values increasing the stiffness of the
void fraction equation. The difficulties in physical interpretation of the rate constant and
inaccurate description of the transient behaviour of cavitation flow forced Chen and
Heister to move to the more sophisticated approach based on non-linear bubble-
dynamics theory (Chen and Heister, 1996). Also, Sauer (2001) concluded that an
accurate description of the collapse of cavitation bubbles requires the Rayleigh-Plesset
equation, while the results of calculations with the Rayleigh model (2.27) could be very
approximate. Alternatively, to improve features of the transport-equation models of
cavitation, bubble-dynamics effects require special examination for the variety of the
stages and regimes of cavitation.
The models of Chen and Heister (1995) Kunz, et al. (2000) and Ahuja, et al.
(2001) are scalable, but purely heuristic (achieving the relaxation nature of the density
change with the pressure). The inability to incorporate the physical effects in this kind
of model, prevent their further development.
Accurate description of the collapse stage and transient effects of cavitation
requires application of the Rayleigh- Plesset equation.
The cavitation models suggested by Singhal, et al. (2001) Alajbegovic, et al.
(1999) and Yuan, et al. (2001) describe the effect of local pressure on the rate of the
Chapter 2.4. Literature survey. Models of cavitation 68

phase transition in a form consistent with Rayleigh theory (2.27). Equations (2.39),
(2.40), (2.45) and (2.49) reveal the linear effect of the length scale of the mean flow l
on the rates of phase transition. Also, in the model developed by Singhal, et al. (2001)
the rate of phase transition (2.45) can vary with the flow velocity scale. The presence of
these dimensional parameters, which require adjustment with measurements, makes
these models of cavitation non-scaleable.
The theory-based models of cavitation (Yuan, et al., 2001; Alajbegovic, et al.,
1999; Singhal, et al., 2001) describe the bubble-dynamics and nuclei effects of
cavitation, but are non-scalable.

The models by Alajbegovic, et al. (1999) and Yuan, et al. (2001) are both derived
form the Rayleigh model (2.27) and result in the source terms, which depend on the
void fraction in a similar way (Fig. 2-32). The deviations of the curves in Fig. 2-32 is
due to the difference in the number density of nuclei, which determines the evaporation
rate constant. The nuclei parameters Ro and n in the models of Alajbegovic, et al.
(1999) and Yuan, et al. (2001), are flow-specific and require adjustment. This happens
because experiments on cavitation do not usually include measurements of nuclei.
The advantage of the models derived from bubble dynamics is that they allow the
nuclei effects on cavitation to be considered. Similarity is governed by the rate constant
of cavitation, which depends on the number density of cavitation bubbles. To develop a
similarity-consistent model of cavitation, the key issue is to determine parameters,
which the number density depends on.

2.4.4.6. Models based on the Rayleigh-Plesset equation


The transport equation models of cavitation suggested by Alajbegovic, et al.
(1999) and Yuan, et al. (2001), utilise the simplistic Rayleigh model (2.27). This model
describes the limiting case of inertia-controlled growth of a spherical bubble in a liquid
under a step variation in pressure of the surrounding liquid. However, this model can
not accurately describe bubble collapse and neglects a number of effects, which
determine the behaviour of cavitation bubbles. To account for the effects of inertia,
viscosity and surface tension of the liquid and compressibility of the gas-vapour
content, the Rayleigh-Plesset equation can be applied (Brennen, 1995):
Chapter 2.4. Literature survey. Models of cavitation 69

3 R&
R R&& + R& 2 = pv + p g pl 4 2
2 R R. (2.53)

This equation describes the dynamics of a spherical bubble in a liquid. It was


derived by neglecting mass transfer due to phase transition, and also assuming thermal
equilibrium and zero velocity slip between the bubble and the liquid (Brennen, 1995).
For a long time the Rayleigh-Plesset equation (2.53) has been used to incorporate
bubble-dynamics effects into one-dimensional models of cavitation flows. Recently, the
progress in the speed of computers made it possible to employ the equation (2.53) in
the simulation of two- and three-dimensional transient cavitation flows (Kubota, et al.,
1990; Chen and Heister, 1996; Delale, 2001; Chanhine, 2004; and Giannadakis, et al.,
2004).
In the following the main achievements in modelling hydrodynamic cavitation
flows using the Rayleigh-Plesset equation are described. This part is aimed at clarifying
the ability of these models to account for nuclei scale effects and describe the
hydrodynamic scaling of cavitation flows.

Chen and Heister (1996)


The model developed by Chen and Heister (1996) employed the concept of the
volume fraction (2.47), and assumed the presence of bubble nuclei of radius R and
fixed number density n , similarly to the models of Alajbegovic, et al. (1999) and Yuan,
et al. (2001). The analysis of Chen and Heister (1996) resulted in the following equation
for the volume fraction of the vapour:

2
D 2 l D
2

~ = f 1 ( ) f 2 ( ) ~ , (2.54)
D t 2 l cav Dt

~
where t = t u / l is non-dimensional time, is the local cavitation parameter (2.51)
and l cav is the cavitation length scale defined as:

l cav = 3 43 n , (2.55)
Chapter 2.4. Literature survey. Models of cavitation 70

and functions f1 and f 2 are:

6 1 / 3 (1 + 1 / 3 + 2 / 3 ) 2
f .1 ( ) = ,
(2 + 1 / 3 )(1 )1 / 3

11 2 / 3 1 / 3 1 1 + 4 1 / 3 + 2 / 3
f 2 ( ) = .
6 (1 ) 6 2 / 3 (2 + 1 / 3 )(1 + 1 / 3 + 2 / 3 )

Initial void fraction o is required for equation (2.54) to start the vapour
formation in the flow.

The models of Alajbegovic, et al. (1999) and Yuan, et al. (2001) (equations (2.30),
(2.45) and (2.49)) can be considered as a limiting case for the model of Chen and
Hiester (1996), when the second-order term is neglected in (2.54). Similarly to the
models of Alajbegovic, et al. (1999) and Yuan, et al. (2001), equation (2.54) contains
the ratio of the cavitation length scale l cav and the hydrodynamic length scale l .

Kubota, et al. (1990), Giannadakis, et al. (2004)

Advanced models of cavitation (Giannadakis, et al., 2004; Chahine, 2004)


employs the equation for dynamics of a cluster of bubbles of a radius R and number
density n derived by Kubota, et al. (1990):

3
3 & 2 p v p p g , o Ro 2 4 R&
RR
&& + R = +
2 R R R (2.56)
(
2 r 2 nR 2 R
&& + n&R 2 R& + 2nRR& 2 )
where p g ,o and Ro are the initial gas pressure and radius of a bubble and is the

polytrophic exponent. The last term in this equation simulates bubble dynamics on a
sub-grid scale r . This feature becomes useful when using coarse meshes, where the
pressure is shared between several bubbles in a computational cell. However, the model
reduces to the classical Rayleigh-Plesset equation (2.53) for a single bubble when the
mesh is sufficiently fine.
Chapter 2.4. Literature survey. Models of cavitation 71

Applying the scales for a cavitation bubble moving in the flow: bubble radius Ro ,

velocity scale u = 2( p1 p 2 ) / , and time scale Ro / u , equation (2.56) may be


rearranged in the dimensionless form:

p g ,o 1 3 l 2 1 ~&
~ ~ && 3 ~& 2 4 R
R R + R = + 2 ~ ~ + ~
2 u / 2 R Ro We R Re R
(2.57)
2 ~ ( ~ ~ && ~ ~& ~ ~&
r 2 n~R 2 R + n~& R 2 R + 2n~R R 2 )
~
where: R = R / Ro , ~
r = r / Ro , n~ = n Ro3 , is defined in (2.51), and Reynolds and
Weber numbers are defined by the equations (2.2) and (2.3).

Equation (2.56) forms the basis of the most sophisticated models of cavitation
flows, applied in modern CFD codes. In the dimensionless form (2.57) the Rayleigh-
Plesset equation contains the parameter l / Ro , which prevents the hydrodynamic

similarity of a liquid with the bubble nuclei of specific radius Ro .


Analysis of different terms in equation (2.56) with a help of the theory of bubble
dynamics, can lead to substantial simplifications and increase in the computational
efficiency of a model. Thus, the model by Yuan, et al. (2001) can be considered as a
limiting case for the Rayleigh-Plesset model.

2.4.4.7. Scaling of the bubble-dynamics models of cavitation


The review of experimental observations on cavitation flows revealed
experimental evidence of the hydrodynamic similarity of cavitation flows. The scale
effects can be classified depending on their physical nature, and therefore can be
described separately in a model of cavitation flow. At the same time, the issue of
scalability has not been considered when developing the models of cavitation from
bubble-dynamics theory. The following analysis identifies the requirements for a
scalable mathematical model of cavitation flow.
Chapter 2.4. Literature survey. Models of cavitation 72

nozzle wall
liquid
Ro

bubble l

l cav ~ n o1 / 3

Fig. 2- 34. Spatial scales in the flow with cavitation bubbles.

The models developed by Kubota, et al. (1990), Chen and Heister (1996),
Alajbegovic, et al., (1999), Yuan et al. (2001) and Giannadakis, et al. (2004) are
similarly characterised by the length scales (Fig. 2-34):
hydrodynamic scale l ,

bubble size scale Ro and

1
bubble density scale l cav ~ .
3 no

They result in two parameters of the cavitation model:


l cav / l and (2.58)

Ro / l , (2.59)
which together with dimensionless numbers CN, Re and We, etc. (2.1) (2.7)
determine the hydrodynamic similarity of cavitation flows. This means that for
hydrodynamic similarity:
l cav / l = idem (2.60)

Ro / l = idem (2.61)
In a real physical experiment it is impossible to match the set of similarity
conditions (2.60) (2.61) all together. Thus, for two cavitation flows of the same liquid
and different hydrodynamic scales, the conditions Re=idem and We=idem can not be
met simultaneously, but can only be approached approximately in the limit of negligible
surface tension effect.
Chapter 2.4. Literature survey. Models of cavitation 73

A scalable model of cavitation flow should be able to match the similarity


conditions (2.60) and (2.61), which originates from the governing bubble-dynamics
equation (2.57) and the initial condition for the state of a cavitation bubble. Fig. 2-34
provides a simple geometrical interpretation for the hydrodynamic scaling conditions
(2.60) and (2.61).
In the models developed by Kubota, et al. (1990), Alajbegovic, et al. (1999) and
Yuan, et al. (2001) the number density of cavitation bubbles n is considered as a
tuneable parameter, which is treated as a fixed number for a given cavitation flow. This
is a common and very simplistic approach, which neglects completely the distribution
of bubbles by radius. In this case parameters Ro and l cav are not linked to each other,
and can be specified independently in the model. The hydrodynamic similarity
conditions (2.60) and (2.61) can be satisfied by specifying the cavitation scales Ro and

l cav in accordance with the hydrodynamic length scale l . Because the hydrodynamic

length scale l is not defined, the model of cavitation is not complete. Besides that, the
model loses the actual features of cavitation nuclei and is difficult to verify with
measurements.
The similarity condition (2.60) requires that the cavitation length scale be
dependent on the hydrodynamics of the flow. The cavitation length scale is determined
by the number density of cavitation bubbles, which is a function of the spectrum of the
bubbles in the liquid. In turn, the population of bubbles in the liquid depends on the
fluid quality and conditions of the flow.
The models of cavitation developed by Giannidakis, et al. (2004) and Chahine,
(2004) consider the population of bubble nuclei in cavitation flow. Because of the
dependence of the population of bubbles on numerous parameters of the flow and liquid
quality effects, the development of a similarity-consistent mathematical model of
cavitation within this approach becomes rather complex task. Therefore, the present
study is aimed at clarifying the flow sale effects on the nuclei distribution function, and
to applying these findings to match the scaling condition (2.60). The liquid quality
effects are considered in an empirical way.
Chapter 2.4. Literature survey. Models of cavitation 74

2.4.5. Conclusions

From the literature survey of experimental studies of cavitation flows it was


concluded that particular structure and hydrodynamic similarity of cavitation flows
depend on the liquid quality and flow scale effects. The present review is focused on the
comparative analysis of cavitation models and identification of methods for description
of the liquid quality and flow scale effects on cavitation flows.
Comparison of cavitation models shows that to describe the cavitation onset, the
hypothesis about the vapour formation in a bulk liquid as a result of reduction in local
pressure below critical level, is commonly applied. This critical pressure is usually
associated with the saturation pressure in liquid. The concept of stress-induced
cavitation (2.13), which can explain the cavitation onset in high-strain viscous flows,
was not applied in cavitation modelling. To predict the pattern of liquid flow at the
cavitation onset, modern models of cavitation take into account the viscous nature of the
flow and average turbulent motion in fluid.
To describe the development of cavitation structures in the flow, two different
concepts were developed, assuming the barotropic of baroclinic nature of the flow.
Algebraic models of cavitation (equation-of state (EOS) models) assume an explicit
effect of pressure on the local volume fraction of the vapour phase. The models which
use the differential equation to describe the local phase content (models based on the
transport equation for the void/mass fraction, and Rayleigh-Plesset equation for the
radius of cavitation bubbles) describe the baroclinic nature of cavitation flows.
Robustness and efficiency of the liquid-flow algorithms resulted in a progress in
development of single-fluid models of cavitation flows.
The review has revealed that the transport-equation based cavitation models are
more robust and popular in CFD, rather than models based on the Rayleigh-Plesset
equation (2.53). The linear models (2.27) and (2.28) produce a time-independent source
term for the volume-fraction transport equation (2.29) and, thus permit a steady-state
solution for cavitation flow. The non-linear effects described in the Rayleigh-Plesset
equation (inertia and viscosity of the liquid, thermal effects, turbulence, bubble
interactions, etc.) causes the rates of bubble growth/collapse to be time-dependent. The
models based on (2.53) seems to be more appropriate for simulation of transient
structures in cavitation flows, however they are quite expensive to compute. To validate
Chapter 2.4. Literature survey. Models of cavitation 75

assumptions about the dynamics of cavitation bubbles, the analysis of growth and
collapse of cavitation pockets is provided in the following chapter.
Analysis shows that many of the transport-equation based models (Kunz, et al.,
2000; Ahuja, et al., 2001; Yuan, et al., 2001; Alajbegovic, et al., 1999; Singhal, et al.,
2002) contain dimensional parameters, which have to be adjusted to account for liquid
quality and flow scale effects on cavitation. The rate parameters of cavitation model
govern the hydrodynamic similarity of cavitation flow.
It is shown that the advantage of the models derived from bubble dynamics theory
is that they describe the nuclei effects on cavitation. The model developed by Yuan, et
al., (2001) reveals that the rate parameters of the transport-equation based cavitation
models depend on the number density of cavitation bubbles. Because measurements of
the population of bubbles in a liquid are rare and the spectrum of nuclei is liquid-
specific, it causes a difficulty in the specification of concentration of bubbles, as a
parameter for the model. On the other hand, analysis of similarity shows that the
hydrodynamic scaling of cavitation flows requires the rate constants to be determined
by the hydrodynamic scale of the flow. This fact leads to the conclusion that
hydrodynamic and liquid quality effects are merged in one parameter, namely, the
concentration of bubbles, which is available to tune the model for changes in either
effect. The model for the number density of cavitation bubbles was not recognised in
the literature and is needed as a closure to the transport-equation based model of
cavitation flow. The following study is focused on the development of such model.
Chapter 2.5. Literature survey. Numerical computation of cavitation 76

2.5. Numerical computation of cavitation flows


2.5.1. Summary of methods of calculation of liquid-vapour flows

In a thermodynamic sense cavitation flow is characterised by two phases, namely,


liquid phase and gas-vapour phase. From the general point of view, the methods for
calculation of free-surface flows on fixed grids can be classified in two groups: surface
tracking (or surface methods, explicit) and surface capturing (or volume methods)
(Ubbink, 1997).
In the surface tracking methods, special marker points are used to mark an
interface. These methods include the particles on interface method, height functions
method, level set method, and the Volume of Fluid Method (VOF), originally developed
by Hirt and Nochols (1981) for incompressible free-surface flows without phase
transition. In the original VOF method (Hirt and Nichols, 1981), the non-divergent form
of the volume fraction equation is applied to describe conservation of the volume
fraction :

D
= + uk =0 (2.62)
Dt t xk

Surface capturing methods, or volume methods, mark the fluid by particles or


marker function to differentiate between the phases which occupy the control volumes
of the domain. In the methods, which use the marker function, the transport (advection)
equation (2.62) is applied. Because of the numerical diffusion, which appears when
using standard differencing schemes for advection terms in (2.62), accurate calculation
of the sharp interface becomes difficult. Therefore, several methods have been
developed to locate a sharp interface and compute its evolution in time. Some of them
have been developed as variations of the VOF method with special techniques used for
the calculation of face fluxes for the inter-phase control volumes (geometric
reconstruction, donor-acceptor, Euler explicit and implicit time-integration schemes,
and high-order differencing schemes).
The disadvantage of equation (2.62) is that it is not rigorously mass-conserving.
To describe a vapour-liquid flow with phase transition, the continuity equation is
applied to an element of mass in the flow. Therefore another group of the surface
Chapter 2.5. Literature survey. Numerical computation of cavitation 77

capturing methods was developed to utilise the conservative form of the volume
fraction equation, which is also applied in this study:

u k
+ =0
t x k (2.63)

This equation originates from the continuity equation for the mixture. When
applying equation (2.63) to describe a sharp interface in gas-liquid systems, the problem
in stability of iterations may arise due to interaction of the incompressible flow in the
liquid with compressible flow in the gas phase (Yabe, et al., 1995). This is caused by
divergence of velocity inherently present in equation (2.63). The problem becomes less
sever when treating both phases as incompressible. However, because the velocity field
in cavitation flows is not divergent-free, special algorithms are required to improve
convergence.
Application to cavitation flows. In models of cavitation the volume fraction
equation appears in conservative form (2.29) as a result of application of the mass
conservation to each of the phases. To resolve a sharp inter-phase boundary within the
conservative formulation (2.29) one way is to employ the incompressible formulation
r
( div u = 0 ) and use the non-conservative form of the advection operator in equation
(2.62). Then front tracking methods can be applied to resolve the interface (e.g. VOF
method). An alternative method is to keep the divergent form of the advection operator
(2.63) and apply surface capturing methods to locate an interface boundary. Because
many CFD packages provide standard routines to solve the transport conservation
equation for a scalar property, the divergent formulation of the void fraction equation
(2.29) becomes beneficial.
When using coarse meshes, application of high-order approximation schemes
becomes essential to improve the accuracy of the calculation of the cell fluxes in the
phase transition region. The high-order schemes, which have been applied in this study
for computation of cavitation flows, are described in the following chapter.
To tackle the convergence problem, which appears when using an incompressible
segregated pressure solver for calculation of a cavitation flow, several methods have
been developed. The aim of these methods is to account explicitly for the divergence of
velocity in the phase-transition region (Sauer, 2000; Senocak and Shyy, 2002).
Chapter 2.5. Literature survey. Numerical computation of cavitation 78

Modifications to the incompressible pressure solution algorithms for modelling of


variable-density flows are discussed in the following sections.

2.5.2. Basic incompressible pressure solvers

This section describes pressure correction algorithms developed for computation


of steady-state incompressible flows. These algorithms form the basis of segregated
pressure solvers in computational fluid-dynamics packages.

2.5.2.1. Incompressible form of pressure corrections equation


Inspection of the set of continuity and momentum conservation equations reveals
that they do not contain any explicit equation for the pressure variable. The target of the
solution procedure is to find the pressure distribution, which will result in the velocity
flow field governed by the momentum equation and consistent with the continuity
equation. Segregated solvers use the prediction-correction method for velocities and
pressures, which are expressed as:

r r r
u = u * + u , (2.64)

p = p * + p , (2.65)

r r
where u * and p * are guessed values and u and p are corrections. Under

convergence, this procedure must result in velocity and pressure fields consistent with
the original momentum and continuity equations.
Integration of the set of mass and momentum conservation equations for arbitrary
geometry configuration of the flow, requires time-spatial discretisation. This section
describes basic ideas of pressure algorithms for calculation of steady-state
incompressible flows, when discretisation is performed using the finite volume method
(Versteeg and Malalasekera, 1995; Ferziger and Peric, 2002).
Chapter 2.5. Literature survey. Numerical computation of cavitation 79

WW ww W w P e E X

Fig. 2- 35. Control cell around the velocity node w for the X -component of
velocity, in one-dimensional consideration.

Using the finite-volume discretisation method, the momentum equation for


velocity u at the point w (Fig. 2-35) can be transformed into a set of linear algebraic
equations (Versteeg and Malalasekera, 1995):

awuw = anbunb Aw ( pP pW ) , (2.66)


nb

where aw = anb and anb are coefficients defined by convection-diffusion


nb

discretisation scheme at the cells neighbouring to cell w (in one-dimensional case


nb = ww, e as shown in Fig. 2-35), Aw is area of the face at node w orthogonal to the

direction of given velocity component u . This equation produces the correlation for the
velocity corrections:

a nb u nb
u w = nb
+ d w ( pW
p P ) ,
aw

where d w = Aw / a w . Patankar and Spalding (1972) have suggested to neglect the first
term in the right-hand side of this equation, which resulted in the following equation for
the velocity correction u w :

u w = d w ( pW
p P ) . (2.67)

The pressure correction equation is derived from the discretised continuity


equation, which expresses the balance between the mass fluxes at cell faces:
Chapter 2.5. Literature survey. Numerical computation of cavitation 80

(u )nb Anb = 0 . (2.68)


nb

Assuming that the flow is incompressible and splitting the velocity according to
(2.64) and using approximation (2.67), the equation for the pressure corrections can be
obtained:

nb Anb d nb ( p P pNB ) = nb u nb* Anb ,


nb 1nb
442443
m& *P

where index NB denote the cells neighbouring to the cell P , indices nb are applied to
the flow variables defined at the cell faces in contact with the control volumes NB , and

m& *P are mass imbalances at the cell P . The pressure correction equation can be
rearranged in a standard form:

a P p P a NB p NB = m*P , (2.69)
NB

where a NB = (Ad )nb and a P = a NB .


NB

2.5.2.2. SIMPLE algorithm


SIMPLE is an abbreviation for the Semi-Implicit Method for Pressure Linked
Equations (Patankar and Spalding, 1972). This is method is based on equation (2.67) for
the velocity corrections and uses the pressure correction equation (2.69). The main steps
of the algorithm are as follows:

1. Guess a pressure field p * .

2. Solve the momentum equation using the guessed pressure field p * :

a w* u w* = a nb u nb
*
(
Aw p *P pW
*
)
nb

3. Calculate mass imbalances and solve the pressure correction equation (2.69).
Chapter 2.5. Literature survey. Numerical computation of cavitation 81

4. Correct the pressure and velocities using equations (2.64) and (2.65).
5. Repeat steps 2 4 until convergence.

2.5.2.3. SIMPLEST algorithm


To improve convergence of the SIMPLE algorithm, the convection terms in the
momentum equation are treated in an explicit way, being put in the source term as
knowns from the previous iteration (SIMPLEST, Spalding, 1980). The difference
between this algorithm and SIMPLE is in splitting the coefficients a nb in the
momentum equation (2.66) into the convection and diffusion parts:

a nb = C nb + Dnb ,

and treating the convection terms explicitly in the source term, using the known values

of velocity from the previous iteration u old :

a w* u w* Dnb u nb
*
(
= Aw pW
*
)
p *P + C nb u nb
old
.
nb nb

Compared with the SIMPLE algorithm, SIMPLEST neglects the convection effect
of changing velocity, and amplifies the diagonal dominance of the matrix coefficients,
which improves the convergence. SIMPLEST is default method in the PHOENICS
code.

2.5.3. Modifications to the pressure-correction algorithm for flows


with variable density

In the homogeneous models of cavitation flows it is assumed that the local liquid-
vapour content, void fraction and density of a mixture are governed by the local
pressure in the flow. This makes the mixture virtually compressible even when the
liquid and vapour phases are treated as incompressible. Because incompressible
formulation of the pressure correction equation (2.69) neglects the pressure-density
coupling, pressure corrections will cause incorrect adjustment of the velocities, resulting
in poor convergence of iterations. To overcome this difficulty two methods can be
Chapter 2.5. Literature survey. Numerical computation of cavitation 82

applied. One way is to utilise the compressible form of the mass conservation and
pressure correction equations. Another approach is to exclude the density from the
pressure correction equation by transferring to the volumetric form of the continuity
equation. Here, these two methods are discussed from the point of view of their
application to calculation of cavitation flows.

2.5.4.1. Compressible form of the equation for pressure corrections


In compressible pressure solvers the density is presented in a form similar to the

velocity and pressure variables, i.e. as a sum of guess * and correction :

= * + (2.70)

To account for the liquid compressibility effect, the density corrections are
considered to be related to the pressure corrections:

= p , (2.71)

1
where = is compressibility of liquid.
p
Substitution of correlations (2.64), (2.65), (2.67), (2.70) and (2.71) into equation
(2.68), leads to:

[(Ad )nb ( p P p NB ) + (u * A)nb p nb ( p P p NB )] =


+ (Ad )nb p nb
nb

(
= u * A nb , ) (2.72)
1nb
44244 3
m& *P

which yields the compressible pressure-correction equation.


Senocak and Shyy (2000) have used the pressure-correction equation (2.72) in
calculation of cavitation flows described by the homogeneous transport-equation based
model. As a result of their study, they have suggested an approximate correlation for
parameter in equation (2.71):
Chapter 2.5. Literature survey. Numerical computation of cavitation 83


= C , (2.73)
p

where C = 4 is numerical constant adjusted to provide smooth convergence and

stabilise iterations, and and p are density and pressure scales. Senocak (2002) has
stated that according to the nature of equation (2.72) the choice of C should not affect

the converged solution, however big values of C can cause the iterations to quickly

become unstable.
In CFD codes, such as PHOENICS, a parameter can be specified by user to
account for the compressibility of the flowing liquid and control the convergence of
iterations.

2.5.4.2. Gas And Liquid Algorithm (GALA)


Another method of calculation of incompressible flows with variable density is
based on the utilisation of the volumetric continuity equation, which eliminates
completely the effect of density variation on the velocity field of the mixture. This
formulation, originally suggested by Spalding (1974) for calculation of flows without
the phase transition, was adopted in simulation of cavitation flows (Sauer, 2000; Yuan,
et al., 2001). Here, the main equations of this method are derived and discussed from
the point of view of their solution using the PHOENICS code.
In the homogeneous models of cavitation, which use the void fraction and assume
the liquid and vapour phases are incompressible ( l = const and v = const ), the full
derivative for the mixture density can be expressed as:

D r r D
+ u = ( v l )
Dt t Dt (2.74)

Using (2.74), the continuity equation (2.21) can be rearranged in a form:

u k l v D
=
x k Dt . (2.75)
Chapter 2.5. Literature survey. Numerical computation of cavitation 84

The advantage of equation (2.75) is that the cell fluxes are continuous at the
interface boundaries with a steep change in density. Taking this into account,
imbalances in the volume flow rates, can be used to determine the pressure corrections.
The substantial derivative of the volume fraction follows from the void fraction
equation (2.29):

D S u
= k
Dt v x k

and after substitution into equation (2.75), gives the source term for the volumetric
continuity equation:

u k l v
= S .
x k l v

This equation can be applied to calculate the pressure corrections in a SIMPLE-


like iterative procedure described in section 2.5.2. For flows without phase
transformations ( S = 0 ) this procedure was developed in PHOENICS under the name

of the Gas And Liquid Algorithm (GALA). To include the volume source term S (due
to cavitation) in the volumetric continuity equation, special coding is needed.

2.5.4. Conclusions

To calculate the distribution of the void fraction in cavitation flow, surface


tracking and surface capturing methods have been reported in the literature. Surface
capturing methods utilise the conservative form of the void fraction equation, which can
be solved using the standard routines available in CFD codes. To achieve a better
resolution of an interface boundary, a high-order spatial discretisation scheme is needed.
To improve the convergence and stability of incompressible pressure-correction
algorithms for simulation of compressible flows, two different approaches have been
developed. One of them explicitly takes into account the compressibility of the flow
(Senocak and Shyy, 2000), while the other eliminates the density gradients in the
Chapter 2.5. Literature survey. Numerical computation of cavitation 85

formulation of the pressure correction equation (Spalding, 1974; Sauer, 2000). Both of
theses methods are designed to estimate the pressure correction more accurately.
The GALA algorithm (Spalding, 1974; Sauer, 2000) utilises a non-divergent
volumetric form of the continuity equation and allows the density to be excluded
completely from the pressure correction equation. Then the source term for the
continuity equation implicitly accounts for the variations in density due to phase
change.
The Senocak method (Senocak and Shyy, 2002) was developed as an extension to
the standard incompressible pressure solvers, based on the mass-conservation form of
the continuity equation. Effectively, it uses the pseudo-compressibility of cavitating
fluid to satisfy mass conservation under variations in the mixture density caused by
phase transition. In modern CFD codes, such as PHOENICS, which utilise the
compressible form of the pressure-correction equation, the Senocak method can be
employed for calculations of cavitation flows.

2.6. Conclusions for chapter 2


The literature survey has reviewed the results of experimental studies of the
hydrodynamic cavitation in nozzles, and numerical models developed for prediction of
cavitation flows.
Experimental studies of cavitation flows in nozzles have revealed that the
geometry of the nozzle, viscosity of the liquid and turbulent motion determine the
pattern of cavitation flow. Visual observations and measurements of cavitation flows
provide the evidence that cavitation starts develop from bubble nuclei in the bulk liquid.
Attempts to find scaling laws for characterisation of cavitation flows in nozzles have
resulted in the conclusion that hydrodynamic similarity requires minimisation of the
flow and liquid quality scale effects. Experimental studies have shown the relative
importance of different scale effects, and resulted in useful data for testing and
validation of models of cavitation.
In modern practice of calculation of cavitation flows to account for the effects of
geometry, viscous nature of the flow and turbulent motion, CFD methods are used.
Application of single-fluid solvers for calculation of cavitation flows is attractive from
the point of view of the apparent economy in the time to complete calculations in
Chapter 2.5. Literature survey. Numerical computation of cavitation 86

comparison to multi-fluid algorithms; however, they require special numerical methods


for calculation of compressible flows with high density gradients.
The transport-equation based models of cavitation derived from bubble dynamics
theory provide a simple basis to account for the bubble nuclei scale effects on the flow.
These models have been shown to be able to predict accurately the characteristics of
cavitation flows, but require specification of the number density of cavitation nuclei,
which affects the rate of cavitation. Numerical studies have revealed that concentration
of cavitation bubbles is affected by the hydrodynamic length scale, and therefore
requires adjustment for each particular flow. This significantly limits the range of
application of the model. To resolve this problem, the liquid quality and flow scale
effects should be decoupled in the model for the number density of cavitation bubbles.
Accurate description of the liquid quality effect requires consideration of the spectrum
of the cavitation nuclei, which determines the concentration of critical cavitation
bubbles. Also the issue of scalability of cavitation models with a void fraction transport
equation derived from bubble dynamics theory, has not been addressed. A study of
scalability is required to develop an adequate model of cavitation, which would provide
similar solutions when the liquid quality and flow scale effects are matched. Also,
analysis of the rate of bubble growth and collapse in a liquid is needed to clarify the
range of the application of the transport equation-based models.
Joseph (1995) has introduced the hypothesis that viscous shear stresses can
increase the critical pressure threshold for the cavitation onset in flowing liquid.
Comparison of the measurements of cavitation flows in small-scale nozzles performed
by Roosen, et al. (1996) and Winklhofer, et al. (2003) gave experimental evidence of
rise in the critical vapour pressure in high-stress flows. However, the literature survey
has not revealed any models of cavitation which would account for the shear-stress
mechanism of cavitation inception. The measurements by Roosen, et al., (1996); and
Winklhofer, et al., (2003) provide useful data for development and validation of such a
model.
In the single-fluid models of cavitation, the vapour content in the flow is a
function of the local pressure, which is returned by a single-fluid solver. Therefore the
results of calculations of cavitation flows depend upon the accuracy of predictions of
cavitation-free flow. However, accurate simulation of flows in Diesel injectors, even in
the absence of cavitation, is a complex task. This requires modelling of the developing
turbulent flows at moderate Reynolds numbers, description of the flow separation in the
Chapter 2.5. Literature survey. Numerical computation of cavitation 87

nozzle entrance and flow reattachment inside the nozzle including formation of the re-
entrant jet motion. Simulation of these phenomena is not straight-forward even in
modern computational fluid dynamics codes. Therefore, the present study applies well-
known and proven methods for calculation of the viscous turbulent flows to capture the
main average pattern of the liquid flow, and also to calculate the mixture flow and
validate the sub-models for the cavitation process.
Chapter 3. Analysis and model development 88

Chapter 3. Analysis and Model Development


3.1. Introduction
It was shown in the Literature Review that the models of cavitation based on
bubble dynamics theory capture the physical nature of cavitation and permit the viscous
and nuclei scale effects on cavitation. Therefore these models were chosen as a basis for
the development of the cavitation model in the present study. Analysis of models
proposed in the literature has revealed that consideration of the scale effects is crucial
when developing a similarity-consistent model of cavitation flow. However,
hydrodynamic similarity requires the parameters of conventional models of cavitation
(Kubota, et al., 1990; Yuan, et al., 1999), which are the number density and initial size
of cavitation bubbles, to be scaled by the hydrodynamic scale of the flow. The current
part is focused on decoupling of the liquid quality and hydrodynamic scale effects in
parameters of cavitation model.
First, the phenomena of growth and collapse of a single cavitation bubble are
reviewed to justify the use of the Rayleigh equation (2.27) for cavitation modelling.
This is followed by the critical analysis of the modelling approach, which describes
cavitation locally for bubbles of one size and assumes the same concentration of
bubbles everywhere in the flow. This is aimed to clarify the nature of cavitation
bubbles, which concentration should account for non-uniform spectra of the bubbles in
real liquids. This analysis results in the concept of critical cavitation nuclei, which
number density may vary with the tension in liquid.
After this, assuming hydrodynamic similarity of cavitation flows and estimating
tension in cavitation region from the hydrodynamics of liquid flow, correlation for the
number density of cavitation bubbles as a function liquid tension is derived. This
correlation is used as a closure to the model of cavitation flow, which is shown to be
similarity-consistent.
To validate the model for concentration of critical cavitation nuclei, measurements
of spectra of cavitation bubbles in liquids are reviewed. To reveal the range of
application of the model for concentration of critical nuclei, the measurements of
cumulative spectra of bubbles in water by Gindroz, et al. (1997), are applied.
Chapter 3. Analysis and model development 89

3.2. Growth and collapse of cavitation bubbles


To achieve reasonable values of phase transition rates when modelling the source
term in the void fraction equation (2.30), analysis of the laws of growth and decay of
cavitation bubbles in a liquid is essential. This chapter summarises main results of a
theory of behaviour of unique spherical vaporous cavitation bubble in an infinite liquid.
The analysis concerns the dynamics of a bubble under a step variation in pressure of the
surrounding liquid pl , . The temperature and pressure are assumed to be uniform

inside the bubble, and correspond to the saturation conditions at the bubble wall.
The mathematical description of the dynamics of a single spherical vapour bubble
in a liquid is based on the Rayleigh-Plesset equation (2.53). Different effects, which
influence growth and collapse of a bubble, can be revealed by consideration of different
terms in the equation (2.53) (and its modifications), completed by the thermal energy
conservation equation. The effect of gas diffusion can usually be neglected in cavitation
modelling due to the very low speed of diffusion process.

3.2.1. Bubble growth

The classical theory of bubble growth has been stated by Mikic, et al. (1970). He
has found a suitable approximation for the radius of a growing bubble, which links
together the solution of Plesset and Zwick (1954) for heat-diffusion controlled growth
and Rayleighs model (2.27) for inertia-driven bubble growth (Fig. 3-1). These results
where completed by Sauer (2001), who have shown that under rapidly varying pressure
conditions inertial effects determine the behaviour of a cavitation bubble and cannot be
neglected. Sauer (2001) also has concluded that the thermal effect significantly changes
the growth rates under small pressure drops and high temperatures in the liquid.
Chapter 3. Analysis and model development 90

R inertia- transient heat-diffusion


R* +
controlled 0 .01 < t < 50 controlled

10
Raylegh
10 (1917)

0.1 Mikic, et al.


(1970)
0.01
Plesset &
0.001 Zwick

Sauer
0.0001
(2001)

0.0001 0.01 1 100 10000 t 2 p


t+ =
R* 3 l

Fig. 3- 1. Models of bubble growth. (Mikic, et al., 1970, completed by Sauers


12 a l 3 l
(2001) solution for the initial stage of growth). Here R* = Ja 2 is the
2 p

l c p.l (Tl Ts ( pl ))
characteristic size of a bubble and Ja = is the Jacob number.
v hlg

3.2.2. Bubble collapse

Analysis of the dynamics of a collapsing bubble was made by Florschuetz and


Chao (1965). They have classified the collapse process as inertia-controlled, heat-
controlled or intermediate, based on the criterion:

inert
2 R* t coll
B eff = ~ heat (3.1)
12 3 Ro t coll
Chapter 3. Analysis and model development 91

where Ro is initial radius of a collapsing bubble, R* is the characteristic size of a bubble


(see notation to Fig. 3-1).

Fig. 3- 2. Typical behaviour of collapsing bubble at different values of criteria


B eff (Florschuetz and Chao, 1965).

When B eff > 10 the collapse is mainly driven by the pressure difference between
bubble and liquid (inertia controlled case). The inertia effects determine the initial
stage of the bubble collapse (Fig. 3-2). An approximation for the initial stage of bubble
collapse, which neglects the thermal, surface tension and viscosity effects yields the
following approximation for the variation of bubble radius with time (Brennen, 1995):

R p pv 2
= 1 t , (3.2)
2 l Ro
2
Ro

where Ro is the initial radius of a bubble. The total time of bubble collapse is:

l
t coll = 0.915 Ro .
p pv
Chapter 3. Analysis and model development 92

Fig. 3- 3. Consideration of thermal effects on the collapse of a bubble in water


( Tl , = 90oC, pl , = 0.981 bar, the degree of vapour subcooling is -10 K) (Sauer, 2001).

R R P indicates application of the Rayleigh-Plesset equation (2.53).

Fig. 3- 4. Effect of inertia on the collapse of a bubble in water, (Ro= 102 m, pl =


105 Pa, Tl = 20oC) (Sauer, 2001). R Ray and R R P indicate application of the Rayleigh

model (2.27) and the Rayleigh-Plesset equation (2.53).

The results of calculations by Sauer (2001) showed the equation (3.2) captures the
asymptotic behaviour of a collapsing bubble, (Fig. 3-3), though the Rayleigh equation
Chapter 3. Analysis and model development 93

(2.27) provides a good approximation of the average rate of bubble collapse R& R P (Fig.
3-4). Because of simplicity of numerical implementation of equation (2.27), it becomes
attractive for modelling of the collapse stage of cavitation bubbles.

p
Saturation curve
p s (T )

p l , (l , )

(v, o )

p s (Tl , )

Tl , Ts ( pl , ) T

Fig. 3- 5. States of liquid (l , ) and vapour (v, o) at the beginning of bubble


collapse in P-T diagram.

According to Florschuetz and Chao (1965), heat transfer controls the collapse
when B eff < 0.5 (approximately). Fig. 3-2 shows that the rate of bubble collapse
increases with the parameter B, which is linked to Jacob number and reflects the degree
of subcooling in liquid. Using the Clausius-Clapeyron equation for the saturation curve
(Fig. 3-5):

dps v hlg
= , (3.3)
dT Ts (1 v / l )

the temperature difference (degree of subcooling) can be related to the pressure


difference in the liquid pl , p s . Substituting this result into equation (3.1) it can be

shown that B eff ~ (T Ts )1.5 .


Chapter 3. Analysis and model development 94

The temperature difference between the bubble and surrounding liquid reduces the
rate of bubble growth and collapse also leads to the rebound of the cavitation bubble,
Fig. 3-2, Fig. 3-3 (bubble growth which follows the collapse stage due to superheating
and pressure rise in a bubble during collapse). The decay of oscillations of radius of a
collapsing bubble in liquid is due to the viscosity and surface tension effects. Except for
the last stages of bubble collapse, the temperature inside the bubble can be assumed to
be uniform (Nigmatulin, 1991).
Because of the liquid compressibility effect and non-spherical shape of bubble
during the last stages of collapse, the results of simplified analysis based on Rayleigh-
Plesset equation may be quite misleading (Brennen, 1995). However, these results
highlight important features of cavitation bubbles to be considered in a model of
hydrodynamic cavitation.

3.2.3. Conclusions
The results of analysis of the rate of growth of a single cavitation bubble in liquid
show that at high pressure drops in a liquid, if the thermal effects can be neglected, the
rate of bubble growth can be described by the Rayleigh equation (2.27).
In contrast to the bubble growth, the bubble collapse is multi-stage process.
The rate and total time of bubble collapse depend, among other parameters, on the
initial radius of bubble and Jacob number.
The Rayleigh model (2.27) can be applied to describe approximately the rate of
growth and the average rate of collapse of cavitation bubble (Sauer, 2001).
Chapter 3. Analysis and model development 95

3.3. Development of model of cavitation flow


3.3.1. Model for liquid-vapour flow

In this study it has been assumed that the liquid-vapour mixture can be described
in terms of two inter-penetrating continua. Thus, the flow may be simulated as either a
multi-phase (separate velocity vectors for the liquid and vapour), or a single phase
(one velocity vector) flow. The following simplifying assumptions have been made,
there is no slip between the continua and the continua are in thermal equilibrium. Thus
the mixture can be considered as a single phase with its physical properties varying
according to the local concentration of liquid and vapour. The governing equations for
the mixture are comprised by the set of conservation equations (2.21) (2.23),
equations for the mixture density and viscosity (2.19) and (2.20), equations for the
cavitation model and the turbulence model.
In the next section the turbulence model applied in the present study is described.
This is followed by development of the model of cavitation (Martynov, Mason, and
Heikal). Numerical simulation of cavitation flows based on their hydrodynamic
similarity). The transport equation for the void fraction and source term in this equation
to describe the rate of growth and decay of the vapour phase are derived.

3.3.2. Turbulence modelling

For the calculation of turbulent flows, the RNG (Re-Normalisation Group)


turbulence model (Yakhot and Orszag, 1986) is applied in the present study. This model
has proven to be efficient for the prediction of turbulent flows in complex geometry
configurations and flows with re-circulations. It was recommended in PHOENICS for
simulation of confined flows with separation.
As a result of decomposition of the flow variables into a mean and fluctuation
components, and consequent application of the averaging procedure, the turbulent

stresses appear ijt = u iu j in the momentum equation. In the k models of

turbulence, to describe the turbulent stresses, the Boussinesq hypothesis is applied:

2 u
ijt = uiu j = 2 t S ij m + k ij . (3.4)
3 x m
Chapter 3. Analysis and model development 96

where the turbulent viscosity t in equation (3.4) is defined as:

k2
t = c . (3.5)

Here c is a dimensionless constant ( c = 0.0845 in the RNG k model), k is

the kinetic energy of turbulence, and is the rate of its dissipation, which are calculated
from the transport equations:

k u j k k
+ = a k ( + t ) + G , (3.6)
t x j x j x j

u j u m
+ = a ( + t ) + c1G c 2 + c3 k , (3.7)
t x j x j x j k x m

u i
where G = ijt is the rate of generation of the turbulent energy, and the set of
x j

constants is: a k = a = 1.39 , c1 = 1.42 , c3 = 0.373 and c 2 is a function of the ratio of

k /
the turbulent and mean-flow time scales = :
2S ij S ij

c 3 (1 / o )
c 2 = 1.68 + ,
1 + 3

here, o = 4.38 and = 0.012 .

In the current study, the RNG k model is applied in its high-Reynolds number
version, using a standard wall function formulation (Versteeg and Malalasekera, 1995).
Since a homogeneous mixture model is used to describe the flow, the turbulence
model is applied to the mixture. The issue of interaction of the turbulent motion with
cavitation is not included in the scope of the present study.
Chapter 3. Analysis and model development 97

3.3.3. Model for cavitation

The present study is based on a single-fluid cavitation model (Yuan, et al., 2001),
which has been developed under several assumptions about the nature of the cavitation
phenomenon. First, the flow is described assuming a local homogeneous mixture of the
vapour and liquid phases. This concept neglects the actual shape of the cavitation
pockets and uses the void fraction to quantify the local content of vapour in the flow.
Second, to calculate the void fraction, a transport equation is introduced (Yuan, et al.,
2001). In this equation the rates of evaporation and condensation are derived using an
analogy with the growth and decay of spherical bubbles in liquid. To make the
homogeneous concept valid, these bubbles should be small, so that their motion relative
to the liquid can be neglected.
Void fraction. The current approach assumes that cavitation can be described
using the analogy with process of growth and collapse of bubbles in liquid. The model
uses the number density of bubbles per unit volume of liquid as parameter, which is
applied to the whole flow. Locally, cavitation is described for bubbles of one size. The
volume fraction of the vapour can then be computed from the number density and
radius of these virtual bubbles:

n 43 R 3
=
1 + n 43 R 3
(2.42)

Conservation of the number of bubbles per unit volume of liquid is used to


account for the effects of coalescence and disintegration of bubbles according to the
variation in the vapour fraction in a computational cell (Yuan, et al., 2003).

Bubble radii. To estimate the rates of growth and collapse of bubbles, the linear
model (2.27), is applied:

dR 2 pv pl
= sign( pv pl )
dt 3 l
, (3.8)
Chapter 3. Analysis and model development 98

where pv is vapour pressure, associated with the pressure inside the bubble, and pl is
pressure in the surrounding liquid, approximately equal to pressure in the
mixture pl p . This model can be considered as a limiting case of the Rayleigh-Plesset
equation, when the effects of surface tension, liquid viscosity and inertia of the bubble
are neglected.
To initiate cavitation, bubble nuclei of radius Ro are assumed to be present in the
liquid.

Concentration of bubbles. According to the original model suggested by Yuan et


al. (2001), the number density of bubbles in equation (2.42) is considered as a fixed
parameter for the whole flow. It is specified per unit volume of the liquid to account
approximately for the effects of coalescence and break-up of bubbles.

Transport equation for the void fraction. Following (Yuan, et al., 2001), the
transport equation for the void fraction is derived by taking the substantial derivative
of the void fraction, defined by (2.42), with respect to time:
d 2 dn dR
= (1 ) 43 R 3 o + 4R 2 no ,
dt dt dt

which under assumption no = const (Yuan, et al., 2001) gives:

d
= (1 ) no 4R 2
2 dR
,
dt dt

Taking into account definition for the mixture density (2.19), and the continuity
equation:

u k 1 d m l v d
= = ,
x k m dt dt

the following equation for the void fraction can be derived (Yuan, et al., 2001):
Chapter 3. Analysis and model development 99

u k d u
+ = + k =
t x k dt x k
(1 ) l no d 4 3
R
m 1 + no 43 R dt 3
3

By using (2.42) to replace R by in this equation one can get:

u k 1/ 3
4 dR
+ = l 3 no (1 ) 4 / 3 2 / 3 , (3.9)
t x k m 3 dt
1444444 424444444 3
S

which agrees with the equations (2.29) and (2.41) described in the Literature Review.

Taking into account the equations (2.42) and (3.8), the source term S in equation

(3.9) can be rearranged in the equivalent form, which involves only the void fraction
and not the bubble radius R :

f ( ) l pv p
S = sign( p v p )
l cav l
. (3.10)

Here, f ( ) = (1 )4 / 3 2 / 3 and l cav is the cavitation length scale, which can be

expressed through the number density of the bubble nuclei in the liquid n :

1
l cav =
6 3 43 n
. (3.11)

Parameters of the model. According to the model of Yuan, et al. (2001),


parameters Ro and n have to be adjusted to describe the liquid quality effects on a

cavitating flow. Equation (2.9) shows that parameter n also accounts for the
hydrodynamic scaling effect. In the present study, the model for the number density of
cavitation bubbles is developed to automatically achieve the scaling condition (2.9). To
Chapter 3. Analysis and model development 100

develop the scalable model of cavitation, the nature of parameters Ro and n is clarified
in the following section.

3.3.3.1. Models of bubbly liquids


The analysis of the stability of a single vapour-gas bubble of radius Ro in a liquid
reveals that to initiate explosive growth of the bubble, the critical tension in liquid is
required (2.12), (Fig. 3-6, a). This critical tension determines cavitation onset. Since it is
overcome, the liquid will start to cavitate and the average rate of the vapour formation
(average in time, or over ensemble of realizations) will be affected by concentration of
nuclei in the liquid.

Rcr n ncr

Ro

pv p Ro R pv p

a) b) c)
Fig. 3- 6. Variation in the critical size of vapour bubbles with the tension in
liquid (a) and actual (b) and cumulative spectra (c) of cavitation bubbles.

For the imaginary mono-disperse bubble fluid, which contain bubbles of a unique
size, the rate of cavitation would depend only on the concentration of bubbles. In real
liquids the spectrum of bubbles is not uniform (Fig. 3-6, b), and all the bubbles from the
spectrum which radii is larger then Ro contribute to cavitation process (Fig. 3-6, a). As
a result, the number of cavitation events varies with the tension in the liquid (Fig. 3-6,
c), which has been confirmed by measurements (Gindroz, et al., 1997; Waniewski and
Brennen, 1999; and Chahine, 2004). To calculate the rate of cavitation in a bubbly
liquid under tension, either the actual spectrum of bubbles should be taken into account
comprised by the model for the critical excitation of bubbles (Fig. 3-6, a, b), or the
spectrum of critical cavitation bubbles can be employed in a form shown in Fig. 3-
6, c.
Chapter 3. Analysis and model development 101

n1
n

nj

nN

Ro,1 K Ro, j K Ro, N R

Fig. 3- 7. Discretisation of spectrum of bubbles in poly-disperse model.

Poly-disperse bubbly fluids


The actual spectrum of cavitation nuclei can be approximated by a discrete
distribution function (Fig. 3-7). Then, the multi-fluid concept can be applied to describe
the evolution of each phase, representing the bubbles of initial radius Ro, j (Fig. 3-7).

Specification of the nuclei number density distribution function upstream the cavitation
region is needed as the input data for the model.
Poly-disperse models of bubbly fluids are not scalable. Actually, consideration of
the nuclei spectrum n(R) , links the parameters l cav ~ n 3 (2.44) and Ro together, and

therefore only one of them, say l cav , become sufficient for analysis. Of course the real

spectra of bubbles cannot be described using only one parameter l cav . However, even
assuming it is so, the similarity condition (2.60) can not be satisfied, as far as the
number density of cavitation nuclei, and thus the scale l cav are influenced by the
upstream pressure.

Mono-disperse bubbly fluids


Another group of models has been developed assuming that cavitation bubble of
one size can represent the whole spectrum. Because this approach describes cavitation
locally for bubbles of only one size, it may be referred as mono-disperse.
The conventional models of cavitation (Yuan, et al., 2001; Alajbegovic, et al.,
1999; Chen and Heister, 1996; Kubota, et al., 1990) assume the number density of
bubbles per unit volume of liquid no is fixed for the whole flow domain. To account for

the non-uniform spectrum of cavitation nuclei, the number density no should be


Chapter 3. Analysis and model development 102

allowed to vary with a typical tension in the cavitating flow. Then, the cavitation
bubble can be associated with the critical cavitation nuclei in liquid under typical
tension. Thus, to achieve the effect of population of bubbles within this mono-disperse
fluid concept, the spectrum of bubbles can be considered implicitly, when the number
density is a function of the liquid tension (Fig. 3-6, c).
For the arbitrary nuclei distribution function n(R) and size of nuclei Ro , the

similarity conditions (2.60) and (2.61) become inconsistent , as they require no ~ Ro3 .
However, the similarity condition (2.61) can be met in the limiting case:

Rcr / l 0 . (3.12)

This is an important condition, which completes the theoretical analysis by


Kubota, et al. (1990) and Yuan, et al. (2001).
The equation (2.10) shows that the condition (3.12) can be achieved at high
tension in the liquid or negligibly small surface tension of liquid. In Diesel injectors, the
injection pressures can be as high as 2000 bar, producing very low tensions in the
cavitating liquid, so that the limit (3.12) could be considered as reasonable
approximation to the reality.

Conclusions
Within the poly-disperse modelling concept, hydrodynamic scaling can not be
achieved. The mono-disperse models of bubbly flows provide an opportunity to
describe the hydrodynamic similarity of cavitation flows. The scalable mono-disperse
model of cavitation requires infinitesimally small radius of cavitation nuclei in the
liquid.

3.3.3.2. Model for the number density of cavitation bubbles

In the present study to describe the cavitation process, an assumption about the
hydrodynamic similarity of cavitation flows, is applied. This assumption implies that
similar cavitation flows can be described by the same set of non-dimensional governing
equations. Using the length and velocity hydrodynamic scales of mean flow l and U ,
Chapter 3. Analysis and model development 103

the dimensionless mass and momentum conservation equations of the homogeneous


cavitation flow mixture can be presented as:

~~
~ u j
~ + ~ =0
t x j
, (3.13)
~ ~~ T u~i u~ j 2
~ u~i u i u j 1 ~
p 1 u~k
+ = + ~
1 + + ij
~
t ~
x j 2 xi Re ~
~ xj ~x j ~xi 3 ~xk
, (3.14)

p p2
where u~ = u / U , ~
p= x = x / l and ~
, ~ t = t U / l are the dimensionless
p1 p 2
velocity, pressure, coordinate and time, Re is the Reynolds number (2.2).

The dimensionless form of the volume fraction equation (3.9), combined with
equations (3.10) and (3.11) is:

u~k ~
~ + ~ = S
t xk , (3.15)
~ l
S = S
U
f ( ) pv p l
= l
sign ( p v p )
l cav l U

1 p p
l
= f ( ) l 2 v
sign ( p v p )
l cav 1 U2
2 l

1 1 pv p2 + p2 p
= C f ( ) ~ sign ( p v p )
2 p1 p 2

C f ( ) p2 pv p p2
= ~ + sign ( p v p )
2 p1 p 2 p1 p 2

which may be rewritten applying the definitions for dimensionless pressure ~


p and
cavitation number CN (2.1):
Chapter 3. Analysis and model development 104

~ C f ( ) 1
S = +~
p sign( pv p )
2 ~ CN , (3.16)

~ p1 p 2 l
here = ,U= and C = is the cavitation rate constant.
l 2l l cav

Dimensionless equations (3.13) (3.16) describe similar cavitating flows for a


given type of geometry, identical initial and boundary conditions, and specified
Reynolds number Re = idem , cavitation number CN = idem, and cavitation parameter
C = idem . While the Reynolds and cavitation numbers are the similarity criteria for the
flow, the cavitation parameter C is based on the cavitation length scale l cav (3.11),

which requires modelling. To satisfy the condition C = idem , the length scale l cav

should follow the hydrodynamic length scale of the flow l :

l cav ~ l (3.17)

This conclusion was reached by Lecoffre and Bonnin (1979) and expressed in
correlation (2.9). As it was shown above, the model for the number density of cavitation
bubbles (or, in other words, cavitation length scale l cav (3.11)), is essential for
similarity condition (2.9). The following analysis is performed to develop such a model.

Cavitation length scale. According to equation (3.11), the length scale l cav is
related to the distance between cavitation nuclei, and is assumed to be an independent
parameter of the fluid, not directly linked to the hydrodynamic scale of the flow. On the
other hand, l cav can be affected by the local tension in the liquid pv p through the
density of cavitation nuclei:

n = f ( pv p)
(3.18)

So, the conclusion can be made that both the cavitation and hydrodynamic length
scales depend on the local pressure. Since the linkage between the pressure and the
Chapter 3. Analysis and model development 105

hydrodynamic length scale is established, the model for l cav can be developed based on
equation (3.17).

Hydrodynamic length scale. The aim of this part of analysis is to find a linkage
between the hydrodynamic length scale and pressure drop in the flow. An integral
hydrodynamic length scale of the flow, which determines the cavitation rate constant C
in (3.16), can be associated with the hydraulic diameter of nozzle:

l ~ D (3.19)

Substitution of the Bernoulli velocity scale U in the definition of Reynolds


number (2.2), and rearranging for D gives:

Re l
D=
2 l ( p1 p 2 )
(3.20)

Using (3.11), (3.19) and (3.20), the similarity condition (3.17) takes a form:

n ~ ( p1 p 2 ) 3 / 2 . (3.21)

Equation (3.21) shows that the bubble number density is proportional to the
pressure drop across the nozzle, but according to equation (3.18) a relationship in terms
of the local liquid tension is required. Applying a simple one-dimensional analysis for
separation flow in a nozzle (Fig. 2-6), an estimate for the pressure drop in the vena
contracta region can be made. This can be applied to determine the average tension
when the liquid cavitates pv pmin and to put the correlation (3.21) in the same form
as equation (3.18).
For a cavitation-free flow in a sharp-entrance nozzle, the pressure drop in the vena
contracta region (Fig. 2-6) can be estimated by applying one-dimensional analysis when
friction losses are neglected (Chisholm, 1983):

p 2 p min = ( p1 p 2 ) ,
Chapter 3. Analysis and model development 106

1
Cc 2 2
where = 2 1 is the nozzle-specific constant, which is
C 2 C 1
c
2
c
1
( )
expressed through the contraction coefficient C and ratio of the cross-section area of
the nozzle and the area of the upstream plenum = A / A1 (Fig. 2-6).

In the limit of high injection pressures p1 >> p2 (that is typical for Diesel

injection conditions, when p1 ~ 108 (Pa) and p2 ~ 106 (Pa)), the above equation gives

an approximate relationship ( pmin ) ~ p1 , which assuming pmin << pv may be


rewritten in a form:

pv pmin ~ p1 p2
.

(At high injection pressures, when p1 >> p2 , the maximum tension in the flow can be

estimated as pv pmin p1 .) Substitution of this correlation into (3.21) gives the

modelling equation:

3/ 2
p p min
n = n* v
pv , (3.22)

where n* is the number density of cavitation sites when pmin = 0 . The value for this
parameter is specified in order to match the output of the model with experimental data
from a particular nozzle. Once this value has been determined then the model can be
used to simulate the flow of the same fluid through different scale nozzles.

3.4. Model scalability


In the present model, in order to evaluate the number density of active cavitation
nuclei (3.22), the maximum value of liquid tension in the cavitation region p v p , is
used. From this value the number density of cavitation nuclei is obtained. This approach
Chapter 3. Analysis and model development 107

could be useful at least for simple quasi-steady-state flows when the cavitation region is
localized in the flow domain.
The resulting equation (3.22) takes into account the effect of local pressure on the
number of cavitation bubbles in the liquid. In comparison to the models (Chen and
Heister, 1996; Yuan, et al., 2001), the present model achieves the effect of
hydrodynamic scaling by equation (3.22), and therefore contain parameters Ro and n* ,

which depend on the liquid quality only. The fact that the present model is consistent
with the requirement of hydrodynamic similarity of cavitation flows can be shown by
substitution of the modelling equation (3.22) into the source term of equation (3.16):

~ Re f ( ) l 1 1
S = * +~
p ~
p min sign( p v p )
l cav 2 ~ pv l CN CN
, (3.23)

1
l cav ,* =
6 3 43 n*
here .

The source term (3.23) does not include any hydrodynamic scales of the flow
which distinguishes the present model from a number of cavitation models found in the
literature (Yuan, et al., 2001; Alajbegovic, et al., 1999; Kubota, et al., 1990; Chen and
Heister, 1996). As a result, the set of equations (3.13), (3.14), (2.44), (2.45) and (3.22)
describe similar cavitation flows for a specified geometry, Re and CN numbers, and
initial and boundary conditions.

3.5. Populations of cavitation bubbles


The number density of nuclei is an important property of a liquid, which can be
measured (Billet, 1985). In real liquids, the nuclei are comprised of small particles, gas
bubbles and impurities and can be described by a distribution function no (R ) . Nuclei of
radius larger than the critical value can grow, while smaller bubbles stay in the liquid
(Brennen, 1995). Description of the effect of nuclei population on cavitation inception
and scaling is a problem of practical importance, which has been the subject of
numerical and experimental studies (Rood, 1991; Gindroz, et al., 1997; Liu and
Chapter 3. Analysis and model development 108

Brennen, 1998). The results of these studies can be applied in a model of cavitation, to
make the number density of cavitation nuclei a flow variable.
This part is aimed to review the measurements of populations of cavitation
bubbles in real liquids, and to identify the shape and parameters of number density
distribution function. Then these results can be applied to identify the reliability and
range of application of equation (3.22).

3.5.1. Measurements of nuclei populations

Fig. 3- 8. Measurements of the number density distribution function for


cavitation nuclei in waters of different purity (Billet, 1985; Sato and Kakutani, 1994).

The population, or spectrum of bubbles in a liquid can be described by the number


density distribution function, N (R ) , which gives the number of nuclei of a radius R
within the range from R to R + dR .

At present, several experimental techniques have been developed to measure the


bubble density distribution function, namely (Billet, 1985) analysis of samples by
Coulter counter; acoustic and light scattering techniques; reconstruction of a hologram
Chapter 3. Analysis and model development 109

of small volume of the flow (the most reliable method); and cavitation susceptibility
meters (the most popular method). Experimental studies of the spectra of bubble nuclei
in liquids were reviewed by Billet (1985), Rood (1991) and Brennen (1995). These
experiments have shown that the population of nuclei can vary from liquid to liquid, and
depends on the presence of small contaminant particles, free and dissolved gases, and
also the temperature and pressure conditions. Most of these studies have been
performed over limited range of bubble nuclei (Fig. 3-8).
Because of the large dispersion in the liquid quality data, errors in measurements
of the number density distribution function can be of the order of its magnitude,
(DAgostino and Acosta, 1991).

Fig. 3- 9. Cavitation nuclei distribution functions in water of different quality


(Gindroz, et al, 1997). T1 large nuclei; T2 high content of medium-size nuclei; T3
low content of medium-size nuclei; T4 strong degassed water.

Sometimes the results of measurements of nuclei population are presented in a


form of the number density of bubbles or registered cavitation events as a function of
applied tension in liquid (cumulative spectrum of nuclei, Lecoffre, 1999). This is the
usual practice when the size of nuclei is not accessible from measurements and/or it is
difficult to identify the nature of the nuclei. Thus, Gindroz, et al. (1997) have measured
the number density distribution function using the liquid susceptibility method for
different qualities of water (Fig. 3-9). This data clarifies the nuclei effect on the tensile
strength of the liquid, showing that the tensile strength of liquid decreases with the size
of nuclei. A similar format is used to describe the data on the number of nuclei sites in
Chapter 3. Analysis and model development 110

boiling flows, when the number density of nuclei is considered as a function of the
degree of superheat of the liquid. For example, Fujimoto, et al. (1994) have
approximated the number density of surface nuclei by the correlation:

5.279
n = 5.757 1012 exp , m3.
T Ts

Specific parameters of this equation have been adjusted to fit the measured nuclei
densities of n-pentane and n-hexane flows in injection nozzles.
Experimental studies by Ling, et al. (1982) and Peterson (1972) (referred by Rood,
1991), Sato and Kakutani (1994) and also Gindroz, et al. (1997) have revealed that the
number density of cavitation nuclei increases with amount of gas dissolved in liquid,
while the system pressure decreases the concentration of cavitation nuclei (obviously,
the system pressure affects the size of the bubbles).

3.5.2. Approximations for the spectrum of nuclei

0 R

Fig. 3- 10. Nuclei distribution function.

As a result of the generalisation of experimental data, Liu and Brennen (1998)


applied the log-normal distribution function to describe the population of bubble nuclei:
Chapter 3. Analysis and model development 111

dno log e 1 log( R / ) 2


= exp
dR 2 R 2
, (3.24)

where , and are parameters. Equation (3.24) is a rather general approximation


for the nuclei distribution function, and can be applied for different fluids by
appropriate choice of parameters , and .
Equation (3.24) is plotted in Fig. 3-10. It accounts for the most important features
of the number density distribution function, such as the presence of the peak at R =
and rapid decrease in the number of nuclei at large radii (Liu and Brennen, 1998).
Parameter describes the sharpness of distribution function. The number density
typically decreases with the size of nuclei (left part of the curve, Fig. 3-10) and
therefore, nuclei of smaller size and higher number density become activated at higher
tensions.
Often, simplistic power-law approximations to the nuclei distribution function are
applied in practice, for engineering and scientific calculations (Billet, 1985; Wang,
2001):

bN
N ( R) = (3.25)
Rm

where bN is a dimensional parameter and m is the shape factor of the distribution


function.
Thus to describe measurements of number density distribution function of
cavitation nuclei in water in the range from Rmin = 10m to Rmax = 200 m , Wang

(2001) have applied equation (3.25) with parameters:

m = 4 , and

3 o
bN = ,
4 (1 o ) ln( Rmax / Rmin )

where the upstream void fraction o takes into account the air content in the liquid.
Chapter 3. Analysis and model development 112

3.5.3. Conclusions

The critical tension (tensile strength) of macroscopic volume of liquid, which


contains high enough number of nuclei, depends on physical properties of the liquid and
the liquid quality, i.e. presence of small bubble nuclei and contaminating particles.
In the current study, the continuous approach is applied to describe the cavitation
flow, and therefore properties of the flowing liquid, including properties of the bubble
nuclei, are considered locally at each particular point in the flow.

3.6. Validation of the model for concentration of cavitation


bubbles

Fig. 3- 11. Concentration of cavitation bubbles in liquid under tension.


Approximation of measured spectra of bubbles at different oxygen contents (ppm) and
system pressures (bar) in water (Gindroz, et al., 1997) (points) by equation (3.22) with
the variable parameter n* (lines).

Measurements of the spectra of cavitation bubbles are commonly characterised by


large dispersion (Brennen, 1995). This is caused by the significant variation in the liquid
quality depending on the system pressure, presence of contaminating particles and gases
in the liquid. The experimental data by Gindroz, et al. (1997) shown in Fig. 3-11
Chapter 3. Analysis and model development 113

describes the measured spectra of cavitation bubbles in water at different system


pressures and different air content. The measured concentrations of cavitation bubbles
represents the whole spectra of bubbles in the liquid of radii larger than the critical
value (3.22), in the liquid under tension pv p . The number density calculated from

the equation (3.22), which considers the average critical bubble from the spectrum
has a similar meaning.
To validate the model (3.22) measurements of bubble spectra are required for the
conditions when the effects of contaminating particles and gases can be neglected. In
the measurements performed by Gindroz, et al. (1997), the effects of contaminating
particles on cavitation are negligible, at least at relatively low system pressures and

liquid tensions ( pv p < 105 Pa). The data by Gindroz, et al. (1997) also describe the
spectra for reduced concentrations of the dissolved air (concentration of the air/ oxygen
at saturation and atmospheric pressure in water is between 9 ppm and 24 ppm).
Therefore, the data by Gindroz, et al. (1997) are suitable for comparison with the model
(3.22), suggested in the present study.
Fig. 3-11 shows that the slope of the curves described by the equation (3.22) fits
well the measured data at low system pressures (0.5 bar). On the other hand, at higher
system pressures (Fig. 3-11), the correlation (3.22) did not match the measured data.
This deviation could be explained by the activation of the heterogeneous mechanism of
nucleation, when the cavitation bubbles start growing from small sub-micron nuclei
attached to the suspended particles in the liquid. This can be inferred from the Fig. 3-11,
which shows that at system pressure of 1.5 bar the number density of cavitation bubbles
is unaffected by the oxygen content. As a result, at high system pressures, the measured
data (Fig. 3-11) could be affected by the population of particles, which at high
concentrations increases the number of registered cavitation events. At low system
pressures (0.5 bar), when the effect of particles on the number of cavitation events is
eliminated, the shape of the nuclei distribution function is in a good agreement with the
measurements, that validates the equation (3.22) for calculation of cavitation flows.
Chapter 3. Analysis and model development 114

3.7. Discussion of the range of the model application


The premature and gas cavitation are not described by the model, since it neglects
the initial nuclei upstream of the cavitation region and their stable growth/collapse due
to pressure variations in the flow. The scalable model of cavitation requires that
condition (3.12) be met, which can only be achieved approximately at high liquid
tensions and when surface tension of liquid is negligible. For the above reasons,
cavitation inception (and flows at low cavitation numbers CN) can not be accurately
predicted.
The model assumes that the bubbles are filled only with the vapour phase.
Therefore, the flow regimes, such as super-cavitation and hydraulic flip, which require
consideration of the air entrainment from downstream of the nozzle, can not be
described.
The model for the number density of cavitation bubbles (3.22) requires an
estimation of the maximum tension in the liquid for the particular cavitation region. The
model achieves the effect of liquid tension on cavitation flow on average (over the
whole flow, not locally). This limits the application of the model to calculation of quasi-
steady-state cavitation flows, in the absence of multiple cavitation structures.
Preliminary calculations of large-scale flows, for which bubbles were detected
experimentally, have revealed insufficient pressure drop inside the nozzle to activate
vaporous cavitation. In large-scale flows at relatively low pressures, gas-filled bubbles,
which are present in the liquid, could perform as active nuclei for the cavitation process.
However, when the tension in the liquid is below the critical level, vaporous cavitation
will not occur. Therefore, accurate simulation of cavitation in large-scale (low-pressure)
flows requires consideration of the free gas content, i.e. the dynamics of gas-filled
bubbles in the flow. Such effects are not described in the present model, and therefore,
experiments at low system pressures were used for validation and testing purposes.

3.8. Conclusions for chapter 3


In the conventional models of cavitation, based on the mono-disperse concept of
bubbly fluid, such as the model developed by Yuan, et al. (2001), the number density of
cavitation bubbles n is specified as a parameter. Computational practice has proved the
validity of the concept n = idem for prediction of the variety of cavitation flows of a
Chapter 3. Analysis and model development 115

given geometry configuration and scale. However, analysis of the phenomenon shows
that the parameter n must be a function of the liquid tension, which, in turn, depends on
the cavitation flow itself. This makes the hypothesis n = idem not accurate for
prediction of cavitation in flows of different hydrodynamic scales.
To permit the hydrodynamic scaling of cavitation model, the number density of
critical nuclei is considered as a function of tension in the liquid (equation (3.22)). The
model is developed from an assumption of hydrodynamic similarity of cavitation flows
that demands the relationship:

l cav ~ l (3.17)

which follows from (2.60) in consistency with the condition (2.9), suggested by
Lecoffre and Bonnin (1979).
The model accounts for the population of cavitation nuclei in the liquid in an
implicit way n( p v p ) . The hydrodynamic similarity for the solution can be achieved

only in the case when the spectrum of cavitation bubbles follows the power-law
approximation n ~ ( p v p ) 3 / 2 .

The advantage of equation (3.22) is that it determines the number density from the
local tension in the liquid, and avoids application of the relationship (3.17) to guess an
appropriate value for l cav . This means that the model accounts separately for the flow

and liquid-quality effects and require only one set of data for the particular type of
liquid to adjust parameter n* in equation (3.22).

A concept n* = idem is different to the conventional assumption n = idem . The

purpose of the following section is to validate the concept n* = idem for simulation of
cavitation flows in nozzles.
Chapter 4. Numerical procedure 116

Chapter 4. Numerical method and procedure


4.1. Introduction
The present study employs the homogeneous-mixture approach to describe
cavitating flows. This approach assumes that within a control volume, the liquid and
vapour phases are uniformly mixed. However, the liquid and vapour may be non-
uniformly distributed over the flow domain. A consequence of this assumption is that
there will be no sharp inter-phase boundaries. The properties of fluid are calculated on
the basis of properties of pure phases, and their local volume fractions in the flow. In
the preceding chapter, the mathematical description of the model of cavitation flow
was given.
To solve the governing equations (2.19) (2.23) for arbitrary initial and
boundary flow conditions, numerical methods are applied. In this study the
PHOENICS CFD code (abbreviation for the Parabolic Hyperbolic Or Elliptic
Numerical Integration Code Series) (Rosten and Spalding, 1986) is employed as an
instrument to solve the fluid dynamics problem using the finite-volume method
(Versteeg and Malasekera, 1995). The PHOENICS package provides a number of
subroutines and algorithms, which user may select to calculate the heat, mass and
momentum conservation equations. The package allows modelling of the flow-related
phenomena described by the transport convection-diffusion equation, which terms can
be specified in the user-defined subroutines. The current version of PHOENICS has
no built-in option to calculate cavitation flows. Therefore the model of cavitation
described in the previous chapter was incorporated into the package. To meet the
demand of calculation of flows of arbitrary geometry configuration, body-fitted non-
orthogonal grids are used.
This chapter describes the main specifications of the numerical model and
solution procedure applied for the calculation of three-dimensional quasi-steady-state
flows with cavitation. This concerns the method applied to solve the set of governing
conservation equations, description of high-order convection discretisation schemes
applied to improve accuracy of computations on coarse meshes, and modifications to
solver for calculation of virtually compressible cavitation flows with large density
gradients.
Chapter 4. Numerical procedure 117

4.2. Finite-volume method


To solve the governing equations (2.19) (2.23) completed by the model of
cavitation and equations for calculation of the turbulent viscosity, and a set of initial
and boundary conditions, a pressure-based solution algorithm together with the finite-
volume discretisation method, is applied (Versteeg and Malalasekera, 1995; Ferziger
and Peric, 2002). The equations are solved for the control hexahedral volumes, which
form the cells of the body-fitted mesh. All the variables are allocated at the centres of
control volumes (collocated grid).

Y
N
L
n
l
W

P e
w E

h
s X

S
Z H

Fig. 4- 1. Control volume associated with the node P in three dimensions.


Neighbouring nodes are E, W, S, N, L and H, and cell faces e, w, s, n, l and h.

For the generalised dependent variable the conservation equation can be


written in the form:

u k
+ = +S (4.1)
t x k x j x j

For the finite-volume discretisation, this equation is integrated over the control
volume for each cell of the domain (Fig. 4-2) (Versteeg, Malalasekera, 1999). After
replacement of the volume integrals by the surface integrals for the convection and
Chapter 4. Numerical procedure 118

diffusion terms, the conservation equation describes the balance between the rate of
change, advection, diffusion and source/sink for the variable :


V P + u n , f A f A f = S VP ,

(4.2)
t P f n f

where summation is made over all six faces f = e, w, s, n, l , h .

r
uf

f
r
P nf

VP F
Af

Fig. 4- 2. Velocity and its normal component at the cell face f . A f is area of

the face f between the cells P and F, and VP is volume of the cell around the node
P.

In the finite-volume method, the terms constructing this equation are discretised
separately. This study is devoted to numerical modelling of quasi-steady state
cavitation flows. Therefore, in the following description, the terms with the time
derivative are not considered.
The central-differencing second-order scheme is commonly applied for
discretisation of the diffusion term in (4.2):

P
A f = (A) f F = D f ( F P ) , (4.3)
n f x F xP

This transforms equation (4.2) to:


Chapter 4. Numerical procedure 119

(C
f
f f D f ( F P )) = S VP (4.4)

where index F denotes the neighbouring node at cell face f , C f = f u f A f ,

D f = f A f /( x F x P ) and f is the property at the cell face.

Because the variable is defined at the nodal points in centres of control


volumes, special convection-diffusion discretisation method is needed to calculate
f . When a convection scheme for the calculation of the cell-face value f is
chosen, equation (4.4) results in a set of linear algebraic equations:

a P P a F F = S P (4.5)
F

where a F and a P = a F are matrix coefficients, defined by the convection


F

discretisation method. These equations completed by the boundary conditions, have


similar structure for the flow variables and can be solved using the appropriate
methods of linear algebra.
In order to preserve the diagonal dominance of the matrix of coefficients that
require a P a F at all nodes and a P > a F at least at one node (Versteeg
F F

and Malalasekera, 1995), the source term in equation (4.5) is treated imlicitly when it
is negative and explicitly when it is positive. The general practice is to represent the
source term in a linearised form:

S P = s P P + bP , (4.6)

where s P 0 . Substitution of this equation into (4.5) gives:

(a P s P ) P a F F = bP . (4.7)
F
Chapter 4. Numerical procedure 120

When the density and transport coefficient vary from node to node, interpolation
schemes are needed to calculate the values f and f at the cell faces. The general

practice is to use the arithmetic mean or harmonic averaging between the values in
adjacent nodes. For flows with sharp density gradients, such as shock waves or flows
with a phase transition, upwind densities are used in approximations for the mass
fluxes (Malin and Sanchez, 1988). For example for the u -component of velocity (Fig.
4-1):

(uA)e = P Ae max(0, u e ) + E Ae max(0,u e ) . (4.8)

Application of collocated grids requires calculation of velocities at the cell faces


(such as u e and u w ). These velocities appear in convection terms in the dicretised

momentum equation, and also expressions for the mass fluxes at the cell faces (4.8).
To evaluate the cell-face velocities the Rhie-Chow interpolation method (Rhie and
Chow, 1983) is applied in PHOENICS.

4.3. Convection discretisation schemes


The convection discretisation scheme is used to calculate the face values f

which would account for the balance between convection and diffusion fluxes.

4.3.1. Hybrid scheme

One popular approach is to use the hybrid discretisation scheme (Spalding,


1972), which switches between the upwind and central-differencing schemes
depending on cell Peclet number Pe f = | C f | / D f .
Chapter 4. Numerical procedure 121

UU U f D

Fig. 4- 3. Nodes involved in the approximation formula for the cell-face value
f in two-point and three-point convection schemes in one dimensional case (D
downwind, U upwind, and UU upwind-upwind).

For one-dimensional flow the hybrid scheme involves two nodes (Fig. 4-3):

U , if Pe f > 2
f = 1
2 (U + D ), if Pe f 2

The hybrid scheme is conservative, transported and bounded (Versteeg and


Malalasekera, 1995). Due to its stability and robustness the hybrid scheme has
become very popular in CFD codes, thus it is the default scheme in PHOENICS.
However, for flows with a dominant convection direction, the hybrid scheme turns
into the first-order up-wind scheme, that leads to high numerical diffusion and loss of
accuracy. To obtain a higher order of accuracy, and at the same time to keep the
stability and boundedness of the approximation, several high-order convection-
discretisation schemes have been developed in the literature and included in
PHOENICS (Malin and Waterson, 1999).
Application of high-order convection schemes becomes essential to obtain a
more accurate solution on moderately fine meshes. This concerns approximation of
the momentum equation when calculating flows with separation. For the mass
fraction, which naturally contains no diffusion term, application of the high-order
schemes is also required.
The non-linear high-order schemes use flux-limiters to enforce the boundedness
criteria (generate cell-face value f within the range of node values, U , and UU ):

f = U + 12 B (r ) (U UU ) ,
Chapter 4. Numerical procedure 122

D U
where B (r ) is the limiter function and r = is the gradient ratio.
U UU

4.3.2. SMART scheme

The SMART convection scheme is a bounded Quick scheme with piecewise


linear interpolation. The limiter function of the SMART scheme is defined as:

B(r ) = max[0; min(2 r ; 0.75 r + 0.25; 4)] (4.9)

The SMART scheme provides the second order of approximation.

4.3.3. Super-bee scheme


The Super-bee scheme of Roe (1986), is constructed of upwind, central-
differencing and second-order upwind schemes:

B(r ) = max[0; min(2 r ; 1); min( r ; 2 )] (4.10)

It provides better results in resolving the sharp gradient regions using a monotonic
second-order upwind scheme. It is known to be over-compressive and will sharpen the
regions with smooth gradients. However, the super-bee scheme is excellent for
reconstruction of discontinuities in scalar variables governed by the diffusion-free
transport equation.

4.3.4. Implementation of high-order schemes

To provide diagonal dominance of the coefficients in the matrix equations, high-


order convection schemes are commonly implemented using the deferred-correction
formulation (Ferziger and Peric, 2002), when the cell-face value is presented as build-
up to the upwind differencing scheme:

fhigh = U + f .
Chapter 4. Numerical procedure 123

and the high-order contributions f are calculated explicitly from the previous

iterations in the source term of equation (4.7):

aF (P F ) = sP C f ( f U ) , (4.11)
F f
14 4
42444
3
deferred correction

where index f denotes a cell face between the nodes P and F (Fig. 4-2), and the

matrix coefficients a F are defined according to upwind scheme:

a F = max(C F ,0) + DF , C F < 0 . (4.12)

Mailn and Waterson (1999) have noticed that piecewise-linear limiters, which
are used in the SMART and super-bee schemes to switch between the linear high-
order schemes and ensure the boundedness criterion, may induce convergence
problems. To improve the stability of calculations, calculations can be restarted from
the solution obtained using a hybrid scheme (Mailn and Waterson, 1999).

4.4. Boundary conditions


The set of governing equations (momentum, continuity, equations of turbulence
and cavitation models) describes subsonic turbulent three-dimensional steady-state
flow of the homogeneous mixture of variable phase content. Solution of the governing
equations can be found for a given geometrical configuration and specified boundary
conditions. The present study is focused on computation of cavitation flows in
nozzles, with the geometry shown schematically in Fig. 2-6. Computations are
performed using the finite volume discretisation method and non-orthogonal body-
fitted meshes built in the computational flow domain. The boundary conditions are set
up for a part of the actual flow domain formed between the inlet, outlet, nozzle walls
and symmetry planes. Here they are described without details about their
implementation in the finite-volume method, which were discussed by Versteeg and
Malalasekera (1995).
Chapter 4. Numerical procedure 124

4.4.1. Inlet/Outlet

At the inlet of the flow domain, upstream of the nozzle entry (point 1, Fig. 2-6),
the velocity of the flow is specified. The turbulent intensity at the flow inlet is set to
zero. At the outlet of the nozzle, the constant pressure boundary condition is applied.
When the inlet velocity is set up according to the measured mass flow rate,
inaccuracies of the calculations will affect the pressure field. At the same time, the
cavitation model is sensitive to local pressures. As a result the mass flow rate was
adjusted in some simulations in order to produce the same pressure drop as that
observed experimentally before comparisons were made between the predicted and
observed cavitation structures.
When modelling supercavitation flows it becomes essential to take into account
the interaction between the vapour pocket and the high-pressure boundary at the exit
of the nozzle (Dumont, et al., 2001). This requires the application of a transient
compressible flow solver to describe more accurately the collapse of the vapour
pockets at the nozzle exit. The current model uses a steady-state approximation for the
rate of collapse of a cavitation bubble. Therefore this study is focused on the
description of developing cavitation flows, when the vapour structures do not extend
to the nozzle exit.

4.4.2. Walls

At the nozzle walls the no slip boundary condition was applied for the velocity
components parallel to the wall, while the velocity in the direction normal to the wall
was set to zero. The boundary conditions for the momentum equation were formulated
using the standard equilibrium wall functions (Versteeg and Malalasekera, 1995). The
method also assumes that the turbulent boundary layer can be subdivided into viscous
(laminar) sub-layer and logarithmic layer, so that the mean velocity profile can be
approximated by:

y+ y + 11.63

u+ = 1 + +
ln( Ey ) y > 11.63
Chapter 4. Numerical procedure 125

where u + = u / ut is the dimensionless velocity, y + = y ut / is the dimensionless

coordinate in the direction normal to the wall, =0.41 is the von-Karman constant,
and E=9.8 is the parameter for a smooth wall. The turbulent velocity scale ut is

defined assuming an equilibrium turbulent boundary layer:

ut = c1/ 4 k .

The wall functions formulation is appropriate for the simulation of the turbulent
flows with separation, when the wall shear stress can turn to zero at the flow
reattachment point. This standard equilibrium log-law wall functions concept is
known to be inaccurate when applied to modelling recirculation flows, which requires
application of the generalised wall functions concept. However, in PHOENICS the
Collocated Covariant Method used in PHOENICS for computation of flows on body-
fitted meshes accept only equilibrium wall functions. This circumstance can reduce, to
some degree, the accuracy of the prediction of the flow reattachment point, but should
not greatly affect the predicted flow pattern.

4.4.4. Symmetry planes

At the symmetry planes, the normal component of the velocity is set to zero, and
for all other variables zero gradients are applied, allowing no flow and no fluxes
across the symmetry plane.

4.5. Mass fraction equation


To calculate the cavitation flow in PHOENICS, the basic set of continuity and
momentum conservation equations are completed by the transport conservation
equation for the mass fraction of the vapour phase. Using the volume fraction
equation (4.11) together with the linear model (2.27) and definition for the mass
fraction f v / the mass fraction equation can be derived:

f f uk
+ = Sv , (4.13)
t x k
Chapter 4. Numerical procedure 126

where the source term is:

1/ 3
| pv p |
Sv = C (1 f ) 4/3
f 2/3
v sign ( pv p )
l l

According to the linearization practice (4.6), it is useful to rearrange the source


term S v in a form:

S v = C * 1 ( f ) f 2 ( p v p ) , (4.14)

(1 f ) 1 f
1/ 3

where 1 ( f ) = f .

f + v (1 f )
l

When the local pressure is above p v and the source term (4.14) is negative
(evaporation stage), it is calculated using the mass fraction from the current iteration
(implicitly). Conversely, when the pressure is below p v , the source term is positive
and calculated from the previous iteration.

4.6. Solution algorithm


To solve the set of governing equations on non-orthogonal body-fitted grids, the
Collocated Covariant Method is available in PHOENICS (Poliakov and Semin, 1994).
This method uses the SIMPLE-like pressure-correction algorithm and Rhie-Chow
interpolation for the velocities at the cell faces. The sets of discretised equations (4.7)
closed by the boundary conditions are solved in sequential manner using the LU-
preconditioned conjugate residuals method. First, guessing the pressure field,
equations are solved for the velocity field. Then the pressure correction equation is
solved and velocity corrections are made. After this, the transport equation for the
mass fraction of vapour is solved and properties of the mixture are renewed. Then
calculation procedure is repeated until convergence.
Chapter 4. Numerical procedure 127

4.7. Consideration of the mixture compressibility


The advantage of the homogeneous mixture approach is that the computational
efforts are minimised by application of a single-fluid solver to calculate the flow. In
the present study both phases are considered as incompressible. This permits
utilisation of incompressible single-fluid algorithms. However, according to the nature
of cavitation flow, the pressure determines the local phase content in the mixture and
that makes it effectively compressible. To provide a stable and convergent numerical
algorithm, these specifics should be taken into account when solving the discretised
equations. Two methods of stabilisation of the pressure correction algorithm for
calculation of flows with steep density gradients were reviewed in chapter 2.4 (the
GALA method by Spalding, 1974; and a compressible form of the pressure correction
equation proposed by Senocak, 2002). Because the GALA algorithm requires solution
of the volumetric continuity equation with non-zero source term, its implementation is
not straight-forward in the conventional frame-work of CFD codes. On the other
hand, PHOENICS uses the compressible form of the pressure correction equation
when the compressibility of the fluid (DRH1DP variable) is set to a non-zero value.
The fluid compressibility can be specified as a number or as a flow-dependent
variable in the Q1 input file (Appendix A) and user-defined GROUND module
(Appendix B).

4.8. Relaxation and convergence


To assess the convergence of iterations, the sums of residuals are calculated for
each flow variable over the whole flow domain. At each computational cell residuals
represent imbalances in the finite-volume equation (4.11):
P = a P P [a F ( F P )] S P . (4.15)
F

The criterion applied to stop iterations for a particular variable is:

P < o
P
Chapter 4. Numerical procedure 128

where o is desired degree of convergence. To ensure a converged solution the

residuals should be reduced and the flow variables should have virtually stopped
changing.
To control the convergence and avoid destabilization of iterations, an under-
relaxation strategy was applied to the flow variables (Rosten and Spalding, 1986).
Two types of relaxation are available in PHOENICS, namely linear relaxation and
false-time relaxation. Relaxation does not alter the final solution, but affect only the
way in which it is achieved.
With linear relaxation the value of the flow variable which is accepted at the
current iteration is calculated as a linear combination of values obtained at the current

step new and values from the previous iteration old :

(
= old + new old , ) (4.16)

where < 1 is the under-relaxation factor for the variable .

Alternatively, false-time relaxation can be used, which modifies the finite-


volume equation (4.5) by adding an additional pseudo-transient term:

PV p
t
( P )
Pold + a P P a F F = S P (4.17)
F

where t is the false time-step for the variable . Because the false-time relaxation

is implemented implicitly in equation (4.17), there are no stability restrictions on the


value of t .

4.9. Conclusions for chapter 4


The PHOENICS package provides a single-fluid incompressible flow solver
with standard routines for the calculation of the transport of a passive scalar and a set
of options to the improve accuracy of the resolution of sharp interfaces and
calculations of variable-density flows. Several high-order linear and non-linear
convection-discretisation schemes are available to the user. The super-bee scheme is
Chapter 4. Numerical procedure 129

recommended for reconstruction of interfaces described by the diffusion-free transport


equation. To calculate flows with phase transition the mass conservation equation can
be solved in PHOENICS. The source term for the mass fraction equation have to be
defined in the user-defined subroutines (GROUND module, Appendix B).
To promote convergence of pressure-solution algorithm for computation of
cavitation flows, compressibility of the fluid can be considered in the formulation of
the pressure correction equation.
Chapter 5. Results 130

Chapter 5. Application
5.1. Introduction
In this chapter the model of hydrodynamic cavitation is applied to simulation of
quasi-steady state flows in models of Diesel injection nozzles (Roosen, et al., 1996;
Winklhofer, et al., 2001). The model uses the homogeneous approach, and is
comprised of a sub-model for the rate of phase transition, sub-model for concentration
of the cavitation bubble nuclei and sub-model for the critical pressure in liquid. The
numerical results demonstrate that the model can describe the features of real
cavitation flows.
In the preceding chapters, the model for the number density of critical cavitation
nuclei (3.22) was derived assuming the hydrodynamic scaling of cavitation flows, and
compared with direct measurements of cavitation bubbles in a liquid. In this part, by
application of equation (3.22) to similar flow regimes in nozzles of different scales,
the scalability of the cavitation model is examined. The model is verified by
comparison between the numerical results and the measurements of cavitation flows
in a small-size injection nozzle at various cavitation numbers CN (Roosen, et al.,
1996). A separate study is performed to establish the initial mass fraction parameter of
the cavitation model.
To account for the effect of the viscous shear stress on cavitation flow, a model
for the critical vapour pressure in the liquid has been developed. To verify the
hypothesis regarding the effect of liquid stress on the critical pressure (2.13),
measurements of cavitation in nozzles at high system pressures have been applied
(Winklhofer, et al., 2001).
The cavitation model was implemented within the framework provided by the
PHOENICS CFD package version 2.2 (Rosten and Spalding, 1986; www.cham.co.uk).

5.2. Calculation of cavitation flows in nozzles


5.2.1. Introduction

The purpose of this study is to develop a model of cavitation, which would be


able to account for the flow scale effects on the cavitation onset and predict pattern of
Chapter 5. Results 131

the developing quasi-steady-state cavitation in nozzles. The mathematical model of


the hydrodynamic cavitation was described in the previous chapters. It includes the
set of momentum and continuity equations (2.21) (2.23), the equations of turbulence
model (3.5) (3.7), the relationships for calculation of the mixture properties (2.19)
and (2.20), and equations for the cavitation model (3.15), (3.16), (3.11) and (3.22).
Solution of these equations can be found for a given geometry of the flow and specific
initial and boundary conditions.
The following section describes the main specifications to the PHOENICS
solver for the calculation of quasi-steady-state turbulent cavitation flows. In
PHOENICS, geometry of the flow and all the parameters of numerical algorithm for
computation of mixture flow are set-up in Q1 input file (Appendix A). User-defined
subroutines for the calculation of the source term in the mass-fraction equation are
defined in the GROUND module (Appendix B).

5.2.2. Strategy of calculation

To calculate steady-state cavitation flows, the following two-step strategy is


applied. First, the pressure and velocity distributions are guessed for a specific flow
condition neglecting cavitation. Positive liquid tensions p v p indicate the regions

of the flow where the cavitation is expected to develop. Using this liquid tension the
number density of cavitation bubbles in the liquid can be estimated from the
correlation (3.22). After this, the calculations are accomplished with the cavitation
model being switched on, starting from the cavitation-free flow field.

5.2.3. Parameters of cavitation model

The present study is focused on the development of the model of vaporous


cavitation, which would account for the nature of cavitation onset and developed
cavitation flow. For this purpose, the model of cavitation is comprised of a sub-model
for the number density of critical cavitation nuclei and a sub-model for the shear-
stress induced cavitation. These sub-models are required to predict cavitation flows
for a variety of hydrodynamic scales and system pressures. The methods of
specification of parameters of these models are described in the following parts of this
chapter.
Chapter 5. Results 132

In the present study simulations of cavitation flows were performed for the tap
water and Diesel fuel. Thermo-physical properties of these liquids are listed in Table
2-1. The physical properties of the vapour phase, namely density and viscosity,
determines the flow behaviour in cavitation region. The coefficients of viscosity of

the vapour were set to v = 5 10 6 m 2 /sec for Diesel fuel and

v = 1 10 5 m 2 /sec for water. The density of the vapour was set to v = 1 kg/m 3 .

5.2.4. Parameters of iteration process

To provide smooth convergence of iterations, an under-relaxation strategy was


used (Rosten and Spalding, 1986). The linear relaxation (4.16) was applied to the
pressure ( p = 0.1 ), the turbulence kinetic energy ( p = 0.1 ) and the rate of its

dissipation ( p = 0.3 ). The false-time step relaxation (4.17) was used for the velocity

components ( t u = t v = t w = 0.1 ) and the mass fraction ( t f = 10 3 ). Because

cavitation flows are characterised by large density variations, linear relaxation was
applied to the mixture density ( = 0.2 ).

To improve the accuracy of calculations and to decrease the effects of numerical


diffusion, high-order differencing schemes are applied for the spatial discretisation of
the momentum equation (2.22) and the diffusion-free mass-fraction equation (4.13).
Several high-order schemes are available to the user of the PHOENICS package
(Malin and Waterson, 1999). Discretisation of the momentum equation was
performed using the SMART scheme (4.9). Preliminary calculations have revealed
that application of a high-order scheme is essential in order to obtain a correct pattern
of liquid flow using coarse grids. For the mass fraction equation, the Super-bee
scheme (4.10) was applied, as the most appropriate scheme recommended in
PHOENICS for resolving sharp interfaces.
To promote convergence of pressure-solution algorithm for computation of

cavitation flows, the compressibility factor was set to a small constant 10 6 10 5


(the variable DRH1DP in the Q1 input file, Appendix A).
To avoid unphysical solutions, the lower and upper limits for the mass fraction
f were set up to 0 and 1.
Chapter 5. Results 133

5.3. Application of the model for the critical number density


of the bubble nuclei
In this section the results of testing and application of the model of cavitation
flow, which account for the variation in the number density of cavitation bubbles in
liquid (equation (3.22)), are described.
First, the model is adjusted to predict the steady-state cavitation flows in a real-
scale model of Diesel injection nozzle (Roosen, et al., 1996). Then to demonstrate the
scalability of the model of cavitation developed, the calculations are performed for
similar flow regimes in a nozzle of a magnified scale.
The study is performed to reveal the effect of the number density of cavitation
bubbles and initial mass fraction on the results of calculations of the flow field and
vapour disribution. The number density parameter n* and the vapour pressure p v in

equation (3.22) are considered as fixed parameters, which depend on the liquid quality
and can not be affected by cavitation number. The conventional concept in simulation
of cavitation flows n = idem is compared to the concept n* = idem , which is
consequence of using equation (3.22).
In this section, it will be shown that the initial mass fraction can be setup to a
small number f o = 10 6 , which has no effect on the resultant flow field, but is needed

to activate the source term in the equation (4.13).


Chapter 5. Results 134

5.3.1. Reference measurements

Nozzle wall

Inlet
L

H
Outlet
Y
X Rin
W
Z

Fig. 5- 1. Shape and main dimensions of planar nozzle (Yuan, et al., 2001).

For the adjustment and testing of the model of cavitation flow, experimental
data on cavitation of tap water in a planar nozzle (Fig. 5-1) from Roosens report
(Roosen, et al., 1996, also described by Yuan, et al., 2001), are used in the present
study. The nozzle has a length L = 1 mm, width W = 0.2 mm and height H =
0.28_mm and radii of inlet corners Rin = 0.028 mm (Yuan, et al., 2001). In this study

two sets of data, described by Yuan, et al. (2001), which include measurements of
pressures in the flow upstream and downstream the nozzle, and images of cavitation
flow, are applied (Table 2-2, Table 5-1).

Table 5- 1. Parameters of cavitation flows of a tap water in planar nozzle (Fig.


5-1) (Pressure measurements by Roosen, et al., 1996).

Inlet Outlet Velocity scale Reynolds number, Cavitation


Flow
lU H
regime
pressure pressure
U = 2 p1lp2 , m/s Re H = l
number,

p1 , bar p 2 , bar CN

Inlet
80 21 108.6 30 416 2.81
cavitation
Super-
80 11 117.5 32 890 6.28
cavitation
Chapter 5. Results 135

5.3.2. Mesh specification

Z
Fig. 5- 2. The structure of body-fitted mesh in the Y-Z plane of the
computational domain (Fig. 5-1).

Discretisation of the differential equations is performed on a body-fitted grid


built for one quarter of a nozzle in between the symmetry planes, back plane wall and
curvilinear nozzle side wall (Fig. 5-2). The calculations were performed using a
coarse mesh with 5 9 52 = 2340 cells.

5.3.3. Convergence of iterations

Tests revealed that in order to minimise the residuals and achieve the flow

fields, which have virtually stopped varying with the iterations 10 4 sweeps (global
iterations of solution algorithm) were sufficient for simulations on a coarse mesh (Fig.
5-2). Calculations for one flow regime on a personal computer with Pentium 4
processor with operating frequency 2.66 GHz took approximately one hour. Most
simulations were started with initial flow fields based upon the converged solution for
liquid only flow. Variation of the initial flow fields was found to have no effect on the
final flow field (as would be expected for a steady state flow), but did significantly
reduced computation time. The results of a mesh dependence study are given in
Appendix C. It was found that the use of high-order spatial discretisation schemes
applied for approximation of advection terms in the mass fraction and momentum
equations, allowed accurate results to be obtained using less fine meshes.
Chapter 5. Results 136

5.3.4. Adjustment of the number density of cavitation bubbles

The model of cavitation developed in the present study requires the specification
of the number density of critical cavitation nuclei in the liquid. To decouple the liquid
quality and flow scale effects, when specifying the number density of cavitation
bubbles, the correlation (3.22) was developed in the present study. In equation (3.22)
parameter n* accounts for the liquid quality and requires adjustment for a specific
fluid.
The number density parameter n* in equation (3.22) can be adjusted using
measurements of the quasi-steady-state vapour pockets for the developing stage of
cavitation. This requires several iterations to determine the number density of
cavitation bubbles, which could match the measured pattern of cavitation flow. Then,
from the calculated liquid tension in the cavitation region, the parameter n* can be
evaluated using equation (3.22).
Matching is performed by comparison of the predicted and visualised vapour
structures in the flow. However, results of calculation of cavitation flow depend not
only on the reliability of the phase transition model, but also accuracy of the single-
fluid solver. Therefore instead of trying to predict a particular flow in detail by
adjusting a set of model constants, the range of concentrations of cavitation bubbles
n , which can be used to approximate quasi-steady-state flow pattern for a given
regime of cavitation, is determined. This can be done for inlet cavitation, formed at
the nozzle entry, and super-cavitation, when the vapour pockets reach the nozzle exit.
To determine the parameter n* for tap water used in Roosens experiments, the
flow regime with inlet cavitation was used first (Table 5-1). After a number of trials,
it was found that number densities from n = 1.6 1013 (m 3 ) to n = 2 1015 (m 3 )
produce vapour pockets at the nozzle entrance, which can be associated with the
cavitation bubble photographed by Roosen, et al. (1996) (Fig. 5-3, (a)). The void
fractions calculated using different number densities of cavitation bubbles are shown
in Fig. 5-3 in comparison with the image of the flow from Roosens report (1996).

This figure shows that an increase in n from n = 1.3 1014 (m 3 ) to n = 2 1015 (m 3 )


caused higher concentrations of the vapour in the cavitation region and large volume
of the vapour pocket.
Chapter 5. Results 137

(a)

(b)

(c)

(d)

0.0 1.0

Fig. 5- 3. Pattern of cavitating flow in a small-scale nozzle (Fig. 5-2)


(photograph by Roosen, et al. (1996) reported Yaun, et al., 2001) at CN=2.81 (a) in
comparison with the results of numerical predictions of the vapour field: (b) -

n = 1.3 1014 (m 3 ) ; (c) - n = 4.4 1014 (m 3 ) ; (d) - n = 2 1015 (m 3 ) .


Chapter 5. Results 138

(a)

(b)

(c)

(d)

Fig. 5- 4. Measured velocity field and shape of the vapour pocket in cavitating
nozzle (Roosen, et al., 1996) at CN = 2.81 (a) in comparison with the results of

numerical calculations using n = 1.3 1014 (m 3 ) (b), n = 4.4 1014 (m 3 ) (c),

n = 2 1015 (m 3 ) (d). The volume of calculated cavitation region is indicated by iso-


surface where the void fraction is = 20% .
Chapter 5. Results 139

The void fraction fields show high concentrations of vapour (more than 50%) in
cavitation region and steep gradients of concentration, which indicates a small
numerical diffusion effect.
Fig. 5-3 reveals that the vapour structures predicted by the model (Fig. 5-3, b, c,
d) are thinner than the vapour pocket observed by Roosen, et al. (1996) (Fig. 5-3, a).
This can be explained by the accuracy of prediction of the flow separation, which may
result from specification of nozzle geometry, model of turbulence and interaction
between the vapour and liquid phases in cavitation flow.
Fig. 5-4 shows results of the calculation of the velocity fields in the middle X-
plane of the nozzle in comparison with the velocity measurements performed by
Roosen, et al. (1996). Shapes of the vapour pockets predicted by the model are
indicated in this figure by iso-surfaces of the void fraction = 20% . At moderate
concentration of cavitation bubbles n , vapour fills the region of recirculation flow,
and the length stays nearly constant (Fig. 5-4 (b, c). When the number density exceeds
a certain level, the vapour pocket and recirculation flow region start to extend towards
the nozzle outlet (Fig. 5-4, d). This behaviour is in agreement with experimental
observations on cavitation flow development (Sato and Saito, 2001; Stutz and
Reboud, 1997; Roosen, et al., 1997), when the cavitation intensity is increased by
increasing the cavitation number CN, as described in Chapter 2.2 (see Fig. 2-7). To
describe correctly the vapour formation in recirculation region, number densities n in

the range from 1.3 1014 (m 3 ) to 4.4 1014 (m 3 ) can be applied. The extent of

cavitation region with n = 2 1015 (m 3 ) (Fig. 5-4, d) gives the best match with the
length of the photographed vapour pocket in Fig. 5-4 (a). However, an increase in n

from 4.4 1014 (m 3 ) to 2 1015 (m 3 ) resulted in a shift of the reattachment point


downstream, which is typical for transitional cavitation (Fig. 2-7) and does not
correspond to the sub-cavitation regime with steady-state inlet vapour pocket in Fig.
5-4 (a).
The number densities above the maximum level n = 2 1015 (m 3 ) lead to a
sharp transition to super-cavitation flow regime. Yuan et al. (2001) have decided use
this maximum number density of cavitation bubbles for simulation of arbitrary flow
regimes in a nozzle. They have found the maximum level at n = 1.5 1014 (m 3 ) . The
discrepancy between this number and the result of this study can be explained by the
Chapter 5. Results 140

difference in specifications for the flow model, and numerical methods applied for
calculation of the void fraction equation. Thus, Yuan, et al. (2001) have performed
simulations for 2D flow, applied the k model of turbulence and used volumetric
continuity equation and VOF method to calculate the void faction field.
Precise tuning of the model parameter n for a specific regime of cavitation
(given cavitation number) requires comparison of the results with the measurements
of the void fraction, which were not described by Yuan, et al. (2001). Therefore the
numerical study for the inlet cavitation can be concluded only with the range for the

number density of cavitation bubbles n = 1.3 1014 2 1015 (m 3 ) .


Comparison of the pressure distributions for the cavitation-free and cavitation
flow conditions (Fig. 5-5) reveals pressure recovery in the cavitation region at the
nozzle entry (Fig. 5-5). Downstream, the pressure in the cavitation region becomes
closer to the vapour pressure pv .

Fig. 5- 5. Pressure distributions along the nozzle center-line (CL) and near the
wall predicted for the inlet cavitation (Table 5-1) using

n = 4.4 1014 (m 3 ) (continuous lines), comparing to the pressure distribution in


cavitation-free flow (stroke lines). Thick dashed line marks the vapour pressure pv .
Chapter 5. Results 141

Because measurements of the discharge coefficient are not available from the
paper by Yuan, et al. (2001), the mass flow rate can not be set up according to the
experiment. Therefore, in the calculations, the velocity of the liquid at the inlet of the
domain was adjusted to u in = 14 m/s to predict the measured pressure drop over the
nozzle. Calculations have revealed that the discharge coefficient of nozzle was not
affected by the inlet cavitation, and kept the same value as for the cavitation-free flow
at a given Reynolds number:

0.5 l (u in Hin ) 2
2 H
0 .5 l u
Cd = =
p1,o p 2 p1,o p 2
, (5.1)
0.5 10 3
(14 0.228 ) 2
0.92
59 10 5

5.3.5 The effect of initial mass fraction

To activate the source term in the transport equation (4.13) the mass fraction f
should not be set to 0 or 1. This means that an initial fraction of vapour is required to
start evaporation in the liquid and some amount of liquid should be present in the
vapour to initialise condensation. For this purpose a small number f o (initial mass
fraction) was introduced in the mass-fraction-dependent term in the equation (4.14)
(parameter MFinit in Q1-file , Appendix A):

(1 f + f o )4 / 3
1 ( f ) =
1
(5.2)

+ v (1 f ) ( f + f o )
1/ 3
f
l

In the previous section, when adjusting the number density of cavitation bubbles
for the inlet cavitation in the Roosens nozzle, the initial mass fraction f o was set

to a small number f o = 10 6 . This part describes the results of study performed to

examine the effect of the parameter f o on the simulation results.


Chapter 5. Results 142

(a) (b) (c) (d)

0.0 1.0
Fig. 5- 6. Effect of the initial mass fractions f o on the vapour field predicted

for inlet cavitation (Table 5-1), n = 4.4 1014 (m3). (a) - f o = 107 , (b) - f o = 106 ,

(c) - f o = 10 5 , (d) - f o = 104 . Plots show the right half of the nozzle, flow is from
top to bottom.

To reveal the effect of the initial mass fraction on the flow field, several
calculations were performed for various flow regimes and different values of the
parameter f o . The results of this study are shown in Fig. 5-6 for the inlet cavitation

flow in Roosen nozzle (Table 5-1). Fig. 5-6 shows that the vapour distribution does
not reveal significant variations when the initial mass fraction varies in the range from

10 7 to 10 5 . For the number density n = 1.3 1014 (m3) all three void fraction
distributions (Fig. 5-6, a, b, and c) matches the pattern of cavitation flow observed by
Roosen (Fig. 5-3, a).
Chapter 5. Results 143

(a) (b) (c)

0.0 1.0
Fig. 5- 7. Effect of the initial mass fractions f o on the vapour field predicted

for super-cavitation flow regime (Table 5-1), n = 4.4 1014 (m3). (a) - f o = 107 , (b)

- f o = 106 , (c) - f o = 10 5 . Plots show the right half of the nozzle, flow is from top
to bottom.

For the super-cavitation flow, the mass fraction f o = 10 5 starts to affect the

shape of the vapour region, as shown in Fig. 5-7, while f o = 10 6 and f o = 10 7


produce effectively the same flow pattern.

From this study it is concluded that the initial mass fraction f o = 10 6 can be
applied for calculations.
Chapter 5. Results 144

5.3.6. Study of the effect of cavitation number

A common strategy in computation of cavitation flows, is to apply a unique


number density parameter n = idem for the calculation of cavitation regimes in a flow
of a given geometrical configuration (see, for example, Yaun, et al., 2001). From the
point of view of the development of a cavitation model of qausi-steady-state
cavitation, it is important to validate the concepts n = idem and n* = idem for
computation of different regimes of cavitation (different cavitation numbers) in a
given nozzle.
Detailed information about developed cavitation in the Roosens nozzle (Table
5-1) at CN = 6.27 is not available, and it is impossible to conclude if it was critical
cavitation (when the vapour pockets just reach the nozzle exit (Winklhofer, et al,
2001)) or developed super-cavitation. Therefore, the study of developed cavitation at
CN = 6.27 is aimed at showing the ability of the model to describe variations in the
length of the vapour region with the cavitation number, rather than to provide proper
validation of the model, which requires more specific experimental information about
the cavitation flow.
To establish the reference point when adjusting the number density of cavitation
bubbles, super-cavitation flow at CN = 6.27 (Table 5-1) was associated with critical
cavitation, when the length of the vapour pocket just reaches the nozzle outlet. To
match the vapour structures observed by Roosen, et al. (1996), the number density of
cavitation bubbles was varied. The calculations were performed starting from the
cavitation-free pressure-velocity field using u in = 15.1 m/s, adjusted for the measured

pressure drop of 69 bar (Table 5-1).


The results of calculations of the void fraction are shown in Fig. 5-8 for two
different number densities n . The number density n = 4.4 1014 (m 3 ) produced the
longest vapour pocket, which nearly reaches the nozzle outlet, although the thickness
of the vapour region is not so large, as it appears on the photograph (Fig. 5-8, a).
Chapter 5. Results 145

(a)

(b)

(c)

0.0 1.0
Fig. 5- 8. Pattern of cavitation flow in a small-scale nozzle (Fig. 5-2) (Yuan, et
al., 2001) at CN=6.27 in comparison with the results of numerical predictions of

cavitation pockets using n = 1.3 1014 (m3) (b) and n = 4.4 1014 (m3) (c).

Higher number densities n also result in super-cavitation vapour structures.

However, concentrations n higher than 2 1015 (m 3 ) produced numerically unstable


results. This instability was caused by collapse of the vapour cavity at the high
pressure specified at the nozzle outlet. To calculate this flow regime, it is necessary to
consider the transient behaviour of cavitation at the nozzle exit, which is beyond the
scope of the present study.
Chapter 5. Results 146

Fig. 5- 9. Pressure distributions along the nozzle center-line (CL) and near the
wall predicted for the super-cavitation flow (Table 5-1), using

n = 4.4 1014 (m 3 ) (continuous lines), comparing to the pressure distribution in


cavitation-free flow (stroke lines). Thick dashed line marks the vapour pressure pv .

Fig. 5-9 shows the pressure distribution along the nozzle. In the cavitation
region the pressure recovers up to a level, which corresponds to the vapour pressure.
Fig. 5-9 reveals an increase in the hydraulic resistance, compared with liquid flow and
inlet cavitation (5.1):

2
0 .5 l u 0.5 10 3 108 2
C dsup er = 0.875 (5.3)
p1,o p 2 76 10 5

In contrast to this result, experiments did not reveal significant changes in the
discharge coefficient under the developing cavitation in nozzles (for example, Fig. 2-
13 shows variations of C d less then 3%). Comparison of discharge coefficients (5.1)

and (5.3) shows that predicted decrease in the discharge coefficient of the nozzle with
cavitation development is not great (about 5%), so the numerical results are in a good
agreement with experimental observations, which did not reveal substantial variations
Chapter 5. Results 147

in hydraulic resistance with the cavitation number for the developing cavitation flows
(CN < CNsuper) (Fig. 2-13, Fig. 2-14).
Increase in hydraulic resistance of nozzle can be explained by variation of
density and viscosity of the flow in the near-wall region. Formation of the vapour
phase reduces the actual thickness of viscous sub-layer:

+
y visc,v y visc v / u ,v v v 10 5 1
= ~ < 1,
+
y visc,l y visc l / u ,l l l 10 6 10 3

+
where y visc 11.63 is dimensional thickness of viscous sublayer, and u = w /

is the turbulent velocity scale based on the wall shear friction w . Thinner laminar

sub-layer enhances the momentum transfer, and results in higher friction losses.
Comparison of the results of calculations performed using n = 4.4 1014 (m 3 )

for two different cavitation numbers permits the concepts n = idem and n* = idem to
be evaluated. Actually, equation (3.22) gives the ratio of number densities of
cavitation bubbles for the initial and super-cavitation flows:

3/ 2
ninlet n*,inlet p p inlet
= v min . (5.4)
nsuper n*,super p p super
v min

From the near-wall pressure distributions shown in Fig. 5-5 and Fig. 5-9, it
follows that the maximum liquid tension p v p min is nearly the same for the initial

and developed cavitation regimes:

pv p min
inlet
p v p min
super
20 bar,

ninlet n*,inlet
which after substitution in (5.4) gives , and shows the equivalence of
nsuper n*,super

the concepts n = idem and n* = idem .


Chapter 5. Results 148

From the known pressure distribution inside the nozzle, one can estimate the
degree of maximum tension in the liquid pv p min , and then calculate the number

density parameter n* in equation (3.22):

3/ 2 3/ 2
pv 2500
3
n* = n 4.4 10 (m )
14
2 1010 (m 3 ) .
pv p min 2 10 6

For inlet cavitation the number density was determined in the range

n = 1.3 1014 2 1015 (m 3 ) , which results in the range of variation for the number

density parameter n* = (0.57 9) 1010 (m 3 ) . The same ranges of n and n* can be


recommended for calculations of super-cavitation flows.

5.3.7. Demonstration of scalability of the model

Table 5- 2. Parameters of the cavitation flows of tap water in a large-scale


planar nozzle (Fig. 5-1).

Inlet Outlet Velocity scale Reynolds number, Cavitation


Flow
lU H
regime
pressure pressure
U = 2 p1lp2 , m/s Re H = l
number,

p1 , bar p 2 , bar CN

Inlet
0.223 0.0775 5.43 30 416 2.81
cavitation
Super-
0.223 0.0505 5.87 32 890 6.28
cavitation

To demonstrate the scalability of the model of cavitation based on equation


(3.22), simulations were performed for a 20 times scaled model of the Roosen nozzle
(Fig. 5-1). This large-scale nozzle has a length L = 20 (mm), a width W = 4 (mm), a
height H = 5.6 (mm) and radii of inlet corners Rin = 0.56 (mm). Parameters of the

flow in the magnified nozzle were determined from the parameters for small-scale
flow (Table 5-1) assuming identity of the Reynolds and cavitation numbers (Table 5-
2). Under these conditions the cavitation structures are assumed to appear identical for
the large and small-scale nozzles.
Chapter 5. Results 149

Fig. 5- 10. Variation in the number density of cavitation bubbles with the
liquid tension. The results of adjustments of the number density parameter n* for

real-scale nozzle n* and application of the equation (3.22) (straight line,

n* = 2 1010 (m 3 ) ) to estimate the number density of bubbles for a large-scale flow.


Lines shows the limits of variation for n established for the inlet cavitation in small-

scale nozzle (dashed line for n* = 9 1010 (m 3 ) and dotted line for

n* = 0.7 1010 (m 3 ) ).

Preliminary calculations have revealed that number densities of cavitation

bubbles of the order 1014 (m 3 ) are too high to describe inlet cavitation in a scaled-up
nozzle. This proves the inability of the concept n = idem to simulation of cavitation
flows at different hydrodynamic scales. Therefore, for the large-scale nozzle the
number density of cavitation nuclei was estimated using the number density for the

small-scale nozzle nsmall = 4.4 1014 (m 3 ) and assuming n* = idem and p v = idem

in equation (3.22):

3/ 2 3/ 2
nlarge p p large p large p large
= v min
~ 1small 2 ~
nsmall p p small p p small
v min 1 2
Chapter 5. Results 150

3 3
u large Dsmall
~ ~ = 1 .
20 3
u small Re=idem Dlarge

Thus, nlarge = 5.5 1010 (m 3 ) . This technique is also illustrated in Fig. 5-10.

Using parameter nlarge = 5.5 1010 (m 3 ) the cavitation structures predicted in

the large scale nozzle appeared similar to the structures predicted for the small-scale
nozzle (Fig. 5-3, c and Fig. 5-8, c).
The results of calculated pressure fields in small and large scale nozzles are
p p2
compared together using the dimensionless pressure ~
p= in Fig. 5-11 and
p1 p 2
Fig. 5-12. Comparison of the curves in these figures show nearly identical pressure
distributions in small and large scale nozzles both for the inlet cavitation (Fig. 5-11)
and super-cavitation flows (Fig. 5-12).

Fig. 5- 11. Pressure fields in small (dashed lines) and large-scale nozzles (solid
lines) predicted for the inlet cavitation at CN = 2.81, comparing to the pressure
distribution in cavitation-free flow (green lines). Solid thick curves are for pressure in
the liquid core, fine curves describe pressure distribution near the wall. Grey dashed
line marks the vapour pressure.
Chapter 5. Results 151

Fig. 5- 12. Calculated pressure distributions in the small-scale (stroke lines)


and large-scale nozzles (continuous lines) predicted for the super-cavitation flow
(Table 5-1) using equation (3.22) with the number density parameter

n* = 2 1010 (m 3 ) . Grey dashed line marks the vapour pressure.

5.3.8. Conclusions

The model for the number density of critical cavitation bubbles in a liquid (3.22)
is applied to describe cavitation flow in a real scale model of a Diesel injection nozzle
(Roosen, et al., 1996).
The results of the calculations of inlet and super-cavitation flows using this
model have revealed the following:
The linear steady-state model for the rate of bubble growth/collapse can
produce the flow patterns of cavitation flows observed in experiments.

The initial mass fraction f o plays the role of the numerical parameter in the

mass fraction equation. When this parameter is set to f o =106, it does not

affect the results of calculations of cavitation flow.

Equation (3.22) can be applied to describe different regimes of cavitation flow


in a nozzle of specific scale.
Chapter 5. Results 152

The conventional concept n = idem failed to predict cavitation flows of


different hydrodynamic scales. To illustrate the scalability of the model of
cavitation, which uses equation (3.22), it was applied to describe similar flow
regimes in real-scale and large-scale models of Diesel injection nozzles.

Adjustment of the liquid-specific number density parameter n* in the equation


(3.22) can not be performed using an image of the flow for one regime of
cavitation, but requires specific information about the flow regime (is it
incipient, developing, critical, or super-cavitation) and measurements of the
void fraction field in cavitation flow. Variations in the number density of
cavitation bubbles n , adjusted for a single regime of cavitation, can vary by
two orders of magnitude.
Chapter 5. Results 153

5.4. Validation of the model for the shear stress induced


cavitation
The model of cavitation originally developed by Yuan, et al. (2001) and
extended in this study (Ch. 4), has been proven to be able to describe both the inlet
and developed (super-cavitation) flows, such as those experimentally observed by
Roosen et al. (1996). However, preliminary calculations (described later in this
section) have revealed that this model does not predict cavitation development for the
conditions of experiments performed by Winklhofer, et al. (2001). The reason for this
could be the shear stress mechanism of cavitation, which was not considered in the
model developed in Chapter 4.
Comparison of the results of experiments by Roosen, et al. (1996) and
Winklhofer, et al. (2001) (Chapter 2.1) revealed similar types of cavitation to those
observed by the authors at similar pressure differences across the nozzles, but at
different levels of absolute pressure downstream the nozzle. Both studies used planar
nozzles of similar scales, but different working liquids (tap water and Diesel fuel).
Considering that the viscosity of Diesel fuel is nearly three times greater than the
viscosity of water an assumption was made that the onset of cavitation can be affected
by the shear stresses in the liquid according to equation (2.13). This hypothesis,
originally suggested by Joseph (1995), is applied in the present study to predict the
cavitation flows observed by Winklhofer, et al. (2001). It is worth to noting that the
literature survey has not revealed any models of cavitation which account for shear
stress induced cavitation.
Chapter 5. Results 154

5.4.1. Description of test cases

Nozzle wall

Inlet
L

H
Outlet
Y
X Rin
W
Z

Fig. 5- 13. Geometry and main sizes of planar rectangular nozzle (Winklhofer,
et al., 2001).

To develop and validate the model for stress-induced cavitation, the


measurements by Roosen, et al. (1996) and Winklhofer, et al. (2001) are applied.
These studies describe steady-state cavitation in plane real-scale models of Diesel
injection nozzles. Two regimes of cavitation are used for the model validation,
namely inlet cavitation and super-cavitation.
The measurements by Roosen, et al. (1996) were discussed in the literature
review (Chapter 2.2, Table 2-2) and described in validation of the model for the
number density of critical cavitation nuclei (Table 5-1). Similar measurements were
performed by Winklhofer, et al. (2001) for flows of Diesel fuel in a planar nozzle of
0.3 mm in width, 0.299 mm in height and 1 mm in length (Fig. 5-13). These data are
described here.
Winklhofer, et al. (2001) have studied cavitation in a model of a real-scale
Diesel injection nozzle formed by two steal sheets placed between a pair sapphire
windows (Fig. 5-13). The nozzle has the following dimensions: length L = 1 (mm),
width W = 0.3 (mm), height H = 0.299 (mm) and radii of the inlet corners Rin = 0.02
(mm). Experiments were performed for Diesel fuel at 30 deg. C.
Chapter 5. Results 155

Table 5- 3. Integral parameters of cavitation flow in a straight nozzle (Fig. 5-


13) (according to the measurements by Winklhofer, et al., 2001).
Inlet Outlet Mass flow Inlet Reynolds Cavitation Discharge
Flow
pressure pressure rate, velocity number, number, coefficient
regime
p1 , bar p 2 , bar G, gm/s u in , m/s Re H CN Cd
Non-
100 43.5 7.14 14.3 7940 1.30 0.82
cavitating
Cavit.
100 42.5 7.22 14.5 8025 1.35 0.82
start (CS)
Critical
100 34.0 7.78 15.6 8650 1.95 0.82
cavit.(CC)

(a) (b)
Fig. 5- 14. Pressure field before start of cavitation (a) and under critical
cavitation conditions (b) (Table 5-3) in planar nozzle (Fig. 5-13), 5 bar steps between
isolines (Winklhofer, et al., 2001).

For model validation, two regimes of cavitation, which reveal variation in the
flow structure with the cavitation number (Table 5-3, Table 2-2), were chosen from
the set of data reported by Winklhofer, et al. (2001). Fig. 2-13 shows the cavitation
structures obsereved inside the nozzle for these flow regimes. The pressure fields
inside the nozzle under cavitation-free and critical cavitation flows are shown in Fig.
5-14, a, b.
Winklhofer, et al. (2001) have measured the pressure fields for flows in nozzles
of different shapes, including straight nozzles, converging nozzles and diverging-
shape nozzles. For cavitation-free flow conditions the pressure field was reported only
for a nozzle of convergent shape. Though the measurements have shown the effect of
the conical shape of the nozzle on the hydraulic resistance of the nozzle, the pressure
field did not reveal significant variations in the pattern of the flow for a nozzle of
Chapter 5. Results 156

slightly convergent shape, as shown in Fig. 5-14 (b) (inlet to outlet ratio of 301:284).
Therefore, the results of numerical calculations of the pressure drop are compared to
the data reported for the straight nozzle (Table 5-1), while the pressure fields are
compared to Fig. 5-14 (b).
First, the velocity and pressure fields were predicted for both nozzles neglecting
the cavitation process. This revealed differences in the pressure fields and shear
stresses formed in the flows. Then, the model for the critical pressure was introduced
and applied to describe the cavitation flows observed by Roosen, et al. (1996) and
Winklhofer, et al. (2001).

5.4.2. Mesh specification

Fig. 5- 15. Mesh structure in Y-Z plane inside the nozzle (shown for half of a
nozzle, Fig. 5-13).

Numerical calculations were performed using structured body-fitted non-


orthogonal meshes with non-uniform distribution of cells in the flow domain (Fig. 5-
15). For the straight part of the nozzle, starting from its inlet corners, a Cartesian grid
was used. Because of the plane symmetry of the flow, calculations were made for one
quarter of the nozzle. In studies, the three-dimensional structure of the flow has been
taken into account, which is essential to predict correctly the hydraulic structure of the
flow. The main results were obtained on a coarse mesh (Fig. 5-15) with 5 10 49 =
2450 cells.
Chapter 5. Results 157

5.4.3. Pattern of cavitation-free flow

From the point of view of understanding of the mechanism of cavitation and the
effects of liquid properties and system pressure, it is interesting to compare the
patterns of liquid flows predicted for cavitation inception the Winklhofers and
Roosens nozzles. In order to clarify the viscous shear-stresses in regions of the flow
with maximum liquid tension where cavitation is expected to occur, clculations were
performed neglecting the cavitation process.
Figs. 5-16 5-21 show distributions for the flow variables under cavitation-free
flows in Roosens and Winklhofers nozzles for the middle ( X = 0 ) cross-section of
nozzle. The calculations were performed for the pressure drops across the nozzle
around 58 bar (inlet cavitations, Table 2-2).
Though the mean flow velocity scales were similar in both cases, as they were
determined by the pressure drop, the flow fields were different in detail. Fig. 5-16
shows the flow separation formed in Winklhofers nozzle (Fig. 5-16, b), which was
not observed for Roosens nozzle (Fig. 5-16, a).
Prediction of the flow separation in Winklhofers nozzle agrees with the
experimental observations (Fig. 5-14, a). For the liquid flow conditions, Fig. 5-14, a
shows contraction of the flow at the nozzle throat (contraction coefficient is about
225/300 = 0.75) and pressure recovery downstream in the vena contracta region. The
calculations (Fig. 5-16, b) predict the length of the separation region to be around one
half of the nozzle height and contraction coefficient about 0.82, which are in a good
agreement with the experimental observations.
The blank spots at the inlet corners in Fig. 5-17 show the area of the flow where
pressure becomes negative (the relative pressure p p 2 drops below the level p 2 ).

These regions indicate the volume of liquid subjected under tension pv p > 0 .
Comparison of Fig. 5-17, (a) and Fig. 5-17, (b) reveals a larger volume of liquid
under tension for the Roosens nozzle, where the flow is characterised by a lower exit
pressure p 2 .
Fig. 5-19 shows distribution of the principal component of the strain rate tensor,
calculated for flows in Roosens and Winklhofers nozzles (Fig. 5-1 and Fig. 5-13)
taking into account the two-dimensional nature of the flow:
Chapter 5. Results 158

1 u z u y
S yz = + . (5.5)
2 y z

This approximation is accurate for the straight part of the nozzles, where the grid has
a Cartesian structure. Comparison of Fig. 5-19, (a) and Fig. 5-19, (b) shows similar
distributions for the rates of strain and approximately the same maximum level of
strains achieved for both nozzles. At the same time, the higher viscosity of the Diesel
fuel results in higher stresses for the flow in the Winklhofers nozzle (Fig. 5-20).
According to the hypothesis (2.13), higher stresses can result in higher critical
pressures for the onset of cavitation.
Under the turbulent flow conditions, both components of the shear stress tensor
(molecular and turbulent) should be included in the criterion (2.13). For developed
turbulent flows, the average turbulent stresses prevail over the laminar stresses in the
core of the flow, which result from high turbulent viscosity (Fig. 5-21). Therefore, the
turbulent stress can make the main contribution to the critical pressure:

pcr = p v + 2 ( + t ) S ijmax . (5.6)


Chapter 5. Results 159

(a) (b)
Fig. 5- 16. Velocity fields in Roosens (a) and Winklhofers (b) nozzles
predicted neglecting the cavitation process.

(a) (b)
Fig. 5- 17. Fields of relative pressure p p 2 predicted for Roosens (a) and
Winklhofers (b) nozzles neglecting the cavitation process.
Chapter 5. Results 160

(a) (b)

Fig. 5- 18. Distributions of the kinetic energy of turbulence k , m 2 /s 2 in


Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process.

(a) (b)

Fig. 5- 19. Distributions of the component of the rate of strain S yz , sec 1 (5.5)

in Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process.
Chapter 5. Results 161

(a) (b)
Fig. 5- 20. Distributions of the shear stresses yz = 2( + t ) S yz , Pa in

Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process.

(a) (b)

Fig. 5- 21. Distributions of coefficient of the turbulent viscosity t , m 2 /sec , in

Roosens (a) and Winklhofers (b) nozzles predicted neglecting the cavitation
process.
Chapter 5. Results 162

Fig. 5-21 (b) indicates that the local maximum in the shear stress takes place at
about one half of the nozzle height downstream the nozzle inlet where the pressure
already recovers to positive values ( p p 2 > p 2 , Fig. 5-17, b). However, degree of
the shear stresses in the flow is not sufficient to overcome absolute pressures and thus
initiate the cavitation. This is schematically shown in Fig. 5-22.

p1
p

p2
pcr ~ p v + 2 t S

pv
0
z

Fig. 5- 22. Variations in absolute pressure and the turbulent component of


shear stresses along the nozzle (schematically).

To activate the shear-stress mechanism of cavitation described in equation (5.6),


an empirical model is developed in the next section. The main assumption of the
analysis is that the critical pressure can be affected by the turbulent shear stress in the
flow.
Before developing the model for the critical pressure, it is interesting to compare
the current method of definition of the critical pressure (5.6) to the model developed
by Singhal, et al. (2002) that accounts for the effect of the turbulent pressure
fluctuations in the following way:

pcr = pv + 0.39 12 k . (2.38)


Chapter 5. Results 163

Equation (2.38) has an empirical nature and contains tuneable constant 0.39.
Inspection of the turbulence kinetic energy k fields in cavitation-free flows in the
Roosens and Winklhofers nozzles (Fig. 5-18) reveals higher level of k reached in
the flow separation region in the Winklhofers nozzle. The kinetic energy of
turbulence k is formed under the effect of stresses in liquid, and therefore correlation
(2.38) can be explained on the basis of the model (5.6).

5.4.4. Model for the critical vapour pressure

To account for the effect of the viscous shear stress on the cavitation onset, the
hypothesis expressed in equation (5.6) is applied in the present study. Here, this
hypothesis is used to formulate the threshold, or critical pressure, for both the
evaporation and condensation stages of cavitation process.
First, to predict the inlet and critical cavitations conditions in the Winklhofers
nozzle (Table 5-3) the method suggested by Yuan, et al. (2001) was applied. The
initial flow field was calculated neglecting the cavitation process. Then the number
density of cavitation bubbles n was varied in order to match the pattern of cavitation
flow observed in the experiments (Winklhofer, et al., 2001). The results of
calculations for different number densities of cavitation bubbles are shown in Fig. 5-
23 for inlet cavitation and in Fig. 5-24 for critical cavitation conditions. Distributions
of the void fraction show that for all the simulation conditions the vapour pocket
occupies a small region at the nozzle entrance. For inlet cavitation conditions, the
vapour is formed only in a few cells near the nozzle inlet corner (Fig. 5-23). For
super-cavitation flow conditions, the length of the vapour pocket does not exceed a
length of two nozzle heights, with increasing number density n (Fig. 5-24). This
results from high system pressure that causes rapid condensation of vapour
downstream the nozzle entry.
Chapter 5. Results 164

(a) (b)

0.0 1.0
Fig. 5- 23. Void fractions predicted neglecting the effect of liquid stress on
critical vapour pressure. Inlet cavitation in Winklhofers nozzle (Table 5-3). (a) - n =

1.6 1016 (m3); (b) - n = 2 1018 (m3) . (Figures show only first section of the right
half of the nozzle).

(a) (b) (c) (d)

0.0 1.0
Fig. 5- 24. Void fractions predicted neglecting the effect of liquid stress on
critical vapour pressure for the critical cavitation flow in Winklhofers nozzle (Table

5-3). (a) n = 1.6 1016 (m3), (b) n = 2 1018 (m3) , (c) n = 1.6 1019 (m3), (d) n =

1.6 10 22 (m3).

To match the shape of cavitation pockets observed in experiments, the model of


cavitation was comprised by the relationship for the critical pressure (5.6). This
Chapter 5. Results 165

equation also was modified, so that the turbulent term was included with empirical
coefficient Ct :


pcr = pv 2 1 + Ct t S yz . (5.7)

where the turbulent viscosity is defined by equation (3.5) and the coefficient Ct is

considered to be a constant. In the following part the coefficient Ct is adjusted using

the measurements of cavitation flows by Roosen, et al. (1996) and Winklhofer, et al.
(2001).

5.4.5. Results

Adjustment of the model constants n and Ct

(a) (b) (c) (d)

Fig. 5- 25. Distributions of the void fraction in the Roosens nozzle under inlet
(a, b) and super-cavitation (c, d) flow regimes (Table 5-1), predicted assuming

pcr = pv (a, c) and using equation (5.7) with Ct = 10 (b, d). n = 4.4 1014 (m3).
(Figures show only right half of the nozzle, the flow runs from top to bottom).
Chapter 5. Results 166

To match the measured patterns of cavitation flows (Table 2-2), the number
density of cavitation bubbles n has been varied, using the procedure described in part
6.3. This number can not be set up according to the adjustments for the Roosens
nozzle, because of the difference in properties and quality of working fluids used by
Roosen, et al. (tap water) and Winklhofer, et al., 2001 (Diesel fuel).
To estimate the constant Ct in equation (5.7), the following method was
developed. First, the model for the critical pressure (5.7) was applied to describe
cavitation flows in the Roosens nozzle (Table 5-1). The purpose of this step was to
determine the maximum value of Ct , which will not affect the patterns of cavitation

flows predicted in section 5.3 assuming pcr = p v . As a result of these calculations, it

was found that application of equation (5.7) with a constant Ct below the

maximum level Ct =10, does not change the solutions for the inlet and developed
cavitation in the Roosens nozzle (Table 5-1). This is shown in Fig. 5-25, which gives
the void fraction fields predicted for the inlet and super-cavitation flows in Roosens
nozzle. Then, Ct =10 was applied to predict the flow regimes in Winklhofers nozzle

(Table 5-3). The objective of this was to reveal the effect of shear stress on the pattern
of cavitation flow. The results of these calculations for inlet and critical cavitation
regimes in the Winklhofers nozzle (Table 5-2) are described in this section. The
study is completed by an analysis of the sensitivity of the results to variations in the
constant Ct .

(a) (b)

0.0 1.0
Fig. 5- 26. Distributions of the void fraction at the nozzle throat for the inlet
cavitation (Table 5-3), predicted taking it into account the effect of shear stress on

critical pressure by the equation (5.7) using n = 2 1018 (m3). (a) Ct = 10 , (b)

Ct = 20 . (Figures show only first section of the right half of the nozzle).
Chapter 5. Results 167

(a) (b) (c) (d) (e)

0.0 1.0
Fig. 5- 27. Distributions of the void fraction at the nozzle throat for the critical
cavitation (Table 5-3), predicted taking it into account the effect of shear stress on
critical pressure by the equation (5.7) with a constant Ct = 10 and various number

densities of cavitation bubbles n (m3): (a) 1.6 1013 , (b) 2 1015 , (c) 1.6 1016 ,

(d) 2 1018 , (e) 1.6 1019 .

(a) (b)
Fig. 5- 28. The shape of the vapour pocket at the nozzle entry (Table 5-3) (iso-
surfaces for the void fraction = 20% (a) and = 50% (b)) predicted for the critical

cavitation using Ct = 10 and n = 1.6 1016 (m3). The entry part of the nozzle is shown
for half of the height and half of the width. Flow from left to right, back wall is shown
in grey.
Chapter 5. Results 168

(a) (b) (c) (d)

0.0 1.0
Fig. 5- 29. Distributions of the void fraction at the nozzle throat for the critical
cavitation (Table 5-3), predicted taking it into account the effect of shear stress on
critical pressure by the equation (5.7) with a constant Ct = 20 and various number

densities of cavitation bubbles n (m3): (a) 1.6 1013 , (b) 2 1015 , (c) 1.6 1016 ,

(d) 2 1018 .

The results of calculations of the vapour distributions in the nozzle for inlet and
critical cavitation conditions using Ct =10, and different number densities of

cavitation bubbles n , are shown in Fig. 5-26 (a) and Fig. 5-27. Fig. 5-28 shows the
shape of the vapour pocket formed at the nozzle inlet corner in three dimensions.
Calculations have revealed that setting Ct =10 does not have a noticeable effect

on the dimensions of the vapour region for the inlet cavitation, so that the vapour
region (Fig. 5-26, a) kept nearly the same as it was predicted by the conventional
model pcr = p v (Fig. 5-23, a).
Under super-cavitation flow conditions the effect of the number density of
cavitation bubbles became more obvious (Fig. 5-27). An increase in the number
density of cavitation bubbles resulted in an elongation of the vapour region and higher

concentrations of the vapour phase. Using n = 1.6 1016 1.6 1019 (m3) it became
Chapter 5. Results 169

possible to match the length of cavitation region observed in experiment (Fig. 5-14,
b).
To reveal the effect of variations in the constant Ct on cavitation flow, the

results of calculations using Ct =20, are show in Fig. 5-26, b and Fig. 5-29. An

increase in Ct resulted in greater amount of vapour being produced in the nozzle.

Thus, for incipient cavitation (Fig. 5-26, b), Ct =20 produced a vapour region, which

was longer, compared to Ct =10 (Fig. 5-26, a), and was a better match with the

experimentally observed vapour field (Table 2-2). For the super-cavitation flow

conditions, number densities higher than n 1.6 1016 (m3) produced long vapour
cavities, which are typical of the super-cavitation flows (Fig. 5-29 c, d).
Analysis of the results of this study show that the constant Ct selected in the

range from 10 to 20 can match the cavitation structures experimentally observed by


Winklhofer, et al. (2001). The number density of cavitation bubbles around

n = 2 1018 (m3) can be applied to describe both the incipient and critical cavitation
flow regimes.

Analysis of tension in the liquid


To reveal the effect of the shear stress on cavitation flow, the values of the

second term in equation (5.7) 2 1 + Ct ( t



) S yz and the local pressure p are

compared in Fig. 5-30. Analysis of the pressure field shows that positive tensions
pv p > 0 and p cr p > 0 appear at the nozzle wall downstream the inlet corner
(Fig. 5-30, a, b). However, application of the criterion (5.7), results in larger volume
of liquid where the pressure drops below the critical level p < p cr (Fig. 5-30, b). This

becomes more obvious for Ct =20 (Fig. 5-31). Also, comparison of the distributions

of p p v and p p cr , shows that the local pressure recovers much faster than the

(
variable p 2 1 + Ct
t

) S yz (Fig. 5-30, c, d). This decreases the rates of

condensation and results in longer vapour pockets being predicted by the model (5.7)
(Fig. 5-27, d) compared to the predictions of the conventional model pcr = p v (Fig.
5-24, b).
Chapter 5. Results 170

(a) (b)

(c) (d)
Fig. 5- 30. Iso-surfaces for the tension p p v (a, c) and local variable

p p cr (b, d), predicted using Ct = 10 and n = 2 1018 (m3). (a) p p v < 0 ; (b)

p p cr < 0 ; (c) p p v < 4 10 5 (Pa); (d) p p cr < 4 10 5 (Pa). (Flow from right to
left around the nozzle inlet corner).

(a) (b)
Fig. 5- 31. Iso-surfaces for the tension p p v < 0 (a) and local variable

p p cr < 0 (b), predicted using Ct =20 and n = 2 1018 (m3). (Flow from right to left
around the nozzle inlet corner).
Chapter 5. Results 171

The number density parameter n*


From the results of calculations using specific number density of cavitation
bubbles n , the number density parameter n* in the equation (3.22) can be determined
for the Diesel fuel used by Winklhofer, et al. (2001). This method was described in
section 5.3 and requires estimation of maximum liquid tension pv p min from the
calculated pressure field in a nozzle. When the critical pressure for the cavitation
onset is defined in (5.7), which differs from saturation pressure in the liquid, the
question arises which pressure to use in equation (3.22), p cr or pv ?
Calculations have revealed that the main effect of shear stresses on the critical
pressure is observed for the region of condensation of the vapour, but the vapour
production is mainly determined by the reduction in the local pressure at the inlet
corners of the nozzle (Fig. 5-22). Therefore, the critical number density of cavitation
bubbles can be calculated applying equation (3.22), where the vapour pressure is
saturation pressure of vapour in liquid and does not account for the effect of viscous
stresses (5.7). For super-cavitation flow, the following estimate can be made for the
number density parameter in equation (3.22):

3 / 2 3 / 2
p p min 3 1.1 10 6 Pa
n* = n v 2 10 (m )
18 9 1012 (m 3 ) .
pv 300 Pa

This number is several orders of magnitude higher than n* predicted for the

water flow in the Roosens nozzle n* (0.07 9) 1010 (m 3 ) . Also, the number
density of cavitation bubbles in the Roosens small-scale nozzle is nearly four orders
of magnitude smaller than that obtained for the Winklhofers flow. This can be
explained by the difference in the liquid quality of the fluids used in these studies.
The accuracy of estimation of the number density parameter n* for equation
(3.22) and further application of this parameter to predict cavitation flows, may be
reduced by uncertainty in the vapour pressure pv , which can vary with the
composition of multi-component liquids, such as Diesel fuel. Therefore, for practical
applications it might be more appropriate to use correlation (3.22) in a form:
Chapter 5. Results 172

n = n* ( p v p )3 / 2 ,

where n* is dimensional parameter, which units are (m3 Pa3/2).

Discharge coefficient of nozzle


The results of calculation for liquid and cavitation flow conditions have not
revealed any significant variations in the discharge coefficient of the nozzle.
According to the numerical predictions, for critical cavitation, the discharge
coefficient is:

2
0 .5 l u 0.5 840 104 2
Cd = 0.90 ,
p1,o p 2 56 10 5

H in 2
where u = u in = 15.6 = 104 .
H 0 .3
The fact that the discharge coefficient is not affected by cavitation when vapour
structures stay inside the nozzle exit agrees well with the experimental observations
(Table 5-3). However, comparison of the numerical predictions of the discharge
coefficient C d with the experimental values from Table 5-3 shows that the
calculations under-predict the hydraulic resistance of the nozzle. This discrepancy
(~10%) can be explained by the accuracy of measurements and also imperfections in
the flow model, as discussed above. Actually, inspection of the pressure field, Fig. 5-
14 (b), reveals that the back pressure downstream the nozzle, which is 42 bar
according to Fig. 5-14 (b), is lower than static pressure in the jet and pressure at the
nozzle outlet p 2 , which have to be specified in simulations. The difference between
the actual pressure at the nozzle exit cross-section and pressure downstream the
nozzle is seen to be around 15 bar, which give an increase in C d of 10%.
Chapter 5. Results 173

5.4.6. Conclusions

The effect of the shear-stress mechanism of cavitation was revealed by analysis


of the patterns of cavitation flows described by Roosen, et al. (1996) and Winklhofer,
et al. (2001).

To account for the effect of shear stresses on cavitation flow the modelling
equation (5.7) for the critical pressure was introduced. This critical pressure was used
to calculate the rates of evaporation and condensation in the flow.

Equation (5.7) includes an empirical constant Ct , which was adjusted for the
conditions of the experiments described by Roosen, et al. (1996) and Winklhofer, et
al. (2001). The results of study revealed that a unique value for the constant Ct can be
applied to predict the cavitation structures for both incipient and critical cavitation
flows in nozzles.

The results of calculations showed that the conventional concept pcr = p v fails
to predict large vapour pockets in the nozzle, which were observed experimentally for
the developing regimes of cavitation flow in the nozzle (Winklhofer, et al., 2001).

Analysis of distributions of the flow variables shows that regions where the
static pressure reaches its minimum and where the shear stress achieves its maximum
are located in different parts of the flow. The calculations showed that the large
amounts of vapour predicted using equation (5.7), comparing to conventional model
pcr = p v , is a result of two effects: 1) an increase in the rate of evaporation at the
nozzle inlet, and 2) a decrease in the rate of condensation downstream the inlet.

The local mechanism of stress-induced cavitation in regions of the flow where


the local absolute pressure may be significantly higher than pv , makes it difficult to
specify the number density of critical cavitation bubbles for the whole flow (an
integral method). More accurate consideration of the shear-stress mechanism of
cavitation would require a local model for the number density of cavitation bubbles.
Chapter 5. Results 174

5.5. Conclusions for chapter 5


The single-fluid model of cavitation was applied for simulation of quasi steady-
state cavitation flows in models of Diesel injection nozzles.

To describe the pattern of the liquid flow in the nozzle more accurately, the
geometry and viscous flow scale effects were considered in the model of the flow. The
calculations were performed taking into account the roundness of the nozzle inlet
corners and three-dimensional nature of the flow. For accurate prediction of turbulent
flow in nozzles with flow separation, the RNG k model was applied. To improve
the accuracy of the numerical calculations using coarse meshes, high-order schemes
were used for discretisation of the convection terms in the mass fraction and
momentum conservation equations.

Validation tests were designed for the sub-models which describe the number
density of cavitation bubbles in the liquid and the effect of shear stress on the critical
pressure for the onset of cavitation. The measurements of cavitation flows in small-
scale nozzles operating under inlet and super-cavitation flow regimes (Roosen, et al.,
1996; Winklhofer, et al., 2001) were used to test these models.

Equation (3.22) for the number density of cavitation bubbles was developed to
decouple the flow scale effects and liquid quality effects. The model assumes that the
liquid quality can be described by the number density parameter n* . The concept
n* = idem was proved for the calculation of developing cavitation in a flow of a
specific geometry scale and for the description of similar flows in nozzles of different
hydrodynamic scales.

The initial mass fraction f o is a numerical parameter of the model, with a small

value ( f o = 10 6 ), which is required to initialise the source term for the mass fraction

equation. It was found that setting f o to a small value had a negligible effect the final
solution.

The model for the critical pressure was developed to describe hydrodynamic
cavitation in high-speed viscous flows at high system pressures. These conditions are
typical for Diesel injectors. The model for the critical pressure describes a reduction
in the local tension in the flowing liquid due to the effect of viscous shear stresses
Chapter 5. Results 175

(equation (5.7)). An empirical constant Ct was introduced to match the cavitation


structures observed by Roosen, et al. (1996) and Winklhofer, et al. (2001). The
numerical studies showed that the same constant Ct 10 20 can be applied for
computation of developing cavitation flows at high and low system pressures.

The results of adjustment of the number density parameter n* show that it may
vary by several orders of magnitude from one liquid to another. The accuracy of an
estimate for the parameter n* can be affected by accuracy of measurements applied
for the model tuning, uncertainty in the saturation pressure pv , and accuracy of
calculations of the pressure field.

The results of calculations of developing cavitation flows in nozzles have not


revealed any significant effect of cavitation on the hydraulic resistance of nozzles
(variation in the discharge coefficient was less than 5%), which is in a good
agreement with measurements.
Chapter 6. Conclusions and recommendations 176

Chapter 6. Conclusions and recommendations


6.1. Conclusions
The phenomenon of hydrodynamic cavitation in Diesel injection nozzles has
been analysed and modelled using the methods of computational fluid dynamics.
From the analysis of experimental observations of cavitation flows it was
concluded that the onset and development of cavitation depend not only on the nozzle
geometry, initial and boundary conditions, and a set of similarity criteria, such as
Reynolds and cavitation numbers, but also on the flow and liquid quality scale effects.
These effects determine the specific features of the cavitation flow. When the liquid
and flow scale effects are minimised, and the similarity criteria are satisfied, the flow
patterns looks similar in cavitation flows at different hydrodynamic scales
(Arcoumanis, et al., 2000).
The review of theoretical studies of cavitation flows has shown that the
cavitation models presented can not describe the hydrodynamic similarity of
cavitation flows. The objective of this work was the development of a scalable model
of cavitation, which took into account the flow and liquid quality scale effects.
To describe the cavitation flow, the concept of a locally homogeneous liquid-
vapour mixture was applied. To calculate the rates of evaporation and condensation in
the flow, the transport equation for the void fraction based on a linear Rayleigh model
for bubble growth and collapse, was adopted. This equation was derived under
assumption that cavitation starts to develop from bubble nuclei and can be described
locally for cavitation bubbles of one size. In contrast with conventional models, the
current study associates these cavitation bubbles with the critical cavitation nuclei in
the liquid. It was shown that by allowing a variation in the number density of critical
nuclei, the liquid quality and flow scale effects can be decoupled. To predict
cavitation flow, the cumulative spectrum of critical cavitation nuclei should be known
for a liquid, and the liquid tension in the region of cavitation must be estimated.
From similarity analysis of the mathematical model of cavitation flow, it was
shown that hydrodynamic scaling requires a certain shape of spectrum of the critical
nuclei and, also, an infinitesimally small initial radius of the bubble nuclei. The
similarity-consistent model for the population of critical nuclei was compared to
direct measurements of cavitation bubbles in water (Gindroz, et al., 1997) and tested
Chapter 6. Conclusions and recommendations 177

for prediction of similar cavitation flows in real and large-scale models of Diesel
injection nozzles.
The model was refined by addition of a sub-model for the effect of the local
shear stress on the critical pressure. This sub-model was based upon the theoretical
hypothesis of Joseph (1995). The refined model was shown to successfully predict
cavitation flows of water and Diesel fuel in real-scale models of Diesel injection
nozzles (Roosen, et al., 1996; Winklhofer, et al., 2001).
To calculate cavitation flows, the methods of computational fluid dynamics
were applied. The model of cavitation, based on the transport equation for the mass
fraction of vapour, comprised by the sub-models for the number-density of cavitation
nuclei and critical pressure in the liquid, was implemented within the framework of
the PHOENICS CFD package and applied for simulation of quasi-steady-state
cavitation flows in nozzles.

6.2. Recommendations for further studies


The results of the present study have shown that the model of cavitation can
describe the main effects and features of quasi-steady-state cavitation flows. To obtain
more accurate results, special attention should be paid to the physical and numerical
aspects of the calculation of cavitation flows. In addition, the inaccuracy of the results
of predictions with the current model may arise from the assumptions made about the
flow and the fact that range of application of the model is not clearly defined due to
the amount of suitable data available for validation.
Single-fluid solver. The reliability of the cavitation model depends on the
accuracy of the single-fluid solver used for prediction of the pattern of the liquid flow,
and properties of the vapour-liquid cavitation flows. The current study applied basic
methods to capture qualitatively the steady-state pattern of liquid flow. The single-
fluid solver used the high Reynolds number form of the RNG k model and the
equilibrium wall functions formulation. The accuracy of calculations of developed
turbulent and transitional flows with separation can be improved by consideration of
the low Reynolds number version of the k model and application of the
generalised wall functions.
Homogeneous model. The single-fluid homogeneous model of cavitation
assumes equality of velocities of liquid and vapour phases, which can be justified only
Chapter 6. Conclusions and recommendations 178

for small cavitation bubbles. To describe more precisely the behaviour of large
cavitation pockets, the slip velocity between the phases should be taken into account.
Additional studies could also be performed to investigate the effect of the correlation
for the mixture viscosity on cavitation flow, and modulation of the turbulent viscosity
by the vapour bubbles in the flow.
Super-cavitation and hydraulic flip. In the present study, the model of
cavitation flow was developed and applied to simulation of developing cavitation
flows in nozzles, characterised by vapour structures, which stay inside the nozzle.
Simulation of super-cavitation and hydraulic flip flows would require consideration of
an interaction of the vapour pockets with the high pressure boundary at the nozzle
exit. This needs special treatment due to the transonic flow at the outlet boundary, and
consideration of the unsteady behaviour of cavitation flow due to periodic collapse of
the vapour pocket at the nozzle exit.
Condensation stage and collapse of cavitation bubbles. The model for the
vaporous cavitation was derived assuming steady-state evaporation and condensation
running at the same rate. This concept can be extended to consider the effect of liquid
sub-cooling on the average rate of the bubble collapse, and, more generally, the
transient nature of bubble collapse. Additional study is needed to develop the model
for periodic shedding of the vapour from the rear part of the cavity under the
transitional cavitation stage.
The current model assumes that the tensile strength of the liquid can be
decreased due to the effect of the viscous shear stresses. The model for the critical
pressure is applied to calculate the rates of evaporation, and the rates of condensation
in the flow. However, this approach has not been proven for the condensation stage,
and requires special studies.
Gaseous cavitation. Cavitation development from gas-filled bubbles was not
described by the model. This effect is important for prediction of large scale flows at
low system pressures, and can be included in the model.
Similarity, nuclei concentration, and critical pressure. An integral concept
for the number density of cavitation bubbles, which assumes that the concentration of
bubbles per unit volume of liquid does not change over the whole flow domain
( n = const ), requires further validation for flows with multiple cavitation regions and
unsteady cavitation flows.
Chapter 6. Conclusions and recommendations 179

Additional studies are needed to validate the application of the model for the
number density of cavitation bubbles for simulation of flows with the shear-stress
mechanism of cavitation.
The scalable model for the number density of cavitation bubbles contains a
liquid specific parameter n* . The estimate of this parameter is affected by uncertainty
in the value of the vapour pressure pv , when it is small. A study is needed to
investigate the sensitivity of the results of simulations to variation in the number
density parameter n* .
Numerical model. To improve the rate of convergence of iterations, when using
the divergent form of the void fraction transport equation, the correlation (2.72)
suggested by Senocak and Shyy (2000) can be useful, especially for calculations of
unsteady cavitation flows. Alternatively, the VOF technique can be used to resolve
inter-phase boundary described by the void fraction equation in non-conservative
form (Spalding, 1974; Sauer, 2000).
Directions for experimental studies. To choose the data for validation of the
model of cavitation developed, numerous experimental studies were reviewed and
their results examined for use in testing the model. This work leads to the conclusion
that measurements of cavitation flows are usually comprised of measurements of the
hydraulic resistance and images of the liquid-vapour cavitation structures.
Measurements of the local velocity, pressure and void fraction fields are extremely
rare even for studies of cavitation in large-scale flows. Measurements of the vapour
concentration in cavitation flows would be very helpful for validation purposes.
Also, experience in the development of the cavitation model and design of
validation tests resulted in suggestions for basic experiments on cavitation flows for
validation of cavitation models and better understanding of the phenomena of
hydrodynamic cavitation:
Measurements of cavitation flows of the same working liquid in nozzles of
different scales operating under similar flow conditions could be useful for
validation of the scalable model for the number density of critical nuclei.
Measurements of cavitation flows of a specific liquid in a given nozzle
operating at similar flow conditions, but different system (outlet) pressures
could be useful to examine the effect of viscous shear stress on cavitation flow.
Chapter 6. Conclusions and recommendations 180

Measurements of cavitation flows of different liquids (this requires control of


the liquid quality) in a nozzle under similar flow conditions, and studies of
cavitation in laminar flows and vortex flows could be useful for testing and
identification of the range of practical application of the model of cavitation,
which accounts for the effect of the viscous shear-stress.

This project is a part of general direction of studies of the fuel injection and
combustion processes in Diesel engine conducted by the ICEG group at the
University of Brighton. The hydrodynamic cavitation provides one of possible
mechanisms of spray break-up, which can also happen under effect of aerodynamic
forces and turbulent stresses acting on the fuel jet. Auto-ignition and combustion of
fuel depend on the spray atomisation and also heating and evaporation of fuel
droplets. From analysis of the problem of transient heat transfer between a spherical
droplet and a stationary gas (Sazhin, et al., paper submitted to the ASME Journal of
Heat Transfer, Appendix D) it was concluded that correction to Newtons law of
cooling is required in order to predict more accurately the rate of droplet heating.
Further studies are needed to account for the effect of cavitation inside injection
nozzles on the spray disintegration in combustion chamber.
References 181

References
Ahuja, V., Hosangadi, A. and Arunajatesan, S. (2001) Simulations of caviting flows
using hybrid unstructured meshes. J. Fluids Engng., vol. 123, pp. 331 339.
Akhatov, I., Lindau, O., Topolnikov, A., et al. (2001) Collapse and rebound of a laser-
induced cavitation bubble. Physics of Fluids, vol. 13, no. 10, pp. 2805 -2819.
Alajbegovic, A., Grogger, H.A., Philipp, H. (1999) Calculation of transient cavitation
in nozzle using the two-fluid model. Proc. ILASS-Americas99 Annual Conf.,
pp. 373 377.
Arai, M., Shimizu, M., and Hiroyasu, H., (1985) Breakup Length and Spray Angle of
High Speed Jet, Proc. of the 3rd International Conference on Liquid
Atomization and Spray Systems (ICLASS), London, 1985.
Arcoumanis, C., Badami, M., Flora, H. and Gavaises, M. (2000) Cavitation in real-
size multi-hole Diesel injector nozzles. SAE paper 2000-01-1249, 2000.
Arcoumanis, C., Gavaises, M., Flora H. and Roth H. (2001) Visualisation of
cavitation in diesel engine injectors. Mcanique & Industries, v. 2, no. 5, pp.
375-381.
Ashley, S. (1997) Diesel cars come clean. Mechanical Engineering, vol. 8, 1997, p. 1.
Available from (28.05.2005):
http://www.memagazine.org/backissues/august97/features/diesel/diesel.html
Avva, R.K., Singhal, A., and Gibson, D.H. (1995) An enthalpy based model of
cavitation. ASME, Hilton Head.
dAgostino, L. and Acosta, A.J. (1991) A Cavitation Susceptibility Meter with
Optical Cavitation Monitoring Part One: Design concepts. Trans. ASME, J.
Fluids Engng., vol. 113, pp. 261 269.
Badock, C., Wirth, R., Fath, A., and Leipertz, A. (1999) Investigation of Cavitation in
Real Size Diesel Injection Nozzles. Int. J. Heat Fluid Flow, v. 20, pp. 538-
544.
Basuki, W., Schnerr, G.H., and Yuan W. (2002) Single-phase and modified
turbulence models for simulation of unsteady cavitation flows. Proc. of
German-Japanese Workshop on Multiphase Flow, August 25-27, 2002.
Karlsruhe, Germany.
References 182

Batchelor, G.K. (1980) An Introduction to Fluid Dynamics. Cambridge Univ. Press,


615 p.
Benajes, J., Pastor, J.V., Payri, R. and Plazas, A.H. (2004) Experiments for the
different values of K-factor. J. Fuids Engineering, vol. 126, p. 63.
Bergwerk, W. (1959) Flow Pattern in Diesel Nozzle Spray Holes. Proc. Inst. Mech
Engrs., 1959, vol. 173, pp. 655 660.
Berthelot, M. (1850) Sur quelques phenomenes de dilation forcee de liquides. Ann.
De Chimie et de Physique, vol. 30, pp. 232 237.
Biesheuvel, A. and Wijngaarden, L. (1984) Two-phase flow equations for a dilute
dispersion of gas bubbles in liquid. Journal of Fluid Mech., vol. 148, pp. 301-
318.
Billet, M.L. (1985) Cavitation nuclei measurements a review. Cavitation and
Multiphase Flow Forum, 1985, Albuquerque, N.M.
Blessing, M., Konig, G., Kruger, C., Michels, U., and Schwarz, V. (2003) Analysis of
flow and cavitation phenomena in Diesel injection nozzles and its effects on
spray and mixture formation. SAE paper 2003-01-1358, pp. 29 39.
Brennen, C.E. (1995) Cavitation and bubble dynamics. Oxford University Press.
Chahine, G.L. (2004) Nuclei Effects on Cavitation Inception and Noise. 25th
Symposium on Naval Hydrodynamics. St. Johns, Newfoundland and
Labrador, Canada, 8-13 August 2004. 14 p.
Chaves, H., Knapp, M., Kubitzek, A. et al. (1995) Experimental study of cavitation in
the nozzle hole of diesel injectors using transparent nozzles. SAE paper No.
950290, pp. 199-211.
Chen, Y. and Heister, S.D. (1995) Two-phase modelling of cavitated flows. In:
ASME Cavitation and Multiphase Forum, Reno, Nevada, 1994. (Comput.
Fluids, vol. 24, no. 7, pp. 799 806)
Chen, Y. and Heister, S.D. (1996) Modelling Hydrodynamic Nonequilibrium in
Cavitating Flows. Tr. ASME J. Fluids Eng., v. 118, pp. 172 178.
Chisholm, D. (1983) Two-phase flow in pipelines and heat exchangers. London:
Godwin, 1983.
Delale, C.F. (2001) Thermal Damping in Cavitating Nozzle Flows. Proc. of the 4th
Int. Symposium on Cavitation: CAV2001, 7 p.
References 183

Dellanoy, Y., and Kueny, J.L. (1990) Two phase flow approach in unsteady cavitation
modelling. In: ASME Cavitation and Multiphase Forum, ASME FED, vol. 98,
pp. 153 158.
Dumont, N., Simonin, O., and Habchi, C. (2000) Cavitating flow in diesel injectors
and atomisation: a bibliographical review. Proc. of the 8th Int. Conf. On
Liquid Atomization and Spray Systems ILASS-2000.
Dumont, N., Simonin, O., and Habchi, C. (2001) Numerical simulation of cavitating
flows in Diesel injectors by a homogeneous equilibrium modelling approach.
4th Int. Symposium on Cavitation: CAV2001.
Ferziger, J. H. and Peric, M. (2002) Computational Methods for Fluid Dynamics. 3rd
ed., 2002, 423 p.
Florschuetz, L.W. and Chao, B.T. (1965) On the Mechanics of Vapor Bubble
Collapse. J. Heat Transfer, vol. 87 pp. 209-220.
Fox, T.A. and Stark, J. (1989) Discharge coefficients for miniature fuel injectors.
Technical note in Proc. Inst. Mech. Engrs, vol.203, part G: Journal of
Aerospace Engng., pp. 75 78.
Franc, J.P. (2001) Partial Cavity Instabilities and Re-Entrant Jet. 4th Int. Symposium
on Cavitation: CAV2001, 21 p.
Frobenius, M., Schilling, R., Bachert, R., Stoffel, B. and Ludwig, G. (2003) Three-
dimensional unsteady cavitation effects on a single hydrofoil and in a radial
pump measurements and numerical simulations. Part two numerical
simulation. Proc. of the 5th Int. Symposium on Cavitation. Paper Cav03-GS-9-
005.
Fujimoto, H.G., Nishikori, T., Hojyo, Y., Tzumakoto, T. and Senda, J. (1994)
Modelling of atomization and vaporization process in flash boiling spray.
Proc. ILASS-1994, Rouen, France, July 1994, paper VI-13, pp. 680 687.
Joseph, D. (1995) Cavitation in a flowing liquid. Phys. Review, E., vol. 51, no. 3, pp.
1649-1650.
Ganippa, L., et al. (2001) Comparison of cavitation phenomena in transparent scaled-
up single-hole diesel nozzles. Proc. of the 4th Int. Symposium on Cavitation:
CAV2001, 9 p.
Giannadakis, E., Gavaises, M., Roth, H. and Arcoumanis, C. (2004) Cavitation
Modelling in Single-Hole Injector Based on Eulerian-Lagrangian Approach.
References 184

THIESEL 2004 Conference on Thermo- and Fluid Dynamic Processes in


Diesel Engines. 14 p.
Gindroz, B., Bailo, G., Matera, F. and Elefante, M. (1997) Influence of the Cavitation
Nuclei on the Cavitation Bucket when Predicting the Full-Scale Behaviour of
a Marine Propeller. Proc. of the 21st Symposium on Naval Hydrodynamics,
1997, pp. 839 850.
Heywood, J.B. (1988) Internal combustion engine fundamentals. McGraw-Hill Co..,
NY, London, 1988.
Hiroyasu, h. (2000) Spray breakup mechanism from the hole-type nozzle and its
applications. Atomization and Sprays, vol. 10, pp. 511-527.
Hirt, C.W. and Nichols, B.D. (1981) Volume of Fluid (VOF) method for the
dynamics of free boundaries. J. Comput. Phys., 39, pp. 201 225.
Hsiao, Chao-Tsung, Chahine, G.L., and Liu, Han-Leih. (2003) Scaling effect on
prediction of cavitation inception in a line vortex flow. Trans. ASME, J. of
Fluids Engineering, 2003, vol. 125, pp. 53 60.
Keller, A., Huber, R. (2001) Cavitation Scale Effects. Empirically found relations and
correlation of cavitation number and hydrodynamic coefficients. Proc. of the
4th Int. Symposium on Cavitation: CAV2001. Available from (28.05.2005):
http://cav2001.library.caltech.edu
Kim, J.-H., Nishida, K., et al. (1997) Characterization of flows in the sac chamber and
the discharge hole of a D.I. Diesel injection nozzle by using a transparent
model nozzle. SAE paper 972942, pp. 11-23.
Knapp, R.T., Daily, J.W., Hammitt, F.G. (1970) Cavitation. (McGraw-Hill).
Kubota, A., Kato, H. and Yamaguchi, H. (1990) Finite-difference analysis of unsteady
cavitation on a two-dimensional hydrofoil. In: Proc. Of the 5th Int. Conf. on
Numerical Ship Hydrodynamics, Hiroshima, Sept. 1990.
Kunz, R.F., Boger, D.A., Stinebring, D.R., et al. (2000) A preconditioned Navier-
Stokes method for two-phase flows with application to cavitation prediction.
Computer & Fluids, v. 29, pp. 849 875.
Laoonual, Y., Yule, A.J, and Walmsley, S.J. (2001) Internal fluid flow and spray
visualization for a large-scale VCO orifice injector nozzle. ILASS-Europe
2001, Zurich 2-6 September, 2001. 6 p.
References 185

Lecoffre, Y. and Bonnin, J. (1979) Cavitation tests and nucleation control. Int.
Symposium on cavitation inception, ASME, 1979, New York, pp. 141 145.
Lecoffre, Y. (1999) Cavitation : bubble trackers. Rotterdam, 1999. 399 p.
Lefebvre, A.H. (1989) Atomization and Sprays. Combustion: An International
Series., Ed. Norman Chigier. Hemisphere Publ. Co., 1989. 421 p.
Lichtarowicz, A., Duggins, R.K and Markland, E. (1965) Discharge coefficients for
incompressible non-cavitating flow through long orifices. J. Mech. Engng.
Sci., vol. 7, no. 2, pp.210 219.
Lichtarowicz, A. and Pearce, I.D. (1974) Cavitation and aeration effects in long
orifices., p. 129 Cavitation. Proc. of the Conference arranged by Fluid
Machinery Group of the Inst. Of Mech. Engineers. Edinburgh, 3-5 Sept, 1974.
Mech. Eng. Publ. Ltd, London, NY. 437 p.
Lindau, J.W., Kunz, R.F., Boger, D.A., et al. (2002) High Reynolds number,
unsteady, multiphase CFD modelling of cavitation flows. Tr. ASME, J. Fluids
Engng., v. 124, pp. 607 616.
Lindau, J.W., Venkateswaran, S., Kunz, R.F., Merkle C.L. (2001) Development of a
fully-compressible multi-phase Reynolds-averaged Navier-Stokes model. 15th
AIAA Computational Fluid Dynamics Conference., AIAA 2001-2648. 13 p.
Liu, Z. and Brennen, C.E. (1998) Cavitation nuclei population and event rates. Trans.
ASME, J. of Fluids Engineering, vol. 120, pp. 728 737.
Malin, M.R. and Waterson, N.P. (1999), Schemes for convection discretisation in
Phoenics. Phoenics J., vol. 12, no. 2, p. 173 201.
Malin, M.R. and Sanchez, L.R. (1988) One-dimensional steady transonic shocked
flow in a nozzle. Phoenics J., vol. 1, no. 2, p. 214-246.
Medvitz, R.B., Kunz, R.F., Boger, D.A., et al. (2002) Performance analysis of
cavitating flow in centrifugal pumps using multiphase CFD. Tr. ASME, J.
Fluids Engng., v. 124, pp. 377-383
Meirer, H.F. et al. (1999) Comparison between Staggered and Collocated Grids in the
Finite-Volume Method Performance for Single and Multiphase Flows.
Computers and Chemical Engng., vol. 23, pp. 247-262.
Mikic, B.B., Rohsenow, W.M., Griffith, P. (1970) On bubble growth rates. Int. J. Heat
Mass Transfer, vol. 13, pp. 657-666.
References 186

Minnaert, M. (1933). Musical air bubbles and the sound of running water. Phil. Mag.,
vol. 16, pp. 235-248.
Nigmatulin, R.I. (1991) Dynamics of Multiphase Media. In 2 vols. NY: Hemisphere.
Nurick , W.H. (1976) Orifice cavitation and its effect on spray mixing. J. Fluids
Engng., 98, pp. 681 689.
Patankar, S.V. and Spalding, D.B. (1972) A Calculation Procedure for Heat, Mass and
Momentum Transfer in Three-dimensional Parabolic Flows. Int. J. Heat Mass
Transfer, vol. 15, p. 1787.
Plesset, M.S. (1949) The dynamics of cavitation bubbles. Trans. ASME, J. Appl.
Mechanics, 16, pp. 228 231.
Plesset, M.S., and Zwick, S.A., (1954) The growth of vapor bubbles in superheated
liquids. Journal of Applied Physics, vol. 25, pp. 493-500.
Poliakov, I. and Semin, V. (1994) Development and Evaluation of New Linear
Equation Solvers for PHOENICS. PHOENICS Journal, Vol. 1, pp34-57.
Pulkrabek, W.W. (1997) Engineering Fundamentals of the Internal Combustion
Engine. Printice-Hall Int. Ltd, London. 411 pp.
Qin, J.-R., et al. (2001) Direct Calculations of Cavitating Flows by the Space-Time
CE/SE Method. AIAA conf. Paper
Rayleigh, Lord (1917) On the pressure developed in a liquid during the collapse of a
spherical cavity. Philosophy Magazine, vol. 34, pp. 94-98.
Rhie, C. M. and Chow, W. L. (1983) Numerical Study of the Turbulent Flow Past an
Airfoil with Trailing Edge Separation. AIAA Journal, 21(11), 1525-1532.
Rood, E.P. (1991) Review mechanism of cavitation inception. Trans. ASME, J.
Fluids Engineering, vol. 113, pp. 163 175.
Roosen, et al. (1996) Untersuchung und Modeleirung des transienten Verhaltens von
Kavitationserscheinungen bei ein- und mehrkomponentigen Kraftstoffen in
schnell durchstromten Dusen. Report of the Institute for Technical
termodynamics. RWTH Aachen (Univ. of Tech.), Germany.
Roosen, P., Unruh, O., Behman, M. (1997) Investigation of cavitation phenomena
inside fuel injector nozzles. Proc. ISATA, Florence, 97LA001, pp. 439 446.
Roosen, P., Unruh, O. (1998) Cavitation-induced flow field inside fuel injector
nozzles. Proc. ISATA, Dusseldorf, 98VR007, pp. 267 273.
References 187

Rosten, H.I. and Spalding, D.B. (1986) Phoenics - Beginners Guide and User
Manual. CHAM TR-100. 116 p.
Rouse, H. and McNown, J.S. (1948) Cavitation and pressure distribution, head forms
at zero angle of yaw. Report of the State Univ. of Iowa, 1948.
Sato, K. and Kakutani, K. (1994) Measurements of cavitation inception. JSME Int.
Journal, Series B, vol. 37, No. 2, pp. 306 312.
Sato, K. and Saito, Y. (2001) Unstable cavitation behaviour in circular-cylindrical
orifice flow. Proc. of the 4th Int. Symposium on Cavitation: CAV2001, 8 p.
Sato, K. and Shimojo, S. (2003) Detailed observations on a starting mechanism for
shedding of cavitation cloud. Proc. of the 5th Int. Symposium on Cavitation:
CAV2003. Osaka, Japan, November 1-4, 2003.
Sauer, J. (2001) Instationr Kavitierende Strmungen Einneues Modell, basierend
auf Front Capturing (VoF) und Blasendynamik. Dr.-Ing. Diss., Karlsruhe
Univ., 2001.
Schmidt, D.P. and Corradini M.L. (2001), The internal flow of diesel fuel injector
nozzles: a review. Int. J. Engine Research, vol. 2, no. 1, pp. 1 22.
Schmidt, D.P., Rutland, C.J., Corradini, M.L., Roosen, P. and Genge, O. (1999)
Cavitation in Two-Dimensional Asymmetric Nozzles. SAE paper 1999-01-
0518, 17 p.
Schmidt, D.P., Rutland, C.J., and Corradini, M.L. (1997) A numerical study of
cavitating flow through various nozzle shapes. SAE paper 971597, 10 p.
Schmidt, D.P., Rutland, C.J., and Corradini, M.L. (1999) A fully compressible two-
dimensional model of high-speed cavitating nozzle. Atomization and Sprays,
v.9.
Schmidt, D.P., Su, Tz.-F., et al. (1995) Detection of cavitation in fuel injector nozzles.
Proc. 8th Int. Symp. on Transport Phenomena in Combustion, San Francisco,
12 p.
Senocak, I. (2002) Computational Methodology for the Simulation of Turbulent
Cavitating Flows. PhD Thesis, University of Florida, 125 p.
Senocak, I. and Shyy, W. (2002) A Pressure-Based Method for Turbulent cavitating
Flow Computations. J. Comp. Phys., v. 176, pp. 363 383.
References 188

Singhal, A.K., Athavale, M.M., Li, H., Jiang Yu (2002) Mathematical basis and
validation of the full cavitation model. Tr. ASME, J. Fluid Eng., vol. 124, pp.
617 624.
Skripov, V.P. (1974) Metastable Liquids. Wiley, New York, 1974.
Soteriou, C., Andrews, R., and Smith, M. (1995) Direct Injection Diesel Sprays and
the Effect of Cavitation and Hydraulic Flip on Atomisation. SAE 950080, pp.
128 153.
Soteriou, C., Andrews, R., Torres, N., Smith, M and Kunkulagunta, R. (2001)
Through the Diesel Nozzle Hole A journey of Discovery II. ILASS-Europe,
Zurich 2-6 September, 2001. 9 p.
Spalding, D.B. (1974) A method for computing steady and unsteady flows possessing
discontinuities of density. CHAM report, 910/2.
Spalding, D.B. (1980) Numerical Computation of Multiphase Fluid Flow and Heat
Transfer. In: Recent Advances in Numerical Methods in Fluids., v.1, eds. C.
Taylor and K. Morgan, Swansea, U.K., pp. 139-168.
Spikes, R.H., and Pennington, G.A. (1959) Discharge coefficient of small submerged
orifices. Proc. Inst. Mech. Engineers, pp. 661 674.
Stinebring, D.R., Billet, M.L., Lindau, J.W. and Kunz, R.F. (2001) Developed
cavitation-cavity dynamics. VKI Lecture Series on Supercavitating Flows.
VKI Press, Brussels, 2001, 20 p.
Stutz, B., and Reboud, J.L. (1997) Experiments on unsteady cavitation. Experiments
in Fluids, vol. 22, pp. 191 198.
Tullis, J.P. (1973) Cavitation Scale Effects for Valves. Journal of the Hydraulics
Division, Trans. ASME., proc. 9874., pp.1109-1128.
Ubbink, O. (1997) Numerical prediction of two-fluid systems with sharp interfaces.
Ph.D. Thesis, Imperial College of Science, Technology and Medicine.
Versteeg, H.K. and Malalasekera, W. (1995) An introduction to computational fluid
dynamics: the finite volume method. Harlow: Longman.
Wallis, G.B. (1969) One-Dimensional Two-Phase Flow, NY: McGraw-Hill.
Wang, G., et al. (2001) Dynamics of Attached Turbulent Cavitating Flows. Progress
in Aerospace Sci., v.37, pp. 551-581.
References 189

Waniewski, T.A. and Brennen, C.E. (1999) The Evolution of Cavitation Events with
Speed and Scale of the Flow. 3rd ASME/JSME Joint Fluids Engineering
Conference, 18-23 July 1999, San Francisco, CA, USA.
Winklhofer, E., Kull, E., Kelz, E., Morozov, A. (2001) Comprehensive hydraulic and
flow field documentation in modl throttle experiments under cavitation
conditions. Proceedings of the ILASS-Europe Conference, Zurich, 2-6
September, 2001, pp. 574 579.
Yabe, T., Xiao, F. and Mochizuki, H. (1995) Simulation technique for dynamic
evaporation process. Nuclera Engineering and Design, vol. 155, pp. 45-53.
Yakhot, V. and Orszag, S. A. (1986) Renormalization Group Analysis of Turbulence:
I. Basic Theory. Journal of Scientific Computing, vol. 1, no. 1, pp. 1-51.
Yount D.E., Kunkle T.D., D'Arrigo J.S., Ingle F.W., Yeung C.M. and Beckman E.L.
(1977) Stabilization of Gas Cavitation Nuclei by Surface-Active Compounds.
Aviat. Space Environ. Med. Vol. 48, issue 3, pp. 185-189.
Yuan, W. Sauer, J., Schnerr, G.H. (2001) Modeling and computation of unsteady
cavitation flows in injection nozzles. Mec. Ind., vol. 2, pp. 383 394.

Publications by the author


Martynov, S.B., Mason, D.J., and Heikal, M.R. Numerical simulation of cavitation
flows based on their hydrodynamic similarity. Accepted for publication in the
International Journal of Engine Research.
Sazhin, S.S., Krutitskii, P.A., Martynov, S.B., Mason, D.J., Heikal, M.R., and
Sazhina, E. M. Transient heating of semitransparent spherical body. Submitted
to the ASME Journal of Heat Transfer.
Martynov, S.B., Mason, D.J., and Heikal, M.R. (2005) Hydrodynamic similarity of
cavitation flows in nozzles. Proc. of the 5th International Symposium on
Multiphase Flow, Heat Mass Transfer and Energy Conversion. Xi'an, China 3-
6 July 2005 (in press).
Sazhin, S.S., Krutitskii, P.A., Martynov, S.B., Mason, D.J., Heikal, M.R., and
Sazhina, E. M. Transient heating of semitransparent droplet. Proc. of the
International Conference on Heat Transfer, Fluid Mechanics and
Thermodynamics HEFAT2005, Cairo, Egypt, 2005 (in press).
References 190

Sazhin, S.S., Abdelghaffar, W.A., Martynov, S.B., Sazhina, E.M. and Heikal, M.R.
and Krutitskii, P.A. (2005) Transient heating and evaporation of fuel droplets:
recent results and unsolved problems. Proc. of the 5th International
Symposium on Multiphase Flow, Heat Mass Transfer and Energy Conversion.
Xi'an, China 3-6 July 2005 (in press).
Martynov, S.B. (2004) Hydrodynamic similarity of cavitation flows. Engineering
Research in Action, Internal Conference, University of Brighton, Brighton, 6
May 2004, pp. 13 16.

Publications by the author before 2002


Kurganov, V.A., Zeigarnik, Yu.A., Maslakova, I.V., Ivanov, F.P. and Martynov, S.B.
The Cooling of Heat-Stressed Surfaces by Chemically Reacting Gases. In:
Heat Transfer in Modern Engineering. Moscow: IVTAN, 1998, p. 48.
Kurganov, V.A., Zeigarnik, Yu.A., Maslakova, I.V., Ivanov, F.P. and Martynov, S.B.
High-Temperature Heat-Shielding Panels with Thermochemical Cooling
Based on the Reaction of Steam Conversion of Methane. High Temperature,
vol. 38, no. 6, 2000, pp. 926-937.
Ivanov, F.P., Zeigarnik, Yu.A., Kurganov, V.A., Martynov, S.B. and Maslakova, I.V.,
Thermochemical Cooling Based on the Reaction of Steam Conversion of
Methane. In: Scientific Basement of the Technologies of XXI century.,
Leont'ev, A.I., et al., Eds., Moscow: UNPTs "Energomash", 2000, 136 p.
Martynov, S.B., The Structure of Pressure Drop in Strongly Heated Gas Flow in
Tube. Proc. of the VI-th All-Russian Conference of Young Scientists.
Novosibirsk, April 25 - 28, 2000.
Kurganov, V.A. and Martynov, S.B. Characteristics of Microporous Catalytic
Coatings. / Proc. of Symposium "Thermochemical Processes in Plazma
Aerodynamics". St.-Petersburg, May 30 - June 3. Radioavionika, 2000.
Kurganov, V.A. Zeigarnik, Yu.A. and Martynov, S.B. On Actual Rates of Steam
Reforming of Methane on Nickel Catalysts./ Proc. of Symposium
"Thermochemical Processes in Plasma Aerodynamics". St.-Petersburg,
September 9 - 12. Radioavionika, 2001.
Kurganov, V.A. and Martynov, S.B. Calculation of Heat Transfer and Drag in Tubes
under Conditions of Laminar Flow of Gases with Varying Properties. High
Temperature, vol. 40, no. 6, 2002, pp. 881-891.
Appendices 191

Appendices
Appendix A. Q1 input file for PHOENICS

TALK=T;RUN( 1, 1);VDU=VGAMOUSE
************************************************************
Group 1. Run Title
TEXT(RooRnd3DM1 SuperC Str0 C0.e5 MF1.e-6 Rv1)
************************************************************
Group 2. Transience
STEADY = T
************************************************************
INTEGER(NyN,NzW,NzT, NxI)
INTEGER(NzI,NzE,NzC,NzN, NzEC1,NzEC2, NzEC12)
INTEGER(NzCext, NzIE,NzIEC,NzCN, NzSum)
REAL(Y1i,Y2i,Y3i,YeW,Y3c,Y3e,Y3n,Y3b, Y3X)
REAL(Z1i,Z2i,Z3i,Zew,Z3c,Z3e,Z3n,Z3b, Z1n,Z1c,Z1e)

REAL(Hin,Hnz,Lin,Len,Lcr,Lnz, DX,DYc,DZc)
REAL(D_X,D_Y,D_Z)

REAL(Uin,Pout, TKEin,EPSin, MFin, RhoIN)


REAL(RhoL,RhoV,ELnu,EVnu,Pvap,Ccav,MFinit)

--- Physical properties of the phases


=== Water (20 deg.C):
RhoL = 1000.; RhoV = 1.
ELnu = 1.e-6; EVnu = 1.e-5
Pvap = 2300.

--- Parameters of cavitation model

Ccav = 0.e5
MFin = 0.
MFinit=1.e-6
RHOin = 1./( (1.-MFin)/RhoL + MFin/RhoV )
--- Boundary conditions
14.0 m/s for pressure drop of 80:21 bars
15.1 m/s for pressure drop of 80:11 bars
Uin = 14.
Pout = 0.
WALLCO=GRND2
TKEin = 0.* (0.02*Uin)**2
EPSin = 0.* 0.09*TKEin**2/(200*ENUL)

--- Reference pressure (at the nozzle outlet)


PRESS0=21.e5

--- Nozzle dimensions:


Hin=1.e-3; Hnz=.14e-3; DX=0.1e-3
Lin=1.e-3; Lnz= 1.e-3; Len=1.e-3
DYc=2.8e-5; DZc=DYc
Lcr=3.*Hnz

--- Number of points for the blocks I, E, C and N


--- I=inlet, E=entry, C=corner and N=nozzle

**** MESH1 ****


NxI = 5
NyN = 9
NzI = 5; NzE = 6; NzC = 3; NzN = 32
NzEC12 = 12
Appendices 192

NzCext= NzEC12+NzC - NzE


NzIE=NzI+NzE; NzIEC=NzIE+NzCext; NzCN= NzC+NzN
NzSum = NzIEC + NzN

* coordinates of points relative to the point P1


Y1i=0.; Y2i=Hin; Y3i=Hin; Y3e=Hnz+DYc; Y3c=Hnz; Y3n=Hnz;
Y3x=Y3e+2.*Hnz
Y3x=Y3i-0.7*Len

Z1i=0.; Z3i=Lin; Z3e=Lin+Len; Z3c=Z3e+DZc; Z3n=Z3e+Lnz


Z2i=0.; Z1e=0.9*Z3i; Z1n=Z3e;
Z1c=Z1e+0.6*Len

Y3b=Y3e-0.7*DYc; Z3b=Z3c-0.7*DZc

YeW=0.757*Hin
ZeW=Z3i+0.652*Len
************************************************************
Group 6. Body-Fitted coordinates
BFC = T
* Define points (indexes i,e,c,n denote Blocks):
GSET(P, P1i, 0., Y1i, Z1i)
GSET(P, P2i, 0., Y2i, Z2i)
GSET(P, P3i, 0., Y3i, Z3i)
GSET(P, PeW, 0., YeW, ZeW)
GSET(P, P3e, 0., Y3e, Z3e)
GSET(P, P3c, 0., Y3c, Z3c)
GSET(P, P1n, 0., Y1i, Z3c)
GSET(P, P1c, 0., Y1i, Z1c)
GSET(P, P1e, 0., Y1i, Z1e)

* Define lines:
GSET(L,L1i, P1i, P2i, NyN, 1.0)
GSET(L,L2i, P2i, P3i, NzI, 1.3)
GSET(V,V1W, P3i, PeW, P3e)
GSET(L,C1W, P3i, P3e, NzEC12,-1.5, ARC,0.,Yew,Zew)

GSET(V,V2w, P3e, P3b, P3c)


GSET(L,C2w, P3e, P3c, NzC, 1.0, ARC,0.,Y3b, Z3b)
GSET(L,L1n, P3c, P1n, NyN, 1.5)
GSET(L,L4c, P1n, P1c,NzCext,-0.7)
GSET(L,L4e, P1c, P1e, NzE, -0.9)
GSET(L,L4i, P1e, P1i, NzI, 1.0)
* Set and match frames:
D_X=DX; D_Y=Hin; D_Z=Z3n
GSET(D,NxI,NyN,NzSum,D_X,D_Y,D_Z)
GSET(F,F1,P1I,-,P2I,P3I.PEW.P3X.P3E,P3C,-,P1N,P1C.P1E)
GSET(F,F1,P1i,-,P2i,P3i.P3e,P3c,-,P1n,P1c.P1e)
GSET(M,F1,+J+K,1,1,1,LAP5)
GSET(C,I:NxI+1:,F1,I1,1,NyN,1,NzIEC,+,D_X,0,0,INC,1)
GSET(C,K:NzSum+1:,F1,K:NzIEC+1:,1,NxI,1,NyN,+,0,0,Lnz,INC,1)

************************************************************
Group 7. Variables: STOREd,SOLVEd,NAMEd

ONEPHS = T
NAME(C1) = MF
NAME(C2) = VF
NAME(C3) = DWDY
NAME(C4) = DVDZ
NAME(C5) = ten2
NAME(C6) = strn
NAME(C7) = strs
SOLVE(P1,U1,V1,W1,MF)
TURMOD(KERNG)
SOLUTN(KE ,Y,Y,Y,N,N,Y)
SOLUTN(EP ,Y,Y,Y,N,N,Y)
Appendices 193

SOLUTN(MF, Y,Y,Y,N,N,Y)
STORE(PRPS, RHO1,UCMP,VCMP,WCMP,UCRT,VCRT,WCRT)
STORE(VF,MEM2, DWDY,DVDZ,ENUT, ten2,strn,strs)
************************************************************
Group 8. Terms (in differential equations) & devices
UCONV=T; DENPCO=T
RG(1)=RhoL; RG(2)=RhoV; RG(12)=RhoL/RhoV - 1.;
RG(3)=ELnu; RG(4)=EVnu-ELnu;
RG(5)=RhoL*(1000.)**2; RG(6)=RhoV*(300.)**2
RG(5)=1.e9; RG(6)=1.e5
RG(7)= 1.
RG(8)= 0.195*RhoL
RG(9)=Pvap; RG(10)=Ccav; RG(11)=MFinit
RG(20)=PRESS0
IG(1)= index for p_cr: 0=pv, 1-pv+0.39/2RhoK, 2,3=pv-2MuS
IG(1)=3

GALA=F

TERMS(MF, N,Y,N,Y,Y,N)
************************************************************
Group 9. Properties of the medium
RHO1- GRND; ENUL= GRND
DRH1DP=1.e-6
=== DRH1DP=GRND
NAMFI='L3D'
L3D - inlet cavitation
LSD - super-cavitation
FIINIT(P1)=READFI
FIINIT(WCRT)=READFI; FIINIT(VCRT)=READFI; FIINIT(UCRT)=READFI
FIINIT(W1) =READFI; FIINIT(V1) =READFI; FIINIT(UCRT)=READFI
FIINIT(KE) =READFI; FIINIT(EP) =READFI
FIINIT(MF)=READFI
************************************************************
Group 13. Boundary & Special Sources
BFCA = RHOin
* Inlet
PATCH(BFCIn, LOW, 1, NxI, 1, NyN, 1, 1, 1, LSTEP)
COVAL(BFCIn ,P1 , FIXFLU , GRND1)
COVAL(BFCIn ,U1 , ONLYMS , GRND1)
COVAL(BFCIn ,V1 , ONLYMS , GRND1)
COVAL(BFCIn ,W1 , ONLYMS , GRND1)
COVAL(BFCIn ,UCRT, ONLYMS , 0.)
COVAL(BFCIn ,VCRT, ONLYMS , 0.)
COVAL(BFCIn ,WCRT, ONLYMS , Uin)
COVAL(BFCIn ,MF , ONLYMS , MFin)
COVAL(BFCIn ,KE , ONLYMS , TKEin)
COVAL(BFCIn ,EP , ONLYMS , EPSin)

* Outlet
PATCH(OUTLET, OUTFLO, 1,NxI, 1,NyN, NzSum,NzSum, 1,LSTEP)
COVAL(OUTLET ,P1 , 1.*RHO1 , Pout)
COVAL(OUTLET ,U1 , ONLYMS , SAME)
COVAL(OUTLET ,V1 , ONLYMS , SAME)
COVAL(OUTLET ,W1 , ONLYMS , SAME)
COVAL(OUTLET ,MF , ONLYMS , SAME)
COVAL(OUTLET ,UCRT, ONLYMS , SAME)
COVAL(OUTLET ,VCRT, ONLYMS , SAME)
COVAL(OUTLET ,WCRT, ONLYMS , SAME)
COVAL(OUTLET ,KE , ONLYMS , SAME)
COVAL(OUTLET ,EP , ONLYMS , SAME)
Appendices 194
* Wall
PATCH(Wall, NWALL, 1, NxI, NyN, NyN, 1, NzSum, 1,LSTEP)
COVAL(Wall, U1, WALLCO, 0.0)
COVAL(Wall, V1, WALLCO, 0.0)
COVAL(Wall, W1, WALLCO, 0.0)
COVAL(WALL, UCRT, WALLCO, 0.0)
COVAL(WALL, VCRT, WALLCO, 0.0)
COVAL(WALL, WCRT, WALLCO, 0.0)
COVAL(Wall, KE, WALLCO, WALLCO)
COVAL(Wall, EP, WALLCO, WALLCO)

PATCH(WX, EWALL, NxI,NxI, 1,NyN, 1,NzSum, 1,LSTEP)


COVAL(WX, W1, WALLCO, 0.0)
COVAL(WX, V1, WALLCO, 0.0)
COVAL(WX, U1, WALLCO, 0.0)
COVAL(WX, KE, WALLCO, WALLCO)
COVAL(WX, EP, WALLCO, WALLCO)

**-- Source term for the evaporation


PATCH(SEVAP3, VOLUME, 1,NxI, 1,NyN, NzIE,NzSum, 1,LSTEP)
COVAL(SEVAP3, MF, FIXFLU, GRND)

**-- Source term for the condensation


PATCH(SCOND3, VOLUME, 1,NxI, 1,NyN, 1,NzSum, 1,LSTEP)
COVAL(SCOND3, MF, GRND, 0.0)
**-- Prevent cavitation upstream the nozzle
PATCH(LiquidI, VOLUME, 1,NxI, 1,NyN, 1,NzIE, 1,LSTEP)
COVAL(LiquidI, MF, FIXVAL, MFin)

************************************************************
Group 15. Terminate Sweeps
SELREF = T; RESFAC = 1.E-03
RESREF(P1)=1.E-6
RESREF(U1)=1.E-6; RESREF(V1)=1.E-6; RESREF(W1)=1.E-6
RESREF(MF)=1.E-6; RESREF(KE)=1.E-6; RESREF(EP)=1.E-6

LSWEEP = 5000
************************************************************
Group 17. Relaxation
(for steady-state sweeps)
RELAX(RHO1, LINRLX,0.2)
RELAX(P1, LINRLX,1.e-01)
RELAX(U1, FALSDT,1.e-00)
RELAX(W1, FALSDT,1.e-00)
RELAX(V1, FALSDT,1.e-00)
RELAX(MF, FALSDT,1.e-02)
RELAX(KE, LINRLX,1.e-01)
RELAX(EP, LINRLX,3.e-01)
KELIN=2

************************************************************
Group 18. Limits
VARMAX(P1) = 1.e10; VARMIN(P1) = -1.e10
VARMAX(U1) = 100.*Uin; VARMIN(U1) = -100.*Uin
VARMAX(V1) = 100.*Uin; VARMIN(V1) = -100.*Uin
VARMAX(W1) = 100.*Uin; VARMIN(W1) = -100.*Uin
VARMAX(MF) = 1.00; VARMIN(MF) = 1.e-10

************************************************************
Group 19.
NAMGRD=CONV

************************************************************
Group 22. Monitor Print-Out
IXMON = 1; IYMON = NyN-1; IZMON = NzIEC+1
NPRMNT = 1
Appendices 195
TSTSWP = -1

NXPRIN=1; IPROF=2
PATCH(Center,PROFIL,1,1,1,1,1,NzSum,1,LSTEP)
PLOT(Center,P1,0.0,0.0)
PLOT(Center,W1,0.0,0.0)
PLOT(Center,V1,0.0,0.0)
PATCH(NearWall,PROFIL,1,1,NyN,NyN,1,NzSum,1,LSTEP)
PLOT(NearWall,P1,0.0,0.0)
PLOT(NearWall,W1,0.0,0.0)
PLOT(NearWall,V1,0.0,0.0)
PLOT(NearWall,VF,0.0,0.0)

************************************************************
CCM=T
NONORT=F; LSG3=T; LSG4=T

--- LSG7 permits CCM-solver to use higher order schemes.


LSG7=T
SCHMBEGIN
VARNAM MF SCHEME SUPERB
VARNAM UC1 SCHEME SMART
VARNAM VC1 SCHEME SMART
VARNAM WC1 SCHEME SMART
SCHMEND

************************************************************

STOP
Appendices 196

Appendix B. GROUND module

C FILE NAME GROUND.FTN--------------------------------140796


SUBROUTINE GROUND
INCLUDE '/d_phoe22/d_includ/satear'
INCLUDE '/d_phoe22/d_includ/grdloc'
INCLUDE '/d_phoe22/d_includ/grdear'
INCLUDE '/d_phoe22/d_includ/grdbfc'

COMMON/GENI/NXNY,IGFIL1(8),NFM,IGF(21),IPRL,IBTAU,ILTLS,IGFIL(15),
1
ITEM1,ITEM2,ISPH1,ISPH2,ICON1,ICON2,IPRPS,IRADX,IRADY,IRADZ,IVFOL
COMMON/DRHODP/ITEMP,IDEN
CXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX USER SECTION STARTS:

C***************************************************************
C
C--- GROUP 1. Run title and other preliminaries
C
1 GO TO (1001,1002,1003),ISC
C
1001 CONTINUE
C Auxilary arrays.......................................... sbm
CALL MAKE(GRSP1)
CALL MAKE(GRSP2)

C***************************************************************
C
C--- GROUP 9. Properties of the medium (or media)
C
C The sections in this group are arranged sequentially in their
C order of calling from EARTH. Thus, as can be seen from below,
C the temperature sections (10 and 11) precede the density
C sections (1 and 3); so, density formulae can refer to
C temperature stores already set.
9 GO TO 91,92,93,94,95,96,97,98,99,900,901,902,903,904,
905),ISC
C***************************************************************
91 CONTINUE
C * ------------------- SECTION 1 ---------------------------
C For RHO1.LE.GRND--- density for phase 1 Index DEN1
IF (RHO1.EQ.GRND) THEN
C.....RG(1)=RhoL,..RG(12)=RhoL/RhoV - 1......................sbm
C............. FN6(Y,X,A,B).....Y = 1/(A + B*X)................
C............. FN25(Y,A)........Y = A*Y........................
C write(14,*) RG(2)
CALL FN6(DEN1,C1,1.,RG(12))
CALL FN25(DEN1,RG(1))
CC CALL FN1(DEN1,RG(1))
C C2 is used for the Volume Fraction of the Vapour phase
C VF = (RhoMix/RhoV) * MF
C........ FN21(Y,X1,X2,A,B).........Y = A + B*X1*X2............
CALL FN21(C2,C1,DEN1,0.,1./RG(2))

ENDIF
RETURN

96 CONTINUE
C * ------------------- SECTION 6 ---------------------------
C For ENUL.LE.GRND--- reference laminar kinematic viscosity
C Index VISL
IF (ENUL.EQ.GRND) THEN
C.......RG(3)=ELnu,..RG(4)=ELnu-EVnu........................sbm
C.......FN2(Y,X,A,B).....Y = A + B*X...........................
CALL FN2(VISL,C1,RG(3),RG(4))
Appendices 197

ENDIF
RETURN

C***************************************************************
C
C--- GROUP 13. Boundary conditions and special sources
C Index for Coefficient - CO
C Index for Value - VAL
13 CONTINUE
GO TO (130,131,132,133,134,135,136,137,138,139,1310,
11311,1312,1313,1314,1315,1316,1317,1318,1319,
1320,1321),ISC
130 CONTINUE
C------------------- SECTION 1 ------------- coefficient = GRND
C...........................................................sbm
IF ((NPATCH.EQ.'OUTLET').AND.(INDVAR.EQ.P1)) THEN
C.......Set coefficient in fixed pressure type of outflow....
C.......boundary to the local fluid density DEN1.............
CALL FN0(CO,DEN1)
ENDIF
C............................................................

IF ((NPATCH.EQ.'SCOND3').OR.(NPATCH.EQ.'SCOND4').
& AND.(INDVAR.EQ.C1)) THEN
C.......Condensation.........................................
C IG(1)= index for p_cr:
C 0=pv, 1-pv+0.39/2RhoK, 2,3=pv-2MuS
C RG: 1-RhoL, 2-RhoV, 3-NuL, 4-NuL-NuV,
C 7-"2" for S=2*Mu*S, 8-(0.39/2.)*RhoL, 9-Pv,
C 10-Ccav, 11-MFmin, 20-PRESS0
C FN0(Y,X): Y = X
C FN2(Y,X,A,B): Y = A + B*X
C FN34(Y,X,A): Y = Y + A*X
C FN33(Y,A): Y = Y + A
C FN25(Y,A): Y = Y * A
C FN26(Y,X): Y = Y * X
C FN46(Y,X,A,B): Y = Y *(A +B*X)
C FN53(Y,X1,X2,A): Y = Y + A *X1*X2
C FN56(Y,X1,X2,X3,A): Y = A * X1*X2/X3
C FN57(Y,X1,X2,X3,A): Y = Y + A* X1*X2*X3
C FN10(Y,X1,X2,A,B1,B2): Y = A + B1*X1 + B2*X2
C FN12(Y,X1,X2,X3,A,B1,B2,B3): Y = A+ B1*X1 + B2*X2 + B3*X3

C.....Calculate strain and stress ...........................


C......C6 = STRAIN = 0.5*abs(dW/dY + dV/dZ)
CALL FN0(C6,C3)
CALL FN34(C6,C4,1.)
CALL FN40(C6)
CALL FN25(C6,0.5)
C......C7 = STRESS = 2*Rho*(VisL + VisT)*S
CALL FN10(C7,VISL,VIST,0.,1.,1.)
CALL FN46(C7,C6,0.,2.)
CALL FN26(C7,DEN1)
C......C5 = Tension =
C = P1 - GRSP2 = P1 - 2*Mu*[1 + (Mut/Mu)]*S
c CALL FN0(C5,P1)
c CALL FN34(C5,C7,-1.)

IF (IG(1).le.1) THEN

C *** GRSP1 = (P+PRESS0) - Pv


CALL FN2(GRSP1,P1,RG(20)-RG(9),1.)

C *** Singhal' model for the critical pressure:


C GRSP1 = (P+PRESS0) - { Pv + (0.39/2)*Rho*KE }
IF (IG(1).eq.1) CALL FN34(GRSP1,KE,-RG(8))

ELSEIF (IG(1).le.3) THEN


Appendices 198
C GRSP1 = (P+PRESS0 - Pv) - 2*(Mu+MuT)*S
C GRSP1 switches between evaporation and condensation
IF (IG(1).eq.2) THEN
C *** S = STRAIN approximately = 0.5*SQRT(GENK)
write(*,*)'model',IG(1)
CALL FN0(GRSP1,GENK)
CALL FN30(GRSP1)
ELSE
C *** S = STRAIN = 0.5*abs(dW/dY + dV/dZ)
c write(*,*)'model',IG(1)
CALL FN0(GRSP1,C3)
CALL FN34(GRSP1,C4,1.)
CALL FN40(GRSP1)
ENDIF
CALL FN25(GRSP1,0.5)
C VisL + VisT
CALL FN10(GRSP2,VISL,VIST,0.,1.,RG(7))
C Rho*(Enu + EnuT)*S
CALL FN46(GRSP2,GRSP1,0.,2.)
CALL FN26(GRSP2,DEN1)

CALL FN10(GRSP1,P1,GRSP2,RG(20)-RG(9),1.,-1.)

ENDIF

C GRSP1 = (P+PRESS0 - Pv) - 2*(Mu+MuT)*S


C GRSP1 switches between evaporation and condensation
C and determines the rate of evaporation
C GRSP2 = P + PRESS0 - Pv
C GRSP2 determines the rate of condenstaion (?)
c CALL FN2(GRSP2,P1,RG(20)-RG(9),1.)

CALL SourceMF(CO,C1,GRSP1,RG(10),RG(1),RG(2),RG(11))

C CO = SourceMF, if p > p_cr


C CO = 0, if p < p_cr
CALL SignDP(GRSP1)
CALL FN33(GRSP1,-1.)
CALL FN25(GRSP1,-1.)
CALL FN26(CO,GRSP1)

ENDIF

RETURN
1311 CONTINUE
C------------------- SECTION 12 ------------------- value = GRND
C...........................................................sbm
IF ((NPATCH.EQ.'SEVAP3').OR.(NPATCH.EQ.'SEVAP4')
& .AND.(INDVAR.EQ.C1)) THEN
C.......Evaporation............................................

IF (IG(1).le.1) THEN
C *** GRSP1 = (P+PRESS0) - Pv
CALL FN2(GRSP1,P1,RG(20)-RG(9),1.)
C *** Singhal' model for the critical pressure:
C GRSP1 = (P+PRESS0) - { Pv + (0.39/2)*Rho*KE }
IF (IG(1).eq.1) CALL FN34(GRSP1,KE,-RG(8))

ELSEIF (IG(1).le.3) THEN


C GRSP1 = (P+PRESS0 - Pv) - 2*(Mu+MuT)*S
C GRSP1 switches between evaporation and condensation
IF (IG(1).eq.2) THEN
C *** S = STRAIN approximately = 0.5*SQRT(GENK)
CALL FN0(GRSP1,GENK)
CALL FN30(GRSP1)
ELSE
C *** S = STRAIN = 0.5*abs(dW/dY + dV/dZ)
CALL FN0(GRSP1,C3)
Appendices 199
CALL FN34(GRSP1,C4,1.)
CALL FN40(GRSP1)
ENDIF
CALL FN25(GRSP1,0.5)
C VisL + RG(7)*VisT
CALL FN10(GRSP2,VISL,VIST,0.,1.,RG(7))

C Rho*(Enu + RG(7)*EnuT)*S
CALL FN46(GRSP2,GRSP1,0.,2.)
CALL FN26(GRSP2,DEN1)

C C5 = Tension (includes RG(7)) =


C = P1 - GRSP2 = P1 - 2*Mu*[1 + RG(7)*(Mut/Mu)]*S
CALL FN0(C5,P1)
CALL FN34(C5,GRSP2,-1.)

CALL FN10(GRSP1,P1,GRSP2,RG(20)-RG(9),1.,-1.)

ENDIF

CALL SourceMF(VAL,C1,GRSP1,RG(10),RG(1),RG(2),RG(11))

CALL SignDP(GRSP1)
C VAL = 0, if p > p_cr
C VAL = SourceMF, if p < p_cr
C FN26: Y = Y * X
CALL FN26(VAL,GRSP1)
C VAL = SourceMF * (MFinit + MF)
C MFinit=RG(11)
C.............FN46(Y,X,A,B).....Y = Y*(A + B*X)................
CALL FN46(VAL,C1,RG(11),1.)

ENDIF

RETURN

C***************************************************************
Subroutine SignDP(K1)
C K1 = (res) = (0.) if K1>0; (1.) if K1<0;
C (arg) = P - Pvap.
COMMON F(1)
COMMON /IGE/IXF,IXL,IYF,IYL,IGFILL(21)

C write(14,*)' SignDP in progress..'


CALL L0F1(K1,I,IADD,'SignDP')

DO IX=IXF, IXL
I=I+IADD
DO IY=IYF, IYL
I=I+1
IF ( F(I).GE.0.) THEN
F(I) = 0.
ELSE
F(I) = 1.
ENDIF
END DO
END DO

End

C***************************************************************
Subroutine SourceMF(K1,K2,K3, Ccav, RhoL,RhoV, MFo)
REAL Ccav, RhoL,RhoV
REAL MFo

C Strain is taken into account in K3


Appendices 200
C Subroutine returns an absolute value of
C the SourceTerm divided by the mass fraction MF.
C The result can be applied when calculating the Value
C for the evaporation source-term (treated explicitly):
C VAL = SourceMF * MF,
C and when calculating the Coefficient for
C the condensation source-term (treated implicitly):
C Co = SourceMF.

C K1 = (rez) source;
C K2 = (arg) mass fraction;
C K3 = (arg) = P1+PRESS0 - 2*Mu*GENK - Pv

COMMON F(1)
COMMON /IGE/IXF,IXL,IYF,IYL,IGFILL(21)
CALL L0F3(K1,K2,K3,I,I2M1,I3M1,IADD,'SourceMF')

DO IX=IXF, IXL
I=I+IADD
CDIR$ IVDEP
DO IY=IYF, IYL
I=I+1
C .......Correct function f(MF) derived from Yuans model....
F(I) = ( (RhoV/RhoL)* ((1.-F(I2M1+I)+MFo)**4. /
& (F(I2M1+I)+MFo))) **0.3334
& / ( F(I2M1+I) + (RhoV/RhoL)* (1.-F(I2M1+I)) )
& * Ccav * SQRT( ABS(F(I3M1+I))/ RhoL)

C........Avoid condensation of small bubbles (nuclei)


C........and evaporation of small droplets:
IF ( F(I3M1+I) .GE. 0.) THEN
IF (F(I2M1+I).LE.MFo) F(I)=0.
ELSE
IF (F(I2M1+I).GE.(1.-MFo)) F(I)=0.
ENDIF

END DO
END DO

End
Appendices 201

Appendix C. Study of grid dependence


The results of a CFD calculation are known to be affected by resolution of
computational mesh, especially in the regions of high gradients of velocity and other
flow variables. In order to select a mesh of sufficient density, grid dependence
analysis should be performed.

Table C- 1. Grids used in the study of mesh dependence.

Mesh Total number of cells Z (mm) Mesh structure in Y-Z plane at the
nozzle inlet corner

3 5 33 =
0 0.05
495

5 9 52 =
1 0.03125
2340

9 16 90 =
2 0.01786
12 960

15 25 149 =
3 0.01042
55 875

The grid dependence study was performed for the water flow in Roosens nozzle
(Fig. 5-1) at the inlet velocity u in = 14 m/s (Table 5-1). The boundary conditions and
numerical parameters are described in section 5.2. Four structured grids of different
resolutions were used in the study (Table C-1). Ten thousand sweeps (global
Appendices 202

iterations of the prediction-correction algorithm) were made to achieve the


convergence. The results of numerical calculations using these grids are compared to
each other for the local pressure fields (Fig. C-1), overall pressure drop across the
nozzle and CPU time consumed (Table C-2).

Fig. C- 1. Pressure distributions along the nozzle centreline (CL) and near the
wall, predicted using different meshes (Table C-1).

Table C- 2. The pressure drops, the minimum pressures at the nozzle throat,
the discharge coefficients of nozzle predicted and the CPU time spent for
computations using different meshes (Table C-1).
p1 p 2 , p min p 2 , CPU time,
Mesh C d (Eq. 2.15)
MPa MPa hours
0 5.74 -3.24 0.933 0.24
1 5.73 -5.19 0.934 0.72
2 5.60 -5.90 0.945 2.27
3 5.50 -6.06 0.953 10.1

Fig. C-1 shows similar pressure distribution along the nozzle, when using the
three meshes. The calculated pressure drop across the nozzle p1 p 2 do not reveal
Appendices 203

significant grid dependence. This results in variation in the discharge coefficient of


nozzle C d with the mesh, when changing from Mesh 0 to Mesh 3, about 2% (Table
C-2). Similar shapes of the pressure curves and close values of the discharge
coefficient predicted using different meshes resulted from application of the high-
order differencing scheme (4.9) for the discretisation of convection terms in the
momentum equation. On the other hand Fig. C-1 shows that the local minimum
pressure p min p 2 is strongly affected by the mesh size.

Fig. C- 2. Effect of mesh size on the pressure drop p1 p 2 , relative minimum


pressure in the nozzle throat p min p 2 .

In order to assess grid dependence of the solution, the behaviour of the pressure
drop p1 p 2 , the local pressure minimum p min p 2 and the pressure residual with
the mesh size, were plotted (Fig. C-2 a, b) as functions of cells size Z (Table C-1).
Fig. C-2 reveals that the pressure drop p1 p 2 does not exhibit any significant grid
dependence, while minimum value of pressure at the nozzle throat decreases with Z
reaching the level around 6 MPa .
Also, Fig C-2 indicates the accuracy of solution. According to equations (2.27)
and (2.53) cavitation is governed by the local pressure. In terms of prediction of the
cavitation onset, it becomes important to describe accurately the maximum liquid
tension in the flow pv p min . Using Mesh 1 (Table C-1, Fig. C-2) the local
minimum in the pressure field can be captured with the accuracy about 10 %, while
using Mesh 2 with 3%. Fig C-1 indicates that for the sharp-entry nozzles the
Appendices 204

minimum pressure p min p 2 is expected to be of the order of magnitude


( p1 p 2 ) . Taking into account that in Diesel injectors

( p1 ~ 10 3 MPa) >> ( p 2 ~ 10 MPa ) , the absolute values of p min are expected to be

much lower than the saturation pressure in liquid pv ~ 10 3 Pa. Consequently, the
accuracy of 10 % is sufficient for prediction of the cavitation events at the degree of
accuracy of measurements. The analysis performed justifies using single-block
structured meshes with similar resolution to Mesh 1 in Table C-1, for the practical
calculations. This choice of mesh optimises CPU time (Table C-2) for required
accuracy.
Appendices 205

Appendix D. Transient heating of a semi-transparent


spherical body

Вам также может понравиться