Vous êtes sur la page 1sur 159

NUMERICAL SIMULATION OF TWO-PHASE ANNULAR FLOW

by
Joseph Michael Rodriguez

A Thesis Submitted to the Graduate


Faculty of Rensselaer Polytechnic Institute
in Partial Fulfillment of the
Requirements for the degree of
DOCTOR OF PHILOSOPHY
Major Subject: MECHANICAL ENGINEERING

Approved by the
Examining Committee:
_________________________________________
Richard T. Lahey, Jr., Thesis Co-Adviser

_________________________________________
Kenneth E. Jansen, Thesis Co-Advisor

_________________________________________
Donald A. Drew, Member

_________________________________________
Michael Z. Podowski, Member

_________________________________________
Gretar Tryggvason, Member

Rensselaer Polytechnic Institute


Troy, New York
July, 2009
(For Graduation August 2009)

CONTENTS
NUMERICAL SIMULATION OF TWO-PHASE ANNULAR FLOW............................ i
CONTENTS ...................................................................................................................... ii
LIST OF TABLES............................................................................................................. v
LIST OF FIGURES .......................................................................................................... vi
ACKNOWLEDGMENT ................................................................................................. xii
ABSTRACT ................................................................................................................... xiii
1. INTRODUCTION ....................................................................................................... 1
2. HISTORICAL REVIEW OF ANNULAR FLOW MODELING................................ 4
2.1

Annular Flow Description.................................................................................. 4

2.2

Multi-Field Two-Fluid Model............................................................................ 5

2.3

Overview of Closure Models ............................................................................. 7


2.3.1

Interfacial Shear ..................................................................................... 8

2.3.2

Droplet Entrainment............................................................................. 11

2.3.3

Droplet Size.......................................................................................... 13

2.3.4

Liquid Film Height............................................................................... 16

2.3.5

Lateral or Levitation Force................................................................... 18

3. OVERVIEW OF INTERFACE TRACKING METHODS....................................... 22


4. DESCRIPTION OF PHASTA-IC ............................................................................. 25
4.1

Finite Element Discretization of the Incompressible Navier-Stokes Equations


.......................................................................................................................... 25
4.1.1

Governing Equations............................................................................ 25

4.1.2

Finite Element Formulation ................................................................. 26

4.2

Generalized Alpha Method .............................................................................. 29

4.3

Level Set Method ............................................................................................. 31


4.3.1

Governing Model ................................................................................. 31

4.3.2

Finite Element Formulation ................................................................. 33

ii

4.3.3
4.4

Volume Constraint on Level Set Redistance Step ............................... 34

Continuum Surface Tension (CST) Model ...................................................... 36

5. CANONICAL TEST PROBLEMS ........................................................................... 39


5.1

Rotating Semicircle.......................................................................................... 39

5.2

Analysis of 2D Dam Break .............................................................................. 42

5.3

Analysis of Solitary Wave ............................................................................... 46

6. PARALLEL MESH ADAPTATION ........................................................................ 50


6.1

Code Description.............................................................................................. 50

6.2

Example Problem-A Convecting Sphere ......................................................... 52

7. SELECTION OF EXPERIMENTAL DATA FOR SIMULATION ......................... 55


7.1

Selection of Data Set........................................................................................ 55

7.2

Reducing the Experimental Data ..................................................................... 56

7.3

Estimating Minimum Turbulent Scales ........................................................... 57

7.4

Selecting the Test Run for the Simulation ....................................................... 58

8. SIMULATION OF TWO-PHASE ANNULAR FLOW ........................................... 62


8.1

Problem Description ........................................................................................ 62


8.1.1

Mesh Description ................................................................................. 63

8.1.2

Initial and Boundary Conditions .......................................................... 64


8.1.2.1 Effective-Viscosity Wall Function ........................................ 66
8.1.2.2 Verification of the Effective Viscosity Wall Function .......... 68

8.1.3
8.2

Simulation History ............................................................................... 68

Data Analysis Procedures ................................................................................ 69


8.2.1

Classification of Phase Fields .............................................................. 70

8.2.2

Evaluation of Data on Exit Plane ......................................................... 71


8.2.2.1 Description of Procedure ....................................................... 71
8.2.2.2 Computed Quantities ............................................................. 73

8.2.3

Evaluation of Wall Shear Stress........................................................... 75

iii

8.3

8.2.4

Evaluation of Total Local Shear Stress................................................ 76

8.2.5

Evaluation of Average Shear Stress Distribution................................. 78

Results and Discussion..................................................................................... 80


8.3.1

Void Fraction ....................................................................................... 87

8.3.2

Field Populations.................................................................................. 92

8.3.3

Field Mass Flow Rates ......................................................................... 94

8.3.4

Film Thickness ................................................................................... 102

8.3.5

Dispersed Field Size and Interfacial Area Density ............................ 107


8.3.5.1 Liquid Droplets .................................................................... 108
8.3.5.2 Steam Bubbles ..................................................................... 113

8.3.6

Wall Shear Stress ............................................................................... 118

8.3.7

Total Shear Stress & Interfacial Shear Stress .................................... 119

8.3.8

Film Thickness vs. Wall Shear........................................................... 126

9. CONCLUSIONS ..................................................................................................... 127


10. RECOMMENDATIONS FOR FUTURE WORK .................................................. 131
11. REFERENCES ........................................................................................................ 133
APPENDIX A: CLASSIFICATION OF PHASE FIELDS........................................... 140

iv

LIST OF TABLES
Table 1: RISO data for adiabatic steam/water tests in test section 20............................. 60
Table 2: Turbulent scales for adiabatic steam/water RISO data in test section 20.......... 61
Table 3: Simulation history ............................................................................................. 69
Table 4: Post processing methods used in evaluating the simulation data ...................... 70
Table 5: Experimentally determined mass flow rates through 30 degree section of
tube for RISO run#602. ....................................................................................... 97
Table 6: Averaged field mass flow rates and velocities. Values averaged over the
last 30 msec of the simulation. ............................................................................ 99
Table 7: Averaged field mass flow rates and velocities adjusted to simulate
numerical equivalent of the RISO test liquid film suction. Values averaged
over the last 30 msec of the simulation. ............................................................ 100
Table 8: Averaged field mass flow rates and velocities adjusted to simulate
numerical equivalent of the RISO test liquid film suction. Values averaged
over the last 30 msec of the simulation. ............................................................ 125
Table A1: Definition of Four Field values .................................................................... 142

LIST OF FIGURES
Figure 1: Two-phase annular flow..................................................................................... 4
Figure 2: Transition region of finite thickness 2 between fluids 1 and 2. ..................... 32
Figure 3: Schematic of two-dimensional level set test problem. Semicircle of
R=0.26 is placed in center of domain with a =2.0 prescribed velocity field.
The domain was 0<x<1.0, -0.5<y<0.5. The grid size and level set parameter
values were varied in the tests. ............................................................................ 40
Figure 4: Initial level set and velocity fields for two-dimensional level set test
problem. Semicircle of R=0.26 is placed in center of domain with a =2.0
velocity field. The domain was 0<x<1.0, -0.5<y<0.5. ....................................... 41
Figure 5: Semicircle interface at t=0 (red line) and after one revolution (black line)
for the implicit/explicit solver type using the 50x50 grid; =0.02, t=0.005
(CFL=0.25), =0.01 (CFL=0.50), ls=0.03 (1.5), lsd=0.03 (1.5), c1=0.0..... 41
Figure 6: Semicircle interface at t=0 (red line) and after one revolution (black line)
for the implicit/explicit solver type using the 100x100 grid; =0.01, t=0.01
(CFL=1.0), =0.005 (CFL=0.50), ls=0.015 (1.5), lsd=0.03 (3.0), c1=1.0... 42
Figure 7: Schematic of two-dimensional dam break problem. Water was initially
confined in a square box. The computational domain was 5a x 1.25a., and a
400x100 uniform mesh was used. ....................................................................... 44
Figure 8: Interface and velocity vectors at various times for the 2-D dam break using
a 400x100 grid. .................................................................................................... 45
Figure 9: 2-D dam break: non-dimensional surge front position versus nondimensional time.................................................................................................. 45
Figure 10: 2-D dam break: non-dimensional water column height versus nondimensional time.................................................................................................. 46
Figure 11: Schematic of 2-D traveling solitary wave. Water was started at rest with
a Bousinesq profile, of height Ao, with a hydrostatic pressure balance. The
computational domain of 20h x 2h was discretized with a 200x120 nonuniform mesh. ...................................................................................................... 48

vi

Figure 12: Snap shots of the solitary wave calculation. The solitary wave is
considered generated 0.6 sec after the free fall of the initial profile. It is
designated with the t=0.0 marker. Grid: 200x120, wave speed=1 m/s. ............ 48
Figure 13: Snap shot of the velocity vectors of solitary wave at t=0.4s. The wave is
traveling to the right at a wave speed=1 m/s. ...................................................... 49
Figure 14: Solitary wave amplitude. (Grid: 200x120)..................................................... 49
Figure 15: Schematic of three-dimensional convecting sphere. A red water/green
water problem with surface tension. The computational domain was 0.5m
x 0.2m x 0.2m. ..................................................................................................... 53
Figure 16: Transport of a sphere through the domain using adaptive-mesh approach:
element size = 0.005m near interface and 0.02m elsewhere, mesh size ~
410,000 tetrahedrons. The black line indicates the interface. Mesh
displayed in upper half of planar cuts while the color spectrum in the lower
half indicates pressure (Pa).................................................................................. 54
Figure 17: RISO test section 20 data plotted on the vertical flow regime map of
Hewitt and Roberts (1969) for thin tubes (1-3 cm) which was validated
against low pressure air/water and high pressure steam/water data. ................... 59
Figure 18: Schematic of the simulation domain. Length = 0.025m = 1/8 w for
RISO Run # 602................................................................................................... 63
Figure 19: Boundary layer structure on wall. .................................................................. 64
Figure 20: Initial condition for RISO run #602 using plug flow velocity as initial
guess (ul=2.64 m/s, ug=4.99 m/s) and applying an imposed perturbation with
wavelength p = c............................................................................................... 65
Figure 21: Simulation solution at 63.9113 msec. (a) 3D interface of steam/water
interface colored by velocity magnitude, (b) phase solution at exit plane. ......... 73
Figure 22: Idealized annular flow.................................................................................... 79
Figure 23: 3D contours of steam/water interface at times in simulation. Contours
are colored by velocity magnitude. The dashed ovals indicate the location of
the large wave primarily responsible for droplet entrainment............................. 82
Figure 24: 3D contours of steam/water interface at times in simulation viewed from
exit plane. Contours are colored by velocity magnitude. ................................... 83
vii

Figure 25: Velocity fringe plots of center xy plane at times in simulation. Black line
depicts steam/water interface............................................................................... 84
Figure 26: 3D contours of steam/water interface at times in simulation. Contours
are colored by field (blue=steam bubble, red=liquid droplet, green=steam
core/ liquid film interface). The dashed ovals indicate the location of the
large wave primarily responsible for droplet entrainment................................... 85
Figure 27: 3D contours of steam/water interface at times in simulation viewed from
exit plane. Contours are colored by field (blue=steam bubble, red=liquid
droplet, green=steam core/liquid film interface). ................................................ 86
Figure 28: Simulation void fraction and associated error................................................ 89
Figure 29: Field area fractions at exit plane. ................................................................... 90
Figure 30: Disperse Field area fractions at exit plane: (a) dispersed liquid (DL), (b)
dispersed vapor (DV), (c) average over 30 msec. ............................................... 91
Figure 31: Maximum and minimum void fraction as a function of averaging time.
Oscillations in average void fraction curve are removed for averaging times
of 30 msec or greater. .......................................................................................... 92
Figure 32: Population of disperse fields in 30 degree sector during simulation: (a)
liquid droplets (DL), (b) steam bubbles (DV). .................................................... 93
Figure 33: Mass flow rates of continuous liquid (CL) and continuous vapor (CV)
fields as a function of simulation time................................................................. 97
Figure 34: Mass flow rate of disperse fields as a function of simulation time: (a)
liquid droplets (DL), (b) steam bubbles (DV). .................................................... 98
Figure 35: Averaged mass flow of fields as a function of simulation time: (a)
continuous fields (CL and CV), (b) disperse fields (DV and DL). Averages
performed over 30 msec. ..................................................................................... 99
Figure 36: Total mass flow rate as a function of simulation time. ................................ 100
Figure 37: Field axial velocity at exit plane as function of simulation time. ................ 101
Figure 38: Averaged axial velocity as a function of simulation time: (a) continuous
fields (CL and CV), (b) disperse fields (DV and DL). Averages performed
over 30 msec. ..................................................................................................... 102

viii

Figure 39: (a) Liquid film (CL) mass flow rate at exit plane as function of
simulation time, (b) Instantaneous and averaged liquid film thickness as
function of simulation time. Averages performed over 30 msec...................... 103
Figure 40: Continuous Liquid (CL) circumferentially-averaged area fraction at the
exit plane for four times at the end of the simulation. Each curve represents
an average over 15 msec.................................................................................... 104
Figure 41: Continuous Vapor (CV) circumferentially -averaged area fraction at the
exit plane for four times at the end of the simulation. Each curve represents
an average over 15 msec.................................................................................... 105
Figure 42: Disperse Liquid (DL) circumferentially -averaged area fraction at the exit
plane for four times at the end of the simulation. Each curve represents an
average over 15 msec......................................................................................... 105
Figure 43: Disperse Vapor (DV) circumferentially -averaged area fraction at the exit
plane for four times at the end of the simulation. Each curve represents an
average over 15 msec......................................................................................... 106
Figure 44: Comparison of predicted y(10%), y(50%), and y(90%) positions to the
experimentally measured values........................................................................ 106
Figure 45: Probability Distribution Functions (PDF) for (a) radial location of
droplets and (b) droplet size in terms of equivalent diameter, DEQ. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec). ..... 109
Figure 46: Probability Distribution Functions (PDF) for radial location of droplets in
steam core for 12 ranges of droplet equivalent diameter, DEQ. The total
probability for each sums to 100 percent. Data tallied from time steps
110,000 130,000 (time = 126.08 156.08 msec). .......................................... 110
Figure 47: Probability Distribution Functions (PDF) for droplet size for 12 ranges of
radial locations. The total probability for each sums to 100 percent. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec). ..... 111
Figure 48: Droplet Interfacial Area Density as a function of simulation history.
Simulation value ( Ai =surface area/volume) are plotted with values
assuming a spherical droplet ( Ai =6/DEQ) where DEQ is the droplet
equivalent diameter............................................................................................ 112
ix

Figure 49: Ratio of actual droplet interfacial area density to that assuming a
spherical droplet. Ratio is plotted against the simulation time. ........................ 112
Figure 50: Probability Distribution Functions (PDF) for (a) radial location of
bubbles and (b) bubble size in terms of equivalent diameter, DEQ. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec). ..... 114
Figure 51: Probability Distribution Functions (PDF) for radial location of bubbles in
liquid film for 9 ranges of bubble equivalent diameter, DEQ. The total
probability for each sums to 100 percent. Data tallied from time steps
110,000 130,000 (time = 126.08 156.08 msec). .......................................... 115
Figure 52: Probability Distribution Functions (PDF) for bubble size for 9 ranges of
radial locations. The total probability for each sums to 100 percent. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec). ..... 116
Figure 53: Bubble Interfacial Area Density as a function of simulation history.
Simulation value ( Ai =surface area/volume) are plotted with values
assuming a spherical bubble ( Ai =6/DEQ) where DEQ is the bubble
equivalent diameter............................................................................................ 117
Figure 54: Ratio of actual bubble interfacial area density to that assuming a
spherical bubble. Ratio is plotted against the simulation time. ........................ 117
Figure 55: Computed wall stress compared to experimental shear stress from RISO
run #602. Value from effective viscosity wall function computed from
Equation (138). Values for fully-developed assumption computed from
Equation (151) using data averaged over 30 msec. ........................................... 119
Figure 56: Total shear stress plotted against radial position for continuous liquid
(blue) and continuous vapor (red) fields over the time steps 65,000 95,000
where the axial pressure gradient was -1600 Pa/m. Each are plotted up to
the radial position of the 50 percent volume fraction. The average total shear
stress (green dashed) for steady, fully-developed annular flow is also shown
along with the steam volume fraction (black dotted). ....................................... 122
Figure 57: Total shear stress plotted against radial position for flow over the time
steps 65,000 95,000 where the axial pressure gradient was -1600 Pa/m.
The average total shear stress (green dashed) for steady, fully-developed
x

annular flow is also shown along with the steam volume fraction (black
dotted). ............................................................................................................... 123
Figure 58: Total shear stress plotted against radial position for continuous liquid
(blue) and continuous vapor (red) fields over the time steps 100,000
130,000 where the axial pressure gradient was -3300 Pa/m. Each are plotted
up to the radial position of the 50 percent volume fraction. The average total
shear stress (green dashed) for steady, fully-developed annular flow is also
shown along with the steam volume fraction (black dotted)............................. 124
Figure 59: Total shear stress plotted against radial position for flow over the time
steps 100,000 130,000 where the axial pressure gradient was -3300 Pa/m.
The average total shear stress (green dashed) for steady, fully-developed
annular flow is also shown along with the steam volume fraction (black
dotted). ............................................................................................................... 125
Figure 60: Illustration of how wall shear trends with film thickness @ exit plane: (a)
gradient of axial velocity w.r.t wall normal near the wall, (b) effective film
thickness. ........................................................................................................... 126
Figure 61: Computed wall shear stress compared to the experimental wall shear
stress for RISO run #602. The arrow shows the projected path of the
simulation assuming a linear increase in computed wall shear with
simulation time. ................................................................................................. 130
Figure A1: Illustration of walk-out procedure used to define a phasic glob. ................ 141

xi

ACKNOWLEDGMENT
I am grateful for the guidance and instruction from my advisors, Richard T. Lahey Jr.
and Kenneth E. Jansen. I have benefited from many hours discussing this research with
them and discussing how to proceed as the scope evolved over the years. This work
certainly would not have been possible without their involvement. Even as the time
passed, they remained committed to the research and inspired me to do the same. I
would like to also thank T. Darton Strayer who played a large role in me beginning this
research. He has provided continued technical input and encouragement from the start
and has been a great mentor and friend throughout the process. I also appreciate the
advice from Donald Drew who sat in on the countless meetings with us and provided
valuable insight. I would also like to thank the rest of my committee, Michael Podowski
and Gretar Tryggvason, who have taken time to evaluate my research and offer advice.

I am especially grateful to my wife, Mindy, and children, Ben and Anya, who have
patiently endured my journey. Never have they given anything but their full support and
I am forever thankful.

xii

ABSTRACT
A numerical simulation of a two-phase annular flow was performed using a threedimensional (3-D) stabilized finite element code, PHASTA-IC, with an implemented
level set method to capture the interface between the liquid and gas phases.

The

problem simulated was run #602 of the experimental tests of Wurtz (1978), which was a
70 bar, adiabatic, steam/water annular flow in a 20mm I.D. tube having a total inlet mass
flux of 500 kg/s-m2 and an exit quality of 0.30. The mean experimental film thickness at
the exit of the tube was measured to be 0.94mm. The simulation modeled a 30 degree
segment of the tube of length 0.025m using a uniform tetrahedron mesh of edge length
size 0.00005m. The simulation, although not yet reaching equilibrium annular flow, was
able to capture the major mechanisms associated with annular flow. This includes
generation of instabilities on the interface between the steam core and liquid film,
formation of liquid ligaments that stretch into the steam core and shear off to form liquid
droplets, deposition of droplets back into the liquid film, the carry-under of steam
bubbles into the liquid film, and the development of large roll waves responsible for
most of these mechanisms. A classification tool was developed that interrogates the 3-D
solution and classified all entities in the domain into one of four fields: continuous liquid
(i.e., the liquid film), continuous vapor (i.e., the steam core), dispersed liquid (i.e., liquid
droplets in the steam core), and dispersed vapor (i.e., steam bubbles in the liquid film).
Various quantities, such as the mass flow rate, volume fraction, velocity, and interfacial
area density, were calculated for each field. In addition, methods were developed to
compute the total shear stress distribution, the interfacial shear stress, and the wall shear
stress. Comparisons were made to the experimental flow rates, shear stresses, and film
thickness.

xiii

1. INTRODUCTION
Multiphase flow occurs in many important industrial applications such as phase-change
heat exchangers, oil well pipelines, fossil-fired boilers and nuclear reactors. Many flow
regimes are possible, but the most common one in power production and utilization
technology is annular flow.

This flow regime is quite complex and, as shown

schematically in Figure 1, it is characterized by a droplet laden gas core and a liquid film
on the conduit wall. The dynamics occurring on the wavy interface between the liquid
film and gas core are important to the process of liquid droplet entrainment into the gas
core, the lateral force responsible for keeping the liquid film on the conduit walls, and
the turbulence structures in the liquid and gas phases which effect the pressure drop.
Due to the difficulty in obtaining detailed local data in annular flow experiments,
researchers currently do not have a fundamental understanding of these interfacial
processes. With the advancement in modern computing capabilities, Direct Numerical
Simulation (DNS) can be used to help us understand the physical mechanisms in annular
flow (and other flow regimes) and generate numerical data to support the development
of physically-based closure laws to be used in advanced-generation 3-D, two-fluid
computational multiphase fluid dynamic (CMFD) models [Lahey (2009)].
As a result of the importance of annular two-phase flow, numerous researchers [Hewitt
and Hall-Taylor (1970), Henstock and Hanratty (1976), Tatterson et al. (1977), Kataoka
and Ishii (1983), Whalley (1987), Fore et al. (2000)] have developed phenomenological
models describing the physical processes of annular flow which were based more on
empiricism than detailed physical understanding. In mechanistically based two-fluid
computational multiphase fluid dynamics (CMFD) model formulations [Siebert et al.
(1995), Kumar et al. (2004), Lahey (2005)], which attempt to capture the dynamics of
two-phase flow, one uses closure laws based on the physical transport process at the
interface. Significantly, having accurate closure laws is important for both steady and,
especially, transient calculations.
Given that detailed experimental methods are not available to fully characterize the
wavy interface, one may use DNS data to support fundamental two-fluid model
development. That is, although DNS is not feasible as a practical tool for the design and
analysis of multiphase systems (due to the large computational expenses), Lakehal
1

(2004) and Lahey (2005) have shown how DNS may be used for closure model
development.

This approach is consistent with a roadmap for basic research in

multiphase flow which was assembled by leading experts in the field [Hanratty et al.
(2003)], which supports the use of DNS (and improved experimental methods) to help
understand the microscopic mechanisms that affect the macroscopic behavior of
multiphase flows. The vast majority of DNS calculations done to date for two-phase
flows have focused on bubbly flow [Bunner and Tryggvason (1999), Tryggvason et al.
(2001), Nagrath et al. (2005)]. In contrast, very little has been done for annular flows
due to the computational costs associated with accurately capturing droplet entrainment
and the wave structure at the interface. Stability limits for stratified two-phase flows
have been investigated by Cao et al. (2004), Lakehal et al. (2003) and Fulgosi et al.
(2003), but these were done for relatively low Reynolds numbers such that wave
breaking and entrainment mechanisms were avoided. Nevertheless, some interesting
annular flow simulations have been performed previously. For example, Li and Renardy
(1999) simulated unsteady, axisymmetric oil/water annular flows using a Volume of
Fluid (VOF) method. Their computation was made at Reynolds numbers less than 10
and with a density ratio near unity.
entrainment processes did not occur.

Consequently, wave breaking and droplet


Also, using a level set method, Fukano and

Inatomi (2003) have simulated the formation of the liquid film around the tube wall in
horizontal air/water annular flow.

They specifically investigated the source of the

levitation force that causes the liquid film to spread to the upper surface of the horizontal
tube. Although their simulation was able to capture a physical mechanism which could
explain the observed liquid film mobility, it was performed on a fairly coarse grid which
was incapable of resolving the droplets entrained in the gas core or any bubbles ingested
into the liquid film.
This investigation uses the PHASTA-IC code, with an implemented Level Set method,
to model two-phase annular flow. PHASTA-IC solves the discretized form of the
Incompressible Navier-Stokes (INS) equations in three dimensions using a stabilized
finite element method (FEM). A second-order accurate and stable generalized- time
integrator [Jansen et al. (2000)] was applied to the INS equations and used to march the
solution in time [Whiting and Jansen (2001)]. PHASTA incorporates the level set

method of Sussman [Sussman et al. (1998), Sussman et al. (1999), Sussman and Fatemi
(1999)] and Sethian (1999) to resolve the interfaces in two-phase flows by modeling an
interface as the zeroth level set of a smooth function. The Continuum Surface Tension
(CST) model of Brackbill et al. (1992) was used to represent the local surface tension
force as a volume force applied over an interface of finite thickness.
Section 2 presents the multi-field two-fluid model and closure models relevant to twophase annular flow that may benefit from numerical data obtained using DNS. Section 3
briefly reviews the most common Interface Tracking Methods (ITM) in use today. The
theory and numerical implementation of the FEM and Level Set methods in PHASTAIC are given in Section 4. Section 5 presents canonical test problems used to verify
PHASTA-IC.

An adaptive mesh refinement methodology of two-phase flows is

presented in Section 6 but, since it could not yet be supported by the Blue Gene
computer at RPI, it was not used in the simulation presented herein. Nevertheless, it
should be important for future simulations. The experimental data, to which the twophase annular flow simulation was compared, is described in Section 7. Finally, the
simulation and methods of reducing the data are discussed in Section 8 along with the
most salient results. Conclusions and recommendations for future work are presented in
Section 9 and 10, respectively.

2. HISTORICAL REVIEW OF ANNULAR FLOW MODELING


2.1 Annular Flow Description
In confined two-phase flows the annular flow regime exists when the liquid flow rate is
low and the gas flow rate is sufficiently high. It has been well investigated over the
years [Woodmansee and Hanratty (1969), Hewitt and Hall-Taylor (1970), Shedd and
Newell (2004)] and can occur in horizontal and vertical orientations. It is characterized,
Figure 1, by a liquid film on the wall containing entrained bubbles and a gas core
containing liquid droplets. In horizontal annular flow, the effect of gravity-induced
drainage increases the thickness of the liquid film on the bottom surface while reducing
it on the top surface. This effect diminishes with increasing gas velocity. Two types of
waves have been reported to exist on the interface: disturbance and ripple waves.
Disturbance waves are large amplitude roll waves that are responsible for the
entrainment of liquid droplets into the gas core. The base liquid film under these waves
is generally much higher than under ripple waves. Ripple waves are the low amplitude
surface waves which create interfacial roughness and are, thus, primarily responsible
for the pressure drop.

Although these waves play a large role in the interfacial

dynamics, not enough is known about how they form and to what extent they influence
other aspects like droplet entrainment, gas and liquid turbulence, and interfacial shear.
Over the years this has lead to researchers constructing closure models based more on
empiricism than physics.

High Velocity
Gas Core
Liquid film
Entrained gas
bubbles in film

Liquid droplets
in gas core
Figure 1: Two-phase annular flow

2.2 Multi-Field Two-Fluid Model


For development of mechanistically based two-fluid codes, the multi-field modeling
approach has proven to be effective in modeling the various flow regimes [Lahey
(2009)].

In this formulation, the flow is represented as a mixture of various

interpenetrating fields, for each of which the governing equations are solved. As the
number of fields increases the ability to model the flow increases; however, the burden
of developing the closure models describing the forces for each field and the interactions
with other fields limits the models to the minimum number of fields needed to capture
the dominant physics. Siebert et al. (1995) have shown that a four-field, two-phase
model is capable of capturing the flow regimes from bubbly to annular flow. Lahey and
Drew (2001) have also presented a four-field model which, without varying model
coefficients, can predict a wide variety of adiabatic and diabatic bubbly flows having
flow passages of various shapes. However, their model was not extended for annular
flows. A four-field implementation models the flow as the following idealized fields:
continuous liquid (CL), continuous vapor (CV), dispersed liquid (DL), and dispersed
vapor (DV). For each field the governing equations and turbulence models are solved
making it computationally costly to add additional fields, but the number of fields can be
reduced when modeling specific flow regimes. Trabold and Kumar (2000), and Faghri
and Sunden (2004) present a three-field approach for analyzing annular flow. They
neglected the role of gas droplets (DV) within the liquid film since there was no wall
heating. Their results compared favorably to local void fraction and droplet velocity
measurements for high void fraction (gas>0.75) adiabatic annular flows using
refrigerant R-134a. The bubbly flow model presented by Kumar et al. (2004) models
only the disperse vapor and continuous liquid fields. They showed good agreement for
high pressure flow boiling in thin gap geometries when the void fraction was within the
bubbly flow regime.
Solving the Navier-Stokes equations for each phase requires a Direct Numerical
Simulation (DNS) of the flow to resolve the interfacial dynamics occurring on a
microscopic level.

This may be feasible on a special basis for closure model

development but the vast amount of computational resources it requires precludes it


from being a tool for practical use [Lakehal (2004), Lahey (2005)].
5

Instead, the

governing equations of mass, momentum, and energy are averaged, effectively


averaging out the microscopic phenomena including capturing interfaces.

The

information lost due to averaging is injected back into the solution via closure models,
the forms of which are dependent upon the physical phenomena being modeled. Of the
three averaging methods of space, time, and ensemble, Drew and Passman (1999), the
ensemble average was selected because it retains the spatial and temporal resolution of
the solution. An ensemble average is an average over all possible realization, , over a
suitably large subset of realizations. The governing equations are ensemble averaged by
first multiplying each equation by the phase indicator function, Xk,
1
X k ( x ,t ) =
0

(1)

and averaging over a large set of realizations. The average of the phase indicator
function is the void fraction for phase k, k.

X k (x , t ) = k

(2)

where x is the coordinate vector and t is time. All variables, f, are density-weighted

according to:

k =
fk =
where is the density.

k X k
Xk

fk k X k
Xk

k X k
k

fk k X k

(3)

k k

Drew and Passman (1999) give the multi-field ensemble

averaged equations for two-phase flow.

Continuity:

( jk k ) + ( jk k v jk ) = jk + m jk
t

(4)

Momentum:

( jk k v j ) + ( jk k v jk v jk ) =
t
Re
jk p jk jk k g + jk jk + jk + M jk M wjk

(5)

jk v jk + m jk v jk
Energy:

( jk k h jk ) + ( jk k v jk h jk ) =
t

jk q jk jk q jk + q jk

)+ D

jk

Dp jk
Dt

(6)

+ qijk Ai+ jk jk + m jk jk

where the subscript jk represents field j of phase k, is the void fraction, is the
density, is the interfacial mass transfer rate due to phase change, m is the mass source
density of fluid from other fields, p is pressure, is the stress tensor, Re is the
Reynolds stress tensor, M and M w represent the interfacial and wall forces per unit
volume, respectively, q is the volumetric heat generation rate, qi is the interfacial heat
T
flux, q is the wall heat flux, q is the so-called turbulent heat flux, Ai is the interfacial

density, D is the dissipation function, and is the internal energy.


The underlined terms in the governing equations represent the interfacial transfer terms
that require closure models. These represent phenomena like interfacial exchange of
momentum through droplet entrainment and deposition, interfacial shear, and phase
change.

2.3 Overview of Closure Models


When the conservation equations are ensemble-averaged, the microscopic physics
associated with the interfaces and wall is lost. This includes such dynamic processes as
droplet entrainment into the vapor core, interfacial and wall shear, effect of the wavy
interface on the liquid film turbulence, and the droplet size distribution. Closure models
describing the relevant physics are typically used to recover the information lost during
averaging. Ideally, a closure model will be mechanistically based and accurately model

the physical phenomena. Due to the complexity of annular flow and the difficulty of
obtaining local information in the flow, most closure models to date are instead
correlations (using non-dimensional numbers) which are deemed to be relevant to the
physics of interest. While correlations are often numerically stable models, they are
suspect when used outside the range of data against which they were correlated.
Furthermore, most are correlated against equilibrium conditions and thus have limited
use in transient calculations.
Some of the closure models for annular flow are discussed below. They have been used
in a variety of computer codes to successfully predict data and so have been an asset to
date. However, their use is restricted by the data range over which they were correlated
and by the assumptions made in developing the models. Future fundamental codes will
require more mechanistically based models that adequately model the physics. It is with
these models that extrapolations beyond experimental data may be made.

The

development of these improved models will require knowledge of local physics that may
be obtained through many simulations of annular flow conditions. This thesis discusses
the ability to perform one of these simulations but does not address formulation of the
improved closure models. However, for completeness, the forms of popular closure
models for annular flow are discussed below.

2.3.1 Interfacial Shear


The wavy interface on the liquid film creates a shear on the high-speed gas core more so
than had the gas core been traveling along a smooth wall. Recognizing this, Wallis
(1969) likened the wavy interface to a roughened wall and modified the single phase
friction factor, fsp, to compute the interfacial friction factor, fi.

He used the non-

dimensional length scale of liquid film height, , to tube diameter, D, to obtain:

f i = f sp 1 + 300 .
D

(7)

Noting the gas velocity to be more significant than the liquid film velocity, the
interfacial shear, i, is computed using the single-phase wall friction model.

i =

1
2
f i gU g
2

(8)

Although many authors have modified Wallis model over the years (e.g., to use relative
velocity rather than Ug), the basis form of Equation (7) has constituted the leading
models in use today. Henstock and Hanratty (1976) noted the difficulty of knowing the
film thickness height, , a priori, and developed a correlation for the film height based
on the gas core Reynolds number, ReG, the liquid film Reynolds number, ReLF, and the
ratio of the liquid to gas densities and viscosities. The interfacial friction factor was
given by:

f i = f sp (1 + 1400 F ) Veritical Flow

f i = f sp (1 + 850 F )

Horizontal Flow

(9)

F = Function(Re LF , Re G , L , L
G
G )

Noting that the tube diameter does not have as much influence on the friction factor as
given in previous models, Asali et al. (1985) used a friction length scale to scale the
single phase friction factor according to /lf instead of /D, where lf is given by:
lf =

g
i

0.5

(10)

Ambrosini et al. (1991) refit Asalis correlation to account for a wide range of air/liquid
flows, flow rates, and pipe diameters. His modified interfacial friction factor is given
by:

0.2
0.6
+
f i = f sp 1 + 13.8WeD Re g mG 200 g .
l

(11)

WeD, ReG, and mG+ are the Weber number based on tube diameter, gas Reynolds
number, and non-dimensional film thickness given by:

gU g 2 D
WeD =

Re G =

UgD

mG =

mU g*

(12)

where Ug* is the friction velocity which is equal to (i/g)0.5. Additional modifications
have been made by Brauner and Maron (1993) who added a memory effect to improve
calculations for transition between smooth and wavy stratified flow , Fukano and
Furukawa (1998) who tested the effects of varying viscosity on interfacial shear and

added a viscosity correction term, and Fore et al. (2000) who shifted the correlation to fit
an expanded database.
Ishii and Mishima (1984) departed from the linear relationship between the interfacial
and single phase friction factors by introducing a power law relationship between the
wall shear, w, and interfacial shear:
R
i w
Rw

(13)

where R and Rw are the local radius and tube radius, respectively. Similarly, Dobran
(1983) chose to use a power wall relationship to compute the effective momentum
diffusivity, eff, in the wavy region:

eff

n
= 1 + C1 + t+ ,

(14)

where l is the momentum diffusivity for liquid only, + is the average film thickness, t+
is the thickness of the liquid film to the wave troughs (.i.e. thickness of the base film),
and C1 and n are constants determined from the data. The underlying assumption is that
the base film can be modeled similar to single phase flow but the wavy top of the liquid
film should be a function of the wavy height and not the distance to the wall. The
different length scales will lead to a modified diffusivity. Dobran derived an expression
for t+ by correlating it to air/water data. The law-of-the-interface approach by Kumar
and Edwards (1996) attempted to reduce the amount of correlation involved with
modeling the interfacial shear by using an analogous law-of-the-wall relation to directly
compute shear from the turbulent kinetic energy. Noting the single-phase shear, , is
related to the turbulent kinetic energy, k, by:

= C k ,

(15)

Kumar and Edwards (1996) assumed a similar relationship for the interfacial shear:

i = Tm (u l u i )
Tm =

g C 0.25 k 0.5 ,

(16)

u+

where the non-dimensional velocity, u+, is given by the log relationship similar to the
law-of-the-wall and the terms have been normalized against the interfacial velocity,

10

shear, and film thickness. ul and ui are the liquid and interfacial velocities. They were
unable to show agreement with data, citing issues with the droplet entrainment model
they used in their model.
2.3.2 Droplet Entrainment

Most researchers acknowledge liquid droplets are formed in the gas core via one of three
mechanisms; breakup of liquid bridges carried over from the transition to annular flow,
disintegration of larger droplets into smaller droplets, and entrainment from the liquid
film. The high speed photographic study of Woodmansee and Hanratty (1969) of
parallel air flow over water found the latter mechanism to be dominant due to a KelvinHelmholtz instability forming on the back of the large disturbance waves.

Small

wavelets were found to form on top of the disturbance waves, accelerate behind the
wave due to the lower pressure, be lifted by the high-speed gas core, and then shatter,
entraining droplets into the core. Considering the key forces acting on the wavelet along
the interface to be the drag and surface tension, Woodmansee and Hanratty selected the
ratio of these forces, the Weber number, to be the critical parameter governing the onset
of entrainment:

l c g (U g C )

We =

(17)

where lc is a characteristic length of the wavelet, Ug the velocity of the gas core, C the
waves celerity, g the gas density, and the surface tension. Their experiment showed
the critical air velocity for the onset of entrainment to increase with surface tension and
decrease with decreasing film thickness. Viscosity was not found to be of primary
importance.
Although some authors have done simplified force balancing and instability analyses to
generate templates for the droplet entrainment model, most are still correlated against
various data sets using various non-dimensional numbers, including We. Based on the
mechanistic model that entrainment occurs by the shearing of wavelets on disturbance
waves, Kataoka and Ishii (1983) and Ishii and Mishima (1984) developed models for
entrainment for three zones of the entrainment region of air/water data. The first zone is
the geometry dependent region in which entrainment can be very sensitive to the inlet

11

condition of the tube but is usually very small. They found this region to exist for z <
160DWe0.25Ref-0.5, where z is the axial coordinate along the tube and D is the tube
diameter. We and Ref are the Weber and liquid Reynolds numbers, respectively, and are
defined by:

g j g 2 D
jD
and Re f = l l ,

We =

l
g

(18)

where jg and jl are the gas and liquid superficial velocities and l is the liquid viscosity.
The last region is the equilibrium region where the entrainment and deposition rates are
equal and begins at:

z = 440D We

0.25

Re f

0.5

(19)

The entrainment fraction, E, defined as the ratio of the droplet volumetric flux, jfe, to the
total liquid volumetric flux, jl, is given by:

j fe
0.25
= E = tanh 7.25 x107 We1.25 Re f
.
jl

(20)

The entrainment rate,  , defined as the mass of the entrained droplets per unit time per
unit interfacial area, is given by:

 D
0.925 g
= 6.6 x10 7 (Re f We )
l
l

0.26

(21)

The region between the equilibrium and geometry dependent regions is the developing
entrainment region where the entrainment rate was correlated to be:
2

 D
E
1.75
0.25
+
= 0.72 x10 9 Re f We(1 E ) 1
l
E
0.925 g
6.6 x10 7 (Re f We )
l

0.26

(22)

(1 E )0.185

Dallman et al. (1984) correlated air/water data with different tube diameters to include
the effect of tube diameter on entrainment. They gave the following correlation for
entrainment fraction in the equilibrium region:

12

W
E = 1 LFC
WL

8
3.6 x10 (D 2m )( g l ) U g

1 + 3.6 x10 8 (D 2m )( g l )0.25 U g 3

3 1 .5

0.25

1 .5

(23)

where WLFC is the critical liquid film flow rate below which entrainment does not occur,
WL is the liquid flow rate, and m is the average film thickness. Lopez de Bertodano et
al. (1995) performed separate effects tests using air/water and Freon as working fluids to
see the effects of surface tension and density ratio. Their correlation for entrainment rate
was a merger of those of Dallman et al. (1984) and Kataoka and Ishii (1983) and is given
by:

 D k A
(Re LF Re LFC )We g l
=
l
4
g

0 .5

(24)

where kA is a dimensionless atomization coefficient, ReLF is the liquid film Reynolds


number, and ReLFC is the critical film Reynolds number below which entrainment does
not occur. Holowach et al. (2002) developed a model, derived from stability analysis
and simplified force balances, which they felt would be better suited for transient
calculations. Their model for entrainment rate, using the symbol SE rather than  , is
given by:
S E = 0.0311 Re film

1.67

7/8

Ventr , w l (U gc U f

3 (1 l )0.5

(25)

Refilm is the film Reynolds number, N is the viscosity number, is the wavelength of
the interfacial wave, Ventr,w is the volume of liquid entrainment from the wave crest,

U gc is the average core velocity, and U f is the average film velocity. Although this
model is mechanistic in origin, Holowach et al. (2002) used correlations for the droplet
drag coefficient and interfacial friction factor in the derivation.
2.3.3 Droplet Size

The majority of the droplet size models are based on the assumption that the droplet size
is set during entrainment via a Kelvin-Helmholtz instability as seen by Woodmansee and
Hanratty (1969). Several variations of the model by Tatterson et al. (1977) have been
generated to correlate with an expanding experimental database.

A smaller set of

authors, like Kocamustafaogullari et al. (1994), assume the droplet size may exceed the
13

maximum entrained droplet diameter during entrainment and that the final size is set by
droplet disintegration mechanisms in the gas core.
Tatterson et al. (1977) did a force balance about a liquid ligament prior to entrainment in
the core and used Kelvin-Helmholtz theory for inviscid flow over a small amplitude
wave to determine the pressure gradient as a function of wave number, k. Using friction
velocity, Ug* = (i/g)0.5 as the characteristic velocity scale and assuming k ~ 1/m for
thin films, where m is the film height, the droplet diameter, d, was given by:
*
d g U g m
=
m

1/ 2

(26)

They found m and i using the correlations from Henstock and Hanratty (1976). Note
there is no dependence on liquid film parameters. As suggested by data, the droplet
distribution could be modeled using an upper limit log normal distribution of the form:
2

d max
d (v )
d
2
ln (a ) ,
exp ln
=

d (d )
d
d

d (d max d )
max

(27)

where v is the volume fraction of droplets having diameters less than d, dmax is the
maximum drop diameter, is a size distribution parameter, and a is a dimensionless
parameter defined by:

a=

d vm
,
d max d vm

(28)

where dvm is the volume droplets median diameter. Droplets with d > dvm occupy 50%
of the volume. Evaluations of current air/water data showed and dmax/dvm are weakly
dependent upon the flow parameters and may be assumed to be constant. Tatterson et al.
assumed the average value from the data: =0.72, dmax/dvm=2.90. They computed the
mean volume diameter as:

g U g2 f sp D
d vm

= 1.60 x10 2

D
2

1 / 2

(29)

Ambrosini et al. (1991) took the same approach at Tatterson et al. when developing a
correlation for a wider range of data that included varying liquids, gases, and pipe
diameters, but used their own correlations for m and i. They also included terms that

14

provided a density ratio correction, accounted for droplet collisions in the gas core, and
corrected the correlation for low core velocity data. Their correlation is given by:

d 32

= 22.0
f U 2m
m
g i g

0.5

0.83

G LE D
99.0
+
exp0.60
.
g U g d 32 WeD

(30)

Note their correlation is for the Sauter mean diameter, d32, for which droplets having
diameters above d32 account for 68% of the droplet volume. GLE is the mass flux of the
entrained liquid and WeD is the Weber number defined in Equation (12).
Other relevant correlations are provided by Ueda (1979) and Kataoka et al. (1983).
Ueda assumed a gamma distribution for the droplet density function and, comparing
against data from air/liquid systems using water, glycerol, and aqueous alcohol
solutions, they developed the following correlation for the maximum droplet diameter:

d max


= 5.8 x10 3
gU g

1.25

0.34

(31)

Kataoka and Ishii (1983) based their correlation solely on air/water data and gave the
following expression for the mean volume diameter, dvm:

d vm = 0.028

g jg 2

Re f

1 / 6

Re g

2/3

1 / 3

2/3

(32)

Ref, Reg, and jg are the film Reynolds number, gas Reynolds number, and gas superficial
velocity, respectively. They found dmax = 3.13dvm which is similar to that of Tatterson et
al. (1977).
A different approach was taken by Kocamustafaogullari et al. (1994). They based their
model on the maximum droplet diameter being set by droplet disintegration in the gas
core and not by the entrainment process. A critical Weber number could be constructed
when doing a force balance between the surface tension force acting to maintain the
droplet and the disrupting shear placed on the droplet by the core flow. Two disrupting
forces were considered and a correlation for dmax derived from each. First, for flows
where the density of the disperse phase, d, is less than the continuous phase density, c,
(i.e. bubbles in liquid medium), the disrupting force in primarily due to turbulent
fluctuations in the liquid film acting on the bubble interface. However, when d > c

15

(i.e., droplets in the gas core) the disrupting force results from the local relative motion
about the droplet interface. For this case, dmax is given by:
d max = 2.609C w
*

4 / 15

Wem

3 / 5

Re g 4

Re f

1 / 15

4 / 15

(33)

where dmax* = dmax/dh is the non-dimensional diameter, dh is the hydraulic diameter of


the flow channel, Wem is the Weber number based on the hydraulic diameter and gas
superficial velocity, and Cw is a coefficient defined by:

0.25

Cw = 1
4 / 15
N

35.34

N > 1
15
N < 1
15

.
N =

f
0.5

(34)

0.5

Kocamustafaogullari et al. (1994) found dmax/dvm=2.93 and dmax/d32=4.01. A comparison


of the volume mean diameter models from Tatterson et al. (1977), Kataoka et al. (1983),
and Kocamustafaogullari et al. (1994) was done by Fore et al. (2002) against high
pressure nitrogen/water droplet data at system pressures on 3.4 and 17 bar. Only the
correlation from Kocamustafaogullari et al. (1994) reasonably predicted dvm for the
nitrogen data; the other correlations under-predicted the data. However, the coefficient
had to be increased to predict the correct Sauter mean diameters.
2.3.4 Liquid Film Height

The height of the liquid film, , has been primarily correlated against the liquid film
Reynolds number, ReLF, with some authors including other terms to account for density
ratio, viscosity ratio, and buoyancy effects. One of the earliest models is by Kosky
(1971) who performed a force balance of the liquid film and, assuming the velocity
profile in the liquid film to be that for single phase turbulent flow, computed the average
film height. Since the turbulent profile used depended on the height of the film, Kosky
derived the following two film thickness equations:

16

+ = 2 Re LF 1 / 2

for + < 25 (low Re LF )

(35)

+ = 0.0504 Re LF 7 / 8 for + > 25 (high Re LF )


ReLF is the liquid film Reynolds number defined as:
Re LF =

lU film Dh _ film
,
l

(36)

where Ufilm and Dh_film are the velocity and hydraulic diameter of the liquid film. + is
the non-dimensional film thickness given by:

+ =

lu *
,
l

(37)

where u* is the frictional velocity. Asali et al. (1985) modified the low flow model of
Kosky to fit their low ReLF data (~20-300) for vertical upward flow using air/water and
air/aqueous glycerin mediums. They proposed the fit

+ = 0.34 Re LF 0.6 .

(38)

Ambrosini et al. (1991) later re-correlated the interfacial friction factor and entrained
droplet size for a wide range of data with varying pipe diameters and working fluids.
They found the data was best fitted using the mean film thickness correlation of Asali et
al. (1985) for ReLF<1000 and the high ReLF model of Kosky (1971) for ReLF>1000.
Other researchers correlated data against additional terms to include effects other than
film Reynolds number. Henstock and Hanratty (1976) correlated the film thickness
against a limited set of horizontal and vertical air-water data. They found the vertical
data (ReLF ~ 10-10000) was best fit using:

6.59F

(1 + 1400F )0.5

(39)

while the horizontal data (ReLF ~ 1000-10000) was best fit using:

6.59F

(1 + 850F)0.5

(40)

F is a flow parameter defined by:

[(0.707 Re
F=

) + (0.0379 Re

LF

Re g

0.90

17

4.0
0.9 2.5

0.5 2.5

LF

l
g

l
g

(41)

and Reg is the Reynolds number based on gas velocity and tube diameter. Although they
found Koskys high ReLF model predicted their data, Fukano and Furukawa (1998)
correlated their air/water and air/glycerol solution data to include the effect of the
buoyancy using the Froude number, Fr = U/(g0.5 D). They correlated the mean film
thickness as:

= 0.0594 exp 0.34 FrGO

0.25

Re LO

0.19

x ,

(42)

where FrGO, ReLO, and x are the Froude number for the gas core, liquid Reynolds
number, and the flow quality defined as:

FrGO =

jg
gD

, Re LO =

jg g
l jl D
, x=
( j g g + jl l )
l

(43)

and jl and jg are the liquid and gas superficial velocities, respectively.
2.3.5 Lateral or Levitation Force

Perhaps the biggest unknown associated with annular is the mechanism by which the
liquid film forms on the walls of the conduit, particularly on the upper surface in
horizontal annular flow. The four predominant theories that have been presented are:

Droplet entrainment and deposition

Secondary gas flow

Wave spreading

Pumping action of disturbance waves

Each theory may be supported by experimental, and sometimes numerical, evidence.


Neglecting the secondary gas flow theory, the other theories agree the disturbance waves
play a leading roll in maintaining the liquid film, whether it be by generating liquid
droplets in the vapor core or causing the film at the bottom to be spread or pumped
upwards. Instantaneous measurements of film velocity and direction are required before
the physical mechanism may be understood.

Given that the current experimental

techniques provide averaged (time and space) data, perhaps the best method is the use of
DNS. For completeness, each theory is discussed below.

18

Droplet Entrainment and Deposition:


Russell and Lamb (1965) hypothesized that for horizontal annular flow the liquid film on
the upper wall drains down the horizontal tubes wall but is continuously replenished by
impacting liquid droplets from the vapor core. Their experiments measured the salt
concentration at circumferential locations around the tube.

Based on the salinity

measurements and the location of the salt injection point, the direction and velocity of
the circumferential flow could be inferred. They found the flow at the top surface to
flow downward on a time-averaged sense; however, there were fluctuations in which the
flow could flow upwards. They reasoned the bulk drainage of the liquid would drain the
liquid film unless droplets replenished the liquid supply. An effective circumferential
diffusivity was used to model the net exchange of liquid though drainage and droplet
supply.
Secondary Gas Flow:
The idea that secondary flows in the vapor core were responsible for pulling liquid up
the wall was proposed by Laurinat et al. (1985) and supported by Lin et al. (1985).
Secondary flows may be generated as a result of either flow traveling through a noncircular cross section generated by a circumferentially varying film thickness or
circumferentially varying interfacial roughness. These secondary flows then shear the
liquid film and act to pull the liquid up the wall. Flores et al. (1995) used vorticity
meters to measure the presence of secondary flows in air-water horizontal annular flows.
At the low air flow rates used, the annular flow was near its inception point, minimizing
the importance of disturbance waves. The circumferential wall shear, s, on the liquid
was proposed and based on the gas velocity, Vg, gas density, g, tube diameter, D, and
surface roughness, :

s = 6.7 E 04 D g 1.5V g 2 [1 + 75(1 )]log[3.33E 4 D(1 )] .

(44)

No conclusions could be made of the importance of secondary flow at high flow rates
where disturbance waves play a larger role. Jayanti et al. (1990) evaluated the timedependent behavior of air/water annular flow based on measurements of circumferential
velocity and frequency. They found the film thickness on the upper surface correlated
with the frequency and strength of the disturbance waves. They also reasoned the
secondary flow to be too weak to produce the film thickness measured.
19

Wave Spreading:
Butterworth and Pulling (1972) proposed that the orientation and velocity of the
disturbance waves at the bottom of the tube result in a net flow of liquid toward the wall.
This results in the fluid spreading up the wall. Gravity waves will travel faster where the
film thickness is thicker the tube bottom. This variation in velocity results in the
disturbance waves having velocity components directed up the tube wall. This effect is
thought to be stronger as the gas velocity increases and disturbance waves become
stronger, in which case the film thickness becomes more uniform around the tube. This
was supported though the experimental work of Jayanti et al. (1990) who found the film
thickness correlated with the strength of the disturbance waves.
Pumping Action of Disturbance Waves:
Fukano and Ousaka (1989) present a model that attributes the levitation force to the
disturbance waves, but not based on the wave spreading theory of Butterworth and
Pulling (1972) but rather based on a pumping action setup by a pressure gradient
induced by the disturbance waves. The height and velocity of the pressure waves is
greatest at the bottom of the tube. Stagnation regions in the gas behind the disturbance
waves are, therefore, greater at the bottom than the top. Considering the film to be
sufficiently thin, and considering the pressure in the film to be that of the gas phase, a
negative pressure gradient is generated in the disturbance waves from the bottom to the
top. This pressure gradient is responsible for pumping the liquid to the top of the tube
when disturbance waves pass. The non-dimensional pressure in the disturbance wave
due to the gas stagnation was given by :

xx

g (u g1 C DW )2 (u g 2 C DW )2
= C1
= C1
,
s
4 s
p

(45)

where g is the gas velocity, ug1 and ug2 are the gas velocities far upstream and just
behind (stagnation region) the disturbance wave, CDW is the velocity of the disturbance
wave, s is the wall shear stress for single phase gas flow, and C1 is a constant. The
experimental investigation of Sutharshan et al. (1995) supports this model and the wave
spreading theory when they optically measured the film velocity in horizontal
air/kerosene annular flows using a dye tracing method. They found the liquid film at the
top surface to be continuously draining down the tube wall except when disturbance
20

waves passed and the liquid moved upward. Secondary flow was deemed not be a major
cause for the film development because had it been a leading factor the liquid film would
have moved upwards the majority of the time. The pumping action theory was further
supported by the DNS results for horizontal annular flow by Fukano and Inatomi (2003)
who was able to simulate the development of disturbance waves and resulting film
development on the upper surface of the tube. They showed the film development was a
result of the pumping action set up by the pressure gradient in the disturbance wave.
Unfortunately, the simulation was performed on a grid too coarse to capture high
frequency instabilities and adequately capture droplet entrainment, thereby preventing
droplet entrainment from establishing the film. Also, they noted the film velocity was
highest near the wall instead of near the interface, which is contrary to the secondary
flow theory.

21

3. OVERVIEW OF INTERFACE TRACKING METHODS


The open literature has an abundance of Interface Tracking Methods (ITM) that track or
capture interfaces for multiphase problems. While each method has its own advantages,
the most prominent methods used are the Volume of Fluid (VOF), Front Tracking (FT),
and Level Set methods. Solving the Navier-Stokes equations for multiple phases has
proven to be very challenging so most computations deal with single or multiple bubble
dynamics and free surface flows. Very little has been done to characterize the interfacial
dynamics of annular flow liquid films in which wave breaking and droplet entrainment
occurs. Categorizing the various ITM is a tedious task but, fortunately, there are good
reviews provided by Lakehal et al. (2002) and Tryggvason et al. (2001). Using the
categorization of Tryggvason et al. most ITMs may be considered to be one of four
types: Front Capturing, Front Tracking, Boundary-Fitted grids, or Lagrangian.
The front capturing schemes solve the flow on a stationary grid and resolve the interface
using a marker function. The most popular front capturing methods are the Marker and
Cell (MAC), Volume of Fluid (VOF), and Level Set (LS) methods. The MAC method
places mass-less particles at the interface and convects them with the flow. Their
position is used to infer the interface location and topology. Perhaps more popular, the
VOF method [Hirt and Nichols (1981) and Rider and Kothe (1998)] advects a marker
function which is the volume fraction of liquid in the cell volume. The function has a
value of 1.0 when the cell volume contains all of phase A, a value of 0.0 when it
contains phase B, and has a steep gradient near the interface where it varies from 0.0 to
1.0. The interface is defined where the function has a value of 0.5. Because the
interface needs to be inferred from the discontinuous marker functions in the MAC and
VOF methods, accurate interface reconstruction is the liability of these methods,
especially with sharp interfaces. Initially, the VOF method used the algebraic Simple
Line Interface Construction (SLIC) technique in which a line was used to mark the
interface in each cell.

Significant improvement has been made recently using the

Piecewise Linear Interface Construction (PLIC) method of Rider and Kothe (1998). The
major advantage of VOF is that it conserves mass, something that is more difficult to
accomplish with the Level Set method. For a more complete review of VOF methods to
date, the reader is referred to the work of Scardovelli and Zaleski (1999).
22

The Level Set method was first introduced by Osher and Sethian (1988) with
improvements by Sussman et al. (1998) and Sussman and Fatemi (1999).

Nice

overviews of the methods are provided by Osher and Fedkiw (2001) and Sethian and
Smereka (2003). In contrast to VOF, the LS method uses a smooth and continuous
marker function, , termed the level set function, to advect the interface. The level set
function is the signed distance from the nearest interface with the interface defined as the
zero level set (=0). The interface is advected through the domain using a simple
advection equation and, since the level set function is spatially differentiable
everywhere, the method can easily handle merging interfaces and can compute interface
curvature. The ability to handle sharp, merging, and fragmenting interfaces is the LS
methods greatest characteristic. To handle the jump in physical properties across the
interface the interface is modeled with a finite thickness over which the fluid properties
smoothly vary. Accordingly, interfacial forces like surface tension are applied using a
Continuous Surface Force model like that proposed by Brackbill et al. (1992). The
smeared interface does give rise to numerical errors that corrupt the distance field which
lead Sussman and Fatemi (1999) to introduce an extra redistancing algorithm to help
preserve the level set field. Although improved, the LS method is tainted with not being
able to definitively conserve mass.

Although mass correction methods [Lakehal et al.

(2002)] have been proposed for the standard LS method, variants on the LS method are
the recent trend. One variant couples the LS and VOF methods [Sussman and Puckett
(2000), Sussman (2003), Son (2005), and van der Pijl et al. (2005)] to merge the
interface resolving ability of LS with the mass conservation of VOF. The improved
mass conservation is achieved with a more complicated numerical method. Another
approach is the particle level set method of Enright et al. (2002) that proposes advecting
mass-less particles with the interface to use to check and correct the zero level set.
Front tracking methods are characterized by solving the flow on a stationary grid while
tracking the interface with a lower order grid. One of the more popular FT methods in
use is described in Unverdi and Tryggvason (1992) and Tryggvason et al. (2001). The
interface grid is comprised of points with elements formed between them. As the
interface deforms and stretches in the domain, the interface grid is restructured to ensure
adequate resolution.

To maintain adequate resolution of the interface, points are


23

removed in regions where they have become crowded and added in regions where they
are spaced too far apart. This restructuring of the front can be complicated for arbitrary
topologies in 3D computations. While physical discontinuities in properties and forces
do exist at the interface, these must be transferred to the fixed grid. This is usually done
via a smoothing operation which effectively transforms the interface forces in to volume
forces to be applied to the fixed grid. This is similar to the way the LS method uses the
CSF to apply surface tension over a finite thick interface.
Boundary fitted methods [Ryskin and Leal (1984) and Fulgosi et al. (2003)]

are

probably the most rigorous methods but their application is restricted to simple
interfaces. The Navier-Stokes equations are first solved for each fluid separately over
their respective subdomain. They are then coupled at the interface via jump conditions
for continuity and momentum with the interface motion computed by solving an
advection equation for the interface height. This method cannot be extended to flows
with strong topological changes that may produce entrainment. For this reason it is not
applicable for modeling annular flow.
With Lagrangian methods [Hu et al. (2001) and Johnson and Tezduyar (1997)] the grid
follows the interface. In explicit Lagrangian schemes, the flow field is first solved using
the particle velocities and used to compute the forces acting on the particles. The
particles velocities are then computed and used to update the particle positions. The
mesh movement scheme used is not trivial with additional complexity added when
modeling colliding particles. The ability to properly follow breaking waves is suspect
and so Lagrangian methods are not well suited for annular flow simulations.

24

4. DESCRIPTION OF PHASTA-IC
PHASTA-IC solves the Incompressible Navier-Stokes (INS) equations in three
dimensions using a stabilized finite element method. The second-order accurate and
stable generalized- time integrator of Jansen et al. (2000) as applied to the INS
equations [Whiting (1999)] is used to march the solution in time.

PHASTA-IC

incorporates the Level Set method [Sussman et al. (1998), Sussman et al. (1999),
Sussman and Fatemi (1999), and Sethian (1999)] to resolve interfaces in two-phase
flows by modeling the interface as the zero level set of a smooth function.

The

Continuum Surface Tension (CST) model of Brackbill et al. (1992) is used to represent
the local surface tension force as a volume force applied over an interface of a finite
thickness.

4.1 Finite Element Discretization of the Incompressible Navier-Stokes


Equations
4.1.1 Governing Equations

The spatial and temporal discretization of the INS equations within PHASTA has been
described in Whiting (1999) and Nagrath (2004). This section follows their discussion.
The strong form of the INS equations is given by:
continuity:

u i ,i = 0

(46)

momentum:

u i ,t + u j u i , j = p , i +ij , j + fi

(47)

where is density which is constant, ui is ith component of velocity, p is pressure, ij is


the stress tensor, and fi represents body forces along the ith coordinate.

For the

incompressible flow of a Newtonian fluid, the stress tensor is related to the strain rate
tensor, Sij, as:

ij = 2 S ij = (u i , j + u j ,i ).

(48)

Using the Continuum Surface Tension (CST) model of Brackbill et al. (1992), the
surface tension force is computed as a local interfacial force density, which is included
in fi .

25

4.1.2 Finite Element Formulation

The INS equations apply over a spatial domain defined by R N where N equals
three for our case. This domain is composed of the interior, , and the boundary, .
Note the boundary is further subdivided into regions with natural boundary conditions,
h, and regions with essential boundary conditions, g, such that = h g . The
finite element formulation is based on the so-called weak form of Equations (46) and
(47). To derive the weak form, continuous function spaces of trial solutions, S, and
weighting functions, W, are defined with square-integrable values and derivatives (H-1
functions) on . The domain is discretized into nel finite elements defining the element
domain, e . Consequently, finite-dimensional approximations, or subspaces, of S and

W, specified as Sh and Wh, respectively, are defined as:

}
= {w w (, t ) H ( ) , t [0, T ], w
P ( ) , w (, t ) = 0 on }
P = {p p (, t ) H ( ) , t [0, T ], p
P ( ) }

S kh = u u ( , t ) H 1 ( ) , t [ 0, T ] , u x Pk ( ) , u ( , t ) = g on g
Whk

x e

k
h

x e

(49)
(50)
(51)

where h represents the characteristic length of the parameterized domain, g is an


approximation of the prescribed boundary condition in the discretized domain, and
Pk( e ) is the piecewise polynomial space defined on the element e for order k 1.
Note the velocity and pressure variables use the same polynomial space, which is
enabled through the stabilized formulation used in PHASTA-IC.
The so-called weak form is generated by dotting Equations (46) and (47) on the left by
the weight functions defined above and integrating over the domain. Following the
stabilized formulation of Taylor et al. (1998) the residuals of the continuity and
momentum equations are added together. The Galerkin finite element formulation is
given by finding u S kh and p Phk , such that:

26

BG ( wi , q; u i , p ) = 0

BG ( wi , q; u i , p ) = q ,i u i + wi {u i ,t +u j u i , j f i }+ wi , j { p ij + ij } d +

qu n
i

(52)

wi ( ij p ij )n j d

for all w Whk and q Phk .


The pressure, viscous stress, and continuity terms are integrated by parts eliminating all
second order derivatives. The boundary integral term in Equation (52) is a by-product of
this and only exists on h. Stabilization terms are added to the standard Galerkin method
to correct known stabilization problems with advection dominated problems.
resulting

stabilized

formulation

used

in

PHASTA-IC

is

given

The
by

finding u S kh and p Phk , such that the residual, B, becomes:


B( wi , q; u i , p ) = BG ( wi , q; u i , p)+ < stabilization terms > 0

(53)

where,

B ( wi , q; u i , p ) = q ,i u i + wi {u i ,t +u j u i , j f i }+ wi , j { p ij + ij } d +

qu n
i

wi ( ij p ij )n j d

[ {u
nel

e =1 e

w u

nel

e =1 e

wi , j + q, i }Li + C wi ,i u i ,i d e +

(54)

u i , j + u j wi , j u k u i ,k d e

for all w Whk and q Phk , where Li represents the residual of the strong form of the ith
momentum equation, Equation (47);
Li =u i ,t +u j u i , j + p, i ij , j f i

(55)

The third line of Equation (54) is the standard Streamwise Upwind Petrov-Galerkin
(SUPG) stabilization terms for incompressible flow [Franca and Frey (1992)]. The last
line are terms added to achieve global conservatism where

u i = M Li

(56)

and u i is the incremental velocity required to achieve conservation. The first term in the
last line was added by Taylor et al. (1998) to overcome the lack of mass conservation
27

introduced as a consequence of the momentum stabilization in the continuity equation.


The second term on this line was introduced to stabilize this new advective term. The
stabilization parameters,

described in detail by Whiting and Jansen (2001), for

continuity, C, and momentum, M are given by:

M =

C
c1
+ c 2 u i g ij u j + c3 2 g ij g ij
t 2

C =

and

1
.
M trace (g ij ) )

(57)

(58)

The stabilization term for the conservation terms, ~ , is given by:

~ =

(59)

c 2 u i g ij u j

The coefficients C, c1, c2, and c3 are defined based on the one-dimensional linear
advection-diffusion equation using a linear finite element basis. gij is the covariant
metric tensor related to mapping for global to parametric coordinates and is defined as gij
= k,ik,j where is the parametric coordinate.
The weight functions and solution variables in Equation (54) are expanded in terms of
the finite element basis. The spatial integrals are evaluated using Gauss quadrature.
Since the weight functions are arbitrary, this results in a system of first-order, non-linear
ordinary differential equations that may be written in the form:
R A (u i ,u i ,t , p ) = 0, A = 1" n s

(60)

where RA represents the residual error, A represents the values of the arbitrary weight
functions, and ui, ui,t, and p are the solution vectors in the discrete space.

The

generalized- time integrator, discussed below, is then used to convert equation (60) to a
system of non-linear algebraic equations. Newtons method is used to linearize the
problem about u, tn +1 and p, tn +1 . The system of equations may now be written as:
RM
K G u n +1

D C p
R

n +1
C

28

(61)

where RC and RM represent the portions of the residual, R, from the continuity and
momentum equation, respectively, and p n+1 represents a modified p n +1 (change in
pressure acceleration) to allow using Backward Euler for pressure while maintaining D =
-GT. K, G, D, and C are tangent matrices of the residual vectors with respect to the state
variables u, tn +1 and p, tn +1 . Approximate forms of these tangent matrices are used to
achieve better convergence. As such, they are defined as:
K

R M
u n +1

R M
p n +1

R C
u n +1

R C
.
p n +1

(62)

The reader may refer to Whiting (1999) and Jansen et al. (2000) for a more complete
discussion of using the generalized- method to solve the stabilized finite element
formulation of the incompressible Navier-Stokes.

4.2 Generalized Alpha Method


PHASTA-IC uses the Generalized- method Jansen et al. (2000) to integrate the
spatially discretized Navier-Stokes equations from time step n to n+1. Whiting (1999)
presents its application to the incompressible Navier-Stokes equations.

This is a

predictor-multi-corrector algorithm in which the solution is related to its time derivative


by:
y n +1 = y n + ty n + t ( y n +1 y n ) ,

(63)

where y is the solution vector, y is its time derivative, t is the time step, is a
weighting parameter, n denotes the previous time step, and n+1 denotes the current time
step being solved. The spatially discretized Navier Stokes equations may be expressed
as:
R ( y , y , p ) = 0 ,

(64)

where R is the residual of the continuity and momentum equations. Given a solution at
time step n, the generalized- method first makes a prediction of the solution and its
time derivative at n+1. A series of correction passes are then made to improve this
initial guess. The prediction used in PHASTA-IC assumes the solution at n+1 matches
the previous solution at n:

29

y ni =+01 = y n
y ni =+01 =

1 ,
y n

(65)

where i is the iteration on correction passes. For each correction pass the solution is
computed at an intermediate time. The solution is computed at a time n+f and the time
derivative is computed at a time n+m and these are used to compute the residual from
Equation (64). The intermediate solutions are defined by:

(
(y

y ni ++1 m = y n + m y ni +11 y n
y ni ++1 f = y n + f

i 1
n +1

yn

)
)

(66)

A Newton type linearization of the residual with respect to the time derivative of the
solution at n+1 yields the following equation:
R i y ni + f
i
i
y n + f y n +1

i
y n +1 = R i .

(67)

Equation (67) represents a matrix problem in which the change in the solutions time
derivative, y ni +1 , at the ith iteration is computed. The solution at n+1 is then corrected
according to:
y ni ++11 = y ni +1 + y ni +1
y ni ++11 = y ni +1 + ty ni +1

(68)

As many correction passes from Equations (66) to (68) are performed to find the
solution at time n+1.
To obtain second order accuracy and satisfy stability, the three parameters, , f, and m
are related in the following manner:

1
1 3
+ m f , m =
2
2 1 +

1
, f =
1 +

= user defined parameter


Note that setting = f = m = 1 results in a Backward Euler solution.

30

(69)

4.3 Level Set Method


4.3.1 Governing Model

The level set method of Sussman [Sussman et al. (1998), Sussman and Fatemi (1999),
Sussman et al. (1999)] and Sethian (1999) involves modeling the interface as the zero
level set of a smooth function, , where represents the signed distance from the
interface. Hence, the interface is defined by = 0. The scalar , the so-called first
scalar, is convected within a moving fluid according to,
D
=
+ u = 0
t
Dt

(70)

where u is the flow velocity vector. Phase-1, typically the liquid phase, is indicated by a
positive level set, > 0, and phase-2 by a negative level set, < 0. Evaluating the jump
in physical properties using a step change across the interface leads to poor
computational results. Therefore, the properties near an interface are defined using a
smoothed Heaviside kernel function, H, given by:
0
, <

1 1
H ( ) = 1 + + sin , <

2

1
, >

where the interface thickness is defined by 2, as shown in Figure 2.

31

(71)

Fluid 1
Fsv

1, p1
n2

n1

interface
2, p2

Fsv

Fluid 2
Figure 2: Transition region of finite thickness 2 between fluids 1 and 2.

The fluid properties are thus defined as:

( ) = H ( ) +

2
(1 H ( ))
1

( ) = H ( ) + 2 (1 H ( ))
1

(72)

Although the solution may be reasonably good in the immediate vicinity of the interface,
the distance field may not be correct throughout the domain due to varying fluid
velocities throughout the flow field which distort the level set contours (i.e., see
Equation (70)). Thus the level set is corrected with a redistancing operation by solving
the following PDE:
d
= S ( )[1 d ] ,

(73)

where d is a scalar that represents the corrected distance field and is the pseudo time
over which the PDE is solved to steady-state. This may be more clearly expressed as the
following transport equation:
d
+ w d = S ( ) ,

where
32

(74)

w = S ( )

d
.
d

(75)

The second scalar, d, is originally assigned the level set field, , and is convected with a
pseudo velocity w, where S() is defined as:
1
, <

1
S ( ) = + sin , < <

1
, >

(76)

Note that, in a continuous solution of Equation (74), the zeroth level set or interface, =
0, does not move since its convecting velocity, w, is zero. Solving the second scalar to
steady-state restores the distance field to d = 1 but does not alter the location of the
interface defined by equation (70). The first scalar, , is updated using the steady
solution of d.
4.3.2 Finite Element Formulation

The level set scalar equations given by Equations (70) and (74) may be cast in the
following form:

, t + Ai , i + S = 0
where Ai is ui for scalar 1, , and S ( )d

(77)

for scalar 2, d. S is zero for scalar 1 and -

S() for scalar 2. The level set equations represented by Equation (77) are solved in the
same manner as for the INS equations given in section 4.1. Temporal discretization is
controlled via the generalized- time integrator while the stabilized finite element
formulation is used for the spatial discretization. The weak form of Equation (77), with
stabilization terms added, is given by:
nel

T
e
(w , t + wAi , i + wS )d + L w ( , t + Ai , i + S )d

e =1 e

where LT is an advective operator = u i

xi

=0,

(78)

, is a stabilization parameter, and w is the

weight function belonging to the weight space w Wh. The solution belongs to the
solution space: Vh. The weight function and solution variable are expanded in terms
of linear basis functions:
33

nnp

w = wB N B
B =1

nnp

= A N A
A=1

nnp

, i = A N A ,i
A=1

nnp

, t = A ,t N A .

(79)

A=1

After evaluating the integrals using Gauss quadrature, Equation (78) may be expressed
as a non-linear ODE given by:
R
n +1

,
Nt

n +1

, t = R ,

(80)

where R is the residual error of equation (78). PHASTA-IC uses the Generalized-
method described in section 4.2 to advance the solution from time n to n+1. This is a
predictor-multi corrector algorithm in which the solution and its time derivative are
computed at intermediate time n+f and n+m, respectively.

The solution is then


n + m

projected to time n+1. Given this, M is evaluated using the chain rule for variables , t
n + f

and .
PHASTA-IC is able to solve Equation (80) implicitly or explicitly for either scalar. A
Forward Euler method is used as the explicit solver. The user selects the type of implicit
solver for scalar 1 by setting the parameter (see section 4.2 on generalized-a method).
When solving the second scalar implicitly, PHASTA-IC uses a Backward Euler
approach ( = f = m = 1). Once the change in the time derivative of the second scalar,

d , is solved the scalar field is corrected using Equation (68):


d n +1 = d n + d

(81)

where dn+1 is the current solution, dn is the old solution, and is the pseudo time step
for the second scalar PDE. The user defines the number of iterations, or time steps, to
solve the re-distance equation to steady-state.
4.3.3 Volume Constraint on Level Set Redistance Step

Sussman et al. (1999) and Sussman and Fatemi (1999) proposed an additional constraint
to be applied during the redistance step to help ensure the interface (=0) does not move.
They found imposing this constraint to improve the convergence of the redistance step
while maintaining the original zero level set. The essence of the constraint is to preserve

34

the original volume of each phase during the redistance step. The volume of each phase
within a cell, Vij , is defined as:
Vijk =

H (d

)dx

(82)

where H is the Heaviside function given in Equation (71) and dk is the value of the
distance field, d, at iteration k. Small changes in the zero level set from pseudo time step
0 to k may be approximated by:

Vijk Vij0 k 0

) H (d

H ' (d )(d
0

dx

(83)

d dx
0

where H' is

H (d )

and is given by:

H ' ( ) =
1 1 1
d
+ cos
2

d >
d

(84)

Hence, to minimize the deviation in volume each new d ijk is constrained to satisfy the
following:

H ' (d )(d
0

d 0 )dx = 0 .

(85)

~
This is accomplished by projecting the value of the current level set, d ijk , onto the new
level set field, d ijk , where d ijk is computed from:
~
d ijk = d ijk + ij ( k 0 )H (d 0 ) .

(86)

The scaling function ij is assumed constant within each cell and is computed from
Equations (85) and (86) to be:
~
d k d0
0

dx

H d k
0

.
ij =
0 2
H d dx

( )

( ( ))

35

(87)

~
Note when the distance field has converged, d ijk = d ijk , and ij = 0. In PHASTA-IC, the
volume constraint correction occurs at the global level after each redistancing iteration.
The calculation of ij occurs at the element level and is projected back to the global level
to be used by Equation (86).

4.4 Continuum Surface Tension (CST) Model


Surface tension is the result of sharp changes in the molecular forces of attraction at an
interface due to the discontinuous changes in properties. It is difficult to accurately
model this local force for complex geometries. Instead, the Continuum Surface Tension
(CST) model of Brackbill et al. (1992) is utilized to model the surface tension force as a
continuous, three-dimensional volume force smoothly varying across a finite interface.
The pressure change across an interface, defined by fluids 1 and 2, is given by the
momentum jump condition at the interface:

(p p + + n n + n n )n +
(n t n t )t = m (v v )
2

where pj, v j , and

(88)

are the pressure, velocity vector, and viscous shear stresses at the

interface for fluid-j, is the surface tension coefficient, is the local curvature of the
interface, m 2 is the mass flux due to phase change for fluid 2, n i is the unit normal
vector to the interface and t is the unit tangent vector along the interface. For the
special case of inviscid and adiabatic conditions and considering only the normal
component for constant surface tension, neglecting tangential forces, Equation (88)
reduces to:

p 2 p1 =

(89)

where local curvature is defined as the divergence of the normal along the interface:

= s n = [I n n ] n .

(90)

The surface force per interfacial area, normal to and at any given position along the
interface, is thus given by:
f s = n (x x i )

36

(91)

where (x-xi) is the Dirac delta function and restricts the surface tension force to the
interface defined to be at x=xi. Consider the interface shown in Figure 2, as defined in
the level set method above, where the interface is given a finite thickness of 2. The
local surface tension force given in Equation (91) may be expressed as a volume force,
Fsv, such that:

lim Fsv dV = f s dA .
0

(92)

A smoothing function is chosen that smoothly varies the properties across the interface.
Note the smooth Heaviside function, H, as defined in Equation (71), is used in
PHASTA-IC for this purpose. The foundation of the CST is to use the gradient of the
smoothing function, H as defined in Equation (84), to vary the surface tension force
across the finite interface. This function must zero out the force at the boundaries of the
finite interface but provide a net force over the volume equivalent to the local area force
fs. Additional choices for smoothing functions are explored in Brackbill et al. (1992)

and Williams et al. (1999). Note in the limit as the interface shrinks, H becomes the
delta function:
lim H ( ) = (x x i ) .
0

(93)

Since the interface is defined in terms of contours of the smoothing function, the unit
normal at the interface may be defined in terms of the smoothing function:
n =

.
=

(94)

Due to the differentiability of the smoothing function, the integral of the surface tension
force in Equation (92) may be expressed as (see Brackbill et al. (1992)):
H ( )

f dA = lim
( ) [H ( )] dV .

(95)

According to Equation (71), the jump in H across the interface, [H], is unity.
Comparing Equation (95) to Equation (92), we see that the volumetric surface tension
force is given by:
Fsv = ( )H ( ) .

When put in non-dimensional form Fsv becomes,


37

(96)

Fsv =

* ( ) * H ( )
We

(97)

where We is the Weber number defined as:


We =

1 LU 2

(98)

and the non-dimensional curvature, *, is given by:


*
( ) = s * .

(99)

Note the gradient operators in Equations (97) and (99) are non-dimensional.
The choice of the smoothing function can impact accuracy and stability of the numerical
computations. A discussion of the desirable and required features of kernels is provided
in Williams et al. (1999). Of these, the following three are essential:
1. It must be sufficiently smooth.
2. It must have a normality property; (i.e.,

K ( x, )dx = 1 , where represents the

finite interface.)
3. It must approach the Dirac delta function, (x), in the limit as |k| 0.

38

5. CANONICAL TEST PROBLEMS


Three canonical test problems were studied to assess the ability of PHASTA-IC to
accurately capture the movement of interfaces. The first problem tracked the interface
of a semicircle as it rotated in a purely rotational velocity field. The level set parameters
were varied to test their sensitivity to interface resolution. The second problem
numerically tracked the front propagation of a 2-D dam break problem that has been
experimentally and analytically studied. The third problem tracked the propagation of a
traveling solitary wave. The results were compared to data and analytical solutions.
PHASTA-IC was able to match the theoretical and experimental results for these
problems while incurring minimal mass error.

5.1 Rotating Semicircle


The relative importance of the various parameters related to the level set solution method
was tested with the simple canonical test problem shown in Figure 3. A semicircle of
radius R = 0.26 was placed in a 1.0 x 1.0 domain with a purely rotational velocity field
having an angular velocity of =2.0 rad/s. Solving only the level set equations, the
semicircle was tracked as it rotated around the center of the domain. Contours of the
initial level set and velocity fields are given in Figure 4. For each calculation, enough
time steps were taken such that the semicircle did a complete revolution.

The

boundaries of the semicircle at t=0 (red line) and after one revolution (black line) are
presented together to show the effect of the level set parameters for each case. Contour
lines of the distance field are also provided.
After performing a sensitivity study in which the grid size, time step and interface
thickness for advecting the interface, the time step and interface thickness for the
redistancing step, the type of solver, and the stability parameter were varied, the various
default values of these parameters were chosen. Figure 5 and Figure 6 give the results
on the 50x50 and 100x100 grids. The suggested default values are listed below.

39

t such that CFL=0.50

ls = 1.5

such that CFL=0.50

lsd = 3.0

Default solver type = implicit / explicit (scalar 1 / scalar 2)

Stability parameter constant, c1 = 0.0

Boundary Conditions
x: periodic
y: periodic
z: periodic
Grids Investigated
Grid
25x25x1
50x50x1
100x100x1

=2.0
y
x
z

Semicircle of R=0.26 placed in domain


such that r=0 is at center of domain.

Element Size
x=y=z=0.04
x=y=z=0.02
x=y=z=0.01

Initial Conditions
Pressure = 0
Velocity = r (=2.0 rad/s)
Radius = 0.26
Properties
No difference in properties
No surface tension
Gravity ignored
Domain
0.0 < x < 1.0
-0.5 < y < 0.5
0.0 < z < x

Figure 3: Schematic of two-dimensional level set test problem. Semicircle of


R=0.26 is placed in center of domain with a =2.0 prescribed velocity field.
The domain was 0<x<1.0, -0.5<y<0.5. The grid size and level set parameter
values were varied in the tests.

40

Level Set Field

Velocity Field

Figure 4: Initial level set and velocity fields for two-dimensional level set test
problem. Semicircle of R=0.26 is placed in center of domain with a =2.0
velocity field. The domain was 0<x<1.0, -0.5<y<0.5.

Figure 5: Semicircle interface at t=0 (red line) and after one revolution
(black line) for the implicit/explicit solver type using the 50x50 grid;
=0.02, t=0.005 (CFL=0.25), =0.01 (CFL=0.50), ls=0.03 (1.5), lsd=0.03
(1.5), c1=0.0

41

Figure 6: Semicircle interface at t=0 (red line) and after one revolution
(black line) for the implicit/explicit solver type using the 100x100 grid;
=0.01, t=0.01 (CFL=1.0), =0.005 (CFL=0.50), ls=0.015 (1.5), lsd=0.03
(3.0), c1=1.0

5.2 Analysis of 2D Dam Break


The 2-D dam break problem was studied experimentally by Martin and Joyce (1952) and
numerically investigated by Kelecy and Pletcher (1997) and Yue et al. (2003). It was
also simulated using PHASTA-IC and the results are given herein. The problem setup,
along with a schematic indicating the free surface before and after the dam break, is
given in Figure 7. The problem consists of a column of water being held in place by a
membrane in a box. At t=0, the membrane was broken allowing the water column to
collapse. The location of the surge front, s, and the height of the water column, h, was
measured as a function of time. The numerical problem has no-slip wall boundary
conditions in the x and y dimensions and periodicity in the z dimension (i.e., PHASTAIC was a 3-D simulation with only one node in the z direction). The air and water were
at atmospheric pressure and 68oF and their velocities were initially zero, with a
hydrostatic pressure profile in the water column. The domain was modeled using a
uniform 400x100 mesh (x=y=0.000714m). The interface thickness, for advection

42

of the interface using the level set method, was 2=0.004~5.5x.


thickness for the redistancing operation was kept at 2=0.002~3x.

The interface
The level set

equations were solved explicitly with the CFL number for all equations not exceeding
0.6.
Snapshots of the free surface position and velocity vectors in the computational domain
are provided at various times in Figure 8. Computations were made to the point where
the water climbs the adjacent wall. At the start of the dam break, the water accelerated
from its initial static condition due to the relatively large pressure difference existing on
the right side of the water column once the membrane was removed. The velocity of the
surge front of the water continued to increase until the opposite wall was impacted. At
each time, the maximum velocity was in the air just ahead of the surge front. The nondimensional surge front location, Z*, and water column height, H*, are plotted along
with the experimental data of Martin and Joyce (1952) in Figure 9 and Figure 10,
respectively. PHASTA-IC did a good job predicting the data and also agrees with the
numerical predictions presented by Kelecy and Pletcher (1997) and Yue et al. (2003).
As can be seen, the predictions begin to deviate a little from the data after T*>2.0. Yue
et al. (2003), who also used a level set method, also found this deviation and attributed it
to using too large an interface thickness, and they found the best numerical prediction
was when using 2=2x (note, the PHASTA-IC computation presented herein used
2=5.5x; however, the actual values of the interface thickness are similar to those used
by Yue et al. (2003) who used a mesh twice as coarse as used herein). In addition,
PHASTA-IC did a good job conserving mass (i.e. within 0.6 percent) even for a
relatively coarse spatial mesh (i.e. 400x100).

43

Boundary Conditions
x: no-slip wall
y: no-slip wall
z: periodic

400x100 Grid
1.25a

Air
a

Water

t>0

y
x

s
5a

a = 2.25 in = 0.05715 m
x = y =7.14375E-4m

Properties
Liquid: =1,000 kg/m3
=1.0E-03 Pa-s
U = 0.0 everywhere (t=0)
Gas: =1.2 kg/m3
=1.8E-05 Pa-s
Surface Tension:
= 0.06 N/m

Level Set Parameters


ls = 0.002 m lsd = 0.001m (~1.5x)
t = 0.0001s = 0.0002s (CFL=0.3)

Figure 7: Schematic of two-dimensional dam break problem. Water was initially


confined in a square box. The computational domain was 5a x 1.25a., and a
400x100 uniform mesh was used.

44

Time = 0.0
T* = 0.0

Time=0.10
T* = 1.31

Time=0.20
T* = 2.621

Time=0.30
T* = 3.931

Figure 8: Interface and velocity vectors at various times for the 2-D dam break
using a 400x100 grid.

2D Dam Break: Surge Front -vs- Time


a=2.25 in

z/a
Z*Z*
= =s/a

6.0
5.0

Martin & Moyce, 1952

4.0

PHASTA 400x100 dt=0.0001

3.0
2.0
1.0
0.0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.5

T* = nt*(g/a)

Figure 9: 2-D dam break: non-dimensional surge front


position versus non-dimensional time.

45

2D Dam Break: Height -vs- Time


a=2.25 in
1.2

Martin & Moyce, 1952

1.0

PHASTA 400x100 dt=0.0001

H* = h/a

0.8

0.6
0.4
0.2
0.0
0.0

1.0

2.0

3.0

4.0

5.0

0.5

= t*(g/a)

Figure 10: 2-D dam break: non-dimensional water column


height versus non-dimensional time.

5.3 Analysis of Solitary Wave


A 2-D traveling solitary wave, as analyzed by Yue et al. (2003), was numerically
simulated and compared to the analytical theory of Mei (1989). The solitary wave was
generated by releasing water, initially at rest, having a Boussinesq profile [Whitham
(1974)] given by:
A( x) =

Ao

3 o x
h
cosh
2
h

(100)

This profile is shown in Figure 11, where h is the depth of the water and Ao is the initial
height of the water on the left wall. A hydrostatic pressure profile was applied across the
domain. As seen in Figure 11, the water will fall under gravity to create a solitary wave.
For the case of zero bulk velocity for air and water, where Ao/h << 1, the dispersion
relation for shallow water waves [Whitham (1974)] gives the wave speed, Cw, as:

C w = gh ~ 1.0m / s .

(101)

For this problem the initial amplitude, Ao, was selected to be 0.4h. The domain was
modeled with a 200x120 non-uniform grid where the x-dimension had 200 elements of
46

uniform width (x=0.1h), while the y dimension had two regions of uniform width, and
a periodic boundary condition was used across a single element in the z direction. The
region containing the solitary wave had 100 elements of width y=0.01h while the
region outside the solitary wave had 20 elements of width y=0.05h. The mesh region
containing the solitary wave was consistent with that used by Yue et al. (2003).
The solitary wave shown in Figure 11 was fully formed at t=0.6 seconds and this is
marked as the t=0.0 waveform in Figure 12. The solution had an amplitude of 0.185h
and a wave speed of 1.0 m/s, which agreed with Equation (101). The velocity vectors at
t=0.4 given in Figure 13 shows the typical vortex generated at the crest of the solitary
wave. According to the perturbation theory of Mei (1989), the amplitude of the solitary
wave decreases with time due to viscous dissipation according to:
Amax

1 / 4

= Ao _ max

1 / 4


+ 0.08356 1 w 3

2
Cw 2 h

C wt
,
h

(102)

where Ao_max was taken to be the amplitude of the solitary wave once fully established at
t=0.6s; hence, Ao_max = 0.185h. The computed wave amplitudes are compared with
Equation (102) in Figure 14 and found to agree rather well, even if surface tension was
not considered. The PHASTA-IC predictions for times greater than t=0.6 seconds
began to be influenced by the opposite wall which is not in the analytical model. As for
the dam break problem, the mass error associated with the calculation was small.

47

200x120 Grid
0.10
0.08

0.5h, y=0.05h

Air

0.06
0.04

0.02

0.5h, y=0.01h

Ao = 0.4h

0.00
-0.02

0.5h, y=0.01h

-0.04
-0.06

-0.08

0.5h, y=0.05h

Water

-0.10
0.0

0.5

x = z = 0.1h = 0.01 m

1.0

1.5

Boundary Conditions
x: no-slip wall
y: no-slip wall
z: periodic
Properties
Liquid: =1,000 kg/m3
=2.0E-03 Pa-s
Gas: =1.2 kg/m3
=3.6E-05 Pa-s
Surface Tension:
= 0.06 N/m
Level Set Parameters
ls = lsd = 0.005 m
t = 0.001 s
2.0 = 0.0005 s

x = 20h,x x=0.1h

Figure 11: Schematic of 2-D traveling solitary wave. Water was started at rest with
a Bousinesq profile, of height Ao, with a hydrostatic pressure balance. The
computational domain of 20h x 2h was discretized with a 200x120 non-uniform
mesh.

Figure 12: Snap shots of the solitary wave calculation. The solitary wave is
considered generated 0.6 sec after the free fall of the initial profile. It is
designated with the t=0.0 marker. Grid: 200x120, wave speed=1 m/s.

48

Figure 13: Snap shot of the velocity vectors of solitary wave at t=0.4s. The wave
is traveling to the right at a wave speed=1 m/s.

2D Solitary Wave: Wave Amplitude


0.30

Amplitude (Amax/h)

0.25

0.20

0.15

Analytical Perturbation Sol'n by Mei

(1989)

0.10

PHASTA - Surf. Tension


PHASTA - No Surf. Tension

0.05

0.00
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Time(s)(s), t
Time

Figure 14: Solitary wave amplitude. (Grid: 200x120)

49

0.8

6. PARALLEL MESH ADAPTATION


The mesh size (i.e., the characteristic size of a mesh cell) necessary to resolve the flow
may vary spatially across the computational domain (i.e., the length scales at the wall
and interface may require smaller mesh size than in the free stream). With this in mind,
using a mesh of variable size, where the size of a mesh cell is small only where it is
needed and larger elsewhere, allows for the best resolution of the flow for a given cell
count. And given the interface moves with time in multiphase problems, such that the
location of the refined and coarse mesh cells will vary in time, an adaptive mesh
adaptation scheme must be employed. The phParAdapt mesh-adaptation code was used
in conjunction with PHASTA-IC to iteratively refine and/or coarsen the computational
mesh based on various correction indicators. The phParAdapt code performs parallel
mesh adaptation in which the local mesh size is constructed using selected correction
indicators like residual error, shear, and/or proximity to the interface. The code also uses
parallel mesh tools developed at RPI [de Cougny and Shephard (1999)] and the parallel
MeshSim libraries of Simmetrix Inc. [Simmetrix (2009)]
Unfortunately, at the time of the simulation, the parallel mesh tools were not supported
on the platform (i.e., RPIs Blue Gene computer) where the two-phase annular flow
simulation was computed. Hence, the simulation presented in this thesis did not use an
adaptive mesh refinement scheme. However, future work should use iterative mesh
refinement to take advantage of the increased flow resolution and/or mesh size
reduction. The phParAdapt code is described below in more detail along with an
example illustrating its use.

6.1 Code Description


The phParAdapt code takes the current parallel mesh along with the PHASTA-IC flow
solution and error estimates as inputs and computes the desired mesh size field required
by the parallel mesh refinement/coarsening procedure to produce a computational mesh
which creates the local element length scales required to adequately resolve the physics
length scales. The process of mesh adaptation utilizes local mesh modification operators
[Li et al. (2005)] to improve the mesh with respect to the desired mesh size field, and
executes over a distributed mesh on a parallel computer [de Cougny and Shephard
50

(1999) and Alauzet et al. (2006)]. In this process, the flow solution on the input mesh is
mapped onto the adapted mesh as the mesh is being modified locally, which provides
updated PHASTA-IC input to continue the simulation. Such a solution strategy allows
all the necessary steps of the simulation to be carried out in a parallel manner without
creating the bottlenecks associated with a serial mesh adaptation process.
There are several options to compute the desired mesh size field. In general, phParAdapt
computes the new mesh size field by comparing the local value of the correction
indicators with the target ones, determined by the user controlled parameters, with the
aim of equi-distributing the errors over the domain [Babuvska and Rheinboldt (1978)].
In locations where the local value of the correction indicators exceeds the target value,
the size field is refined, whereas in regions where these values are lower than the target,
the size field is coarsened.
The correction indicators used to determine the desired mesh size field for the domain
are either imported from the error indicators output by PHASTA-IC or computed within
phParAdapt. Typical methods of generating the size field are:
1. Error indicators output from PHASTA-IC provide the residual error in the
momentum equations and the shear stress divergence (which can be related to
shear stress jumps across elements) in the three coordinate directions.

The

analysis of a plunging liquid jet has indicated good success using these type error
indicators [Galimov (2007)].
2. Interpolation-based error indicators are determined by considering the lowestorder error term which dominates when interpolating a function. When using
linear basis functions in the calculations, the interpolation errors are dictated by
the second derivatives that can be reconstructed to extract error information
[Sahni et al. (2006)]. Furthermore, directional error indicators can be derived by
using second derivatives in the form of a Hessian matrix [Sahni et al. (2006)].
Such a scheme, referred to as a Hessian strategy [Kunert (2002)], is often applied
in flow problems involving anisotropic solution features like shock waves and/or
boundary layers.
3. A proximity indicator, E(), which indicates proximity to the interface, such as:

51

1 1
1 1 + + sin
E ( ) = 2

, <

, >

(103)

where is the level set field indicating the signed distance to the interface.
4. A ramp proximity indicator that allows the user to specify the desired mesh size
field continuously.

The general logic sets the size near the interface to a

specified low value, min_size, and set the size away from the interface to a
specified max value, max_size:

<
min_size
,
(min_size + max_size )
, < < 3
Size =
2

max_size
,
> 3

(104)

where the thickness of the region of small elements around the interface is given
by .
It is also possible to use combinations of the above strategies (e.g., use of a proximity
detector like (3) as a conditional sample for error indication based on strategies (1)
and/or (2) as was done by Galimov (2007).

6.2 Example Problem-A Convecting Sphere


A simple convecting sphere in a plug flow problem, shown in Figure 15, was simulated
to assess the grid refinement and coarsening capability of the phParAdapt code. This
was a so-called red water / green water problem where there was surface tension, but
the properties of the two fluids were the same. The domain, a 0.5m x 0.2m x 0.2m box,
was periodic in the axial direction (i.e., z direction) and had slip walls elsewhere. The
properties were for water at standard temperature and pressure and were constant
throughout; however, since surface tension was applied at the interface there was a
pressure rise consistent with the Laplace equation within the sphere. Two approaches
were used to track the sphere for one evolution through the domain. The first approach,
termed the uniform approach, used a uniformly sized mesh (size=0.005m) where no
mesh adaptation was performed. The second approach, termed the adaptive-mesh
approach, used phParAdapt to keep a mesh having a size of 0.005m around the sphere
52

interface but used a mesh of size 0.02m elsewhere. For the adaptive-mesh approach, a
PERL script controlled the iterations between PHASTA-IC and phParAdapt with
PHASTA-IC automatically halting when the interface (i.e., zeroth level set) reached the
limits of the fine mesh region.

The uniform mesh had a mesh size of 1,776,240

tetrahedrons which is more that four times the average size of the adapted mesh of
410,000 tetrahedrons.
The time step, t, was set such that CFL=1.0, was set such that CFL=0.25, ls=2,
lsd=2.5, where represents the local mesh size. The specified size field method given
in Equation (104), with = 0.04m, min_size = 0.005m, and max_size = 0.02m, was used
in the mesh adaptation process. Figure 16 shows 3-D contours and planar cuts of the
sphere at the initial, middle, and final times during the initial transport of the sphere
through the domain using the adaptive-mesh approach. At time step 405, as the sphere
was crossing through the periodic plane, the interface reached the coarse elements within
the specified tolerance of 0.010m (i.e. 2min). Figure 16 shows the mesh before and after
adaptation at time step 405, where coarse regions near the interface were refined and fine
regions away from the interface were coarsened. At step 868, the solution has again
progressed to the point where it must be halted and adapted before it can again be
advanced further.

Boundary Conditions
x: periodic
y: slip wall
z: slip wall
Initial Conditions:
U=<1,0,0> m/s

Tetrahedron Mesh

Properties
=1,000 kg/m3
=1.0E-03 Pa-s
= 0.05 N/m

Level Set Parameters


ls = 2
lsd = 2.5
t = (CFL=1.0)
= (CFL=0.25)
Sphere:
Centered @ (0.25,0.0,0.10) m
Rb=0.075 m

Figure 15: Schematic of three-dimensional convecting sphere. A red


water/green water problem with surface tension. The computational domain
was 0.5m x 0.2m x 0.2m.

53

3D View of Sphere

Time
Step

Planar Cut of Sphere

405

405 after
adapt

868

Figure 16: Transport of a sphere through the domain using adaptive-mesh


approach: element size = 0.005m near interface and 0.02m elsewhere, mesh size ~
410,000 tetrahedrons. The black line indicates the interface. Mesh displayed in
upper half of planar cuts while the color spectrum in the lower half indicates
pressure (Pa).

54

7. SELECTION OF EXPERIMENTAL DATA FOR SIMULATION


7.1 Selection of Data Set
The flow conditions for the DNS of annular flow must be established based upon the
experimental test condition being modeled.

Two experimental data sets were

considered: the air/water experiments of Owen (1986) and the adiabatic steam/water
experiments of Wurtz (1978).

Descriptions of both data sets are given below for

completeness.
Both data sets capture vertical annular flow and the annular flow is considered to be
steady, fully-developed, and in equilibrium based on pressure drop measurements.
However, the steam/water data of Wurtz provides film thickness and interfacial wave
frequency measurements that can be used in validating the DNS results. Although the
critical wavelengths for Helmholtz instabilities are very similar between the two data
sets, meaning the maximum element size required to resolve instabilities are similar, the
lower density ratio for steam/water data reduces the discontinuity in properties at the
interface making it easier to resolve interfacial instabilities. Thus the simulation done in
this thesis was based on Run 602 from Wurtz (1978).
Air/Water Data of Owen (1986):
Owen performed a multitude of two-phase flow tests in the LOTUS facility at the
Harwell Laboratory. The experiment consisted of a 0.03175m (1.25 inch) diameter tube
with a length to diameter ratio of L/D = 600 which operated at system pressures between
1.5 to 3.8 bar. The two-phase flows spanned several flow regimes with most being in
annular flow. Air was forced at mass fluxes between 4.4 and 296 kg/s-m2 and water at
fluxes between 5.3 and 1000 kg/s-m2. Near the exit of the tube, the pressure gradient,
shear stress, and liquid film mass flux were measured. Unfortunately, the liquid film
thickness was not measured. The density ratio for these air/water experiments are
between 200 and 530. This high density ratio has two adverse effects for the numerical
simulation. First, the sharp discontinuity in density at the interface is smoothed over a
finite thickness in the Level Set method. The thickness of this interface is controlled, in
part, by the stability of the level set method and is usually on the order of 3 to 5
elements. The large gradient in the density function over this interface makes it difficult

55

to capture the physical instabilities at the interface brought about by the density
difference.

The second effect is the small element size needed to capture the

instabilities.

As a first order estimate, the critical wavelength for a Helmholtz

instabilities is used. For air/water this is about 1.5 cm.


Steam/water Data of Wurtz (1978):
The steam/water experiments of Wurtz were conducted at the RISO National Laboratory
in Denmark. The experiments consisted of two-phase annular flows inside several tube
and annuli tests sections. Of particular interest are the adiabatic annular flow tests in test
section 20 which was a 20 mm diameter tube having a length of 9 m (i.e. L/D = 450).
Pressure taps were placed axially at equidistant locations near the exit allowing for
comparison of pressure gradient data to ensure equilibrium annular flow was established.
The flow was sucked off with a porous sinter at the exit and passed through a heat
balance unit to determine the liquid film flow rate. In addition, a needle contact probe
was used to measure both the mean film thickness and capture the roll wave frequencies
along the interface. The tests were performed at a system pressure of 70 bar with total
mass fluxes between 500 and 2000 kg/s-m2. The small density ratio of 20 at this
pressure (relative to the 530 density ratio for the air/water data of Owen (1986) results in
a much smaller discontinuity in physical properties over the interface region making it
more likely to capture interfacial instabilities. The reduction in the surface tension,
however, counteracts the reduced density ratio to keep the critical wavelength, c,
around 1 cm, which is similar to the air/water data of Owen (1986).

7.2 Reducing the Experimental Data


Table 1 provides the global conditions and exit quality for each run. It also provides the
measured axial total pressure gradient, dp

dz

, and frictional component, dp ,over


dz f

the last meter of the test section, the mean liquid film thickness, m, and the mean
wavelength, w, of the disturbance waves computed from the contact probe data. The
global void fraction, , was computed by assuming m represents all the liquid in the
domain (.i.e., liquid entrainment is neglected). Hence, a geometrical relation yields:

56

R m
=

(105)

where R is the tubes outer radius. Now the mean velocity for the water, u l , and steam,
u g , may be computed as:

wl = (1 xout )win ,

ul =

wl
,
l (1 )AXS

w g = x out win

ug =

wg

(106)

g AXS

A force balance can be used to determine the wall shear stress, w, from the axial
frictional pressure gradient, (dp/dz)f:
dp D
,

dz f 4

w =

(107)

where D is the tube diameter. We note for upward flowing annular flow that w may be
positive, negative, or zero depending on how the density head compares to the total
pressure gradient [Lahey and Moody (1993)]. With the experimental shear known, the
friction velocity, u, and turbulent Reynolds number, Re, may, when w 0 , be
computed as

u =

m u
, Re = ,

where is the density and is the kinematic viscosity (/).

(108)
The mean liquid film

thickness can be expressed in wall units as:

m+ =

m u
= Re .

(109)

7.3 Estimating Minimum Turbulent Scales


It is difficult to know a priori the necessary resolution needed to capture the relevant
physics of annular flow. However, one can get a sense of the magnitude of the smallest
element size by computing the Kolmogorov length scale, , which is an estimate of the
smallest turbulent eddy in the flow. The Kolmogorov length (), time (), and velocity
() scales are defined as:

57

( )

=
3

( )

, =

, = ( )

(110)

Wilcox (1998) estimates that the average turbulent dissipation, , in a channel is given
by:

2u u l

(111)

where u l is the mean flow in the channel and is the channels width. Applying this to
the liquid film, the Kolmogorov scales can be computed for each run. Table 2 provides
the turbulent parameters and Kolmogorov scales for each run. The minimum length
scales vary from 0.3 to 3.0 m and the time scales can be as low as 1 sec. These values
are more relevant in the liquid film making the liquid film the heaviest meshed region in
the domain.

7.4 Selecting the Test Run for the Simulation


There were 21 adiabatic tests conducted in test section 20. The details of each are listed
in Table 1 and Table 2. The runs, plotted on the vertical flow regime map of Hewitt and
Roberts (1969) in Figure 17, are in the annular flow regime. The flow regime map plots
the superficial mass flux of the liquid, ljl2, against the superficial mass flux of the vapor,
vjv2, where jl and jv are the superficial velocities of the liquid and vapor, respectively,
and are defined by:
jl =

G (1 x )

and

jv =

Gx

(112)

G is the mass flux entering the test section and x is the flow quality.
When considering which run number to choose our simulation, the size of the domain
and the required grid resolution were of primary importance. The grid resolution is
directly related to the Kolmogorov length scale, , given in Table 2. Comparing the
data, one can see that decreases as the exit quality increases and as the bulk flow rate
increases as a result of the increased wall shear. This requires a finer mesh to resolve the
smaller length scales. However, the mean wavelength of the large disturbance waves,
w, decreases as the exit quality increases making the domain shorter. The deciding
trend is the increased velocity associated with increased quality which will result in
58

smaller time steps. Although it is not obvious which test condition will require the least
computational resources to simulate, it was decided to maximize and minimize the
phasic velocities to limit computational resource usage. Desiring a low quality run at a
low mass flow rate, run #602, having an exit quality of 30 percent and a disturbance
wave wavelength below 0.2m, was chosen as the test to simulate.

Figure 17: RISO test section 20 data plotted on the vertical flow regime
map of Hewitt and Roberts (1969) for thin tubes (1-3 cm) which was
validated against low pressure air/water and high pressure steam/water
data.

59

Table 1: RISO data for adiabatic steam/water tests in test section 20.
Run
No.

60

601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621

Press

Gin

win

xout

dp/dz

(dp/dz)f

wg

wl

in

ul

ug

(bar)
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70
70

(kg/m2-s)
500
500
500
500
500
500
750
750
750
750
750
750
1000
1000
1000
1000
1000
1000
2000
2000
2000

(kg/s)
0.1571
0.1571
0.1571
0.1571
0.1571
0.1571
0.2356
0.2356
0.2356
0.2356
0.2356
0.2356
0.3142
0.3142
0.3142
0.3142
0.3142
0.3142
0.6283
0.6283
0.6283

(-)
0.2
0.3
0.4
0.5
0.6
0.7
0.2
0.3
0.4
0.5
0.6
0.7
0.2
0.3
0.4
0.5
0.6
0.7
0.2
0.3
0.4

(kPa/m)
3.6
3.3
3.5
3.7
4.0
4.4
4.1
4.4
4.9
5.5
6.1
7.0
5.4
6.1
6.9
8.0
9.2
10.6
11.8
14.1
17.3

(kPa/m)
1.3
1.6
2.1
2.5
3.1
3.6
2.0
2.8
3.7
4.5
5.3
6.4
-4.6
5.8
7.0
8.5
10.0
10.0
12.7
16.1

(m)
1570
940
620
500
420
320
1250
820
580
410
265
185
1380
850
450
250
170
105
985
270
110

(m)
0.257
0.183
0.139
0.122
0.094
0.092
0.173
0.152
0.113
0.087
--0.175
0.121
0.094
---0.375
---

(kg/s)
0.031
0.047
0.063
0.079
0.094
0.110
0.047
0.071
0.094
0.118
0.141
0.165
0.063
0.094
0.126
0.157
0.188
0.220
0.126
0.188
0.251

(kg/s)
0.126
0.110
0.094
0.079
0.063
0.047
0.188
0.165
0.141
0.118
0.094
0.071
0.251
0.220
0.188
0.157
0.126
0.094
0.503
0.440
0.377

(-)
0.711
0.821
0.880
0.903
0.918
0.937
0.766
0.843
0.887
0.920
0.948
0.963
0.743
0.837
0.912
0.951
0.966
0.979
0.813
0.947
0.978

(m/s)
1.87
2.64
3.37
3.46
3.28
3.22
3.46
4.51
5.39
6.30
7.75
8.29
4.20
5.81
9.21
13.67
16.02
19.39
11.53
35.49
74.06

(m/s)
3.84
4.99
6.21
7.57
8.93
10.21
5.35
7.29
9.24
11.14
12.97
14.89
7.35
9.79
11.98
14.37
16.97
19.53
13.45
17.32
22.35

dp/dz = total axial pressure drop over the last 1m


(dp/dz)f = frictional component of axial pressure drop over
the last 1m
m = mean liquid film thickness
w = wave length of disturbance waves based on measured
mean central frequency

in = steam fraction based on water volume implied by mean

u l = mean water velocity


u g = mean steam velocity

Table 2: Turbulent scales for adiabatic steam/water RISO data in test section 20.
Run
No.

w_exp

m+

Re

ul

61

(m)

(Pa)

(m/s)

(-)

(-)

(m/s)

(m /s )

(m)

(s)

(m/s)

601

1.57E-03

6.5

9.37E-02

1192.5

1192.5

1.87

2.09E+01

3.08E-06

7.69E-05

4.01E-02

602

9.40E-04

1.04E-01

792.1

792.1

2.64

6.07E+01

2.36E-06

4.51E-05

5.23E-02

603

6.20E-04

10.5

1.19E-01

598.5

598.5

3.37

1.54E+02

1.87E-06

2.83E-05

6.61E-02

604

5.00E-04

12.5

1.30E-01

526.7

526.7

3.46

2.34E+02

1.68E-06

2.30E-05

7.33E-02

605

4.20E-04

15.5

1.45E-01

492.6

492.6

3.28

3.28E+02

1.55E-06

1.94E-05

7.97E-02

606

3.20E-04

18

1.56E-01

404.5

404.5

3.22

4.89E+02

1.40E-06

1.59E-05

8.81E-02

607

1.25E-03

10

1.16E-01

1177.6

1177.6

3.46

7.47E+01

2.24E-06

4.06E-05

5.51E-02

608

8.20E-04

14

1.38E-01

914.1

914.1

4.51

2.08E+02

1.73E-06

2.44E-05

7.12E-02

609

5.80E-04

18.5

1.58E-01

743.2

743.2

5.39

4.65E+02

1.42E-06

1.63E-05

8.70E-02

610

4.10E-04

22.5

1.74E-01

579.4

579.4

6.30

9.35E+02

1.19E-06

1.15E-05

1.04E-01

611

2.65E-04

26.5

1.89E-01

406.4

406.4

7.75

2.09E+03

9.73E-07

7.68E-06

1.27E-01

612

1.85E-04

32

2.08E-01

311.8

311.8

8.29

3.87E+03

8.34E-07

5.64E-06

1.48E-01

614

8.50E-04

23

1.76E-01

1214.5

1214.5

5.81

4.25E+02

1.45E-06

1.70E-05

8.51E-02

615

4.50E-04

29

1.98E-01

722.0

722.0

9.21

1.60E+03

1.04E-06

8.77E-06

1.19E-01

616

2.50E-04

35

2.17E-01

440.6

440.6

13.67

5.17E+03

7.76E-07

4.88E-06

1.59E-01

617

1.70E-04

42.5

2.40E-01

330.2

330.2

16.02

1.08E+04

6.45E-07

3.38E-06

1.91E-01

618

1.05E-04

50

2.60E-01

221.2

221.2

19.39

2.50E+04

5.24E-07

2.22E-06

2.36E-01

619

9.85E-04

50

2.60E-01

2075.0

2075.0

11.53

1.58E+03

1.04E-06

8.83E-06

1.18E-01

620

2.70E-04

63.5

2.93E-01

641.0

641.0

35.49

2.26E+04

5.37E-07

2.34E-06

2.30E-01

621

1.10E-04

80.5

3.30E-01

294.0

294.0

74.06

1.46E+05

3.37E-07

9.18E-07

3.67E-01

613

Average dissipation computed from Wilcox (1998): ~

2u2 u l /

(average dissipation for channel flow)

Kolmogorov scales: length()= (3/)0.25 , time()=(/)0.5 , velocity()= ()0.25

8. SIMULATION OF TWO-PHASE ANNULAR FLOW


8.1 Problem Description
The simulation of two-phase annular flow was based on run #602 of the experimental
steam/water annular flow tests of Wurtz (1978).

Run #602 was an adiabatic test

conducted at 70 bar and the associated saturated mixture temperature, had an inlet mass
flux of 500 kg/s-m2, and an exit quality of 30 percent. The flow had a liquid film
turbulent Reynolds number, Re = um/, of 792, where u is the friction velocity =
(wall/)0.5, wall is the experimentally measured wall shear (8 Pa), is the liquid density
(741 kg/m3), m is the experimentally measured mean liquid film thickness (0.94 mm),
and is dynamic viscosity of the liquid (= 9.13E-5 Pa-s). The computational domain,
shown in Figure 18, was a 30 degree wedge of the 0.02m diameter tube used in the RISO
test. Ideally, the domain length would be at least the experimentally measured mean
wavelength, w, of the disturbance waves, which for run #602 was approximately 0.2m;
however, the domain length was limited to w/8 in our simulation to keep the problem
size tractable. Periodicity was enforced in the axial and circumferential directions. The
centerline was not included in the domain due to singularity issues that arise in the flow
solver. That is, the model was clipped at a radius of 0.0005m. An effective-viscosity
turbulent wall function was used along the tube wall, where the first layer of the
structured mesh along the wall was within the viscous sublayer to prevent the need to
know the log-law region coefficients.

62

Boundary Conditions
x: periodic (circumferential)
y: wall & symmetry
z: periodic (axial)

Initial Conditions:
Flow: Phasic plug flow
Level Set: m (with perturbation)
Imposed experimental pressure
gradient.
Properties
R = 0.01m, =30
-0.00259m < x < 0.00259m
0.00048m < y < 0.01m
-0.025m < z < 0.0m
o

z
x
y

(kg/m3)
: (Pa-s)
: (N/m)

Water
740
9.13E-5
0.0177

Steam
36.53
1.90E-5

Figure 18: Schematic of the simulation domain. Length = 0.025m = 1/8 w for
RISO Run # 602.
8.1.1 Mesh Description

The simulation was performed on an isotropic tetrahedron mesh, in which elements were
of size 0.00005m (y+~42), with an imposed boundary layer structure on the tube wall to
improve turbulence resolution. The boundary layer structure is shown in Figure 19
where the first layer is within the viscous sublayer. Hence, the wall function will impose
u+=y+ instead of u+=ln(y+)/+C. Note that Vassallo (1999) and Azzopardi (1999) found
that the turbulence in the liquid film, although enhanced by the effect of the interfacial
waves, still obeys the general form of the law of the wall. However, the coefficients for
the log law region in the liquid film are not well known. Placing the so-called off-thewall node in the viscous sublayer eliminates the need to know the values of these
coefficients. The mesh on the negative z plane was matched to the mesh on the positive
z plane to support axial periodicity. Likewise, the mesh on the circumferential planes
was matched to support circumferential periodicity. The resultant mesh consisted of
59.6 million tetrahedrons and 10 million nodes.

63

y+ = 35.5, y = 4.21E-5 m
y+ = 21.8, y = 2.59E-5 m
y+ = 12, y = 1.42E-5 m
y+ = 5 ,
y = 5.93E-6 m

Figure 19: Boundary layer structure on wall.


8.1.2 Initial and Boundary Conditions

As discussed above, periodic boundaries were enforced axially (z direction) and


circumferentially. A symmetry boundary condition was placed along the clipped inner
radius (r=0.0005m) by enforcing zero traction and no normal flow.

An effective

viscosity turbulent wall function, discussed in Section 8.1.2.1, was enforced at the wall
(r=0.01m) where the off-the-wall node was at y+=0.5 based on the experimental wall
shear listed in Table 2 for RISO run #602. Two body forces were imposed on the flow.
The first was acceleration in the negative z direction to simulate gravity. The other was
a pressure gradient body force in the positive z direction equal to the experimentally
measured pressure gradient. Only the frictional component was imposed in the first part
of the simulation (i.e., the first 95,100 time steps), but this was later increased to the total
axial pressure gradient, as shown in Table 3. Each phase was initially assigned a flat
(i.e., plug) velocity profile with a magnitude equal to the average velocity of the phase,
where the average velocity can only be determined once the void fraction is known. For
this simulation the global void fraction was computed using the experimentally
measured mean film thickness, m, and the assumption of negligible liquid droplet
entrainment. For RISO run #602 the mean liquid and steam velocities were 2.64 m/s and
4.99 m/s, respectively.
An initial perturbation, m, was applied to the interface to expedite problem start-up

64

2
m = cos

z ,

(113)

where is the assumed amplitude, p is the wavelength of the perturbation, and z is the
axial coordinate.

The perturbations wavelength, p, was set equal to the critical

wavelength, c, for Helmholtz instability:

c =

( g )g 2
c = l
,

(114)

where c is the critical wave number, g is gravity and is the surface tension. For RISO
run #602, c = 0.01m and c = 624.4 m-1 which results in two and one half wavelengths
over the truncated domain length (L=0.025m). The amplitude, , was set equal to two
element sizes.

For =0.00005m this relates to =0.0001m.

Hence the initial

perturbation on the interface was:


2
z .
m = 0.0001 cos
0.01

Figure 20 shows the initial condition for the simulation.

Figure 20: Initial condition for RISO run #602 using plug flow
velocity as initial guess (ul=2.64 m/s, ug=4.99 m/s) and applying
an imposed perturbation with wavelength p = c.

65

(115)

8.1.2.1 Effective-Viscosity Wall Function

The effective viscosity wall function used in the simulation uses the Spalding single
formula for the law of the wall [Spalding, (1961)] to produce a wall shear that supports
the flow velocity near the wall. It sets a zero velocity Dirichlet boundary condition on
the wall node and modifies the molecular velocity at the wall to produce the required
wall shear stress.
PHASTA-IC finds the closest node that exists off the wall to each wall node. These are
termed the off-the-wall nodes (otwn) that correspond to the wall nodes. For a given wall
node, nwall is normal vector to the wall and uotwn is the velocity of the off-the-wall node
which is known from the previous iteration or initial condition. The parallel component
of the off-the-wall node, u||, is computed by subtracting the projected component on the
wall normal vector from the velocity vector as shown by:
u || = u otwn (u otwn n w )n w .

(116)

The normal distance, dn, of the off-the-wall node to the wall node is given by:
d n = (x wall x otwn ) n w ,

(117)

where xwall and xotwn are the coordinates of the wall and off-the-wall nodes, respectively.
An iterative procedure is then used to compute the friction velocity, u, that supports the
parallel velocity, u||, at the specified distance, dn, from the wall. This procedure uses the
Spalding formula for the law of the wall region [Spalding, (1961)], as defined by:

( )

+
y + = u + + e e u 1 u + 1 u +
2

( ) ),

1 u +
6

(118)

which describes the relation between the non-dimensional wall distance, y+, to the non
dimensional velocity, u+, through the viscous sublayer and log-law regions.

The

constants and are defined as =0.4 and =5.5. y+ and u+ are defined by:
y+ =

yu

, u+ =

u
u

(119)

where is the molecular viscosity and is the fluid density. Note the Spalding formula,
Equation (118), reduces to the laminar profile u+=y+ in the viscous sublayer.
Finding the friction velocity that satisfies Equation (118) for the off-the-wall nodes
distance, dn, and velocity, u || , requires an iterative search procedure. At the start of the

66

search, an arbitrary value for u = 0.04 is specified. Using y=dn and u= u || in Equation
+
+
(119), the non-dimensional distance, y otwn
, and velocity of the off-the-wall node, u otwn
,

are computed.

+
, from the Spalding formula is
The non-dimensional distance, y spalding

computed using Equation (118). The error, f, is defined as the difference between
+
+
y spalding
and y otwn
as given by:
+
+
f = y spalding
y otwn

( ))

( )

+
+
+
= u otwn
y otwn
+ e e u 1 u + 1 u +
2

1 u +
6

(120)

The error, f, can be differentiated with respect to u to determine the correction in u


needed to correct the error f. The derivative of f with respect to u is given by
df
f u +
f y +
= +
+ +
du u u y u

((

( )

+
1 +
=
u + y + + e u + e u u + u +
u

( ) )}

1 u +
2

(121)

The correction in the friction velocity, u is then given by

u =

f
f

(122)

The current guess on friction velocity is adjusted according to:

new

= u

old

u .

(123)

The new value of u is fed into Equations (119) through (123) to obtain an improved
guess on friction velocity. This iterative scheme is continued until f < 1.0E-06. The
resulting value of the friction velocity is used to compute the effective viscosity, eff,
needed at the wall to support the existence the parallel velocity, u||, at the specified
distance, dn, from the wall. The effective viscosity is defined as:

eff = wall

dn
u||

u 2 d n
u ||

The effective viscosity replaces the molecular velocity for the off-the-wall nodes.

67

(124)

8.1.2.2 Verification of the Effective Viscosity Wall Function

As a check on the ability of the effective viscosity to compute the correct friction
velocity, the wall function was tested using the experimental data from RISO run #602.
Table 1 shows the frictional pressure gradient over the last one meter of the test section
to be 1600 Pa/m. Using Equations (107) and (108), the shear stress, w, and friction
velocity, u, are computed to be 8 Pa and 0.1039 m/s, respectively. The first layer of the
boundary layer structured mesh is 5.93E-06 m off the wall. Using the values of u and
the liquid density and viscosity listed in Figure 18, this corresponds to a y+ of 5.0 using
Equation (119). Since this corresponds to the viscous sublayer where y+ = u+, the
corresponding u+ is 5.0. From Equation (119), the parallel velocity, u = u || , is 0.52 m/s.
Now, using the values of y=5.93E-06 m and u || =0.52 m/s, the effective viscosity wall
function described in Section 8.1.2.1 is used to compute the required friction velocity.
After seven iterations, the wall function converges on u = 0.105 m/s which is within one
percent of the experimental value.
8.1.3 Simulation History

The flow was solved implicitly using a time step that maintained a CFL number around
unity. The first level set scalar was also solved implicitly at the flow time step, while the
second level set scalar was solved explicitly using a pseudo time step of 5.0E-06 seconds
which resulted in a CFL number near 1.0. The thickness of the interface for both scalars
was set at two element sizes or 1.0E-4 m. Table 3 provides the history of the simulation.
The simulation time range, time step size, t, and imposed axial pressure gradient are
listed for each range of time steps. Also listed is the non-dimensional transport time
range which is the simulation time range normalized by the average time it took for the
liquid film to transport through the domain. For an average liquid film velocity of 2.64
m/s and a domain length of 0.025 m, the average transport time was 9.47 msec.

68

Table 3: Simulation history


Time

Time
Step t

Imposed
Axial
Pressure
Gradient

NonDimensional
Transport
Times

(msec)

(msec)

(Pa/m)

(-)

0 - 6000

0.0 6.00

1.0E-6

1600

0 0.634

6,001 9,400

6.0 7.70

5.0E-7

1600

0.634 0.813

9,401 18,200

7.70 14.30

7.5E-7

1600

0.813 1.510

18,201 47,000

14.30 31.58

6.0E-7

1600

1.510 3.335

47,001 95,100

31.58 - 103.73

1.5E-6

1600

3.335 10.954

95,101 130,000

103.73 156.08

1.5E-6

3300

10.954 16.482

Time Steps

8.2 Data Analysis Procedures


Post processing of the simulation data can be categorized into three primary procedures.
These procedures, along with a brief description and key outputs of each, are presented
in Table 4. The first procedure was classification of the phase fields to distinguish
between dispersed and continuous phases in the domain. This method characterized the
number, size, volume, and surface area of each field. Once this was done, data at the
exit plane could be evaluated for instantaneous mass flow rates and area fractions of the
fields. This data could also be averaged circumferentially to provide radial profiles of
the area fractions.

Finally, by averaging the solution over many time steps, average

velocity and void profiles could be computed along with the necessary averaged
quantities to compute the Reynolds stress terms. However, care was taken to perform
these averages for each field. The sections below discuss each procedure in more detail
and provide details on how the data was processed.

69

Table 4: Post processing methods used in evaluating the simulation data


Post
Processing
Procedure

Domain
Evaluated

Description

Key output

Domain
volume

Method identifies all groupings of cells


(termed globs) of common phase in
the volume and classifies them as either
continuous liquid (CL), dispersed
liquid (DL), continuous vapor (CV), or
dispersed vapor (DV). Each glob is so
characterized.

Glob count, size,


volume, velocity, and
surface area. Can
compute global
volume fraction and
interfacial area.

Evaluation of
exit plane
data

Exit plane

The velocity and area of each field are


computed at the exit plane. Void
fraction data is averaged in the
circumferential direction to provide
radial profiles.

Evaluation of
turbulence
quantities

Domain
volume

Flow data is averaged and then


collapsed in homogenous directions for
each field.

Classification
of phase
fields

Field mass flow rates


and area fractions.
Radial void profiles
indicate liquid film
profile (base and
wavy components).
Averaged and
fluctuating pressures
and velocities. Wall
shear, interfacial shear
and Reynolds stresses.

8.2.1 Classification of Phase Fields

The simulation data only distinguishes between liquid and steam via the sign of the level
set scalar, .

Utilizing mesh database tools from Simmetrix (2009), a tool was

developed that marches through the domain, identified globs of common phase, and
classified them consistent with a 3-D, two-fluid, four-field computational multiphase
fluid dynamic (CMFD) approach [Lahey (2009)]. This approach distinguished between
disperse and continuous fields of each phase. Hence, there are four fields: continuous
liquid (CL) representing the liquid film on the wall, continuous vapor (CV) representing
the steam core, dispersed liquid (DL) representing the liquid droplets contained within
the steam core, and dispersed vapor (DV) representing the steam bubbles entrained into
the liquid film. A more detailed description of the tool is provided in Appendix A.
Summarizing the information from Appendix A, the classification tool computed the
following information for each phasic glob in the domain at each time step:

70

Number of vertices and regions comprising the glob

Volume of glob, Vglob

Surface area of glob, SAglob

Average velocity vector of glob, uglob

Extremes in each coordinate direction

Field classification (CL, CV, DL, or DV)

The field information was used for all other post-processing steps to distinguish data
associated with disperse and continuous fields.
The volume fraction of each field, j, could be computed based on the volume fraction
of field j to the total domain volume as given by:

(V )
Nj

j =

glob i

i =1

(125)

Vtotal

where Nj is the number of globs of field j and Vtotal is the domain volume.

The

interfacial area density, Ai , of each glob was computed knowing the volume and surface
area as given by:
Ai =

SAglob
V glob

(126)

This value could be averaged over all globs of field j to get the average interfacial area
density for field j, Aij .
8.2.2 Evaluation of Data on Exit Plane
8.2.2.1 Description of Procedure

Analysis of the data on the exit plane was useful in that it is an existing location of
solution nodes contained within a common plane and was much faster than processing
the entire volume domain. Using a glob definition procedure, similar to that described in
Appendix A, the size and velocity of the phase fields were determined. From these
phasic mass flow rates, area-based void fractions, circumferentially-averaged void
fraction profiles, and effective liquid film thickness could be computed.

71

A similar post processing code was used to evaluate the planar data at the exit of the
domain as was used for the classification of phase fields described in Appendix A. The
exception is that the field classification was already known and the globs were areas
rather than volumes. The procedure marched out from an arbitrary vertex and found all
vertices and faces comprising the glob. It was important to know a priori the globs field
classification because it was impossible to know whether a glob was dispersed or
continuous based on just a planar cross section. This point is illustrated in Figure 21
where a liquid glob at the exit plane could be mistaken for a liquid droplet when it was
actually part of a ligament attached to the liquid film and, thus, should have been treated
as part of the continuous liquid film (field CL).
The code identified those faces that were contained completely within the glob and
stored them in a list, faceIn, with the total number being Nface_in. It also found the faces
that crossed the glob boundary and stored them in a list, faceCross, with the total number
being Nface_cross. Interrogating the faces contained within these lists, the glob area and
velocity could be computed. The code also tracked the extremes in the coordinate
directions for each glob where only coordinates from vertices contained within the glob
are considered.

72

Attached to
liquid film

droplet

(a)
3D Contour of steam/water interface
colored by velocity magnitude

(b)
Exit plane colored by phase field

Figure 21: Simulation solution at 63.9113 msec. (a) 3D interface of steam/water


interface colored by velocity magnitude, (b) phase solution at exit plane.
8.2.2.2 Computed Quantities

The area of the glob, Aglob, is defined as the sum of areas of the faces contained within
the glob area on the exit plane and a fraction of the areas of faces that cross the glob
boundary as given by:
Aglob =

(A ) +

N face _ cross

N face _ in
i =1

face i

i =i

N vert _ in
N vert _ face

(A ) ,
face i

(127)

where Nvert_in is the number of vertices in face i that are contained within the glob and
Nvert_face is the total number of face vertices. To compute the average velocity, uglob, of
the glob, an area-weighted average of all face-averaged velocities within the glob is
computed. The average velocity is given by:

(u

N face _ in

u glob =

i =1

face A face )i +

(u

N face _ cross

Aglob

i =i

face

A face )i
,

(128)

where Aglob is the glob area defined in Equation (127), Aface is the area of a given face,
and uface is the average velocity of the face. The mass flow rate of each field, wj, where j
73

is the field (CL, CV, DL, or DV), is computed knowing the area and velocity for all
globs for field j as given by:
w j = j (Aglob u z )i ,
Nj

(129)

i =1

where j is the density of field j, Nj is the number of globs of field j, and uz is the axial
component of velocity of uglob which is perpendicular to the exit plane. The volume
fraction of each field, j, can also be computed based on the area fraction of field j to
total exit plane area as given by:

(A )
Nj

j =

glob i

i =1

(130)

Atotal

where and Atotal is the exit plane area. A circumferential average of the axial velocity,
n

u z , of the boundary layer nodes may be computed as given by:

uz

1
=
Nnodes

Nnodes

u
1=1

n
z

(131)

where n is the layer off the wall and Nnodes is the number of nodes circumferentially
around the tube. A quick estimate of the wall shear may be computed at each layer by
computing the axial velocity gradient in the direction of the wall normal as the
difference between the layers average axial velocity and the wall velocity (which is 0
m/s for the effective wall function) over the distance to the wall, as given by:

u
u z
uz
z wall
r
r
rn
n

wall

) = (0 u ) ,
n

) (R r )
z

(132)

where R is the tube radius and rn is the radius at layer n of the boundary layer structure.
By multiplying Equation (132) by the molecular viscosity, , an estimate of the wall
shear may be computed. However, because the effective wall function uses an enhanced
molecular viscosity at the wall, the enhanced viscosity will be required to compute the
wall shear using the velocity gradient at the wall.

74

8.2.3 Evaluation of Wall Shear Stress

The wall shear stress can be computed using several approaches. One approach, and the
most accurate, is using a consistent method for computing the boundary fluxes using a
finite element method such as outlined in Hughes et al. (1987). This approach reverses
the problem by treating the boundary flux as unknowns and the internal flow solution as
a boundary condition. The boundary fluxes are then solved for using the finite element
method.

Another approach is to sum the wall force computed from the effective

viscosity wall function. This approach has less accuracy than the consistent method but
is fairly close in value, relatively quick to compute, and lends itself well to post
processing. The wall shear stresses reported within used the latter approach due to the
ease of computation. Future work can incorporate the consistent method as an in situ
computation within the flow solver.
The effective viscosity wall function, described in Section 8.1.2.1, computes a friction
velocity, u, to support the off-the-wall parallel velocity. From Equation (108), this
enforces a local wall shear stress, w. Knowing the area to which the local shear stresses
apply, the local shear forces can be found and summed to give the total shear force on
the wall. A post processing code was developed that could spin through any number of
stored result files and compute the wall shear stress at each time step. For each time
step, the flow solution file is read and the wall surface mesh extracted. For each node on
the wall mesh, the closest off-the-wall node is identified. The parallel velocity, u || , of
the off-the-wall node to the wall is computed, as well as the normal distance, dn, to the
wall. The same procedure, described in Section 8.1.2.1, used in PHASTA-IC for the
effective viscosity wall function is used to compute the friction velocity, u. Using
Equation (108), the shear stress, ( w )i , for node i is given by:

( w )i = (u )i 2 .

(133)

A loop over the element faces on the wall is then performed and the average shear stress
over each face j computed as:

w _ avg

1
=
nnp j

nnp j

( )
i =1

75

w i

(134)

where nnpj is the number of vertices on face j. The shear force, (Fw ) j , over face j is then
computed by:

(Fw ) j = ( w _ avg ) j Area j ,

(135)

where Areaj is the area of face j. The total shear force, Fw_tot, over the wall is then
obtained by summing the shear force from each face on the wall, as given by:
Fw _ tot =

nfaces

(F )
w

j =1

(136)

where nfaces is the number faces comprising the wall mesh. Likewise, the total area of
the wall mesh, Aw_tot, is computed by
Aw _ tot =

nfaces

Area
j =1

(137)

The total wall shear stress is given by:

w _ tot =

Fw _ tot
Aw _ tot

(138)

8.2.4 Evaluation of Total Local Shear Stress

The instantaneous local velocity, ui, in the i-direction is composed of a mean


component, u i , and fluctuating component, ui , as given by:
u i = u i + u i .

(139)

When averaging the Navier-Stokes equations, the total shear stress, Tij, is comprised of
the shear stress due to molecular forces, ij, and the Reynolds stress due to turbulent
fluctuations, Rij, as given by:
Tij = ij + Rij .

(140)

The local molecular shear stress, ij, is given by:

ij =

u i
,
x j

(141)

and Rij , the Reynolds stress tensor, defined by:

Rij = u iu j .

76

(142)

Consistent with Pope (2000), the Reynolds stress can be expressed as:
u iu j = u i u j u i u j .

(143)

By averaging the right hand side terms of Equation (143) over a number of increment
time steps, the components of the mean Reynolds stress maybe found. By computing
the following averages at each node over a series of time steps, the local Reynolds stress
tensor can be constructed:

1 nsteps
ui =
ui
nsteps n =1
ui u j =

(144)

1 nsteps
ui u j ,
nsteps n =1

(145)

where nsteps is the number of time steps being averaged. Equations (144) and (145) are
appropriate for a homogenous flow field. When dealing with discrete phase fields (i.e.,
CL, CV, DL, and DV) care must be taken to average the terms separately for each field.
Hence, contributions to the summations in Equations (144) and (145) are made only if
the field is the field being averaged. Considering this, Equations (144) and (145) are
more appropriately written as:
ui =

nsteps

n =1

ui u j =

nsteps

n =1

nsteps

u u

nsteps
n =1

field

(146)

field

n =1

field ,

(147)

field

when gathering averages for a specific field. field is the phase indicator function which
is 1.0 if the field is present or 0.0 otherwise. Including it in Equations (146) and (147)
ensures only solutions of the field of interest contribute to the average. By averaging

field the volume fraction for the field, field, can be obtained as:

field =

1 nsteps
field .
nsteps n =1

(148)

For our simulation, the terms for Equations (146) through (148) were averaged at each
node over a series of time steps. This averaged solution was transferred to a uniform
hexahedron mesh with an ordered node numbering that simplified the spatial averaging
77

of the data in homogenous directions. The hexahedron mesh matched the element size
of the tetrahedron mesh and had the same boundary layer structure near the wall. Radial
profiles of the averaged terms are produced when the averaged data is further averaged
in the homogenous directions (axial and circumferential at a given radial position). By
applying a finite difference stencil in the radial direction, the molecular shear stress, zy,
can be computed as:

zy i +1 =

i +1

i +1

u z
u
uz
,
z i +1
i +1
(y y i )
y
i

(149)

where uz is the axial velocity (in z-direction), y is the coordinate along the radial
direction, i+1 is the radial location of interest and i represents the solution at the
previous radial location.

Using the averaged data at each location i+1, the local

Reynolds stress, Riji+1, is computed from Equations (142) and (143),. The total shear
stress in the radial plane, Tzy, is then computed as:
Tzy = zy + u z u y ,

(150)

where uz is the axial velocity and uy is the radial velocity.


An alternative method of computing the total shear stress is given next.
8.2.5 Evaluation of Average Shear Stress Distribution

The shear stress distributions at each radial position may be computed locally at each
time step and then averaged to compute a mean value at a given location. This option
requires algorithms that compute the local velocity gradient and the local fluctuations so
that total local shear stress can be computed. The values at a given lateral position can
then be averaged to give an overall average at a given radial position, r (including r=R,
the wall). This is a good topic of future research, but is not presented within. Another
option is using a steady, fully-developed force balance for annular flow, as presented by
Lahey and Moody (1993), which neglects surface tension and assumes no spatial and
temporal acceleration (i.e., fully-developed, steady state conditions). Given the idealized
annular flow presented in Figure 22, a force balance on the flow at radial position, r,
yields an expression for the total shear stress (i.e., viscous plus Reynolds), (r), as given
by:

78

r dp g sin ( )

0 l (1 ) + g rdr
2 dz
r
r

(r ) =

(151)

where, neglecting droplets and bubbles, is the total volume fraction of the gas phase
(i.e., = CV + DV). The first term on the right hand side is the total pressure drop
across the domain while the second term represents the elevation head. The lateral
distribution of shear stress can be obtained by averaging the flow across many time
steps. Since the location of the interface is not explicitly known either the mean film
thickness or base film thickness (thickness at which solution is liquid 90 percent of the
time) may be used to obtain the interfacial shear stress. Note that since the void fraction
varies across the radius due to entrainment and the wavy interface, the integral in the
elevation term, Equation (151), must be numerically evaluated. Given the averaged
solution is known at discrete points, i, where the void fraction, i, is valid over the range
of ri- to ri+, the integral term may be analytically evaluated over each discrete region as
follows:
1
[ (1 ) + ]rdr = 2 [ (1 ) + ](r

ri +1 / 2

2
i +1 2

ri 21 2 ) .

(152)

ri 1 / 2

In this way the average shear stress distribution (and the wall and interfacial shear) can
be easily evaluated.

w
i
(r)

r
R

p+(dp/dz)z

Figure 22: Idealized annular flow.

79

8.3 Results and Discussion


Based on the initial mean liquid velocity of 2.64 m/s, the simulation was run almost 11
transport times at the lower friction (only) pressure gradient of 1600 Pa/m and then, to
date, another 5.5 transport times at the experimental pressure gradient of 3300 Pa/m, as
shown in Table 3. The simulation captured the major mechanisms associated with twophase annular flow: interfacial instabilities, ligament formation on the interface with
eventual liquid droplet entrainment, droplet deposition, and steam bubble carry under
into the liquid film. Even with the domain being eight times shorter than the mean
experimental wavelength of disturbance waves (~0.20m for RISO run#602), a large
disturbance wave continually formed in the domain and was the primary source of
droplet entrainment. This is consistent with experimental results [Hewitt and HallTaylor (1970)].
A series of snapshots of the steam/water interface over time steps 105,000 through
125,000 is provided in Figure 23. The contours are colored by the velocity magnitude.
An axial view of the interface from the tube exit is provided in Figure 24 at the same
time steps. After many transport times through the domain, a single large disturbance
wave, highlighted by the dashed oval in Figure 23, existed on the interface. Away from
the interface, small ripples waves also existed; however, on the crest of the large
disturbance wave surface instabilities formed which grew to form ligaments and
eventually sheared off to form droplets. Most of the ligaments deposited back into the
liquid film but some were transported in the steam core and began accelerating towards
the steam cores velocity. Droplets that were too small to be resolved with the 0.0005m
tetrahedron elements eventually vanished due to numerical dissipation. The rest existed
in the steam core until they collided with a breaking wave which typically occurred
within two transport times. It was also along the leading edge of the large wave that
steam bubbles were captured and carried under into the water film by the rolling action
of the waves ahead of the large wave. Fringe plots of the velocity magnitude on the
center XY plane for the same time steps are provided in Figure 25 where a solid black
line indicates the position of the steam/water interface. The steam core velocity reached
velocities around 12 m/s while the liquid film maintained velocities of 3 m/s or less. The
instantaneous location of the large wave is more obvious in the XY plots than in the 3D
80

contour plots in Figure 23, as are the droplets entrained from the crest of the large wave.
Significantly, the predicted phenomena looked very similar to the seminal experimental
observations of Hewitt et al. at AERE Harwell [Hewitt and Hall-Taylor (1970)].
Using the field classification procedure described in Section 8.2.1, the field classification
for each node was computed for each saved time step. This procedure made it much
easier to distinguish between the liquid droplets in the core and the steam bubbles in the
liquid film. It also made it easier to recognize ligaments still attached to the film, as
opposed to liquid droplets, that spanned across periodic boundaries. To illustrate this,
the same 3D images of the steam/water interface in Figure 23 and Figure 24 are provided
in Figure 26 and Figure 27 but colored according to the field classification.

81

82
time step =105,000
time = 118.53 msec

time step = 110,000


time = 126.03 msec

time step = 115,000


time = 133.03 msec

time step = 120,000


time = 141.03 msec

time step = 125,000


time = 148.53 msec

Figure 23: 3D contours of steam/water interface at times in simulation. Contours are colored by velocity magnitude. The
dashed ovals indicate the location of the large wave primarily responsible for droplet entrainment.

83
time step =105,000
time = 118.53 msec

time step = 110,000


time = 126.03 msec

time step = 115,000


time = 133.03 msec

time step = 120,000


time = 141.03 msec

time step = 125,000


time = 148.53 msec

Figure 24: 3D contours of steam/water interface at times in simulation viewed from exit plane. Contours are colored by
velocity magnitude.

84
time step =105,000
time = 118.53 msec

time step = 110,000


time = 126.03 msec

time step = 115,000


time = 133.03 msec

time step = 120,000


time = 141.03 msec

time step = 125,000


time = 148.53 msec

Figure 25: Velocity fringe plots of center xy plane at times in simulation. Black line depicts steam/water interface.

85
time step =105,000
time = 118.53 msec

time step = 110,000


time = 126.03 msec

time step = 115,000


time = 133.03 msec

time step = 120,000


time = 141.03 msec

time step = 125,000


time = 148.53 msec

Figure 26: 3D contours of steam/water interface at times in simulation. Contours are colored by field (blue=steam
bubble, red=liquid droplet, green=steam core/ liquid film interface). The dashed ovals indicate the location of the large
wave primarily responsible for droplet entrainment.

86
time step =105,000
time = 118.53 msec

time step = 110,000


time = 126.03 msec

time step = 115,000


time = 133.03 msec

time step = 120,000


time = 141.03 msec

time step = 125,000


time = 148.53 msec

Figure 27: 3D contours of steam/water interface at times in simulation viewed from exit plane. Contours are colored by field
(blue=steam bubble, red=liquid droplet, green=steam core/liquid film interface).

8.3.1 Void Fraction

The 30 degree wedge model was run using periodic boundary conditions and, therefore,
the volume of each phase was fixed throughout the simulation. This means the global
volume fraction, , and static quality, xs, are constants. The static quality is the mass
fraction of the gas phase to the total mass as defined by:
xs =

mg
m g + ml

(153)

where mg and ml are the mass of the gas and liquid phases, respectively. The flow
quality, x, is defined as the ratio of the mass flow rate of the gas phase to the total mass
flow rate as given by:
x=

wg
w g + wl

g u g Ag
,
g u g Ag + l u l Al

(154)

where w, u, , and A are the mass flow rate, velocity, density, and cross sectional area,
respectively, and the subscripts g and l denote the gas and liquid phases, respectively.
The flow quality and static quality are related [Lahey and Moody (1993)] by:

x
x
=S s ,
1 x
1 xs

(155)

where S is the slip ratio defined as the ratio of the gas velocity to the liquid velocity as
given by:
S = u g ul .

(156)

Although the static quality is fixed, the flow quality may vary as the slip between the
phases may vary. For the initial mean liquid film thickness of 940 m the initial mean
void fraction was 0.820. With periodic boundaries, the simulation should conserve
volume fraction while adjusting mass flux to match the imposed pressure gradient. The
deviation from the initial condition is related to the ability of the level set method to
conserve mass. The global void fraction, , is the sum of the disperse and continuous
steam fields as given by:

= CV + DV ,

87

(157)

where CV and DV are computed from Equation (125). Figure 28 shows the change in
the global void fraction, , during the simulation and the percent error relative to the
initial value of 0.820. The volume fraction fluctuated in the early part of the simulation
but settled on an increasing trend. Mass error can occur either through numerical error
in advancing the level set scalar or numerical dissipation of under-resolved entities like
liquid droplets, liquid ligaments, or steam bubbles. Experience has shown that phasic
entities with less than 10 mesh regions between opposing interfaces are not well resolved
and can suffer numerical dissipation. Consequently, for this simulation, any regions
with curvatures greater then 2000 m-1, such as liquid droplets, steam bubbles, or the
wavy interface between the steam core and liquid film, are prone to mass loss. For this
simulation, the mass loss can be largely attributed to the inability of the mesh to properly
resolve small liquid droplets. As the droplets vanished, the computational cells switched
from liquid to steam phase resulting in an increase in global void fraction. The growth
rate in volume fraction also increased after the imposed axial pressure gradient was
increased to 3300 Pa/m, at t = 103.7 msec, as the number of droplets sheared off into the
steam core also increased. The same affect would occur for under-resolved steam
bubbles in the liquid film which would tend to reduce the global void fraction as steam
phase was lost to liquid phase. However, as discussed in Section 8.3.2, there were
nearly twice as many liquid droplets as steam bubbles so the net affect was an increase
in global void fraction. For this simulation, the error in global void fraction of 1.23
percent was acceptable but could easily have grown to unacceptable levels if the
simulation was much longer without proper resolution of small droplets.
Although the global void fraction remained fairly constant, locally the volume fractions
varied. The local variation at the exit plane was computed but expressed as an area
fraction, Equation (130), rather than a volume fraction. The area fractions for each field
are plotted in Figure 29 over the history of the simulation. The continuous liquid (CL)
and continuous vapor (CV) area fractions oscillated about their initial values of 0.18 and
0.82, respectively, whereas the disperse fields made up less than 0.4 percent of the
global void fraction. The disperse fields are plotted on an expanded scale in Figure 30
along with their average over a 30 msec period, which is more than three times what it
took for the film to transport through the domain based on the initial film velocity.
88

Although the droplets and bubbles could have area fractions upwards of 5 and 2 percent,
respectively, their time averaged area fractions were much smaller at 0.3 and 0.05
percent, respectively, near the end of the simulation.

The selection of a 30msec

averaging window was guided by Figure 31 which suggests that this is the minimum
time to average-out fluctuations in the void fraction.

Global Void Fraction


Time Steps 1 - 130000

Void Fraction,

0.845

Void Fraction

0.840

Error

2.0%

0.835
1.5%

0.830
0.825

1.0%

0.820
0.815

0.5%

0.810
0.805
0.800
0

20

40

60

80

100

120

140

0.0%
160

Time (msec)

Figure 28: Simulation void fraction and associated error.

89

Volume Error (%)

2.5%

0.850

Field Area Fractions @ Exit Plane


1.0
0.9
0.8

Area Fractions

0.7
0.6
Pressure Gradient
increased to 3300 Pa/m

0.5
0.4
0.3
0.2
0.1
0.0
30

40

50

60

70

80

90

100

110

120

130

140

150

Time (msec)
CL

DV

DL

CV

Initial CL

Initial CV

dp/dz increase

Figure 29: Field area fractions at exit plane.

90

160

(a)

DL Liquid Fraction

DL Area Fraction @ Exit Plane


0.05
0.04
0.03
0.02
0.01
0.00
30

40

50

60

70

80

90

100 110 120 130 140 150 160

DV Area Fraction @ Exit Plane

(b)

DV Void Fraction

0.020
0.015
0.010
0.005
0.000
30

40

50

60

70

80

90

100

110

120

130

140

150

160

Average Area Fraction @ Exit Plane


Average over 30 msec

(c)

Area Fractions

0.005
0.004

Pressure Gradient
increased to 3300 Pa/m

0.003
0.002
0.001
0.000
30

40

50

60

70

80

90

100 110 120 130 140 150 160

Time (msec)
DL

DV

dp/dz increase

Figure 30: Disperse Field area fractions at exit plane: (a) dispersed liquid (DL),
(b) dispersed vapor (DV), (c) average over 30 msec.

91

Variation in Averaged Exit Plane Area Fraction


(starting at timestep 95100, time=103.7 msec)
1
Min Area Fraction

Area Fraction

0.95

Max Area Fraction

0.9
0.85
0.8
0.75
0.7
0.65
0.6
0

10

20

30

40

50

60

T, averaging time (msec)

Figure 31: Maximum and minimum void fraction as a function of averaging time.
Oscillations in average void fraction curve are removed for averaging times of 30
msec or greater.

8.3.2 Field Populations

The field classification procedure, discussed in Appendix A, used a volume criterion to


distinguish between disperse and continuous fields. This resulted in only one continuous
glob for each phase; all the remaining globs were classified as dispersed.

It was

assumed that any liquid touching the liquid film was part of the film. This includes
ligaments stretching out into the steam core but still physically attached to the film. In
some instances, a liquid droplet would temporarily contact a portion of the liquid film.
During that time, it would have been considered as part of the film, but become disperse
once the contact was broken.
The population of liquid droplets, NDL, and steam bubbles, NDV, were tallied as part of
the classification procedure. Figure 32 provides the droplet and bubble populations as a
function of simulation time for the 30 degree domain. The droplet count was in the
range of 10 to 20 at the lower pressure gradient and increased to a range of 20 to 40 after
the imposed pressure gradient was increased. This is consistent with the increase in
mass flow rates and interfacial shear that accompany the larger pressure gradient. The

92

increased interfacial shear tends to form larger surface instabilities and promote
increased droplet entrainment. A similar, but less dramatic, trend was seen for the
bubble population curve. The number of bubbles present in the liquid film tended to be
below 10 at any given time at the lower pressure gradient but increased to about 15 at the
higher pressure gradient.
For equilibrium annular flow the population curves would oscillate about a steady value
as the droplet entrainment and deposition mechanisms balance. The population curves
in Figure 32 do not exhibit this behavior so it is reasonable to assume the simulation did

Number of Droplets, NDL

not yet reach an equilibrium state.

Population of Droplets
40

Pressure Gradient
increased to 3300 Pa/m

30
20
10
0
30

50

70

90

110

130

150

130

150

(a)

Number of Bubbles,
N DV

Population of Bubbles
25
Pressure Gradient
increased to 3300 Pa/m

20
15
10
5
0
30

50

70

90

110

Time (msec)

(b)

Figure 32: Population of disperse fields in 30 degree sector during


simulation: (a) liquid droplets (DL), (b) steam bubbles (DV).

93

8.3.3 Field Mass Flow Rates

The experimental mass flow rates of the water and steam were computed using Equation
(106) knowing the total inlet mass flow rate into the test section and the exit quality. For
RISO run #602, the water and steam mass flow rates were 0.110 kg/s and 0.047 kg/s,
respectively. The flow rates for a 30 degree segment are reduced to 0.009 kg/s and
0.004 kg/s for the water and steam, respectively. The RISO experiment suctioned the
liquid flow off the wall at the end of the test section. By varying the suction rate and
performing a heat balance of the suctioned flow, Wurtz (1978) bounded the liquid film
flow rate. It was assumed that the rest of the total liquid mass flow rate was attributed to
the liquid droplets. Table 5 provides the range of mass flow rates from RISO run #602.
The test did not distinguish between disperse and continuous steam fields. Interestingly,
even when using the upper bound on the film flow rate, the flow rate of the droplets
exceeded the flow rate of liquid on the wall. This implies the procedure used to suction
off the liquid film did not conservatively upper bound the liquid film flow rate and the
reported liquid film flow rates should be higher.
The field mass flow rates and velocities from the simulation were computed at the exit
plane according to Equation (129) and Equation (128). The mass flow rates for the
continuous fields are shown in Figure 33 as a function of the simulation time. Included
on the plot are the experimental water and steam flow rates. Figure 34 shows the mass
flow rates for the disperse fields. The mass flow rate of the steam core (CV) remained
steady during the simulation and was slightly below the experimental steam flow rate at
the lower pressure gradient while a little above after the increase in pressure gradient.
The mass flow rate of the liquid film (CL), however, varied with the peaks being as
much as six times greater than the troughs and was generally within the range measured
(Table 5) or higher. The peaks corresponded to times when the large waves passed
through the exit plane. The simulation data must be averaged before being compared to
the experimental data from Table 5 since the experimental data is representative of an
average. The time-averaged field mass flow rates are provided in Figure 35, where the
simulation data was averaged over a 30 msec window. Whereas the other fields seem to
be fluctuating about a mean value, the continuous vapor mass flow rate is steadily
increasing; again, showing the simulation did not yet reach a steady state. The averaged
94

values for the last 30 msec are provided in Table 6. Although a comparison to the
experimental mass flow rates (Table 5) is flawed in that the experimental results are
based on an equilibrium annular flow while the simulation is still developing, the
following conclusions are made at this point in the simulations time:

The simulation predicted a liquid film mass flow rate which was about 12 percent
greater than the experimentally measured upper bound.

With the flow still

adjusting to the imposed pressure gradient, this value would likely get larger.
This is consistent with a potential under-measurement of the liquid film flow rate
from the RISO test. It also is consistent with the fact the simulation domain did
not support the mean wavelength for disturbance waves as measured from the
RISO experiment. The simulation was not able to support the larger disturbance
(i.e., roll) waves that might have produced larger droplet entrainment rates.

The simulation predicted a steam mass flow rate within 14 percent of the
experimental value. This value would likely get higher if the simulation reached
an equilibrium annular flow state.

Assuming the deficit of liquid mass flow rate from the experiment, after
accounting for the film mass flow rate, is related to droplets, the simulation
under-predicted the mass flow rate of droplets by more than an order of
magnitude. This large difference can be attributed to four issues. First, as stated
above, the simulation domain was not able to support the wave length for the
dominant disturbance waves which can lead to an under prediction of droplet
entrainment. Second, the experimental value for droplet entrainment may be too
high due to a potential under-measurement of the experimental liquid film flow
rate. Third, the inability of the mesh to resolve small droplets lead to a decrease
in droplet mass flow rate.

Last, the simulation was still responding to the

imposed pressure gradient, so an equilibrium between droplet entrainment and


deposition had not yet been reached. Hence, the final droplet flow rate was not
yet known.

An estimate on the improvement in liquid droplet mass flow rate may be


provided by simulating the potential under-measurement of experimental liquid
95

film flow rate and assuming all the mass error in the simulation is from underresolved droplets. Assuming the liquid film suction procedure to only be able to
extract the mean film thickness, the portion of the liquid film above of the mean
film thickness of 0.936 mm is classified as DL. This results in an increase in the
droplet mass flow rate, and corresponding decrease in the liquid film flow rate, of
0.00083 kg/s. Also, if the mass error from the simulation, Figure 28, is assumed
to be due to under-resolved droplets vanishing and that mass is added back to the
liquid droplet population as a constant area droplet spanning the length of the
domain, the increase in mass flow rate of DL would be 0.0009 kg/s. The
adjusted values in predicted liquid film (CL) and droplet (DL) mass flow rates
are provided in Table 7 which shows the droplet mass flow rate to increase by a
factor of 12. The liquid film flow rate is now under-predicted by 6 percent and
the droplet flow rate under-predicted by 59 percent.

This level of under-

prediction should decrease once the simulation reaches steady state. However,
the amount is not known at this point.

Looking at the total mass flow rate of the simulation compared to the experiment,
see Figure 36, the simulation under-predicted the total mass flow rate by more
than 25 percent. Accounting for the mass error and assuming the lost mass were
droplets traveling at the average droplet velocity the error reduces to 18 percent.
However, the averaged mass flow rate curve in Figure 36 was increasing,
showing the flow was still accelerating, which would reduce this error once the
simulation reached equilibrium annular flow.

The axial velocity of each field at the exit plane is given in Figure 37 with the averaged
value, over a 30 msec period, provided in Figure 38. Prior to the increase in axial
pressure gradient, the field velocities seem to have reached a steady mean value. With
the increase in pressure gradient to the experimental value the core velocity (CV) was
steadily increasing up to the end of the simulation. This indicates the simulation had not
reached a steady state. One interesting point was the reduction in droplet (DL) velocity
after the increase in pressure gradient.

With the increase, the droplet population

increased with most droplets existing near the film and few escaping toward the center

96

of the tube and attaining higher velocities. This resulted in a decrease in the average
droplet velocity.

Table 5: Experimentally determined mass flow rates through 30 degree section of


tube for RISO run#602.
Field

Minimum Mass Flow


Rate

Maximum Mass
Flow Rate

(kg/s)

(kg/s)

2.93E-03
4.61E-03

4.56E-03
6.23E-03

Water Film (CL) (1)


Droplets in Core (DL) (2)
Steam (CV) (3)

3.93E-03

(1)

The RISO test suctioned off the liquid film on the wall. By varying the suction
rates the researches bounded the film mass flow rate between the minimum and
maximum values listed.
(2)
The droplet mass flow rates are the balance of the liquid flow rate after
subtracting off the film mass flow rate.
(3)
The steam mass flow rate is computed from Equation (106). There was no
range given so only a single value is listed.

Mass Flow Rates @ Exit Plane of Continuous Fields

Mass Flow Rate (kg/s)

1.4E-02

Pressure Gradient
increased to 3300 Pa/m

1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
30

40

50

60

70

80

90

100

110

120

130

140

150

160

Time (msec)
CL

CV

RISO 602 Exp. Liquid

RISO 602 Exp. Steam

dp/dz increase

Figure 33: Mass flow rates of continuous liquid (CL) and continuous vapor (CV)
fields as a function of simulation time.

97

Droplet (DL) Mass Flow Rate


Mass Flow Rate
(kg/s)

5.0E-03

Pressure Gradient
increased to 3300 Pa/m

4.0E-03
3.0E-03
2.0E-03
1.0E-03
0.0E+00
30

40

50

60

70

80

90

100

110

120

130

140

150

160

120

130

140

150

160

Time (msec)

(a)
Bubble (DV) Mass Flow Rate
Mass Flow Rate
(kg/s)

5.0E-05
Pressure Gradient
increased to 3300 Pa/m

4.0E-05
3.0E-05
2.0E-05
1.0E-05
0.0E+00
30

40

50

60

70

80

90

100

110

Time (msec)

(b)

Figure 34: Mass flow rate of disperse fields as a function of simulation


time: (a) liquid droplets (DL), (b) steam bubbles (DV).

98

Average Mass Flow Rates of Continuous Fields


Mass Flow Rate
(kg/s)

1.0E-02
Pressure Gradient
increased to 3300 Pa/m

8.0E-03
6.0E-03
CL

4.0E-03

CV

2.0E-03

dp/dz increase

0.0E+00
30

40

50

60

70

80

90

100

110

120

130

140

150

160

140

150

160

(a)
Average Mass Flow Rates of Disperse Fields
Mass Flow Rate
(kg/s)

1.0E-02

Pressure Gradient
increased to 3300 Pa/m

1.0E-03
1.0E-04

DL
DV

1.0E-05

dp/dz increase
1.0E-06
1.0E-07
30

40

50

60

70

80

90

100

110

120

130

Time (msec)

(b)

Figure 35: Averaged mass flow of fields as a function of simulation time:


(a) continuous fields (CL and CV), (b) disperse fields (DV and DL).
Averages performed over 30 msec.

Table 6: Averaged field mass flow rates and velocities.


Values averaged over the last 30 msec of the simulation.
Average Mass Flow
Rate

Average
Velocity

(kg/s)

(m/s)

CL

5.10E-03

1.466

CV

4.48E-03

5.742

DL

1.58E-04

3.810

DV

5.03E-07

1.805

Field

99

Table 7: Averaged field mass flow rates and velocities


adjusted to simulate numerical equivalent of the RISO test
liquid film suction. Values averaged over the last 30 msec
of the simulation.
Field

Average Mass Flow


Rate

Average
Velocity

(kg/s)

(m/s)

CL

3.87E-03

CV

(1)

1.466

4.48E-03

DL

2.29E-03

DV

5.742

(1)

3.810

5.03E-07

1.805

(1)

All liquid film (CL) flow above the mean film thickness
of 0.936 mm is assumed not to have been suctioned off
and so would be classified as liquid droplets (DL). This
raises the DL flow rate from Table 6 by a factor of 5.23.
Also, the simulation mass error of 1.23 percent is added
back into the liquid droplet (DL) field as a constant area
droplet spanning the length of the domain. This raises the
DL flow rate from Table 6 by an additional factor of 5.75.

Total Mass Flow Rate

Mass Flow Rate (kg/s)

0.018
0.016
0.014
0.012
0.010
0.008
0.006
0.004

Pressure Gradient
increased to 3300 Pa/m

0.002
0.000
30

40

50

60

70

80

90

100

110

120

130

140

150

Time (msec)
Instantaneous

Average over 30 msec

RISO run #602

dp/dz increase

Figure 36: Total mass flow rate as a function of simulation time.

100

160

Field Axial Velocity @ Exit Plane


Pressure Gradient
increased to 3300 Pa/m

8.0E+00

Axial Velocity (m/s)

7.0E+00
6.0E+00
5.0E+00
4.0E+00
3.0E+00
2.0E+00
1.0E+00
0.0E+00
30

40

50

60

70

80

90

100

110

120

130

140

150

Time (msec)
CL

CV

DV

DL

dp/dz increase

Figure 37: Field axial velocity at exit plane as function of simulation time.

101

160

Axial Velocity
(m/s)

Average Axial Velocity of Continuous Fields


8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0

Pressure Gradient
increased to 3300 Pa/m

CL

CV
dp/dz increase
30

40

50

60

70

80

90

100

110

120

130

140

150

160

140

150

160

(a)

Axial Velocity
(m/s)

Average Axial Velocity of Disperse Fields


8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0

Pressure Gradient
increased to 3300 Pa/m

DL
DV
dp/dz increase

30

40

50

60

70

80

90

100

110

120

130

Time (msec)

(b)

Figure 38: Averaged axial velocity as a function of simulation time: (a)


continuous fields (CL and CV), (b) disperse fields (DV and DL). Averages
performed over 30 msec.

8.3.4 Film Thickness

The average thickness of the liquid film on the wall was extracted using the exit plane
data. Assuming a perfectly smooth film of constant thickness, f, the area occupied by
the film, Afilm, is given by

A film =

Ro 2 (Ro f )2
12

],

(158)

where Ro is the tube outer radius and the factor of 12 accounts for the simulation only
modeling 30 degrees of the tube. Solving for the f :
12 A film

f = Ro Ro 2

102

0.5

(159)

Using Equation (159), the instantaneous film thickness at the exit plane was computed
knowing the area of the continuous liquid. This is plotted in Figure 39(b) along with the
average film thickness, over a 30 msec period, which is consistently just below 1 mm.
The mass flow rate of the liquid film is provided in Figure 39(a) to show that the liquid
film mass flow rate was highest when the film thickness was highest. This confirms that
the large waves carried most of the mass through the domain.

Liquid Film Mass Flow Rate @ Exit Plane


Pressure Gradient
increased to 3300

Mass Flow Rate (kg/s)

1.4E-02
1.2E-02
1.0E-02
8.0E-03
6.0E-03
4.0E-03
2.0E-03
0.0E+00
30

40

50

60

70

80

90

100

110

120

130

140

150

160

150

160

Film Thickness (mm)

(a)
Effective Thickness of Liquid Film @ Exit Plane
Pressure Gradient
increased to 3300 Pa/m

2.5
2.0
1.5
1.0
0.5
0.0
30

40

50

60

70

80

90

100

110

120

130

140

Time (msec)
Instantaneous

Averaged over 30 msec

dp/dz increase

(b)

Figure 39: (a) Liquid film (CL) mass flow rate at exit plane as function of
simulation time, (b) Instantaneous and averaged liquid film thickness as
function of simulation time. Averages performed over 30 msec.

Using a uniform grid of 116 equidistant points in the radial direction and 42 equidistant
points in the circumferential direction, the area fraction on the exit plane was
interpolated onto the grid and then averaged in the circumferential direction to produce a
103

radial area fraction profile. These were then averaged over a 15 msec period. Figure 40
through Figure 43 provides these radial profiles of area fraction for the CL, CV, DL, and
DV fields, respectively, at four times at the end of the simulation. The spike in area
fraction for DL at the inner radius of 0.0005 m is an anomaly due to a droplet that
migrated to the centerline and became numerically fixed at that position.
The RISO tests used a needle probe to produce similar radial profiles of void. The radial
locations where the probe was in contact with the liquid film 10, 50, and 90 percent of
the time were recorded.

These were known as the y(10%), y(50%), and y(90%)

positions, respectively. These were computed for the simulation by interpolating the CL
radial area fraction profile. Figure 44 compares the y(10%), y(50%), and y(90%) values
from the experiment and simulation. As can be seen, the simulation does a good job
predicting these once the axial pressure gradient was increased to the experimental value
at 103.7 msec.

CL Area Fraction, CL, at Exit Plane


1.0

CL Area Fraction, CL

0.9
0.8
0.7

Time Step 100000, t=111.08 msec


Time Step 110000, t=126.08 msec
Time Step 120000, t=141.08 msec
Time Step 130000, t=156.08 msec

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.000

0.002

0.004

0.006

0.008

0.010

Radial Coordinate (m)

Figure 40: Continuous Liquid (CL) circumferentially-averaged area fraction at


the exit plane for four times at the end of the simulation. Each curve represents
an average over 15 msec.

104

CV Area Fraction, CV, at Exit Plane


1.0

CV Area Fraction, CV

0.9
0.8
0.7
0.6
0.5
Time Step 100000, t=111.08 msec
Time Step 110000, t=126.08 msec
Time Step 120000, t=141.08 msec
Time Step 130000, t=156.08 msec

0.4
0.3
0.2
0.1
0.0
0.000

0.002

0.004

0.006

0.008

0.010

Radial Coordinate (m)

Figure 41: Continuous Vapor (CV) circumferentially -averaged area fraction at


the exit plane for four times at the end of the simulation. Each curve represents
an average over 15 msec.

DL Area Fraction, DL, at Exit Plane


0.01

DL Area Fraction, DL

0.009
0.008
0.007

Time Step 100000, t=111.08 msec


Time Step 110000, t=126.08 msec
Time Step 120000, t=141.08 msec
Time Step 130000, t=156.08 msec

0.006
0.005
0.004
0.003
0.002
0.001
0
0.000

0.002

0.004

0.006

0.008

0.010

Radial Coordinate (m)

Figure 42: Disperse Liquid (DL) circumferentially -averaged area fraction at the
exit plane for four times at the end of the simulation. Each curve represents an
average over 15 msec.

105

DV Area Fraction, DV, at Exit Plane

DV Area Fraction, DV

0.007
0.006

Time Step 100000, t=111.08 msec


Time Step 110000, t=126.08 msec
Time Step 120000, t=141.08 msec
Time Step 130000, t=156.08 msec

0.005
0.004
0.003
0.002
0.001
0
0.000

0.002

0.004

0.006

0.008

0.010

Radial Coordinate (m)

Figure 43: Disperse Vapor (DV) circumferentially -averaged area fraction at the
exit plane for four times at the end of the simulation. Each curve represents an
average over 15 msec.

Liquid Film Thickness


2500
dp/dz=1600 Pa/m

dp/dz=3300 Pa/m

RISO y(10%)

Thickness (microns)

2000

1500

1000
RISO y(50%)

500
RISO y(90%)

0
70

80

90

100

110

120

130

140

150

160

Simulation Time (msec)


y(90%) - RISO 602
y(90%) - Simulation, 15 msec Avg.
Increased Pressure Gradient

y(50%) - RISO 602


y(50%) - Simulation, 15 msec Avg.

y(10%) - RISO 602


y(10%) - Simulation, 15 msec Avg.

Figure 44: Comparison of predicted y(10%), y(50%), and y(90%) positions to the
experimentally measured values.

106

8.3.5 Dispersed Field Size and Interfacial Area Density

The size, location, and interfacial area of the droplets and bubbles were characterized
using the glob information produced from the field classification tool.

Probability

distribution functions (PDF) for the radial location and equivalent diameter, DEQ, for
droplets and bubbles were produced to give insight into where the disperse fields were
generated and their sizes. Using the computed volume and surface area, the interfacial
area density for the droplets and bubbles were computed and compared to that computed
when making a spherical shape assumption.
For a known glob volume, Vglob, the equivalent spherical diameter, DEQ, is given by
DEQ

6Vglob
=

(160)

The interfacial area density, Ai , is the ratio of the surface area, Aglob, to the volume of a
glob as defined by,
A i =

A glob
Vglob

(161)

For spherical globs, this is simply expressed in terms of the glob diameter, D, defined
by,
A i =

6
D

(162)

A comparison was made of the true interfacial area density, given by Equation (161),
and that often assumed in modeling, Equation (162), when globs (droplet and/or
bubbles) are assumed spherical. For this case, the equivalent diameter, DEQ, given in
Equation (7) was used in Equation (162). Only globs with volumes greater then 6.5E-11
m3 were considered when computing the interfacial area density. Assuming spherical
globs, this volume corresponds to globs that are well resolved by having at least 10
elements across their span.

In addition, the volume error from the four field

classification method (assumed 50% volume of all elements crossing the glob boundary
is part of the glob) was minimized since the glob had more elements completely
contained within than crossing the boundary.

107

8.3.5.1 Liquid Droplets

The equivalent diameter for the droplets ranged from 0.043 mm to 2.18 mm. Probability
distribution functions (PDF) for the droplet radial location and size are provided in
Figure 45(a) and Figure 45(b), respectively. The droplets were concentrated primarily
between radial locations of 6 and 9 mm. This was the region just above the liquid film,
whose average thickness from Figure 39(b) was about one mm. The droplets tended to
have equivalent diameters between 0.1 and 0.5 mm. Unfortunately, this was the region
of diameters where the mesh did not properly resolve the droplet and thus the droplet
was at risk of being lost.
Probability distribution functions for droplet radial location are provided for twelve
ranges of droplet diameters in Figure 46. The smaller diameter droplets (0.5 mm or less)
tended to be more evenly distributed from radial positions of 6 mm to 9 mm. The
population would probably have existed nearer the tube center but they either collided
with larger and/or waves of the film surface or were numerically lost due to being under
resolved. The larger diameter droplets were largely restricted to be near the interface
and usually deposited back into the film. Figure 47 provides probability distribution
functions for droplet size at specific radial locations. No plots were provided for radial
positions less than 4 mm due to the absence of droplets in this region. The smaller
droplets (<0.5 mm) existed in all radial locations greater than 4 mm. The larger droplets
(>7 mm), although small in number, existed in the intermediate radial range of 6 mm to
8.5 mm. This was largely due to the way the larger droplets were created. Often, the
larger droplets were produced from ligaments that pinched off the crest of the large roll
wave. This positioned them within this radial band. These droplets were short lived and
usually deposited back into the film soon after creation.
A plot of the actual droplet interfacial area density from Equation (161) throughout the
simulation is provided in Figure 48 for droplets that are resolved by the mesh (volume >
6.5E-11 m3). The interfacial area computed from Equation (162) assuming a spherical
droplet is also plotted. A plot of the ratio of the actual to spherical interfacial area
densities is provided in Figure 49. Considering a sphere has the largest surface area per
volume, the spherical droplet assumption results in lower interfacial area densities than
actual by an average of about 20 percent.
108

Droplet Location
30%

PDF

25%
20%
15%
10%
5%
0%
0.0

2.5

5.0

7.5

10.0

Radius (mm)

(a)
Droplet Size
30%

PDF

25%
20%
15%
10%
5%
0%
0.0

0.5

1.0

1.5

DEQ (mm)

(b)

Figure 45: Probability Distribution Functions (PDF) for (a) radial


location of droplets and (b) droplet size in terms of equivalent
diameter, DEQ. Data tallied from time steps 110,000 130,000
(time = 126.08 156.08 msec).

109

Droplet Location

Droplet Location

DEQ range: 0.2 to 0.3 mm

DEQ range: 0.3 to 0.4 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%
20%

40%
20%

0%
2.5

5.0

7.5 10.0

0%

0.0

5.0

7.5 10.0

0.0

2.5

5.0

7.5 10.0

Radius (mm)

Radius (mm)

Droplet Location

Droplet Location

Droplet Location

DEQ range: 0.5 to 0.6 mm

DEQ range: 0.6 to 0.7 mm


100%

80%

80%

80%

60%

60%

60%

40%

40%

20%

20%

0%

0%

2.5

5.0

PDF

100%

PDF

100%

0.0

7.5 10.0

40%
20%
0%

0.0

Radius (mm)

2.5

5.0

7.5 10.0

0.0

Radius (mm)

2.5

5.0

7.5 10.0

Radius (mm)

Droplet Location

Droplet Location

Droplet Location

DEQ range: 0.7 to 0.8 mm

DEQ range: 0.8 to 0.9 mm

DEQ range: 0.9 to 1.0 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

40%

20%

20%

0%

0%

0.0

2.5

5.0

PDF

100%

PDF

PDF

2.5

Radius (mm)

DEQ range: 0.4 to 0.5 mm

7.5 10.0

40%
20%
0%

0.0

Radius (mm)

2.5

5.0

7.5 10.0

0.0

Radius (mm)

2.5

5.0

7.5 10.0

Radius (mm)

Droplet Location

Droplet Location

Droplet Location

DEQ range: 1.0 to 1.1 mm

DEQ range: 1.1 to 1.2 mm

DEQ range: 1.2 to 1.3 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

40%

20%

20%

0%

0%

0.0

2.5

5.0

7.5 10.0

Radius (mm)

PDF

100%

PDF

PDF

40%
20%

0%

0.0

PDF

PDF

100%

PDF

PDF

Droplet Location
DEQ range: 0.1 to 0.2 mm

40%
20%
0%

0.0

2.5

5.0

7.5 10.0

Radius (mm)

0.0

2.5

5.0

7.5 10.0

Radius (mm)

Figure 46: Probability Distribution Functions (PDF) for radial location of


droplets in steam core for 12 ranges of droplet equivalent diameter, DEQ. The
total probability for each sums to 100 percent. Data tallied from time steps
110,000 130,000 (time = 126.08 156.08 msec).

110

Droplet Size

Droplet Size

R = 4.5 to 5.0 mm

R = 5.0 to 5.5 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

40%

20%

20%

0%

0%

0.0

0.5

1.0

PDF

100%

PDF

PDF

Droplet Size
R = 4.0 to 4.5 mm

1.5

20%
0%

0.0

1.0

1.5

0.0

0.5

1.0

Droplet Size

Droplet Size

Droplet Size

R = 6.0 to 6.5 mm

R = 6.5 to 7.0 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

20%

20%

0%

0%

0.5

1.0

PDF

100%

40%

1.5

40%
20%
0%

0.0

DEQ (mm)

0.5

1.0

1.5

0.0

DEQ (mm)

0.5

1.0

Droplet Size

Droplet Size

Droplet Size

R = 7.5 to 8.0 mm

R = 8.0 to 8.5 mm
100%

80%

80%

60%

60%

60%

40%
20%

0%

0%

0.0

0.5

1.0

PDF

100%

80%

PDF

100%

20%
1.5

40%
20%
0%

0.0

DEQ (mm)

0.5

1.0

1.5

0.0

DEQ (mm)

0.5

1.0

Droplet Size

Droplet Size

Droplet Size

R = 9.0 to 9.5 mm

R = 9.5 to 10.0 mm
100%

80%

80%

60%

60%

60%

PDF

100%

80%

PDF

100%

20%

40%
20%

0%
0.5

1.0

DEQ (mm)

1.5

40%
20%

0%

0.0

1.5

DEQ (mm)

R = 8.5 to 9.0 mm

40%

1.5

DEQ (mm)

R = 7.0 to 7.5 mm

40%

1.5

DEQ (mm)

R = 5.5 to 6.0 mm

0.0

PDF

0.5

DEQ (mm)

PDF

PDF

DEQ (mm)

PDF

40%

0%

0.0

0.5

1.0

DEQ (mm)

1.5

0.0

0.5

1.0

1.5

DEQ (mm)

Figure 47: Probability Distribution Functions (PDF) for droplet size for 12 ranges
of radial locations. The total probability for each sums to 100 percent. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec).

111

(m /m )

i
Ai'''
Interfacial Area Density, A

DL Interfacial Area Density


22000
20000
18000
16000
14000
12000
10000
8000
6000
4000
2000
100

110

120

130

140

150

160

Time (msec)
Simulation: Ai'''
Ai = Area/Vol

Spherical Assumption: Ai'''


Ai = 6/Dia

Figure 48: Droplet Interfacial Area Density as a function of simulation history.


Simulation value ( Ai =surface area/volume) are plotted with values assuming a
spherical droplet ( Ai =6/DEQ) where DEQ is the droplet equivalent diameter.

DL Interfacial Area Density Error


Ai'''actual / Ai'''spherical_assumption

2.0
1.9
1.8
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1.0
100

110

120

130

140

150

160

Time (msec)

Figure 49: Ratio of actual droplet interfacial area density to that assuming a
spherical droplet. Ratio is plotted against the simulation time.

112

8.3.5.2 Steam Bubbles

The equivalent diameter for the bubbles ranged from 0.02 mm to 0.88 mm (see Figure
50). Probability distribution functions (PDF) for the bubble radial location and size are
provided in Figure 50(a) and Figure 50(b), respectively. The primary method for bubble
generation was the carry-under of steam by the collapsing of the waves just ahead of the
large roll wave. The bubbles existed primarily in the base film but could also exist
within the large roll wave. 93 percent of the bubbles had equivalent diameters of 0.5
mm or less. As with the droplets, this is the size range that was not well resolved by the
mesh and, as such, the bubbles would likely vanish if they did not collide with the film
interface first.
Probability distribution functions for bubble radial location are provided for nine ranges
of bubble diameters in Figure 51. The smaller diameter bubbles (0.5 mm or less) could
exist anywhere within the liquid film, including within the large roll wave. The large
bubbles, however, remained within the base film and would eventually vanish by
pushing through the film/core interface. Figure 52 provides probability distribution
functions for bubble size at specific radial locations. No plots were provided for radial
positions less than 5.5 mm due to the absence of bubbles in this region. The PDFs show
that only small bubbles (< 0.4 mm) existed at radial locations outside the base film
thickness and larger bubbles only existed within the base film.
A plot of the actual bubble interfacial area density from Equation (161) throughout the
simulation is provided in Figure 53 for bubbles that are resolved by the mesh (.i.e.,
bubble volume > 6.5E-11 m3). The interfacial area computed from Equation (162)
assuming a spherical bubble is also plotted. A plot of the ratio of the actual to spherical
interfacial area densities is provided in Figure 54. The spherical droplet assumption
results in lower interfacial area densities than actual by an average of about 10 percent.

113

Bubble Location
100%

PDF

80%
60%
40%
20%
0%
0

10

Radius (mm)

(a)

Bubble Size
100%

PDF

80%
60%
40%
20%
0%
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

DEQ (mm)

(b)

Figure 50: Probability Distribution Functions (PDF) for (a)


radial location of bubbles and (b) bubble size in terms of
equivalent diameter, DEQ. Data tallied from time steps 110,000
130,000 (time = 126.08 156.08 msec).

114

Bubble Location
DEQ range: 0.2 to 0.3 mm

100%

100%

100%

80%

80%

80%

60%

60%

60%

40%
20%

PDF

Bubble Location
DEQ range: 0.1 to 0.2 mm

PDF

PDF

Bubble Location
DEQ range: 0.0 to 0.1 mm

40%
20%

0%

20%

0%

10

0%

Radius (mm)

10

Bubble Location

DEQ range: 0.4 to 0.5 mm

DEQ range: 0.5 to 0.6 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

20%

20%

0%

0%

PDF

DEQ range: 0.3 to 0.4 mm


100%

40%

10

40%
20%
0%

Radius (mm)

10

Radius (mm)

Bubble Location
DEQ range: 0.8 to 0.9 mm

100%

100%

100%

80%

80%

80%

60%

60%

60%

40%
20%

0%

0%

Radius (mm)

10

PDF

Bubble Location
DEQ range: 0.7 to 0.8 mm

PDF

Bubble Location

20%

10

Radius (mm)

DEQ range: 0.6 to 0.7 mm

40%

10

Radius (mm)

Bubble Location

PDF

PDF

Radius (mm)

Bubble Location

PDF

40%

40%
20%
0%

Radius (mm)

10

10

Radius (mm)

Figure 51: Probability Distribution Functions (PDF) for radial location of bubbles
in liquid film for 9 ranges of bubble equivalent diameter, DEQ. The total
probability for each sums to 100 percent. Data tallied from time steps 110,000
130,000 (time = 126.08 156.08 msec).

115

Bubble Size

Bubble Size

R = 6.0 to 6.5 mm

R = 6.5 to 7.0 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%
20%

40%
20%

0%

0%

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

DEQ (mm)

DEQ (mm)

DEQ (mm)

Bubble Size

Bubble Size

Bubble Size

R = 7.0 to 7.5 mm

R = 7.5 to 8.0 mm

R = 8.0 to 8.5 mm
100%

80%

80%

60%

60%

60%

40%

PDF

100%

80%

PDF

100%

40%

20%

20%

0%

0%

40%
20%
0%

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

DEQ (mm)

DEQ (mm)

DEQ (mm)

Bubble Size

Bubble Size

Bubble Size

R = 8.5 to 9.0 mm

R = 9.0 to 9.5 mm

R = 9.5 to 10.0 mm

100%

100%

80%

80%

80%

60%

60%

60%

40%

PDF

100%

PDF

PDF

40%
20%

0%

PDF

PDF

100%

PDF

PDF

Bubble Size
R = 5.5 to 6.0 mm

40%

20%

20%

0%

0%

40%
20%
0%

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

0.0 0.2 0.4 0.6 0.8 1.0

DEQ (mm)

DEQ (mm)

DEQ (mm)

Figure 52: Probability Distribution Functions (PDF) for bubble size for 9 ranges
of radial locations. The total probability for each sums to 100 percent. Data
tallied from time steps 110,000 130,000 (time = 126.08 156.08 msec).

116

14000
12000
10000
2

(m /m )

i
Interfacial Area Density, AAi'''

DV Interfacial Area Density

8000
6000
4000
2000
100

110

120

130

140

150

160

Time (msec)

Ai = 6/Dia
Spherical Assumption: Ai'''

i = Area/Vol
Simulation: A
Ai'''

Figure 53: Bubble Interfacial Area Density as a function of simulation history.


Simulation value ( Ai =surface area/volume) are plotted with values assuming a
spherical bubble ( Ai =6/DEQ) where DEQ is the bubble equivalent diameter.

DV Interfacial Area Density Error


Ai'''actual / Ai'''spherical_assumption

2.0
1.9
1.8
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1.0
100

110

120

130

140

150

160

Time (msec)

Figure 54: Ratio of actual bubble interfacial area density to that assuming a
spherical bubble. Ratio is plotted against the simulation time.

117

8.3.6 Wall Shear Stress

The wall shear stress was computed using the procedures outlined in Sections 8.2.3 and
8.2.5. By summing, over all mesh faces on the wall, the wall force computed from the
effective viscosity wall function and summing the face areas, Equation (138) was used to
compute the wall shear. This procedure was performed at each time step during the
simulation. The wall shear was also computed using Equation (151), for r=R (i.e., the
wall), which is based on a force balance for steady, fully-developed annular flow. As
described in more detail in Section 8.3.7, it was computed at two times during the
simulation using data averaged over 30 msec. A plot of the wall shear stress as a
function of the simulation time is given in Figure 55 along with the experimental shear
stress computed from Equation (107) which has a value of 8 Pa. After the initial rise
during the simulation startup, the wall shear began a steady decline. The reason for this
was the experimental frictional pressure gradient of 1600 Pa/m was incorrectly applied
as the body force instead of the total experimental pressure gradient of 3300 Pa/m. A
pressure gradient of 1600 Pa/m results in a 40 Pa pressure drop across the simulation
domain of length 0.025m. For about a 0.82 void fraction steam/water flow, the elevation

head, given by p elev = gz ( l (1 ) + g ) , was somewhat less than 40 Pa. For the
case where the pressure gradient is less than the elevation head, the wall shear for steady,
fully-developed annular flow (Equation (151)) is negative [Lahey and Moody (1993)].
This is seen in Figure 55 where the wall shear stress for the fully-developed assumption
is -7.9 Pa. The difference between the actual wall shear computed from the effective
viscosity wall function and the fully-developed assumption are the non-zero spatial and
temporal acceleration terms that are neglected in the steady, fully-developed model. The
imposed axial pressure gradient was corrected at t = 103.7 msec, as marked by the
vertical dashed line, but it took almost 40 msec before the wall force began to increase.
Clearly the flow did not reach a statistically stationary state as the flow was still reacting
to the imposed pressure gradient; however, the agreement in computed wall shear at the
end of the simulation between the effective viscosity wall function and the steady, fullydeveloped annular flow model had improved.

118

Wall Shear Stress (Pa)

Wall Shear Stress


8
6
4
2
0
-2
-4
-6
-8
30

50

70

90

110

130

150

Time (msec)
Simulation, effective viscosity wall function

Simulation, Fully-Developed Assumption

RISO Run #602

Increased dp/dz

Figure 55: Computed wall stress compared to experimental shear stress from RISO
run #602. Value from effective viscosity wall function computed from Equation
(138). Values for fully-developed assumption computed from Equation (151) using
data averaged over 30 msec.

8.3.7 Total Shear Stress & Interfacial Shear Stress

The total shear stress in each field is computed as described in Sections 8.2.4 and 8.2.5.
The flow field was averaged over time steps 65,000 to 95,000 (time = 58.58 to 103.58
msec), where the imposed axial pressure gradient was -1600 Pa/m, and 100,000 to
130,000 (time = 111.08 to 156.08 msec) where the imposed axial pressure gradient was 3300 Pa/m. For each time range, the time-averaged data was averaged in the azimuthal
and axial directions, producing radial profiles of the averaged quantities. The radial
profile of the mean axial velocity, u z , was used to compute the molecular shear stress,
zy, using Equation (149). The averaged values of the instantaneous radial velocity, axial
velocity, and their product were used to compute the Reynolds stress, Rzy, using
Equations (142) and (143). From Equation (150), the total local shear stress parallel to
the wall, Tzy, was computed for the continuous liquid (CL) and continuous vapor (CV)
fields. The average shear stress distribution, assuming steady, fully-developed annular

119

flow, was computed using Equation (151) and the average radial void fraction profile
resulting from these time-averaged solutions.
Figure 56 shows the total local shear stress for the CL and CV fields across the tube
radius along with the CV volume fraction for data averaged over time steps 65,000
95,000 (pressure gradient = -1600 Pa/m). The total local shear stress is only shown for
the CV field for CV>0.50 and for the CL field for CV<0.50. Beyond these constraints,
the data sampling is actually too small to gather reliable statistics. Also shown in Figure
56 is the average radial shear stress distribution computed assuming the steady, fullydeveloped annular flow model. The relatively good agreement between the total local
shear stress for the CV field and the steady, fully-developed average shear stress model
show the steam had adjusted to the low pressure gradient.

In contrast, the poor

agreement between for the CL field indicates the high inertia liquid film was still
adjusting from the initial condition.

For this pressure gradient (-1600 Pa/m), the

imposed pressure gradient is somewhat less than the elevation head. Hence, the nearwall average wall shear in the liquid film becomes negative, as indicated in Figure 56.
Neglecting the droplets and bubbles, the total local shear stress for the two-phase
mixture, TzyM , can be computed as:

TzyM TzyCL (1 ) + TzyCV ,

(163)

where TzyCL and TzyCV are the total local shear stress for the CL and CV fields,
respectively. Figure 57 plots the total shear stress for the two-phase mixture, using
Equations (150) and (163), and the total average shear stress distribution across the tube
radius, using Equation (151), for the lower pressure gradient. As with Figure 56, the
good agreement between the curves in the steam core and poor agreement in the liquid
film show the steam core had adjusted to the lower pressure gradient but the liquid film
was still developing.
Similarly, Figure 58 shows the total local shear stress for the CL and CV fields across
the tube radius along with the CV volume fraction for data averaged over time steps
100,000 130,000 (pressure gradient = -3300 Pa/m). As before, the total local shear
stress is only shown for the CV field for CV>0.50 and for the CL field for CV<0.50.
Beyond these constraints, the data sampling is considered to be too small to gather

120

reliable statistics. The agreement between the total local CV shear stress and the steady,
fully-developed average shear stress is not as good for the lower pressure gradient,
indicating the steam core is still responding to the higher imposed pressure gradient.
Even though the flow was still developing, the total shear stress had a linear shape across
the steam core (CV > 0.90) but not near the core/film interface where the influence of
the interfacial waves and droplet entrainment was seen. The liquid film (CL) agrees
fairly well with the average shear stress curve; this is likely due to the fact that it had not
deviated much from the experimentally determined initial condition, due to its high
inertia, during the earlier part of the simulation at the lower pressure gradient. The shape
is somewhat linear for most of the profile except near the wall where there is a sharp
jump in total shear stress. This is a result of both the flow still adjusting to the pressure
gradients and the effective viscosity wall function that enforces and enhanced viscosity
at the wall which can distort the true axial velocity gradient. The shape would be
expected to be linear in the base film region where CL > 0.80, which, for this
simulation, represented a 0.350 mm thickness at the wall that was resolved by 10 mesh
points. Unfortunately, the low mesh resolution leads to reduced accuracy. Finally,
Figure 59 shows the mixture total local shear stress curve and the total steady, fullydeveloped average shear stress profile. As noted before, the flow is obviously still
responding to the new imposed axial pressure gradient.
The interfacial shear was computed by interpolating the total local shear stress and
average steady, fully-developed shear stress curves, for the time ranges, at the radial
position indicative of the mean film thickness. From Figure 39, we see that the mean
liquid film thickness over the last 30,000 time steps is about 0.936 mm, relating to a
radius of r = 0.00906m. Table 8 shows the interpolated shear stress values from the
local and average steady, fully-developed shear stress curves in Figure 57 and Figure 59.
Again, it is clear that the flow is still developing.

121

CV Total Shear Stress Across Radius


Data averaged over time steps 65,000 - 95,000
1.0

0.8

2
0

0.6

-2
0.4

-4

Volume Fraction

Total Shear Stress, Tzy


(Pa)

0.2

-6
-8
0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.0
0.010

Radial Position, (m)


CL Local Shear

CV Local Shear

Average Shear Stress

CV volume fraction

Figure 56: Total shear stress plotted against radial position for continuous liquid
(blue) and continuous vapor (red) fields over the time steps 65,000 95,000 where
the axial pressure gradient was -1600 Pa/m. Each are plotted up to the radial
position of the 50 percent volume fraction. The average total shear stress (green
dashed) for steady, fully-developed annular flow is also shown along with the steam
volume fraction (black dotted).

122

CV Total Shear Stress Across Radius


Data averaged over time steps 65,000 - 95,000
1.0

0.8

2
0

0.6

-2
0.4
-4
0.2

-6
-8
0.000

Volume Fraction

Total Shear Stress, Tzy (Pa)

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.0
0.010

Radial Position, (m)


Total Local Shear

Average Shear Stress

CV volume fraction

Figure 57: Total shear stress plotted against radial position for flow over the time
steps 65,000 95,000 where the axial pressure gradient was -1600 Pa/m. The
average total shear stress (green dashed) for steady, fully-developed annular flow is
also shown along with the steam volume fraction (black dotted).

123

CV Total Shear Stress Across Radius


Data averaged over time steps 100,000-130,000
1.0

12

0.8

8
0.6
6
0.4
4

Volume Fraction

Total Shear Stress, Tzy


(Pa)

10

0.2

2
0
0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.0
0.010

Radial Position, (m)


CL Local Shear

CV Local Shear

Average Shear Stress

CV volume fraction

Figure 58: Total shear stress plotted against radial position for continuous liquid
(blue) and continuous vapor (red) fields over the time steps 100,000 130,000
where the axial pressure gradient was -3300 Pa/m. Each are plotted up to the
radial position of the 50 percent volume fraction. The average total shear stress
(green dashed) for steady, fully-developed annular flow is also shown along with the
steam volume fraction (black dotted).

124

CV Total Shear Stress Across Radius


Data averaged over time steps 100,000-130,000
1.0

12

0.8

8
0.6
6
0.4
4

Volume Fraction

Total Shear Stress, Tzy


(Pa)

10

0.2

2
0
0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.0
0.010

Radial Position, (m)


Total Local Shear

Average Shear Stress

CV volume fraction

Figure 59: Total shear stress plotted against radial position for flow over the time
steps 100,000 130,000 where the axial pressure gradient was -3300 Pa/m. The
average total shear stress (green dashed) for steady, fully-developed annular flow is
also shown along with the steam volume fraction (black dotted).

Table 8: Averaged field mass flow rates and velocities adjusted to


simulate numerical equivalent of the RISO test liquid film suction.
Values averaged over the last 30 msec of the simulation.
Interfacial Shear Stress
Mixture
Average
Shear Stress Local Shear
Stress
Profile

Time Range

Axial Pressure
Gradient

(msec)

(Pa/m)

(Pa)

(Pa)

58.58 to 103.58
111.08 to 156.08

-1600
-3300

3.06
9.97

4.38
4.84

125

8.3.8 Film Thickness vs. Wall Shear

An interesting trend seen with the air/water experiments of Gottmann (1997) is that the
wall shear trends with the liquid film thickness.

To make this comparison, the

instantaneous wall shear must be computed rather than one averaged over the domain.
This is done using Equation (132) that is based on data averaged circumferentially
around the tube but not axially averaged. Figure 60 shows a plot of the instantaneous
gradient of the axial velocity with respect to wall normal at the exit plane, based on the
velocity in the fourth layer of the boundary layer structure, along with a plot of the
effective film thickness at the exit plane computed from Equation (159). Although the
velocity gradient was still developing, as the flow was still adjusting to the imposed
experimental pressure gradient, it fluctuated in a similar manner to that of the film
thickness curve. This clearly shows the importance of the disturbance waves on wall

Axial Velocity
Gradient
(m/s/m)

shear.

Near-Wall Axial Velocity Gradient @ Exit


Plane
-10000
-15000
-20000
-25000
-30000
-35000
-40000
100

110

120

130

140

150

160

(a)
Axial Velocity
Gradient (m/s/m)

Effective Film Thickness @ Exit Plane


3.0
2.5
2.0
1.5
1.0
0.5
0.0
100

110

120

130

140

150

160

Time (msec)

(b)

Figure 60: Illustration of how wall shear trends with film thickness @ exit plane:
(a) gradient of axial velocity w.r.t wall normal near the wall, (b) effective film
thickness.

126

9. CONCLUSIONS
Closure laws based on fundamental physics are required for accurate predictions of twophase annular flows using two-fluid CMFD codes. However, obtaining the local data
required to formulate these laws is difficult so Direct Numerical Simulation (DNS) has
been investigated as a tool to help us better understand the physical mechanisms in
annular flow and to generate numerical data to support the development of physicallybased closure laws. Due to the computational cost associated with accurately resolving
annular flows, no comprehensive simulations have been performed to date that have
attempted to capture the various mechanisms associated with annular flow: interfacial
dynamics and instabilities, liquid ligament creation with eventual droplet entrainment
into the vapor core, bubble carry-under into the liquid film, and the existence of large
disturbance waves. Nevertheless, with the increase in modern computational resources,
the analysis of problems of this nature are becoming achievable within a reasonable time
frame. The research discussed herein has successfully demonstrated this by simulating a
two-phase annular flow, using a three-dimensional, stabilized finite element code,
PHASTA-IC, with an implemented level set method, to accurately capture the interface
and by developing tools and methods for processing these data.
The simulation, based on an experimental steam/water annular flow condition, computed
the flow in the liquid film and steam core and, to date, has tracked the interface for more
than 16 transport times through the computational domain. Although there are areas in
which the simulation can be further improved, the simulation was able to capture the
major physical mechanisms seen in annular flow. In particular, it was able to capture
instabilities on the steam/water interface that eventually formed liquid ligaments that
stretched into the steam core and were sheared off to form liquid droplets. It was able to
capture the carry-under of steam bubbles into the liquid film due to the action of waves
just ahead of the large roll wave. It was also able to support the formation of large
disturbance waves, which performed in a similar manner to how roll waves are known to
behave.

Moreover, these large waves were primarily responsible for the droplet

entrainment and bubble carry-under. This is important since roll waves are known to
influence most aspects of annular flow, such as droplet entrainment, wall shear, and

127

liquid film formation around the conduit perimeter. The simulation also showed a strong
correlation between wall shear and liquid film height that has been seen experimentally.
Since the simulation was not yet statistically steady, the agreement between the
simulation and experimental results varied. The use of periodic boundaries effectively
made the global volume constant. Hence, the good agreement with the global void
fraction is not surprising, and the low mass error means the level set method performed
well. The simulation did a good job of matching the liquid film thickness and wave
heights measured from the experiment. It was also able to predict the mass flow rates of
the vapor core and liquid film within 15 percent, but grossly under-predicted the liquid
droplet flow rate.

Although there are good reasons to doubt the accuracy of the

experimental value of the droplet flow rate, there are known issues with the simulation
that support a low prediction (e.g., the loss of small liquid droplets due to underresolution). Not surprisingly, with the flow still developing, the predicted wall shear
was less than half the experimental value but was increasing toward the experimental
value.
Perhaps one of the biggest benefits of this research was the development of the tools and
methods required to process the large quantities of numerical data generated.

classification tool was developed that walked through the domain, at each time step,
and classified all the globs of a common field and phase. Although the classification
was based on the four-field approach in this research (continuous liquid (CL),
continuous vapor (CV), dispersed liquid (DL), or dispersed vapor (DV)), the tool is
general enough to support arbitrary classification schemes. This enabled the distinction
between disperse and continuous phases when computing various quantities such a mass
flow rate, velocity, void fraction, and interfacial area density.

Methods were also

developed to average the data to compute quantities based on profiles averaged in the
axial and azimuthal directions. These include quantities such as Reynolds stress, film
thickness, and shear stress. It was important to distinguish between the various fields
when performing these averages.
This research has been beneficial in that is has demonstrated that a detailed numerical
simulation of two-phase annular flow is possible and has developed tools and methods
for analyzing these data. However, several things can be done to improve the accuracy

128

of the results. First, the flow is still developing so the simulation will require more time
to reach an equilibrium annular flow state. Agreement with the experimental data
should not be expected until this is accomplished. Assuming the current rate of increase
in the wall shear stress computed from the simulation, Figure 61 shows the simulation
will match the experimental wall shear stress in approximately another 125 msec,
requiring another 83,000 time steps based on the current simulation time step of 1.5E-06
seconds. It should be noted that the required time steps may be reduced if the liquid film
is accelerated by placing an additional body force on the liquid phase.

Once the

computed wall shear stress is close to the experimental wall shear stress, the additional
body force may be removed and the flow allowed to adjust to the experimental pressure
gradient. Second, the loss of under-refined droplets and the under-resolution of the
liquid film can be improved via adaptive mesh refinement. Of particular interest is
whether the mean size of the liquid droplets will change as the mesh around the interface
is refined, as seen by Fuster et al. (2009). At some point, the refinement level must be
set and the level of mass error assessed. In any event a method of re-introducing the
mass of lost droplets (i.e., due to under-resolution) must be developed to support
practical problems or a hybrid front capturing method employed such as a coupled
VOF/level set approach that is inherently mass conservative [Sussman and Puckett
(2000), Sussman (2003), Son (2005), and van der Pijl et al. (2005)]. The VOF/level set
approach was cited as being critical to the resolution of small droplets in the atomization
simulation of Lebas et al. (2009). Thirdly, the reduced length of the current simulation
does not support the mean wavelength of the experimentally measured disturbance
waves. A domain at least the length of the mean disturbance wave is required. For the
current simulation, it means a factor of eight increase in domain length.
additional data evaluations should be performed.

Finally,

After a steady-state condition is

reached, the data should be assessed against existing annular flow closure models and, if
the models do not predict the data, new fundamental models developed (however, this
will require more runs at different conditions). For a properly resolved liquid film, a
comparison can be made of the turbulence profile and wall shear to that for single-phase
turbulent flow, and the effect of the interfacial waves quantified.

129

To increase the credibility of the DNS data, additional simulations and investigations
need to be conducted. First, a study should be performed that quantifies the effect of
curvature on mass error so the limitations of the numerical interface tracking method
used within are better understood. Curvature seems to be a good metric to use as it is
relevant to liquid droplets, steam bubbles, and the liquid film/steam core interface. It
can also be used as the criteria during adaptive mesh refinement such that only regions
where curvature is above a threshold value are refined. Second, the validity of using a
turbulent wall function with a tetrahedron mesh should be further investigated. This can
be accomplished by comparing numerical simulations of single phase flow, using a wall
function, to DNS data [Kim et al. (1987)]. Sensitivity of the number of layers in the
boundary layer mesh and the element sizes to matching the DNS data can be
investigated.

Wall Shear Stress


Wall Shear Stress (Pa)

9
8
7
6
5
Simulation
RISO Run #602
Increased dp/dz

4
3
2
30

80

130

180

230

280

Time (msec)

Figure 61: Computed wall shear stress compared to the experimental


wall shear stress for RISO run #602. The arrow shows the projected
path of the simulation assuming a linear increase in computed wall
shear with simulation time.

130

10. RECOMMENDATIONS FOR FUTURE WORK


The ability of computational multiphase fluid dynamics (CMFD) codes to predict a wide
variety of bubbly flows using fundamental closure models was demonstrated by Lahey
and Drew (2001). Here, analytical forms of the closure models were possible using drag
law and potential flow theory. Unfortunately, for the annular flow regime, the closure
models are not easily developed. Functional forms may be constructed via simplified
analysis, but detailed experimental or DNS data is required to develop meaningful
models of the interfacial force densities, M and Mw in Equation (5), that can be applied
over a wide range of parameters [Lahey (2005)]. Specifically, models for the various
interfacial force densities, which may be partitioned into drag and non-drag terms, can
be formulated using the instantaneous conditions around the interfaces to formulate the
required closure laws for CMFD models. This may involve finding the values or
functional dependence of coefficients in the individual closure laws or, in some cases,
modification of the functional model to better represent the local data.

Ideally,

numerical simulations of simplified problems that isolate specific closure models (e.g.,
flow through a convergent nozzle will isolate the virtual mass force) may be performed
to simplify closure model development. Closure models developed in this manner (i.e.,
which capture the relevant physical mechanisms of the interfacial forces) can then be
applied to general problems that may have complex geometries and may extend well
beyond the parameter ranges of the data from which the models were developed.
This thesis demonstrated that DNS of two-phase annular flow is achievable and provided
methods of extracting meaningful data; however, this is one step in a larger process of
developing CMFD codes capable of predicting the two-phase flows for all flow regimes.
Once DNS has been shown to accurately resolve complex two-phase flows, DNS may be
used to develop the interfacial closure models. Ideally, simple geometry problems that
isolate specific interfacial forces can be simulated to reduce the effect of other forces on
the data. However, as in experimental testing, multiple DNS runs should be performed
to improve the generality of the closure model to parameters like Reynolds number, fluid
properties, and geometry. Once developed, these CMFD closure models can be assessed
against data to determine their validity.

131

Using the DNS strategy presented herein, additional problems may be simulated to gain
understanding of physical mechanisms in two-phase annular flow. In some instances,
these problems may be much smaller than the DNS presented in this thesis.

For

example, the levitation force that spreads the liquid film around a horizontal conduit may
be investigated on a relatively coarse mesh as done by Fukano and Inatomi (2003).
Multiple simulations will provide the sensitivity of the levitation force to flow and
geometry parameters. The simulation performed in this thesis may also be simplified
and still provide physical understanding.

One approach would be to reduce mesh

resolution in the steam core but increase the resolution in the liquid film. Although the
liquid droplets would not be resolved, the effect of interfacial wave dynamics on the
liquid film turbulence could be captured. Once a subgrid turbulence model for the liquid
film turbulence is developed, the mesh resolution in the liquid film can be reduced while
increasing the resolution of the liquid droplets. This would allow investigation of
transient and equilibrium droplet entrainment and deposition rates. DNS can also be
used to gain physical understanding of other two-phase problems. One example is the
simulation of the flooding phenomena in vertical annular flow in which liquid down
flow in the conduit is restricted by the upflow of vapor in the conduit. Of particular
interest is the counter current flow limit (CCFL) in which the liquid down flow prevents
the escape of vapor from the conduit which can result, for heated conduits, in a thermal
limit being reached. Naturally, this problem is closely related to the churn-turbulent
flow regime which is quite important but is currently not well understood.

132

11. REFERENCES
Alauzet, F., X. Li, E. Seol and M. Shephard (2006). "Parallel anisotropic 3D mesh
adaptation by mesh modification." Engineering with Computers 21(3): 247-258.
Ambrosini, W., P. Andreussi and B. J. Azzopardi (1991). "A Physically Based
Correlation for Drop Size in Annular-Flow." International Journal of Multiphase
Flow 17(4): 497-507.
Asali, J. C., T. J. Hanratty and P. Andreussi (1985). "Interfacial Drag and Film Height
for Vertical Annular-Flow." AICHE Journal 31(6): 895-902.
Azzopardi, B. J. (1999). "Turbulence modification in annular gas/liquid flow."
International Journal of Multiphase Flow 25(6-7): 945-955.
Babuvska, I. and W. C. Rheinboldt (1978). "Error Estimates for Adaptive Finite Element
Computations." SIAM Journal on Numerical Analysis 15(4): 736-754.
Brackbill, J. U., D. B. Kothe and C. Zemach (1992). "A continuum method for modeling
surface tension." Journal of Computational Physics 100(2): 335-354.
Brauner, N. and D. M. Maron (1993). "The Role of Interfacial Shear Modeling in
Predicting the Stability of Stratified 2-Phase Flow." Chemical Engineering
Science 48(16): 2867-2879.
Bunner, B. and G. Tryggvason (1999). "Direct numerical simulations of threedimensional bubbly flows." Physics of Fluids 11(8): 1967-1969.
Butterworth, D. and D. J. Pulling (1972). "A Visual study of Mechanisms in Horizontal
Annular Air-Water Flow." UKAEA Report No. AERE-M2556.
Cao, Q., K. Sarkar and A. K. Prasad (2004). "Direct numerical simulations of two-layer
viscosity-stratified flow." International Journal of Multiphase Flow 30(12): 14851508.
Dallman, J. C., J. E. Laurinat and T. J. Hanratty (1984). "Entrainment for Horizontal
Annular Gas-Liquid Flow." International Journal of Multiphase Flow 10(6): 677690.
de Cougny, H. L. and M. S. Shephard (1999). "Parallel refinement and coarsening of
tetrahedral meshes." International Journal for Numerical Methods in Engineering
46(7): 1101-1125.
Dobran, F. (1983). "Hydrodynamic and Heat-Transfer Analysis of 2-Phase AnnularFlow with a New Liquid-Film Model of Turbulence." International Journal of
Heat and Mass Transfer 26(8): 1159-1171.
Drew, D. A. and S. L. Passman (1999). Theory of Multicomponent Fluids. Boulder,
netLibrary.
Enright, D., R. Fedkiw, J. Ferziger and I. Mitchell (2002). "A hybrid particle level set
method for improved interface capturing." Journal of Computational Physics
183(1): 83-116.
Faghri, M. and B. Sunden (2004). Heat and fluid flow in microscale and nanoscale
structures. Southampton, UK WIT Press, Billerica MA.
Flores, A. G., K. E. Crowe and P. Griffith (1995). "Gas-Phase Secondary Flow in
Horizontal, Stratified and Annular 2-Phase Flow." International Journal of
Multiphase Flow 21(2): 207-221.

133

Fore, L. B., S. G. Beus and R. C. Bauer (2000). "Interfacial friction in gas-liquid annular
flow: analogies to full and transition roughness." International Journal of
Multiphase Flow 26(11): 1755-1769.
Fore, L. B., B. B. Ibrahim and S. G. Beus (2002). "Visual measurements of droplet size
in gas-liquid annular flow." International Journal of Multiphase Flow 28(12):
1895-1910.
Franca, L. P. and S. L. Frey (1992). "Stabilized Finite-Element Methods .2. The
Incompressible Navier-Stokes Equations." Computer Methods in Applied
Mechanics and Engineering 99(2-3): 209-233.
Fukano, T. and T. Furukawa (1998). "Prediction of the effects of liquid viscosity on
interfacial shear stress and frictional pressure drop in vertical upward gas-liquid
annular flow." International Journal of Multiphase Flow 24(4): 587-603.
Fukano, T. and T. Inatomi (2003). "Analysis of liquid film formation in a horizontal
annular flow by DNS." International Journal of Multiphase Flow 29(9): 14131430.
Fukano, T. and A. Ousaka (1989). "Prediction of the Circumferential Distribution of
Film Thickness in Horizontal and near-Horizontal Gas-Liquid Annular Flows."
International Journal of Multiphase Flow 15(3): 403-419.
Fulgosi, M., D. Lakehal, S. Banerjee and V. De Angelis (2003). "Direct numerical
simulation of turbulence in a sheared air-water flow with a deformable interface."
Journal of Fluid Mechanics 482: 319-345.
Fuster, D., A. Bagu, T. Boeck, L. Le Moyne, A. Leboissetier, S. Popinet, P. Ray, R.
Scardovelli and S. Zaleski (2009). "Simulation of primary atomization with an
octree adaptive mesh refinement and VOF method." International Journal of
Multiphase Flow 35(6): 550-565.
Galimov, A. (2007). An analysis of interfacial waves and air ingestion mechanisms.
MANE. Troy, NY, Rensselaer Polytechnic Institute.
Gottmann, M. (1997). Local Wall Shear Stress and Interface Behavior of Adiabatic AirWater Flows in Rectangular Ducts (Liquid Films). Tuscon, Arizona, University
of Arizona: 216.
Hanratty, T. J., T. Theofanous, J.-M. Delhaye, J. Eaton, J. McLaughlin, A. Prosperetti, S.
Sundaresan and G. Tryggvason (2003). "Workshop Findings." International
Journal of Multiphase Flow 29(7): 1047-1059.
Henstock, W. H. and T. J. Hanratty (1976). "Interfacial Drag and Height of Wall Layer
in Annular Flows." AICHE Journal 22(6): 990-1000.
Hewitt, G. F. and N. S. Hall-Taylor (1970). Annular two-phase flow, Oxford, New York,
Pergamon Press.
Hewitt, G. F. and D. N. Roberts (1969), "Studies of Two-Phase Flow Patterns by
Simultaneous X-ray and Flash Photography"UKAEA, AERE-M2159
Hirt, C. W. and B. D. Nichols (1981). "Volume of Fluid (Vof) Method for the Dynamics
of Free Boundaries." Journal of Computational Physics 39(1): 201-225.
Holowach, M. J., L. E. Hochreiter and F. B. Cheung (2002). "A model for droplet
entrainment in heated annular flow." International Journal of Heat and Fluid
Flow 23(6): 807-822.

134

Hu, H. H., N. A. Patankar and M. Y. Zhu (2001). "Direct Numerical Simulations of


Fluid-Solid Systems Using the Arbitrary Lagrangian-Eulerian Technique."
Journal of Computational Physics 169(2): 427-462.
Hughes, T. J. R., L. P. Franca, I. Harari, M. Mallet, F. Shakib and T. E. Spelce (1987).
Finite Element Method for High-Speed Flows: Consistent Calculation of
Boundary Flux. Proceedings of AIAA 25th Aerospace Sciences Meeting, Reno,
Nevada, USA.
Ishii, M. and K. Mishima (1984). "2-Fluid Model and Hydrodynamic Constitutive
Relations." Nuclear Engineering and Design 82(2-3): 107-126.
Jansen, K. E., C. H. Whiting and G. M. Hulbert (2000). "A generalized-alpha method for
integrating the filtered Navier-Stokes equations with a stabilized finite element
method." Computer Methods in Applied Mechanics and Engineering 190(3-4):
305-319.
Jayanti, S., G. F. Hewitt and S. P. White (1990). "Time-dependent behaviour of the
liquid film in horizontal annular flow." International Journal of Multiphase Flow
16(6): 1097-1116.
Johnson, A. A. and T. E. Tezduyar (1997). "3D Simulation of fluid-particle interactions
with the number of particles reaching 100." Computer Methods in Applied
Mechanics and Engineering 145(3-4): 301-321.
Kataoka, I. and M. Ishii (1983). Entrainment and Deposition Rates for Droplets in
Annular Two-Phase Flow. ASME-JSME Thermal Engineering Joint Conference,
Honolulu, Hawaii.
Kataoka, I., M. Ishii and K. Mishima (1983). "Generation and Size Distribution of
Droplet in Annular 2-Phase Flow." Journal of Fluids Engineering-Transactions of
the ASME 105(2): 230-238.
Kelecy, F. J. and R. H. Pletcher (1997). "The development of a free surface capturing
approach for multidimensional free surface flows in closed containers." Journal
of Computational Physics 138(2): 939-980.
Kim, J., P. Moin and R. Moser (1987). "Turbulence Statistics in Fully-Developed
Channel Flow at Low Reynolds-Number." Journal of Fluid Mechanics 177: 133166.
Kocamustafaogullari, G., S. R. Smits and J. Razi (1994). "Maximum and Mean Droplet
Sizes in Annular 2-Phase Flow." International Journal of Heat and Mass Transfer
37(6): 955-965.
Kosky, P. G. (1971). "Thin Liquid Films under Simultaneous Shear and Gravity Forces."
International Journal of Heat and Mass Transfer 14(8): 1220-&.
Kumar, R. and D. P. Edwards (1996). Interfacial Shear Modelling in Two-Phase Annular
Flow. Experimental multiphase flows: Proceedings of the ASME Heat Transfer
Division Vol 3, Atlanta; GA, ASME; 1996.
Kumar, R., C. C. Maneri and T. D. Strayer (2004). "Modeling and numerical prediction
of flow boiling in a thin geometry." Journal of Heat Transfer-Transactions of the
ASME 126(1): 22-33.
Kunert, G. (2002). "Toward anisotropic mesh construction and error estimation in the
finite element method." Numerical Methods for Partial Differential Equations
18(5): 625-648.

135

Lahey, R. T., Jr. (2005). "The simulation of multidimensional multiphase flows."


Nuclear Engineering & Design 235(10-12): 1043-1060.
Lahey, R. T., Jr. (2009). "On the Computation of Multiphase Flows." Journal of Nuclear
Technology 167(1): 29-45.
Lahey, R. T., Jr. and D. A. Drew (2001). "The analysis of two-phase flow and heat
transfer using a multidimensional, four field, two-fluid model." Nuclear
Engineering and Design 204(1-3): 29-44.
Lahey, R. T., Jr. and R. J. Moody (1993). The Thermal-Hydraulics of a Boiling Water
Nuclear Reactor. La Grange Park, IL, American Nuclear Society.
Lakehal, D. (2004). DNS and LES of turbulent Multifluid Flows. 3rd International
Symposium on Two-Phase Flow Modelling and Experimentation, Pisa, Italy.
Lakehal, D., M. Fulgosi, G. Yadigaroglu and S. Banerjee (2003). "Direct numerical
simulation of turbulent heat transfer across a mobile, sheared gas-liquid
interface." Journal of Heat Transfer-Transactions of the ASME 125(6): 11291139.
Lakehal, D., M. Meier and M. Fulgosi (2002). "Interface tracking towards the direct
simulation of heat and mass transfer in multiphase flows." International Journal
of Heat and Fluid Flow 23(3): 242-257.
Laurinat, J. E., T. J. Hanratty and W. P. Jepson (1985). "Film Thickness Distribution for
Gas-Liquid Annular-Flow in a Horizontal Pipe." Physicochemical
Hydrodynamics 6(1-2): 179-195.
Lebas, R., T. Menard, P. A. Beau, A. Berlemont and F. X. Demoulin (2009). "Numerical
simulation of primary break-up and atomization: DNS and modelling study."
International Journal of Multiphase Flow 35(3): 247-260.
Li, J. and Y. Renardy (1999). "Direct simulation of unsteady axisymmetric core-annular
flow with high viscosity ratio." Journal of Fluid Mechanics 391: 123-149.
Li, X., M. S. Shephard and M. W. Beall (2005). "3D anisotropic mesh adaptation by
mesh modification." Computer Methods in Applied Mechanics and Engineering
194(48-49): 4915-4950.
Lin, T. F., O. C. Jones, R. T. Lahey, Jr., R. C. Block and M. Murase (1985). "Film
Thickness Measurements and Modeling in Horizontal Annular Flows."
Physicochemical Hydrodynamics 6(1-2): 197-206.
Lopez de Bertodano, M. A., C. J. Jan and S. G. Beus (1995). Droplet Entrainment
Correlation for High Pressure Annular Two-Phase Flow. Heat transfer, Portland;
OR, ANS; 1995.
Martin, J. C. and W. J. Joyce (1952). "Part IV. An Experimental Study of the Collapse of
Liquid Columns on a Rigid Horizontal Plane." Philosophical Transactions of the
Royal Society of London Series A, Mathematical and Physical Sciences 244:
312-324.
Mei, C. C. (1989). The Applied Dynamics of Ocean Surface Waves. Singapore, River
Edge N.J.
Nagrath, S. (2004). Adaptive stabilized finite element analysis of multi-phase flows
using level set approach. MANE, Rensselaer Polytechnic Institute: 130.
Nagrath, S., K. E. Jansen and R. T. Lahey, Jr. (2005). "Computation of incompressible
bubble dynamics with a stabilized finite element level set method." Computer
Methods in Applied Mechanics and Engineering 194(42-44): 4565-4587.

136

Osher, S. and R. P. Fedkiw (2001). "Level set methods: An overview and some recent
results." Journal of Computational Physics 169(2): 463-502.
Osher, S. and J. A. Sethian (1988). "Fronts Propagating with Curvature-Dependent
Speed - Algorithms Based on Hamilton-Jacobi Formulations." Journal of
Computational Physics 79(1): 12-49.
Owen, D. G. (1986). An Experimental and Theoretical Analysis of Equilibrium Annular
Flows. Department of Chemical Engineering, University of Birmingham.
Pope, S. B. (2000). Turbulent flows, New York.
Rider, W. J. and D. B. Kothe (1998). "Reconstructing volume tracking." Journal of
Computational Physics 141(2): 112-152.
Russell, T. W. F. and D. E. Lamb (1965). "Flow Mechanism of 2-Phase Annular Flow."
Canadian Journal of Chemical Engineering 43(5): 237-245.
Ryskin, G. and L. G. Leal (1984). "Numerical-Solution of Free-Boundary Problems in
Fluid-Mechanics .1. The Finite-Difference Technique." Journal of Fluid
Mechanics 148(NOV): 1-17.
Sahni, O., J. Mller, K. E. Jansen, M. S. Shephard and C. A. Taylor (2006). "Efficient
anisotropic adaptive discretization of the cardiovascular system." Computer
Methods in Applied Mechanics and Engineering, John H. Argyris Memorial
Issue. Part II 195(41-43): 5634-5655.
Scardovelli, R. and S. Zaleski (1999). "Direct numerical simulation of free-surface and
interfacial flow." Annual Review of Fluid Mechanics 31: 567-603.
Sethian, J. A. (1999). Level Set Methods and Fast Marching Methods, Cambridge
University Press.
Sethian, J. A. and P. Smereka (2003). "Level set methods for fluid interfaces." Annual
Review of Fluid Mechanics 35: 341-372.
Shedd, T. A. and T. A. Newell (2004). "Characteristics of the liquid film and pressure
drop in horizontal, annular, two-phase flow through round, square and triangular
tubes." Journal of Fluids Engineering-Transactions of the ASME 126(5): 807817.
Siebert, B. W., C. C. Maneri, R. F. Kunz and D. P. Edwards (1995). A Four-Field Model
and CFD Implementation for Multi-Dimensional, Heated Two-Phase Flows.
Multiphase flow '95, Kyoto; Japan, Kyoto; Kyoto University; 1995.
Simmetrix. (2009). "http://www.simmetrix.com."
Son, G. (2005). "A level set method for incompressible two-fluid flows with immersed
solid boundaries." Numerical Heat Transfer Part B-Fundamentals 47(5): 473489.
Sussman, M. (2003). "A second order coupled level set and volume-of-fluid method for
computing growth and collapse of vapor bubbles." Journal of Computational
Physics 187(1): 110-136.
Sussman, M., A. S. Almgren, J. B. Bell, P. Colella, L. H. Howell and M. L. Welcome
(1999). "An adaptive level set approach for incompressible two-phase flows."
Journal of Computational Physics 148(1): 81-124.
Sussman, M. and E. Fatemi (1999). "An efficient, interface-preserving level set
redistancing algorithm and its application to interfacial incompressible fluid
flow." Siam Journal on Scientific Computing 20(4): 1165-1191.

137

Sussman, M., E. Fatemi, P. Smereka and S. Osher (1998). "An improved level set
method for incompressible two-phase flows." Journal of Computers & Fluids
27(5-6): 663-680.
Sussman, M. and E. G. Puckett (2000). "A coupled level set and volume-of-fluid method
for computing 3D and axisymmetric incompressible two-phase flows." Journal of
Computational Physics 162(2): 301-337.
Sutharshan, B., M. Kawaji and A. Ousaka (1995). "Measurement of Circumferential and
Axial Liquid-Film Velocities in Horizontal Annular-Flow." International Journal
of Multiphase Flow 21(2): 193-206.
Tatterson, D. F., J. C. Dallman and T. J. Hanratty (1977). "Drop Sizes in Annular GasLiquid Flows." AICHE Journal 23(1): 68-76.
Taylor, C. A., T. J. R. Hughes and C. K. Zarins (1998). "Finite element modeling of
blood flow in arteries." Computer Methods in Applied Mechanics and
Engineering 158(1-2): 155-196.
Trabold, T. A. and R. Kumar (2000). "Vapor core turbulence in annular two-phase
flow." Experiments in Fluids 28(2): 187-194.
Tryggvason, G., B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han, S.
Nas and Y. J. Jan (2001). "A front-tracking method for the computations of
multiphase flow." Journal of Computational Physics 169(2): 708-759.
Ueda, T. (1979). "Entrainment Rate and Size of Entrained Droplets in Annular 2-Phase
Flow." Bulletin of the JSME-Japan Society of Mechanical Engineers 22(171):
1258-1265.
Unverdi, S. O. and G. Tryggvason (1992). "A Front-Tracking Method for Viscous,
Incompressible, Multi-Fluid Flows." Journal of Computational Physics 100(1):
25-37.
van der Pijl, S. P., A. Segal, C. Vuik and P. Wesseling (2005). "A mass-conserving
Level-Set method for modelling of multi-phase flows." International Journal for
Numerical Methods in Fluids 47(4): 339-361.
Vassallo, P. (1999). "Near wall structure in vertical air-water annular flows."
International Journal of Multiphase Flow 25(3): 459-476.
Wallis, G. B. (1969). One-dimensional two-phase flow. New York, McGraw-Hill.
Whalley, P. B. (1987). Boiling, condensation, and gas-liquid flow, Oxford University
Press.
Whitham, G. B. (1974). Linear and nonlinear waves, New York Wiley.
Whiting, C. H. (1999). Stabilized finite element methods for fluid dynamics using a
hierarchical basis. MEAE. Troy, NY, Rensselaer Polytechnic Institute: 135.
Whiting, C. H. and K. E. Jansen (2001). "A stabilized finite element method for the
incompressible Navier-Stokes equations using a hierarchical basis." International
Journal for Numerical Methods in Fluids 35(1): 93-116.
Wilcox, D. C. (1998). Turbulence Modeling for CFD. La Canada, DCW Industries.
Williams, M. W., D. B. Kothe and E. G. Puckett (1999). Accuracy and Convergence of
Continuum Surface-Tension Models. Fluid Dynamics at Interfaces. W. W. Shyy
and R. Narayanan. Cambridge, U.K. ; New York :, Cambridge University Press,:
xv, 461 p. :.

138

Woodmansee, D. E. and T. J. Hanratty (1969). "Mechanism for the removal of droplets


from a liquid surface by a parallel air flow." Chemical Engineering Science
24(2): 299-307.
Wurtz, J. (1978), "An Experimental and Theoretical Investigation of Annular
Steam/Water Flow in Tubes and Annuli at 30 to 90 bar"RISO National
Laboratory, RISO Report No. 372
Yue, W. S., C. L. Lin and V. C. Patel (2003). "Numerical simulation of unsteady
multidimensional free surface motions by level set method." International Journal
for Numerical Methods in Fluids 42(8): 853-884.

139

APPENDIX A: CLASSIFICATION OF PHASE FIELDS


A1 DESCRIPTION OF METHOD
The level set method provides a distinction between the phases (e.g., steam or water),
but additional algorithms are required when subdividing each phase into fields.
Consistent with a 3-D, two-fluid, four-field computational multiphase fluid dynamic
(CMFD) approach [Lahey (2005)], each phase is divided into a continuous and disperse
field. The liquid film on the wall is considered to be a continuous liquid (CL) while the
liquid droplets in the steam core are considered to be dispersed liquid (DL). Similarly,
the steam core is considered to be a continuous vapor (CV) while the bubbles entrained
in the liquid film are considered to be dispersed vapor (DV). This classification scheme
results in three steam/water interfaces: CL-CV, CL-DV, and CV-DL.
Use of the MeshSim libraries [Simmetrix (2009)] allows the mesh to be interrogated in
an efficient manner. One can progress through the mesh by walking along edges or
faces. A code, called fourfield, was developed that uses the MeshSim libraries to walk
through the mesh and identify all groupings of vertices, termed a glob, of common
phase. As illustrated in Figure A1, it does this by starting with an arbitrary vertex in the
glob and performing a walk-out procedure along adjacent edges that finds all phase
boundaries of the glob. The code outputs a new PHASTA-IC solution file that has eight
solution variables (pressure, velocity vector, temperature, level set scalar 1 (), level set
scalar 2 (d), and field classification). Figure A1(a) shows a glob (for example, a bubble
in a liquid film) that is represented by the solid black line. The first vertex encountered
in the glob is represented by the solid black circle. The fourfield code then finds all
edges connected to the initial vertex and the vertices on the opposite ends of the edges
(open circles). This is termed a walk-out step. Using the new vertices as new starting
points, fourfield then performs additional walk-out steps.

This continues until all

vertices within the glob are identified. Figure A1(b) shows the results after the glob has
been fully identified. The solid circles represent vertices contained inside the glob while
the dashed open circles represent the vertices outside the glob that are connected to
edges that cross the glob boundary. The level set value determines the phase of the glob.

140

The glob volume is then used to classify it as disperse or continuous. Ultimately, all
globs are classified as continuous liquid (CL), disperse vapor (DV), disperse liquid (DL),
or continuous vapor (CV). For each glob the code computes the volume, surface area,
average velocity vector, extremes in each coordinate direction, number of vertices
comprising the glob, and the number of regions (i.e., computational cells) comprising the
glob. This is done for a series of time steps where the solution at each time step is saved
to a PHASTA-IC solution file.

Initial vertex in glob


New vertices found in glob

Vertices in glob
Vertices outside glob

After
completing
all walk-outs

Initial vertex is selected.


Walk out along edges to find
other vertices defining glob.

Walk out along edges to find


other vertices. Continue
until all edges that cross
glob boundary are found.
(b)

(a)

Figure A1: Illustration of walk-out procedure used to define a phasic glob.

The algorithm for the fourfield code is provided in section A3. The algorithm makes a
couple of simplifying decisions. First, for those regions that existed only partially in a
glob (i.e., in the regCross list), it was assumed that half the region existed in the glob
and the other half outside the glob. This is obviously not true but the cumulative error is
expected to be small. For simplicity and to aid post-processing time, the 50 percent
assumption was made and each region in the regCross list contributed half of its volume
to the glob volume. The second simplifying decision dealt with distinguishing between
141

disperse and continuous phases. Again for simplicity, a volume criteria was established.
For the RISO data simulation, it was assumed that any glob whos volume that was less
then 10 percent of the initial phasic volume was a disperse glob; otherwise the glob was
continuous. In the end, this logic created only one continuous liquid and one continuous
vapor glob with the others being disperse.
The value of the field classification variable was intentionally ordered as indicated in
Table A1 for post processing purposes.

Neighboring fields correspond to natural

interfaces that occur in two phase annular flow. The 0-1 boundary represents the DVCL interface, or bubbles in a liquid film. The 1-2 boundary represents the CL-CV
interface, or the interface of the liquid film with the vapor core. Finally, the 2-3
boundary represents the DL-CV interface, or the droplets in the vapor core. Ordering
the field values like this allowed easier viewing of fields in the Paraview post processor.
Table A1: Definition of Four Field values
Classification

Field Value

Dispersed Vapor (DV)

Continuous Liquid (CL)

Continuous Vapor (CV)

Dispersed Liquid (DL)

A2 COMPUTED QUANTITIES
Section A3 provides the algorithm for the fourfield code. This section provides the
mathematical representation of the quantities computed therein. For each phasic glob,
fourfield computes the number of vertices, Nverts, contained within the glob, the number
of regions, Nreg_in, contained completely within the glob, and the number of regions,
Nreg_cross, that cross the glob boundary. The glob volume, Vglob, is defined as the sum of
all region volumes completely within the glob and half the total volume of all regions
that cross the region boundary:
V glob =

(V ) + 0.5(V ) ,
N reg _ cross

N reg _ in
i =1

reg i

i =i

reg i

(A1)

where Vreg is the volume of a region. For each region that crosses the glob boundary, the
intersection plane of the glob boundary (i.e., zero level set) with the region can be
142

determined. In three dimensional space, the intersection plane will intersect either three
or four edges of the region; hence, there will be only three or four points comprising the
intersection plane. For a three-point intersection plane, the intersection area is computed
using Herons formula for the area on an arbitrary triangle. For a triangle of ordered
vertices 1, 2, and 3 and edge lengths L1-2, L2-3, and L1-3, Herons formula states the area
of the triangle is, Atri, is given by:

Atri = [s(s L12 )(s L23 )(s L13 )] ,


0.5

(A2)

where s is half the triangles perimeter and is given by:

s=

L1 2 + L23 + L13
.
2

(A3)

The area of four-point intersection planes is computed by splitting the four-point plane
into two three point triangles and using Herons formula for each triangle.

The

intersection plane area, Aint_plane, may be expressed as the summation of triangle areas
over the number of triangles, Ntri, making up the intersection plane:
N tri

Aint_ plane = ( Atri )i ,

(A4)

i =i

where Ntri equals one for three-point intersection planes and equals two for four-point
intersection planes. Summing the area of these intersection planes gives the glob surface
area, Aglob:
Aglob =

(A

N reg _ cross

(A5)

int_ plane i

i =i

To compute the average velocity, uglob, of the glob, a volume-weighted average of all
region-averaged velocities within the glob is computed. The average velocity is given
by:

(u

N reg _ in

u glob =

i =1

reg

Vreg )i +

0.5(u

N reg _ cross
i =i

V glob

reg

Vreg )i
,

(A6)

where Vglob is the glob volume defined in Equation (A1), Vreg is the volume of a given
region, and ureg is the average velocity of the region where only the velocities of vertices
within the region are used in the averaging.

143

The fourfield code also computes the extremes in the coordinate directions for each glob.
Only coordinates from vertices contained within the glob are considered.

A3 CODE ALGORITHM
The algorithm for the four-field classification code is given below:

Load mesh in MeshSim database


Associate mesh with solid model (Parasolid)
Loop over solution files (i.e., time steps)
o Read PHASTA solution and attach to vertices in mesh database
o Loop over vertices in mesh
Set globID = 0
For each unclassified vertex (i.e., vertex has not been assigned to current
or previous glob) then characterize glob and classify:
Clear contents of lists: globVerts, newVerts, lastVerts
Add vertex to globVerts (list of vertices in glob) and newVerts (list of
new vertices added to globVerts)
Attach globID to vertex
Get phase of vertex from level set value (this will be the phase of the
glob)
March = TRUE
While (march)
o March = FALSE (will be changed if new vertices of same phase
are found)
o lastVerts = newVerts
o clear contents of newVerts
o loop over vertices in lastVerts
Loop over edges attached to vertex
overtex = vertex on opposite end of edge
If overtex is of the same phase and has not been assigned
a globID (i.e., a glodID has not been attached to the
vertex) then
o add overtex to newVerts list
o add overtex to globVerts list
o set march=TRUE
end loop over vertex edges
o end loop over lastVerts list
number of vertices in glob = size of globVerts list
Create list of regions (or cells) completely in glob, called regIn, and
list of regions partially in glob, called regCross:
o Loop over all vertices in globVerts list
Loop over all regions associated with vertex

144

If all vertices are of the same phase as the glob, then add
region to the regIn list.
Else, region must straddle glob so add to regCross list.
End loop over regions
Record the maximum and minimum coordinate values while
looping over vertices.
o End loop over globVerts vertices
Compute glob volume and volume-weighted average velocity:
o Loop over regions in regIn list
globVolume[globID] = globVolume[globID] + region volume
Compute average of coordinate velocities
x velocity: globAvgVelocity[globID][0] =
globAvgVelocity[globID][0] +
average_x_velocity*region_volume
y velocity: globAvgVelocity[globID][1] =
globAvgVelocity[globID][1] +
average_y_velocity*region_volume
z velocity: globAvgVelocity[globID][2] =
globAvgVelocity[globID][2] +
average_z_velocity*region_volume
o end loop over regIn regions
o Loop over regions in regCross list
globVolume[globID] = globVolume[globID] + 0.5*region
volume
Compute average of coordinate velocities of only those
vertices in glob
x velocity: globAvgVelocity[globID][0] =
globAvgVelocity[globID][0] +
average_x_velocity*0.5*region_volume
y velocity: globAvgVelocity[globID][1] =
globAvgVelocity[globID][1] +
average_y_velocity*0.5*region_volume
z velocity: globAvgVelocity[globID][2] =
globAvgVelocity[globID][2] +
average_z_velocity*0.5*region_volume
o end loop over regCross regions
o Compute average velocity in each direction:
globAvgVelocity[globID][i] = globAvgVelocity[globID][i] /
globVolume[globID] for i=0,1,2
Compute glob interfacial area:
o Loop over regions in regCross list
Compute number and coordinates of intersection points where
zero level set intersects region
Order intersection points in clockwise direction
Use Herons formula to compute intersection plane area:
intArea
145

globIntArea[globID] = globIntArea[globID] + intArea


o end loop over regCross regions
Classify glob
o If phase = liquid (level set > 0)
If globVolume > cldlVolumeLimit
glob is cl
set vertField = 1
numFields[1]++
else
glob is dl
set vertField = 3
numFields[3]++
o else
If globVolume > cvdvVolumeLimit
glob is cv
set vertField = 2
numFields[2]++
else
glob is dv
set vertField = 0
numFields[0]++
Tag glob vertices with classification (from vertField variable)
o Loop over all vertices in globVerts list
Attach to vertex vertField value
o End loop over globVerts vertices
Increase glob counter: globtotal++
Else vertex is part of existing glob evaluate next vertex
o End loop over vertices
o Output glob count for each field for time step
o Output stats (volume, velocity, field) for each glob.
o Clean all attached data so clean slate for next time step evaluated
End loop over solution files
Release mesh and model
End of algorithm

146

Vous aimerez peut-être aussi