+ All Categories
Home > Documents > Numerical Modeling of Natural Gas Two-Phase Flow Split at ...

Numerical Modeling of Natural Gas Two-Phase Flow Split at ...

Date post: 16-Jan-2022
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
170
The Pennsylvania State University The Graduate School College of Earth and Mineral Sciences NUMERICAL MODELING OF NATURAL GAS TWO-PHASE FLOW SPLIT AT BRANCHING T-JUNCTIONS WITH CLOSED-LOOP NETWORK APPLICATIONS A Dissertation in Petroleum and Natural Gas Engineering by Doruk Alp c 2009 Doruk Alp Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2009
Transcript

The Pennsylvania State University

The Graduate School

College of Earth and Mineral Sciences

NUMERICAL MODELING OF NATURAL GAS TWO-PHASE FLOW SPLIT

AT BRANCHING T-JUNCTIONS

WITH CLOSED-LOOP NETWORK APPLICATIONS

A Dissertation in

Petroleum and Natural Gas Engineering

by

Doruk Alp

c© 2009 Doruk Alp

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

December 2009

The dissertation of Doruk Alp was reviewed and approved* by the following:

Luis F. Ayala

Assistant Professor of Petroleum and Natural Gas Engineering

Dissertation Adviser

Chair of Committee

Turgay Ertekin

Professor of Petroleum and Natural Gas Engineering

George E. Trimble Chair in Earth and Mineral Sciences

Undergraduate Program Officer of Petroleum and Natural Gas Engineering

Robert Watson

Associate Professor Emeritus of Petroleum and Natural Gas Engineering

Mirna Urquidi-Macdonald

Professor of Engineering Science and Mechanics

John Mahaffy

Associate Professor Emeritus of Nuclear Engineering

R. Larry Grayson

Professor of Energy and Mineral Engineering

George H., Jr., and Anne B. Deike Chair in Mining Engineering

Graduate Program Officer of Energy and Mineral Engineering

*Signatures are on file in the Graduate School.

Abstract

Branching T-junctions are essential components of small and large scale piping systems found in

natural gas and oil pipeline networks, as well as various industrial applications. Two-phase flow

through branching tees (T-junctions) result in pronounced hydraulic losses and uneven phase sepa-

ration following the split of flow stream; ultimately causing profound effects on system performance

and quality of delivered fluids. Current state-of-the-art modeling software available to oil and gas

industry do not address uneven phase separation (route preference) issue and typically do not

provide a distinct T-junction component for modeling purposes. A finite-volume (FVM) based

two-fluid model has been developed for one-dimensional, steady-state analysis of two-phase flow split

and uneven phase separation at branching T-junctions of natural gas networks using steady-state

Euler equations and outlet pressure specifications. Based on a comprehensive review of available

branching T-junction and phase separation models in the literature, and classification of modeling

efforts included in this text, Double Stream Model (DSM) of Hart et al. (1991), essentially a

mechanistic phase separation sub-model for low liquid loading conditions (holdup < 0.06), is applied

at the junction control volume to capture uneven phase separation. In order to have a consistent

model, phasic momentum equations are replaced with gas phase Bernoulli equations at the junction

cell and required gas phase loss coefficients (K-factors) are calculated using Gardel correlations.

Generalized Newton-Raphson method is applied for simultaneous solution of governing equations,

for all the control volumes in the computational grid. Along with the staggered solution of governing

momentum equations at CV edges, this allows accounting for the impact of outlet (downstream)

pressures on flow split as well as closed-loop network applicability. Analysis is focused on regular,

horizontal T-junctions with 90o branching angle. Results from air-water and hydrocarbon mixture

studies are in agreement with literature; model captures pressure rise in the main line following

the flow split (the Bernoulli effect) and uneven phase separation with changing liquid flow rates

and outlet pressures. An important observation is that while pressure in the main line right after

the split (entrance pressure of the run) is always higher than the entrance pressure of the branch

due to Bernoulli effect, this is not necessarily the case for run and branch outlet pressures. When

branch outlet pressure is specified higher than the run outlet pressure, less fluid is diverted into the

branch. However, the turning of flow causes branch entrance pressure to be always smaller than run

entrance pressure. Slight difference in overall hydrocarbon mixture composition in run and branch

arms, after flow split at the tee, suggests that for small holdup values compositional change can be

ignored for practical purposes.

iii

Contents

List of Tables vii

List of Figures viii

Nomenclature x

Acknowledgements xii

Chapters

1 Introduction 1

2 Flow In Conduits 10

2.1 Single-Phase Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.1.1 Single-Phase Steady-State Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2 Two-Phase Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.2.1 Flow Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.2.2 Two-Phase Flow Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Homogeneous Equilibrium Model . . . . . . . . . . . . . . . . . . . . . . . . . 20

Separated Flow Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Drift-Flux Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Two-Fluid Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2.3 Mechanistic Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.2.4 Marching Algorithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3 Flow Split at T-Junctions 26

3.1 Two-Phase Flow Split . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.2 Phenomenological and Mechanistic Junction Models . . . . . . . . . . . . . . . . . . 31

3.2.1 Pressure Change Sub-Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.2.2 Phase Separation Sub-Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3.2.3 The Double Stream Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.2.4 Correction Factor and Loss Coefficient Correlations . . . . . . . . . . . . . . 42

3.2.5 Film-stop Phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

iv

4 Model Conceptualization 46

4.1 Conservative Property: FVM vs FDM . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4.2 Single-Phase Flow Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.2.1 Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2.2 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.2.3 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.3 Two-Fluid Model Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.3.1 Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.3.2 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.3.3 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.4 Closure Relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

4.5 Primary Unknowns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.6 Advective Property: Donor Cell Scheme . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.7 Staggered Grid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.8 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4.9 Order of Accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.9.1 Higher Order Approximations . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4.10 Transient Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.11 Numerical Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.11.1 Newton-Raphson Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.11.2 Phase Appearance-Disappearance . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.11.3 Flow Pattern Transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

4.11.4 Alignment of Residuals and Primary Unknowns . . . . . . . . . . . . . . . . . 80

4.12 Two-Fluid Junction Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.13 Fluid Composition After Split . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

5 Results and Discussion 91

5.1 Single-Phase Compressible Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5.2 Two-Phase Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.3 Two-Phase Flow and Secondary Phase Appearance . . . . . . . . . . . . . . . . . . . 97

5.4 Flow Split and Phase Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.4.1 Hydrocarbon Mixture Uneven Phase Separation . . . . . . . . . . . . . . . . 108

5.5 Open Network Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

5.6 Single-Phase Closed-Loop Network Application . . . . . . . . . . . . . . . . . . . . . 113

6 Concluding Remarks 116

6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

References 120

v

Appendices

A Single-Phase Flow in Pipe Networks 127

A.1 Kirchhoff Analysis for Flow in Networks . . . . . . . . . . . . . . . . . . . . . . . . . 129

A.2 Steady-State Analysis of Pipeline Systems . . . . . . . . . . . . . . . . . . . . . . . . 130

A.2.1 Open Network Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

A.2.2 Closed Network Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Cross Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Renouard Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

Stoner Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132

A.3 Transient Analysis of Pipeline Networks . . . . . . . . . . . . . . . . . . . . . . . . . 133

A.3.1 Finite Difference Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

A.3.2 Method of Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

A.4 Description of the Network Model for Numerical Solution . . . . . . . . . . . . . . . 135

B Double Channel T-Junction Model 137

B.1 Double Channel Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

B.2 The Buffer Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

B.3 Simplifying Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

B.4 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

B.4.1 Mass Balances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

B.4.2 Momentum Balances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

B.4.3 Parallel Split . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

B.4.4 Reduction of Unknowns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

C Fluid Properties 151

C.1 Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

C.2 Density calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

C.3 Fugacity calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

C.4 Enthalpy calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

D Hydrocarbon Data 156

vi

List of Tables

2.1 Conservation principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.1 Correction factor and loss coefficient correlations for a 90o branching tee . . . . . . . 42

3.2 Models compared by Walters et al. (1998) . . . . . . . . . . . . . . . . . . . . . . . . 45

4.1 Flow pattern dependent parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.2 Primary unknowns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.1 Single-phase compressible flow study data . . . . . . . . . . . . . . . . . . . . . . . . 92

5.2 Two-Phase flow study data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.3 Air-water flow split data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

5.4 Hydrocarbon mixture flow split data . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5.5 Hydrocarbon mixture flow change in overall composition after flow split at the

T-junction; pavg = 12636.81 kPa, λG = 0.55 and λL = 1 . . . . . . . . . . . . . . . . 110

5.6 Air-water flow open network data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

5.7 Single-phase network data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.8 Comparison of pressure results of of the single-phase network . . . . . . . . . . . . . 115

D.1 Hydrocarbon mixture composition and critical properties . . . . . . . . . . . . . . . 156

D.2 Hydrocarbon mixture Peng-Robinson EOS specific parameters . . . . . . . . . . . . 156

D.3 Hydrocarbon mixture binary interacting coefficients . . . . . . . . . . . . . . . . . . 156

D.4 Hydrocarbon mixture Passut and Danner (1972) coefficients . . . . . . . . . . . . . . 157

vii

List of Figures

1.1 Branching T-junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Typical phase envelope for natural gases . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 Route preference phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4 Sample 3D computational grids for T-junctions (from Adechy and Issa, 2004; Lahey,

1990) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.1 Typical flow patterns observed in near horizontal gas condensates pipelines . . . . . 18

2.2 Sample flow regime map for horizontal pipes (Shoham, 2006) . . . . . . . . . . . . . 19

3.1 Three-arm junctions: tee and wye . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2 T-Junction classification based on flow directions . . . . . . . . . . . . . . . . . . . . 28

3.3 Pressure profiles after single-phase flow split at a branching T-junction . . . . . . . . 28

3.4 Recirculation zones of a branching tee . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.5 Dividing streamlines for two-phase flow in a branching tee . . . . . . . . . . . . . . . 37

3.6 Zones of influence for two-phase flow in a branching tee . . . . . . . . . . . . . . . . 38

3.7 Comparison of K-factors by various correlations . . . . . . . . . . . . . . . . . . . . . 43

4.1 Finite control volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.2 Control volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.3 Force Balance in an Elementary Pipe CV . . . . . . . . . . . . . . . . . . . . . . . . 52

4.4 Energy Balance in an Elementary Pipe CV . . . . . . . . . . . . . . . . . . . . . . . 54

4.5 Control Volume for Stratified Two-Phase Flow . . . . . . . . . . . . . . . . . . . . . 59

4.6 Control Volume for Stratified Two-Phase Flow . . . . . . . . . . . . . . . . . . . . . 60

4.7 Control Volume for Stratified Two-Phase Flow . . . . . . . . . . . . . . . . . . . . . 63

4.8 Coincidental Grid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.9 Staggered Grid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.10 Control Volumes for a T-Junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

5.1 Single-phase compressible flow study pressure profiles . . . . . . . . . . . . . . . . . 92

5.2 Single-phase compressible flow study numerical error profile . . . . . . . . . . . . . . 93

5.3 Two-phase RK and NR pressure profiles . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.4 Two-phase RK and NR temperature profiles . . . . . . . . . . . . . . . . . . . . . . . 95

5.5 Two-phase RK and NR gas velocity profiles . . . . . . . . . . . . . . . . . . . . . . . 95

5.6 Two-phase RK and NR liquid velocity profiles . . . . . . . . . . . . . . . . . . . . . . 96

5.7 Two-phase RK and NR holdup profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 96

viii

5.8 Phase appearance p profiles - 20 CVs . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

5.9 Phase appearance T profiles - 20 CVs . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.10 Phase appearance αG profiles - 20 CVs . . . . . . . . . . . . . . . . . . . . . . . . . . 98

5.11 Phase appearance αG vs vk profiles - 20 CVs, HTC ∼= 5.6783W K−1m−2 . . . . . . 99

5.12 Phase appearance αG vs vk profiles - 20 CVs, HTC ∼= 1.0W K−1m−2 . . . . . . . . 100

5.13 Phase appearance density and velocity profiles - 20 CVs, HTC ∼= 5.6783W K−1m−2 100

5.14 Phase appearance density and velocity profiles - 20 CVs, HTC ∼= 1.0W K−1m−2 . . 101

5.15 Hydrocarbon mixture p− T profiles - 20 and 40 CVs, HTC ∼= 5.6783 [W K−1m−2] . 102

5.16 Phase appearance αG vs cum. mass transfer profiles - 20 and 40 CVs, HTC ∼=5.6783W K−1m−2] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.17 Phase appearance velocity profiles - 20 and 40 CVs, HTC ∼= 5.6783W K−1m−2 . . . 103

5.18 Phase appearance αG and mass transfer per CV profiles - 20 and 40 CVs, HTC∼= 5.6783W K−1m−2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.19 Air-water flow pressure profile at T-junction for mL = 0.0057 kg/s . . . . . . . . . . 105

5.20 Air-water flow αG profile at T-junction for mL = 0.0057 kg/s and prun = 102.125 kPa105

5.21 Air-water flow (∆p)tee vs λL plots for different water flow rates . . . . . . . . . . . . 106

5.22 Air-water flow (∆p)out vs λL plots for different water flow rates . . . . . . . . . . . . 107

5.23 Air-water flow λG vs λL plots for different water flow rates . . . . . . . . . . . . . . 107

5.24 Air-water flow (∆p)out vs λG plots for different water flow rates . . . . . . . . . . . . 108

5.25 Air-water complete phase separation curves . . . . . . . . . . . . . . . . . . . . . . . 109

5.26 Hydrocarbon mixture flow λG vs λL plots for different average flow pressures . . . . 110

5.27 Air-water open network pressure profile . . . . . . . . . . . . . . . . . . . . . . . . . 112

5.28 Air-water open network αG profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

5.29 Single-phase closed-loop network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.30 FVM representation of the single-phase closed-loop network . . . . . . . . . . . . . . 115

A.1 Connection Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

A.2 An Open Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

A.3 Closed Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

B.1 The double channel model control volumes . . . . . . . . . . . . . . . . . . . . . . . 139

ix

Nomenclature

Roman letters:

A Cross-sectional area [m2]

cp Heat capacity [J kg−1K−1]

cU Heat transfer coefficient [J s−1m−2K−1]

D Pipe diameter [m]

e Intrinsic energy [J kg−1]

e∗ Internal energy [J kg−1]

f Darcy-Weisbach friction factor [dimensionless]

g Gravitational acceleration [ms−2]

gc Gravitational constant [dimensionless for SI]

h Enthalpy [J kg−1]

J Drift Flux [ms−1]

k Momentum correction factor [dimensionless]

K Mechanical energy loss coefficient [dimensionless]

p Pressure [Pa]

Re Reynolds number [dimensionless]

q Volumetric flow rate [m3 s−1]

Q Pipe heat input [J m−2]

si Interfacial area, flow pattern dependent si = ωi∆x [m2]

swL Pipe area wetted by liquid, flow pattern dependent swL = ωL∆x [m2]

T Temperature [K]

T∞ Surrounding (ambient) temperature [K]

u Superficial velocity [ms−1]

v Velocity [ms−1]

V Volume [m3]

z Elevation [m]

Z Gas compressibility factor [dimensionless]

Greek letters:

αG Void fraction [dimensionless]

αL Holdup [dimensionless]

β T-junction branching angle [rad]

δk Distance of phase k from pipe wall [m]

x

ε Pipe roughness [m]

η Joule-Thomson coefficient [K Pa−1]

γG Gas gravity (or gas specific gravity) [dimensionless]

ΓG Rate of mass transfer from liquid-to-gas phase, per unit volume of CV, negative for

condensation and positive for evaporation [kgm−3 s−1]

ΓL Rate of mass transfer from gas-to-liquid phase, per unit volume of CV, positive for

condensation and negative for evaporation [kgm−3 s−1]

λ Branch to inlet mass intake ratio [dimensionless]

Λ Phasic volume fraction [dimensionless]

ωi Phasic interface wetted perimeter [m]

ωL Liquid wetted perimeter [m]

φk Phasic flow multiplier [dimensionless]

φ1 Heat transfer factor [m−1]

φ2 Potential and kinetic energy factor [Km−1]

Φ Two-phase loss multiplier [dimensionless]

ρ Density [kgm−3]

τi Interfacial stress [N m−2]

τw Pipe wall shear [N m−2]

θ Pipe inclination [rad]

σ Radial angle for zone of influence on pipe cross-section [rad]

Subscript:

G Gas phase

L Liquid phase

irr Irreversible

rev Reversible

xi

Acknowledgements

I would like to begin by expressing my deepest gratitude to my advisor Dr.Luis Ayala for bringing

me to Penn State and for his financial support. Needless to say, the four years I have been at Penn

State has been a totally life changing experience which has opened such a path for me that I could

not have thought of otherwise. For this great opportunity, many things I have learned from him and

through the course of this research that have fundamentally changed my understanding of numerical

modeling; for his guidance, understanding and patience with me; I am forever in his debt.

I am truly grateful to Dr.John Mahaffy from whom I have learned a great deal on the subject;

through two courses I took from him and our discussions, as well as additional references he provided

me with during early stages of the study. To a large extend, the approach taken in the study is

based on my learnings from him.

I would like to thank the committee members Dr.Turgay Ertekin, Dr.Robert Watson and Dr.Mirna

Macdonald for their time, interest in the study, valuable comments and input.

I would like to extend my special thanks to Dr.Turgay Ertekin for his understanding, support and

guidance through these years.

I would also like to thank EME staff assistants, and Penn State staff in general, for making my stay

smooth and free of obstacles.

Finally, I would like to express my deep gratitude to my parents Zerrin and Mustafa Alp for their

understanding, continuous and unconditional support overseas through these four years. Their

emotional support and financial assurance has always put my mind at ease and allowed me to focus

my full attention on my studies. As always, I am in their debt.

xii

1

Chapter 1

Introduction

Gaseous hydrocarbon mixtures are brought from wellheads to surface facilities via ‘gathering’ pipe

systems. After necessary treatment, natural gas is delivered to industrial or residential customers

several hundred kilometers away via ‘transmission’ pipelines. Main transmission lines are connected

to a ‘distribution’ grid at the gates of cities or industrial zones. Hence, from wellbore to the end

consumer, natural gas is delivered via complex and integrated network of pipes.

Two-phase flow of hydrocarbon mixtures in a network of pipes could be either a deliberate choice of

the operator or a result of inevitable flow conditions; such as ‘retrograde condensation’ of heavier

hydrocarbons in the case of natural gas lines. Two-phase flow is typically expected in the wellbore

and through out the surface gathering system when oil and gas, water and gas, or oil and water are

initially present and produced simultaneously from the reservoir. On the other hand, depending

on reservoir type, production stage and wellbore conditions liquid hydrocarbons (or water) may

condense in the pipeline as the second phase (or water may condense as the third phase, i.e. water

in emulsion with liquid hydrocarbons); even if inlet stream is single phase.

For on-shore operations, once separation and treatment is completed in centralized facilities close to

well sites, it is preferred to transport oil and gas in separate lines, as single phase fluids. In off-shore

operations, space limitations might impose a need for simultaneous transportation of oil and gas to

on-shore facilities located a considerable distance from the well site.

Branching T-junctions are essential components of small and large scale piping systems found in nat-

ural gas and oil pipeline networks, as well as various industrial applications. Two-phase flow through

branching tees (T-junctions) result in pronounced hydraulic losses and uneven phase separation

following the split of flow stream; ultimately causing profound effects on system performance and

quality of delivered fluids. Furthermore, for the practical purposes of system design and analysis,

almost all other types of junctions joining three or more pipes, as well as specialized sink/source

terms such as wellheads or supply/demand nodes, could then be represented with appropriate

arrangement of consecutive three-arm junctions; in particular tees. Consequently, quantifying

two-phase flow associated hydraulic losses and phase separation through T-junctions is a matter of

serious concern for the design and analysis of various pipe systems.

Industrial applications where two-phase flow split is typically observed include:

• Two-phase flow of refrigerants through distribution headers of multi-pass evaporators or

2

Figure 1.1: Branching T-junction

multi-system air-conditioners (Tae and Cho, 2006).

• Process plants where two-phase streams of chemicals, steam or air-water mixtures are delivered

to various locations in the plant via networks of pipes (Fouda and Rhodes, 1974).

• Conventional (fossil-fueled) power plants where steam is the ‘prime mover’.

• Nuclear reactor coolant loops of various design; i.e. pressurized (PWR), light (LWR) or boiling

(BWR) water reactors (Lahey, 1986) and associated small break discharge scenarios (Smoglie

et al., 1987). Two-phase flow is encountered within reactor coolant systems when water boils

due to large pressure drops; typically caused by accidental bursts of pipes in the primary

coolant loop, known as the loss of coolant accident (LOCA, Todreas and Kazimi, 1990).

• Circulation of geothermal fluids for heating purposes or transportation to geothermal power

plants via pipeline networks (Shoham, 2006).

• Steam injection networks for enhanced oil recovery (EOR) applications (Berger et al., 1997;

Jones and Williams, 1993).

• Condensate (water or liquid hydrocarbon) formation in natural gas gathering and transporta-

tion pipelines (Oranje, 1973).

• Functional phase separators and slug catchers for off-shore platforms and sea floor wellheads

(Margaris, 2007; Azzopardi and Rea, 2000).

It is important to note that through out this text, term ‘phase separation’ is distinctively used

to signify individual distribution of each incoming phase to the outgoing arms of a tee (Fig. 1.1),

while the term ‘phase distribution’ denotes the tendency of gas phase (void fraction) to occupy

certain parts of the pipe cross-section (e.g. due to gravity) –pronounced only in multidimensional

analysis because such details are lost with averaging in one-dimensional analysis (Lahey, 1990; cf.

Chapters 2 and 3). Finally, the term ‘phase split’ stands for the formation of the secondary phase

as a result of phase behavior response to changing pressure and temperature conditions.

In many industrial applications involving two-phase flow, formation of a lighter secondary phase (i.e.

gas) is expected, following a pressure drop in the flow stream that is initially all liquid. For instance,

3

gas phase evolves following pressure drops in oil pipelines. Hence, typically high liquid loading

conditions; i.e. predominantly liquid phase flow, is observed. On the other hand, condensation

of water in steam injection or distribution systems and formation of hydrocarbon liquids (the

condensate) in natural gas lines are low liquid loading conditions (predominantly gas phase flow)

where formation of an heavier secondary phase is observed (the liquid hydrocarbon).

In natural gas pipelines, according to gas composition, and particularly if heavier components are

abundant in the stream (i.e. wet gas), the phase envelope of the hydrocarbon mixture can partially

or entirely enclose the pipe operational region. If operational region is completely within the phase

envelope then two-phase flow would begin right at the very inlet of the system. If operational region

is partially enclosed by the phase envelope then two-phase flow is encountered in the sections of

pipe system where flowing pressure falls inside the phase envelope (Fig. 1.2).

Figure 1.2: Typical phase envelope for natural gases

Therefore, even though single phase conditions are prescribed at the inlet, and thus transportation

commences in entirely gaseous phase, (undersaturated) natural gases tend to drop hydrocarbon

liquids once pressure goes below hydrocarbon dew point conditions in the pipe. Thus, multi-phase

flow can prevail in natural gas transmission lines as well as gathering systems due to retrograde

condensation.

Simultaneous flow of natural gas and condensate is not efficient because presence of a secondary

phase (the condensate) decreases the deliverability of primary phase (gas), due to increased flow

(hydraulic) losses associated with decreased flow area available for the primary phase. Besides;

density, volume and calorific value of both phases vary along the pipeline as flow conditions change

because hydrocarbon phases in contact will continuously alter composition as a result of mass

transfer between the phases. Therefore, mass transfer and fluid re-distribution (i.e. phasic flow path

preference) in the pipeline network have significant influence on fluid and flow properties as well as

overall system performance.

There are two parts to the analysis of two-phase flow problem associated with condensate formation

in natural gas pipelines:

• Prediction and modeling of two-phase flow in straight sections of the network.

4

• Calculation of flow (hydraulic) losses and phase separation at T-junctions.

Within the context of pipe systems, one-dimensional inviscid conservation equations, namely the

Euler equations provide sufficient information for two-phase flow analysis, particularly for straight

pipe sections. Although a full set of multidimensional Navier-Stokes equations can always be

employed to get more detailed information on the flow (i.e. velocity profile, turbulence), such level

of detail is almost never required and would be computationally too demanding for the purposes of

studying large and complex pipeline networks.

In oil and gas industry, hydraulic losses through pipes are typically computed using ‘marching

algorithms’ that propagate the solution from pipe inlet to outlet. Governing one-dimensional Euler

equations are typically simplified and arranged in the form of either a pressure gradient ODE

(Shoham, 2006 – also see Sec. 2.1) or non-conservative set of ODEs (Ayala and Adewumi, 2003)

solved simultaneously at each marching step. However, marching algorithms are not suitable for

modeling closed (looped) networks for which flow related information downstream of the conduit, as

well as flow information from other pipes, should better be incorporated by means of simultaneous

solutions (cf. Sec. 2.2.4 and introduction of Chapter 3). Furthermore, particularly non-conservative

ODE based methods may suffer severe conservation problems with increased marching step size

(Ayala and Alp, 2008).

On the other hand, hydraulic losses at junctions are typically considered negligible in the grand

scheme of large and complicated oil and gas networks where junction volumes are insignificant

next to very long pipelines. Therefore, junctions are simply treated as pressure nodes of indefinite

volume, over which only mass conservation is satisfied and same ‘nodal’ pressure is prescribed as

the inlet or outlet pressure for all the connecting pipes of the junction.

In essence, a pipe network is treated like an electrical circuit where boundary conditions and pipe

‘resistance’ to flow (geometry and surface properties of the pipe: diameter, length, elevation change

and roughness) practically have the sole control on the flow split at junctions and, ultimately, on

the solution (cf. Sec. A.1). For example, at a diverging T-junction, flow split is merely dominated

by specified outlet pressures and flow resistance of the diverging arms while effects of split angle

and tee geometry are ignored. Hence, for example, if both outgoing arms have the same resistance

then flow is assumed to split equally provided that outlet pressures are equivalent too, regardless

of junction geometry and angle of split. This is apparent because, if exit pressures and outgoing

pipe resistances are equal and since the junction node (the ‘knot’) pressure (thus inlet pressures) of

the pipes is same, then same pressure drop over identical pipes can only be attained if flow rates

through the pipes are same too.

In fact, flow split at a T-junction is driven by competing (1) inertial forces (momentum) –in the

axial (inlet) direction, and (2) centripetal forces –in the branch direction, acting on the fluid stream.

The centripetal force is governed by junction geometry and pressure drop that draws the fluid to

the branch. Therefore, a slight difference in flow rates are in order even for the case of identical

outgoing arms and equal outlet pressures, that is unless split is parallel (i.e. no split angle). Again,

this difference in flow rates is practically insignificant for the purposes of single-phase network

5

analysis and typically tee-to-outlet pressure drops1 outweigh other factors, ultimately dominating

single-phase flow split at the tee.

For the case of two-phase flow, however, phases tend to split unevenly with changing proportions

at different flow conditions; inducing significant difficulty for the prediction of hydraulic losses.

Because of (1) the difference in inertia (density and viscosity), (2) associated difference in phasic

wall frictions and due to (3) contributing factors such as interfacial drag; phases are accelerated

differently under same pressure gradient. These differences also govern the inherent ability of a

phase to accomplish a change in flow direction. Thus, phase mass fractions going into branch can

differ significantly as disproportionate phase separation is driven by relative ease of phases to change

direction (Ballyk and Shoukri, 1990; Hwang et al., 1989; Shoham et al., 1987).

Because axial momentum of the lighter phase (i.e. gas) is lower, it responds readily to the pressure

drop in branch direction. While lighter phase is inclined to flow into the branch preferentially;

heavier phase (i.e. liquid) tends not to split at all and flow past straight to the run due to its

relatively high momentum. Consequently, higher fraction of the lighter phase goes into the branch.

However, with a slight change in the conditions (i.e. slight increase in the flow rate of the lighter

phase) heavier phase could entirely flow into the branch thus exhibiting the ‘flip-flop effect’ and

ensuing in the ‘route preference’ phenomenon (Fig. 1.3).

Figure 1.3: Route preference phenomenon

This sensitivity to flow conditions requires that impacts of overall tee geometry, wall surface

properties, inlet (or tee) flow mode and flow pattern on the flow split be properly accounted for in

order to accurately predict (1) disproportionate phase separation after the split and (2) inlet-to-run

and inlet-to-branch pressure changes. Here, it is important to underline that, while outlet pressures

still dominate overall flow split at a tee, afore mentioned factors significantly impact disproportionate

phase separation at the tee. Consequently, different phase separations due to a slight change in

1The pressure drop along the straight section, along the pipe that is connecting the ‘knot’ (tee node) to the outletnode

6

(any one of these) contributing factors can cause different inlet-to-run and inlet-to-branch pressure

drops at the tee, however outlet pressures remain the same.

The flip-flop effect attributed to the route preference of condensate in natural gas pipeline systems

is first reported by Oranje (1973) who studied uneven phase separation problem in natural gas lines.

Oranje (1973) concluded that the change in the direction of lighter phase (gas) entering the branch

induced a pressure drop inside the branch thus creating a suction force driving the secondary phase

(liquid) into the branch.

Later, Hong (1978) observed that as inlet gas flow rate increases (for a fixed liquid viscosity and

flow rate), the centripetal force acting on a unit mass of the liquid increases (perhaps associated

with a higher pressure drop). Consequently more liquid is drawn into the branch. While, for a

fixed gas flow rate, an increase in liquid rate raises the inertial force of the liquid without much

change in the centripetal force. Thus, less liquid tends to turn into branch. Also, an increase in the

liquid viscosity slows down the liquid, lowering axial momentum (inertial force) of the liquid phase;

thus, generally ensues in an increased branch liquid intake. Furthermore, Azzopardi and Whalley

(1982) reported that gas entering the branch drags the liquid film with it in the case of annular flow.

Hence, pressure drop and interfacial drag constitute the centripetal force acting on the fluid stream.

In other words, when branch gas intake is small, centripetal force acting on the liquid phase is also

small compared to the inertial force (the momentum in the inlet direction) of the liquid stream.

Therefore, liquid mostly flows straight through the T-junction directly into the run. When mass

fraction of the branch gas intake exceeds a threshold value (critical gas intake), liquid tends to enter

the branch preferentially.

It is well established with reference to the Bernoulli equation that, when single phase (also two-phase)

split takes place at a T-junction, flow continuing straight in the axial direction to the run experiences

a pressure rise (pressure spike) due to flow area expansion. On the other hand, a pressure drop

is typically experienced by the flow turning to the branch because the magnitude of associated

‘irreversible’ losses are usually greater than the Bernoulli type pressure rise due to flow expansion in

the branch.

Presence of recirculation zones (secondary flows) both in the run and branch, as well as the two-

dimensional nature of flow right after the split (at the T-junction), by default, demand that full

set of Navier-Stokes equations should be employed for multidimensional (i.e. 2D or 3D) analysis

and rigorous definition (and solution) of flow through T-junctions (Fig. 1.4). In fact, following the

incredible progress of computational power and modeling capabilities with Navier-Stokes equations,

two-fluid based multidimensional CFD is likely to become the de facto state-of-the-art for the

modeling of flow through T-junctions. Besides in-house two-dimensional Ellison et al., 1997;

Hatziavramidis et al., 1997 and three-dimensional Adechy and Issa, 2004; Issa and Oliveira, 1994,

flow split and phase separation problem has been studied with general purpose commercial CFD

software as well (Al-Wazzan, 2000; Lahey, 1990; Kalkach-Navarro et al., 1990). Typically, satisfactory

agreement with experimental observations is reported, at least in terms of capturing ‘broad features’

of the split. However, it has also been pointed out that certain details were not captured accurately

7

due to fact that particular phenomena is not represented inherently by the governing equation set

(Adechy and Issa, 2004).

Figure 1.4: Sample 3D computational grids for T-junctions (from Adechy and Issa, 2004; Lahey,1990)

Nevertheless, despite the enormous computational power of the day, modeling all the junctions of a

large, complex pipe network using 3D CFD is a computationally demanding task, and probably

not going to be a feasible approach in the near future for the purpose of modeling large networks.

Moreover, such degree of detail is seldom, if not at all, necessary for the purpose of modeling

network flow. Therefore, proper representation of two-phase flow split mechanism at junctions using

one-dimensional analysis is required.

An important aspect of modeling hydrocarbon condensate route preference in natural gas networks is

the inherent challenge of accounting for the compositional change of the stream. Overall composition

of delivered fluids could differ significantly from original composition after consecutive uneven phase

separations along the way to the delivery station.

Martinez and Adewumi (1997) developed a steady-state open network model by incorporating a

marching algorithm based two-phase flow model with Bernoulli equation based Double Stream

Model (DSM) of Hart et al. (1991) for T-junctions. Changes in stream composition after flow

split at junctions are accounted for using a so called ‘separator approach’ (cf. Sec. 4.13). DSM is

essentially a phase separation sub-model (see Secs. 3.2.1 and 3.2.3) for T-junctions and computes

outgoing phasic flow rates based on inlet conditions of the tee. In that regard, although a separate

model is employed at the tee, ‘forward’ workflow aspect of DSM is similar to marching algorithms

and works effectively in the analysis of open networks. In their study, Martinez and Adewumi (1997)

specified inlet pressures and all inlet/outlet gas flow rates as boundary conditions. Moreover, gas

phase split at each junction is either provided by the user or computed from already specified flow

rate data. The overall network solution has two parts; first, marching algorithm propagates the

solution from pipe inlet to a tee. Then, DSM determines the liquid flow rates to be prescribed as

inlet conditions for outgoing arms of the junction (in addition to specified gas rates). Afterwards,

again the marching algorithm advances the solution along the outgoing arms until a new tee or

a delivery station is reached. Consequently, approach is limited to the analysis of open networks.

Furthermore, in its original form, DSM is not completely adequate in addressing the route preference

phenomenon since flow split, in fact, depends on downstream conditions (pressures) as well.

8

Ottens et al. (2001) developed a finite-difference based transient T-junction model by combining

transient equations that relate pipe liquid level to phasic superficial velocities (derived from phasic

mass and momentum equations) with an advanced version of DSM.

In a recent study, Singh (2009) developed a finite-volume based one-dimensional two-fluid model

(see Chapter 4) for steady-state analysis of isothermal air-water split at T-junctions. Benefiting from

the FVM staggered grid, Euler equations are easily extended over the T-junction control volume

(Spore et al., 2001). Nevertheless, study of Singh (2009) had two inherently contentious aspects: (1)

Over specification of boundary conditions – void fractions are specified at the outlets, in addition to

the pressures. This is attributed to a possible misconception regarding the calculation of two-phase

pressure gradient term in the two-fluid model momentum equations. Please see pertinent discussions

regarding the use of void fraction, or holdup for that matter, when writing phasic pressure gradient

terms for the derivation of phasic momentum equations both in PDE and algebraic forms (cf.

Sec. 4.3.2). (2) Direct substitution of Gardel2 mechanical energy loss coefficients (intended for use

with Bernoulli equations) as momentum correction factors in the phasic momentum equations (cf.

Secs. 3.2.3 and 4.12). Utilized loss coefficients inherently account for the branching angle effect

(which is not accounted for by the non-directional Bernoulli equations otherwise) while momentum

equations already accounted for the directional momentum change with an additional term involving

the branching angle.

In this study, one-dimensional steady-state analysis of two-phase flow split at branching T-junctions

is formulated using the two-fluid model based integrated pipe approach. Governing Euler equations

are discretized over a staggered grid using the inherently conservative Finite Volume Method

(FVM, cf. Chapter 4). The staggered grid arrangement allows seamless extension of Euler equations

over the T-junction control volume and adjoining pipe flow equation set with the Double Stream

Model of Hart et al. (1991), essentially a phase separation sub-model, applied at the junction control

volume. In order to have a consistent model, phasic momentum equations of the tee CV are replaced

with gas phase Bernoulli equations at junction cell volume and required loss coefficients (K-factors)

are calculated using Gardel correlations (cf. Secs. 3.2.3). Focus is kept on the analysis of flow split at

horizontal, regular, 90o angle branching tee arrangements with single-phase, mist and stratified flow

patterns separately. Phases are assumed to be in thermodynamic equilibrium; implying mechanical,

thermal and chemical equilibrium between the phases. As a requirement of steady-state analysis

overall mixture composition remains constant along straight sections of the pipe system; i.e. along

each pipe and each arm of a T-junction. Interaction between the phases is represented by mass and

momentum transfer terms and interfacial friction terms. Due to thermal equilibrium assumption

an overall energy equation is utilized. Peng-Robinson EoS based thermodynamic model is used

for phase behavior predictions (Appendix C). Using a generalized Newton-Raphson technique,

numerical solution is obtained for all discrete points (control volumes) on all three arms of the

junction simultaneously, hence the name ‘integrated pipe’ approach. This approach proves to be

essential for closed-loop network analysis as it circumvents the need for marching through the flow

domain in one direction and readily accounts for changes in downstream conditions, particularly the

2Gardel, A. (1957). Les pertes de charge dans les ecoulements au travers de branchementes en te. BulletinTechnique de la Suisse Romande, 9,122 and 10, 143. (Ottens et al., 2001)

9

outlet pressures in this case.

In Chapter 2, solutions of single-phase steady-state flow in conduits and two-phase flow models are

introduced. In Chapter 3, flow split mechanism and uneven phase separation is discussed along

with a survey of available methods with emphasize on phenomenological and mechanistic junction

models. Appendix A presents conventional single-phase network analysis in oil and gas industry.

In Chapter 4, one-dimensional two-fluid model equations are derived based on FVM. Staggered

grid, boundary conditions, T-junction momentum equations and double stream model integration

is discussed along with the handling of phase appearance-disappearance. Appendix B has an

alternative derivation for a dividing streamline based ‘double channel’ T-junction model for FVM.

Chapter 5 includes results from numerical model comparison with single-phase steady-state analytical

solution, single-phase to two-phase transition handling in a single pipe and two-phase flow split

results for air-water and hydrocarbon two-phase flows through T-junctions.

10

Chapter 2

Flow In Conduits

Analysis of flow in conduits is fundamental to various engineering applications of different scales.

Evidently, quantitative description of flow and predicting response to perturbations (i.e. changes

in flow conditions) are essential for engineering designs. This requires determining the values of

governing flow parameters and fluid properties; such as flow rate or velocity, pressure, temperature

and fluid density. Study of flow in conduits is based on the principles of conservation of mass

(continuity), momentum, and energy. Several methods are available for the description of fluid flow

in pipes, with varying levels of detail, complexity and accuracy. Depending on the complexity of

the problem and level of detail desired for the analysis; an entire suite of conservation equations, or

a subset of them, is written along with appropriate simplifying assumptions in order to describe

the flow. It is then possible to quantify parameters governing the flow by solving the conservation

equations. Three-dimensional flow field is governed by Navier-Stokes equations; a comprehensive set

of coupled, nonlinear partial differential equations (PDEs) expressing conservation principles over

the flow domain; the continuum of conduit, and account for viscous and local volume (fluid particle)

forces. For the practical purposes of flow analysis in pipeline networks, where flow property changes

in the radial direction are not as important as property changes along the axial direction, one-

dimensional statements of mass, momentum, and energy conservation provide sufficient detail and

information on the flow. Typically, conservation statements can be generalized for one-dimensional

analysis through the following conservation differential form:

∂t

[Conserved

property p.u.v.

]+

∂x

[Flux of conserved

property p.u.v.

]=

External forcing function p.u.v.

(source/sink)(2.1)

Where; p.u.v. stands for ‘per unit volume’ and ‘flux of conserved property’ is defined as the amount

crossing the unit area, per unit time [(unit of conserved property) L−2 t−1].

Table 2.1 illustrates the relationship among the three conservation principles; where, intrinsic

energy of fluid is sum of its internal, kinetic, and potential energy. It is the energy associated

with fluid at each point of the pipe and accounts for fluid’s internal energy stored at molecular

level (e∗) and additional energy associated due to its velocity (kinetic energy) and location in the

gravitational field (potential energy). Hence, intrinsic energy per unit mass of the working fluid is:

e = e∗ +v2

2+ g∆zel (2.2)

11

Table 2.1: Conservation principles

Mass Momentum Energy

Conserved quantity m P = mv E = me

Conserved quantityp.u.v.

ρ v e

Flux ρv (ρv)v (ρe)v

Source/Sink –Sum of Forces

(Pres. + Fric. + Grav.)Heat input

2.1 Single-Phase Flow

For practical purposes, single-phase flow in wellbores and pipelines is considered unidirectional and

advective where primary concern is the change along the axial direction of flow. Three-dimensional

flow is typically approximated through a one-dimensional field by averaging characteristic flow

properties over the conduit cross-section. Averaging ensues in the omission of viscous and local

volume forces. Hence, Navier-Stokes equations reduce to one-dimensional Euler equations of inviscid

flow. In fact, viscous forces are most significant within the thin layer (the ‘boundary layer’) near

solid surfaces. Nevertheless, thickness of this layer is much smaller compared to the characteristic

length of the solid surface (i.e. pipe length); and volume forces are negligible next to global forces

driving the flow. Consequently, Euler equations of inviscid flow do not account for any dissipative or

diffusive transport (due to viscosity, mass diffusion, and thermal conductivity) along the direction of

flow and transport is assumed to be fully advective. However, effects of fluid viscosity and wall shear,

and wall heat transfer are accounted for through use of appropriate sink/source terms; momentum

and energy forcing functions.

One-dimensional Euler equations of inviscid flow are:

Mass Continuity:dρ

dt+

d

dx(ρv) = 0 (2.3)

Momentum Balance:d

dt(ρv) +

d

dx(ρvv) = −dp

dx− τw

πD

A− ρg sin θ (2.4)

Energy Balance:d

dt(ρe) +

d

dx(ρev + pv) = Q (2.5)

12

Energy balance is usually written in terms of enthalpy flux for an open system:

d

dt(ρe) +

d

dx

[h+

v2

2+ g∆zel

]v

)= Q (2.6)

Where;

h = e∗ +p

ρ(2.7)

Please note that no shaft work is considered for pipe flow. Also, no boundary work is present when

a control volume approach is adopted.

Euler equations could be solved analytically only in the limiting case where a perturbation in

the flow is assumed to occur instantaneously everywhere along the conduit (i.e. the rigid column

flow; Larock et al., 2000) or for steady-state conditions. For other flow conditions, Euler equations

must be solved simultaneously, which requires implementation of appropriate numerical schemes.

Numerical solutions provide approximate values of the flow parameters at discrete points in the

flow domain, which is generally sufficient for most practical purposes.

2.1.1 Single-Phase Steady-State Flow

It is customary to introduce simplifications in the conservation equations based on steady-state

assumptions and expected fluid phase behavior in the system. For steady-state conditions, the

governing equations simplify to:

Mass Continuity:d

dx(ρv) = 0 (2.8)

Momentum Balance:d

dx(ρv v) = −dp

dx− τw

πD

A− ρ g sin θ (2.9)

Energy Balance:d

dx

[ρ v

(e+

p

ρ

)]= cU

πD

A(T∞ − T ) (2.10)

Starting from steady-state momentum and mass balances, one obtains:

d

dx(ρv v) = ρ v

dv

dx+ v

0︷ ︸︸ ︷d

dx(ρ v) = −dp

dx− τw

πD

A− ρg sin θ (2.11)

From where following steady-state pressure gradient (pressure drop) equation is obtained:(dp

dx

)total

=

(dp

dx

)fric

+

(dp

dx

)elev

+

(dp

dx

)acc

(2.12)

13

Where:(dp

dx

)fric

= −τw πDA Pressure loss due to friction(dp

dx

)elev

= −ρ g sin θ Pressure loss due to elevation change (gravity)(dp

dx

)acc

= −ρ v dvdx Pressure loss due to acceleration (or kinetic energy change)

Equation 2.12 is customarily used to predict pressure profiles in pipelines when conditions of flow

are assumed to be isothermal. For non-isothermal flows, the energy balance is implemented. For

steady-state conditions, the energy balance yields:

d

dx

[ρ v

(e+

p

ρ

)]= ρ v

d

dx

(e+

p

ρ

)+

(e+

p

ρ

) 0︷ ︸︸ ︷d

dx(ρ v) = cU

πD

A(T∞ − T ) (2.13)

or

ρ vd

dx

(h+

v2

2+ g∆zel +

p

ρ

)= cU

πD

A(T∞ − T ) (2.14)

From where the steady-state enthalpy-gradient equation is formulated as:(dh

dx

)total

=

(dh

dx

)acc

+

(dh

dx

)elev

+

(dQ

dx

)(2.15)

with:(dh

dx

)acc

= −v dvdx Energy loss due to acceleration

(dh

dx

)elev

= −g sin θ Energy loss due to elevation change (gravity)

(dQ

dx

)= cU

πD(T∞−T )m Energy loss to the environment

For the case of oil (liquid) flow, density of oil is typically assumed constant (ρo ≈ constant), which

is largely true for liquids when no significant temperature change is expected. In this case, mass

conservation imposes v ≈ constant for flows through pipes of constant cross-sectional area or

v1A1 = v2A2 when pipe cross sectional area changes.

Since both density and velocity remain constant throughout an oil pipe of constant cross-sectional

area, there are no pressure losses due to kinetic energy changes and Eqn. 2.12 becomes:

dp

dx= −τw

πD

A− ρ g sin θ (2.16)

14

which, when integrated between two points separated by a distance L, becomes:

(pout − pin) = −πf8

q2LρLD5

L− ρLg∆z (2.17)

Or, re-arranged as:

qL =

√8

π

√1

f

√pin − pout + ρLg∆z

ρLLD

52 (2.18)

Where:

f = 4τw2

ρv2Darcy-Weisbach friction factor and f = 4fFanning

Equations 2.17 and 2.18 are typical pressure-loss and flow rate equations used for liquid-system

calculations. When the definition of friction factor for laminar flow conditions (f = 64/Re) is

substituted, the Poiseuille’s equation for laminar liquid flow through a pipe of uniform (circular)

cross-section is obtained.

For the case of natural gas (compressible) flow, fluid density is a strong function of pressure and

temperature. However, it is still customary to neglect contribution of kinetic energy change in the

pressure gradient equation for gas flow because kinetic energy change contribution is typically very

small compared to the friction and elevation terms; hence, one ends up with the Eqn. 2.16 again.

When pressure-gradient equation for gas flow is integrated between two points separated by a

distance L, the density dependency with pressure is accounted for by means of the real gas law

equation. For isothermal flow conditions, this integration yields a stronger dependency of flow rate

on pressure; i.e. squared pressures (Kumar, 1987):

qG = C

(Tscpsc

)√1

f

√p2in − p2

out es

γGTavgZavgLeD

52 (2.19)

Where:

s =2MWair γG ∆z

1000TavgZavgR

g

gcPipe inclination dependent coefficient, s = 0 for horizontal pipe

[dimensionless]

Le = Les − 1

sThe ‘equivalent’ length definition based on s [m]

C =

√1000 gcπ2R

16MWairA unit system dependent constant [m2s−2K−1]

e The base of natural logarithm [≈ 2.718282]

γG Gas gravity (or gas specific gravity) [dimensionless]

15

On the basis of Eqn. 2.19, several gas flow equations have been proposed throughout the years.

Steady-state empirical gas flow equations such as Spitzglass, Weymouth, Panhandle-A, Panhandle-B,

and the IGT/AGA differ in the functional form simply because each equation utilizes a different

expression for the calculation of friction factor in Eqn. 2.19.

The analytical expressions derived above (Eqn. 2.18 and 2.19) allow calculation of pipe pressure

drops for isothermal, single-phase flow systems. However, flow of liquid and gases may not be

well represented by an isothermal model, especially when fluid inlet temperature greatly differs

from ambient temperature, which creates non-adiabatic conditions throughout the length of the

pipe. Even during adiabatic conditions, natural gas pipelines tend to cool with distance (the‘Joule-

Thomson cooling’ effect), while oil lines tend to heat – because of the very distinct Joule-Thomson

coefficient values of liquids and gases. Therefore, for non-isothermal, steady-state, single-phase flow,

the enthalpy-gradient equation is implemented:(dh

dx

)total

=

(dh

dx

)acc

+

(dh

dx

)elev

+

(dQ

dx

)(2.20)

which, integrated between two points in the pipeline separated a distance x, yields the following

explicit expression for the calculation of temperature change as a function of pressure drop:

T (x) = T∞ + (T0 − T∞)e−φ1x + (1− e−φ1x)

φ1

dp

dx− φ2

φ1

)(2.21)

Where:

T0 Fluid temperature at the pipe inlet [K]

T∞ Surrounding (ambient) temperature [K]

φ1 =cUcp

πD

mHeat transfer factor [m−1]

φ2 =v

cp

dv

dx+g

cp

dzeldx

Potential and kinetic energy factor [Km−1]

Equation 2.21 gives the explicit dependency of temperature on (1) pressure drop (dp/dx), (2) heat

transfer from the environment (φ1), and (3) potential and kinetic changes (φ2). When last two

factors two are neglected, Eqn. 2.21 collapses to the following equation (Coulter, 1979):

T (x) = T∞ + (T0 − T∞)e−φ1x + (1− e−φ1x)η

φ1

dp

dx(2.22)

When one further assumes that thermodynamic changes on fluid pressure do not affect the tempera-

ture of the fluid (i.e. Joule-Thomson coefficient is close to zero); Eqn. 2.22 can be further simplified

to:

ln

(T (x)− T∞T0 − T∞

)= −φ1x (2.23)

Equation 2.23 neglects the changes in fluid enthalpy due to pressure, which can be a reasonable

assumption for liquids. For the derivation of Eqn. 2.23, fluid’s enthalpy changes are implicitly

16

evaluated through the simplified expression dh = cpdT instead of the more rigorous dh = cpdT −ηcpdp.

For simultaneous prediction of pressure and temperature changes along a pipeline, one may resort

to using an analytical flow equation for dp/dx calculations and the analytical energy equation for

dT/dx calculations, working concurrently. However, such approach decouples the combined effect

that pressure and temperature changes can have on fluid density. This might be acceptable for oil

or liquid flows but not for natural gas flows.

2.2 Two-Phase Flow

Two-phase flow occurs during production and transportation of oil and gas, where formation of

the condensed liquid (hydrocarbon or water) is determined by the overall mixture composition and

local P-T couple, according to thermodynamic phase behavior.

In order to adequately describe two-phase flow, new variables need to be defined in addition to the

fundamental single-phase flow parameters pressure, velocity and temperature:

Holdup is the liquid phase volume fraction within a volume element.

αL =VL

VL + VG=ALA

(2.24)

Void fraction, on the other hand, is the gas phase volume fraction in a given volume element.

αG =VG

VL + VG=AGA6= qGqL + qG

(2.25)

In that regard, these two parameters are the analogs of liquid and gas phase saturations in reservoir

engineering.

αL + αG = 1 (2.26)

Phasic superficial velocity is defined as the ratio of phasic volumetric flow rate to the whole

conduit cross-sectional area.

uk =qkA

(2.27)

Phasic velocity (intrinsic) is the ratio of phasic volumetric flow rate to the cross-sectional area

available for the phase in conduit.

vk =ukαk

(2.28)

Mixture velocity is the total volumetric flow rate of both phases per unit area.

vmix =qG + qLA

(2.29)

17

Slip velocity is the velocity of gas phase relative to liquid phase.

vslip = vG − vL (2.30)

Drift velocity, however, is the velocity of each phase relative to mixture velocity.

(vk)drift = vk − vmix (2.31)

Drift flux is the per unit area flow rate of each phase, relative to mixture velocity.

Jk = αk(vk − vmix) (2.32)

Phasic volume fraction is distinguished from holdup or void fraction as the ratio of phasic

superficial velocity to the mixture (sum of phasic) superficial velocities.

Λk =ukumix

=uk

uG + uL=

qGqL + qG

(2.33)

Please note that holdup (or void fraction) is equal to phasic volume fraction (αk = Λk) only for

no-slip condition (i.e. vG = vL).

Quality is the ratio of gas mass flow rate to the total mass flow rate.

x =mG

mG + mL=mG

m(2.34)

2.2.1 Flow Patterns

Gas-liquid two-phase flows are typically categorized according to the governing ‘flow pattern’;

distinct geometric arrangements of phasic volumes, the structure and continuity of the phasic

interface in the conduit. Various flow patterns (Fig. 2.1) fall under three major flow types (Shoham,

2006):

1. Dispersed flows: Bubbly (gas bubbles) or mist (liquid droplets) flow patterns, where particle

size secondary phase is ‘dispersed’ in the continuous phase.

2. Transitional flows: Slug, Churn turbulent or dispersed annular flow patterns.

3. Separated flows: Stratified, Annular flow patterns, where phases are assumed to flow in their

separate channels in the conduit.

Flow pattern is a function of phasic flow rates, fluid properties (i.e. density, viscosity and surface

tension), and pipe geometry (i.e. diameter, inclination angle). Thus, different flow patterns appear

at different inclination angles of the pipe. Fortunately, it is possible to establish a family of flow

regimes applicable at different inclination angles with little or no changes at all. However, since

pipeline networks are practically composed of near horizontal pipes and liquid phase (oil) volume

18

fraction is typically well below 50% (low liquid loading condition) in gas condensate lines, a limited

class of flow regimes is adequate for the purpose of this study. Figure 2.1 shows the flow patterns

near horizontal gas condensate lines commonly feature; (1) stratified smooth, (2) stratified wavy,

(3) annular mist and (4) mist (spray) flow.

Figure 2.1: Typical flow patterns observed in near horizontal gas condensates pipelines

Stratified flow occurs when both phases are flowing at slow rates and separated by gravity. The

interface between the phases becomes wavy as gas flow rate increases and establishes stable waves;

i.e. the stratified wavy pattern.

Annular flow appears at high gas flow rates such that gas phase presses down the liquid flowing

at the bottom of the pipe to spread as a thin film over the entire pipe wall. Liquid phase is usually

thicker at the bottom of the pipe and gas phase flows at the center of pipe; gas core, surrounded by

this liquid film. Some liquid may entrain gas phase as droplets, establishing the annular mist flow

pattern.

Mist flow occurs at very high gas flow rates with turbulence, and liquid phase completely entrains

the continuous gas phase. High gas flow rate prevents liquid droplets from precipitating at the

bottom of the pipe to establish the stratified pattern.

A major issue of two-phase flow modeling is the occurrence of diverse flow patterns throughout a long

conduit, and associated characteristic effects on the flow. Accurate prediction of flow patterns in

the conduit is essential because several flow parameters; such as frictional pressure drop, interfacial

area and associated drag-slip, heat and mass transfers are significantly affected by the governing

flow pattern.

Flow patterns are typically identified according to visual observations and flow regime maps are

prepared based on mass flow rates, void fractions and velocities as well as dimensionless variables

and correction factors, in order to make the maps more general. In fact, term ‘flow pattern’ denotes

the visual distinction among various flow topographies while ‘flow regime’ indicates how the flow

19

patterns affect the system. Then, different flow patterns does not necessarily require a change

in the adopted two-phase flow model (or utilized closure relation ships for that matter) unless a

flow regime change is present (Kleinstreuer, 2003); hence the name ‘flow regime’ map. However,

throughout this text both terms are used interchangeably.

Flow regime maps used in the petroleum industry are detailed and comprehensive for a variety of

flow conditions with different mapping criteria. These maps, such as the one presented in Fig. 2.2,

are typically a function of the superficial velocities of each phase and the inclination angle of the

pipe.

Figure 2.2: Sample flow regime map for horizontal pipes (Shoham, 2006)

Downside to the flow regime maps is that they are usually valid for specific flow conditions (i.e.

certain range of inclination angle or flow rates) and it might be necessary to switch maps with

varying flow conditions in a long pipe. Also, transition zones (or interpolation regions) may not

always be present on the map; hence, non-physical, abrupt changes in the flow regime could be

admissible according to the regime map.

2.2.2 Two-Phase Flow Models

In the early designs of gas pipelines, additional pressure drop due to presence of a second phase

was accounted for by introducing an efficiency factor in Eqn. 2.19; the steady-state, analytical flow

equations such as Weymouth, Panhandle etc.

Later on, efforts were focused on extending single-phase pressure gradient equation (Eqn. 2.12, cf.

Sec. 2.1.1) for two-phase flows using empirical correlations. Such methods include the homogeneous

equilibrium model (HEM), the separated flow model (SFM) and the drift-flux model;

all of which, through simplifying assumptions, treat the two-phase flow problem using methods

developed for single-phase flow analysis and do not leave much room for extensive analysis of

20

different flow patterns. Consequently, such approaches have limited applicability and an analytic

description of the process is required for modeling purposes.

Following the progress made in the nuclear industry, efforts to define two-phase flow using funda-

mental conservation laws gained momentum; leading to the development of the two-fluid model

which considers the phases separately using two sets of conservation equations and allows for the

analysis of transients as well (Ishii and Hibiki, 2003; Todreas and Kazimi, 1990).

In fact, HEM, SFM and Drift-Flux models form the group of ‘flow mixture models’ rather suited for

well-mixed dispersed flows (pseudo single-phase, or pseudo two-phase flows for that matter). On the

other hand, the two-fluid model, while has the best overall applicability, is especially required for

the analysis of separated flows. The two-fluid model constitutes the bulk of ‘mechanistic’ modeling

effort which stands for the ultimate, flow regime dependent modeling approach (cf. Sec. 2.2.3).

Homogeneous Equilibrium Model

Homogeneous equilibrium model (HEM, or the homogeneous no-slip model) treats two-phase flow

as if both phases are combined into a homogeneous, pseudo single-phase with gas-liquid mixture

properties; i.e. mixture density and viscosity, based on the assumption that phases are well-mixed

and in thermodynamic equilibrium. Hence, flow is modeled using a modified single-phase pressure

gradient equation (Crowe, 2006; Shoham, 2006).

− dp

dx= −

(τwπD

A

)mix

− ρmix g sin θ + ρmixvmixdvmixdx

(2.35)

The model assumes no-slip condition (vG = vL = vmix) so that phases move at the same velocity.

At no-slip condition αk = Λk. Fluid properties of the pseudo single-phase are volume fraction

dependent average of both phases. Thus, gas-liquid mixture (homogeneous) density is defined as:

ρmix = ΛGρG + ΛLρL = αGρG + αLρL (2.36)

And, gas-liquid mixture (homogeneous) viscosity is:

µmix = ΛGµG + ΛLµL (2.37)

Then, pressure gradient equation becomes:

− dp

dx= −

(τwπD

A

)G

−(τwπD

A

)L

− (αGρG + αLρL) g sin θ +d

dx

[ρG

u2G

αG+ ρL

u2L

αL

](2.38)

Similar modifications are applied for the derivation of mixture energy equation (Shoham, 2006).

21

Separated Flow Model

Separated flow model (or the Lockhart and Martinelli Model) observes two-phase flow problem as

simultaneous flow of each phase in their separate channels (cross-sectional area occupied by each

phase) within the conduit, based on the hydraulic diameter concept. Model is typically limited to

the calculation of frictional pressure drops in horizontal pipes. Two-phase pressure drop is calculated

by multiplying phasic superficial pressure gradient with a two-phase flow multiplier. (Shoham, 2006;

Crowe, 2006).

(dp

dx

)fric

= φ2L

(dp

dx

)S L

= φ2G

(dp

dx

)S G

(2.39)

and (dp

dx

)fric

=

(dp

dx

)k

= φ2k

(dp

dx

)Sk

(2.40)

Where:(dp

dx

)S k

Phasic superficial pressure gradient(dp

dx

)k

Pressure drop induced by phase k

φ2G = 1 + CX +X2 Gas phase multiplier [dimensionless]

φ2L = 1 +

C

X+

1

X2Liquid phase multiplier [dimensionless]

X =φGφL

The Lockhart and Martinelli parameter [dimensionless]

Correlations for X are derived from empirical data. Constant C takes on different values according

to flow types of both fluid, i.e. C = 12 for laminar liquid flow and turbulent gas flow; and holdup is

calculated by an appropriate correlation; i.e. αL = 1− (1 +X0.8)−0.378 (Chisholm, 1967). φk is a

function of phasic hydraulic diameter (flow area available for phase k).

Drift-Flux Model

Although two-phase stream is treated as an homogeneous mixture, based on local flow conditions,

the drift-flux model assumes either a constant or correlation dependent slippage between phases.

Therefore, flow is a function of void fraction and velocity difference between the phases, which gives

the model a better handle on flow pattern dependent modeling.

Pauchon et al. (1994) studied the two-phase flow capabilities of the drift-flux-based code TACITE,

developed for the analysis of transients in multiphase pipeline networks. The code employs separate

continuity equations for gas and liquid phases while mixture momentum and mixture energy equation

are utilized. To account for the missing momentum information, steady-state closure relationships

are used based on flow pattern. Flow regime transitions are based on the continuity of variables so

that solution for different flow regimes are identical through the transition region.

22

Two-Fluid Model

A rigorous approach to the two-phase flow problem is the two-fluid model where a separate set

of conservation equations the Eulerian frame are employed for individual phases; allowing each

phase its own velocity (and pressure or temperature) field based on the ‘interpenetrating continua’

assumption. Mass and heat transfer between phases is represented with appropriate terms in the

governing equations of both phases, without the presence of such interfacial exchange terms, phases

are essentially independent. These interaction terms, in fact, determine the degree of coupling

between phases. Strongly coupled phases tend to be in mechanical and thermal equilibrium (Ishii

and Hibiki, 2003). However, this does not necessarily imply a chemical equilibrium between the

phases. Therefore, utilization of a kinetic model is required in order to account for the kinetics of

mass transfer between the phases, depending on the problem; fluid properties and flow conditions.

Most important aspect of the two-fluid model is that it can account for transient (dynamic) and

non-equilibrium (mechanical and/or thermal, in addition to chemical) conditions by describing phasic

pressure and temperature fields. Furthermore, model is applicable to particle scale three-dimensional

and two-dimensional analysis using Navier-Stokes equations as well as larger scale one-dimensional

pipe flow analysis using Euler equations.

Although better suited for separated flows, the two-fluid approach inherently considers the dispersed

(particle) phase a ‘continuum’ as well, which may create conceptual difficulties for certain flow

conditions. Regardless of the analysis scale, some sort of averaging method is required for the

derivation of governing equations; i.e. volumetric averaging. The averaging process, while bridging

the molecular scale with the continuum model scale, generates new unknowns. In order to close

the equation set supplementary information is required; i.e. suitable constitutive equations that

describe the dynamics of individual phases and their interactions. (Kleinstreuer, 2003). Success of

the model greatly depends on required constitutive equations (closure relations) which incorporate

flow pattern associated effects.

The two-fluid model can be extended to account for multiple phases (hence the multi-fluid model) or

can account for entrained liquid droplets in the gas phase (or gas bubbles in the liquid phase) using

additional field equations; i.e. a third mass and/or momentum balance equation. In the nuclear

industry, typically a 8 equation version of the two-fluid model is utilized for the analysis of coolant

loops; accounting for 3 mass, 3 momentum and 2 energy conservation equations. Additional mass

and momentum equations describe the liquid droplets traveling (entrained) within the gas phase

(Frepoli et al., 2003; Spore et al., 2001). The most comprehensive commercial code available in the

petroleum industry; two-fluid model based OLGA, uses a similar approach and employs a 7 equation

version of the two-fluid model with a single (mixture) energy equation for analysis of various pipe

systems such as wellbores, oil and gas transmission lines, and networks (Bendiksen et al., 1989).

Derivation of governing equations for the one-dimensional, five equation two-fluid model (single

pressure and temperature; assuming mechanical and thermal equilibrium) used in this study are

presented in Sec. 4.3.

23

2.2.3 Mechanistic Modeling

‘Mechanistic’ modeling effort signifies using different (one-dimensional) flow models along a pipe,

based on predicted flow pattern for a particular section of the pipe. General approach is first to

determine the flow pattern for a section and then utilize the best suited model for accurate pressure

drop calculation.

The two-fluid model by itself could be considered a ‘mechanistic model’ through use of proper closure

relations that account for different flow patterns. However, homogeneous or drift-flux models might

be better suited for dispersed flows under certain conditions; or, a specialized slug flow model might

be necessary in order to accurately calculate pressure drop through a particular section of the pipe,

for instance. This is because, success of the seemingly general-purpose ‘mechanistic’ two-fluid model

depends on the availability of proper closure equations. Then, in a broader sense, a ‘mechanistic’

approach may as well require using different flow models (i.e. two-fluid, drift-flux etc.) together.

Overall aim of mechanistic modeling is to provide proper tools better describing the physics,

the hydrodynamics of individual flow patterns, focusing on the accuracy of pressure drops and

void fraction (or hold up) predictions. A comprehensive review of the current state-of-the-art of

mechanistic modeling in the petroleum and natural gas industry has been recently presented in the

text by Shoham (2006) for two-phase flow in oil and gas pipelines.

The fact that different two-phase flow models are to be used side by side requires that adopted

solution method allows appropriate transition between different models as well as prediction of flow

patterns along the pipe. The ‘marching algorithm’ approach, discussed next in Sec. 2.2.4, is suited

for this purpose and used frequently for steady-state, mechanistic analysis of two-phase flow in oil

and gas pipelines and wellbores.

2.2.4 Marching Algorithms

Marching algorithms solve for pressure drops over short increments of pipe, propagating the solution

step-by-step, hence ‘marching’ from inlet all the way to the outlet (Shoham, 2006).

The need for marching algorithms is three-fold:

1. Allows for the implementation of a mechanistic modeling approach that utilize separate flow

models; i.e. first, the flow pattern is determined for next pipe increment and then appropriate

model is used for pressure drop calculation

2. Because solution is focused only on a single increment, computational requirements (i.e.

storage) are significantly low (compared to methods involving simultaneous solution of several

points in a pipe). This becomes especially important considering that transportation pipelines

can be quite long.

3. Transitions in flow patterns are more precisely traced as solution marches forward with small

steps, albeit, at the expense of longer computational times.

24

Ayala and Adewumi (2003) presented a detailed account of how marching algorithm calculations

performed for two-fluid model. Governing PDEs (Euler equations) of one-dimensional, steady-state

two-fluid model are typically arranged into ordinary differential equations (ODEs, non-conservative

form) by expanding the derivatives. ODEs are discretized using finite difference method (FDM) and

numerically integrated over short increments of pipe, in a sequential way, using the Runge-Kutta

technique.

As opposed to typical FDM based approaches where solution is obtained for all the discrete points

simultaneously, Runge-Kutta (RK), ODE solver numerical procedures are suited for space marching

algorithms. In a RK numerical computation, governing equations are rearranged in the explicit

dU/dx ODE form:dU

dx+dF

dx= C (2.41)

Where, U is the vector of primary unknowns (that numerical solution is obtained for). For single-

phase gas flow calculation, for example, U = [p, vG, T ] and then dU/dx becomes:

d

dx

p

vG

T

=

vG

(∂ρG∂p

)T

ρG vG

(∂ρG∂T

)p

1 ρGvG 0

−ρGvG ηG (cp)G ρGv2G ρGvG (cp)G

−1 0

−FfG − FgGQG

(2.42)

In Eqn. 2.42, equations are properly arranged in order to eliminate zeros in the diagonal of

coefficient matrix. Primary unknowns for five-equation formulation of the two-fluid model are

typically U = [p, vG, vL, αk, T ].

The calculation begins at the inlet of the pipe, where U = U0 and RK algorithm calculates the

outlet conditions for the end of first pipe segment based on the following 5th order estimate:

U(x+ ∆x) = U0 +6∑i=1

aFi FFi (2.43)

Where, quantities FFi are computed as functions of vector dU/dx:

FFi = ∆x · dUdx

∣∣∣∣U0 +

i−1∑j=1

dFijFFj

1 ≤ i ≤ 6 (2.44)

Where, aFi and dFij are constant coefficients of the RK method.

Once outlet conditions for a given block are calculated, those values are assigned as the inlet

conditions for the next block and the algorithm keeps on going until end of the pipe is reached. This

constitutes the so called ‘marching algorithm’ via the RK method. This approach can suffer mass

and energy conservation problems since conservative property of the equations has been altered

with the non-conservative ODE arrangement (Ayala and Alp, 2008).

25

Marching algorithms (also known as the pressure traverse computing algorithms) form the basis of

most steady-state two-phase flow software and are presented by textbooks as the standard techniques

used in steady-state two-phase flow analysis for oil and gas pipelines (Shoham, 2006).

While applied successfully for the analysis of typical two-phase, steady-state flow in single pipes

and perhaps open networks (i.e. without loops; Martinez and Adewumi, 1997), space marching

algorithms are not suitable for two-phase single pipe transients, modeling counter-current flow

(re-flood) analysis where both downstream and upstream flow information in the conduit should be

accounted for via simultaneous solution. Moreover, two-phase, steady-state closed network (i.e with

loops) analysis, as well as open network analysis with marching algorithms can be a challenging

task as flow split and uneven phase separation is greatly influenced by pressures downstream of a

T-junction.

26

Chapter 3

Flow Split at T-Junctions

Two-phase flow in a single oil and gas pipeline or wellbore has been thoroughly studied over the

years; a comprehensive text is made available by Shoham (2006), but the problem of multiphase

flow in networks of multiple pipes and loops is still considered largely unresolved due to challenge of

uneven phase splitting at pipe junctions.

Single-phase, compressible or incompressible flow in networks is typically modeled using one-

dimensional, steady-state, analytical expressions which are closed form equations (hydraulic models)

derived from fundamental principles of mass and energy conservation applied over the entire pipeline

length (Sec. 2.1.1). With an analogy to electric circuits, these flow equations are coupled with

Kirchhoff laws to obtain steady-state solutions. Kirchhoff laws stand for conservation of mass at

nodes (junctions of pipes) and conservation of energy around loops. A more detailed review of

single-phase flow in pipe systems is given in Appendix A.

In such network models based on single-phase analysis (i.e. Mucharam and Adewumi, 1991),

two-phase flow effects are loosely accounted for by inducing higher pressure drops in the two-phase

lines via definition of loss coefficients (or pipe ‘efficiency’ factors) or simplified two-phase flow models

(i.e. Beggs and Brill, 1973). Nevertheless, such methods are known to be inadequate for accurate

assessment of overall system performance; besides neglecting secondary phase ‘route preference’ due

to uneven phase separation at junctions .

Use of marching algorithms, within the system wide iteration scheme, for pressure drop calculations

between the nodes (junctions) of a network could be another option. However, this would probably

require ‘marching’ each pipe over again at every system wide iteration step towards satisfying

Kirchoff laws, several times before system converges to a solution. Regardless, this approach

would again fail to acknowledge establishment of preferential secondary phase paths and associated

compositional (or quality) changes of the fluid stream unless the issue of uneven phase separation

at junctions is properly addressed. More elaborate discussion on this matter (why uneven phase

split problem still not accounted for) follows next in this chapter.

A viable approach is the simultaneous solution of two-fluid model based governing equations for

all the discrete points, or control volumes to be more precise, over the computational grid of pipe

network, with special attention to junction control volumes. A good example to such modeling

efforts is the thermal-hydraulic code TRAC which provides a distinct T-junction component used

for the analysis of nuclear reactor coolant loops (Spore et al., 2001).

27

Treatment at junctions is vital to the overall performance of network model since uneven phase

separation and associated ‘route preference’ is governed by split mechanisms at junctions. For the

practical purposes of system design and analysis, almost all other types of junctions joining three or

more pipes, as well as specialized sink/source terms such as wellheads or supply/demand nodes,

could then be represented with appropriate arrangement of consecutive three-arm junctions; in

particular tees.

There are two types of three-arm junctions (Fig. 3.1):

1. T-junctions (tee), where a second line is connected to the body of a straight main line; i.e.

two of the arms are parallel (make 180o angle) with each other,

2. Y-junctions (wye), where no straight main line is present but all the arms make an angle less

than 180o with each other.

tee wye

Figure 3.1: Three-arm junctions: tee and wye

Also, based on the directions of flow through its arms, there are two kinds of three-arm junctions:

1. Dividing (also diverging or separating)

2. Combining (also merging or joining)

In particular, a dividing T-junction is called a ‘branching tee’ when flow enters the main line

through the ‘inlet’, exits through both the ‘run’ (other end of the main line) and the ‘branch’

(side arm). If fluid is injected into the side arm as well then it is called ‘combining tee’. On the

other hand, if two opposing streams coming from both ends of the main line are combined to exit

through the side arm then it is a ‘counter–combining tee’. Finally, if flow enters the junction

only through the side arm and exits through both ends of the main line it is an ‘impacting tee’

(or counter–dividing tee, Fig. 3.2).

Branching T-junctions are essential components of various piping systems of different scales. Two-

phase flow through branching tees result in pronounced hydraulic losses and uneven separation of

phases with the split of flow stream; ultimately causing profound effects on system performance and

quality of delivered fluids. Consequently, quantifying two-phase flow associated hydraulic losses and

phase separation through T-junctions is a matter of serious concern for the design and analysis of

various pipe systems.

A counter-intuitive result of the flow split at branching tees is a sharp pressure increase in the

run direction, right after the split (Hart et al., 1991; Lahey, 1990; Ballyk and Shoukri, 1990).

Figure 3.3 shows typical pressure profile at a branching T-junction following the single-phase

28

Figure 3.2: T-Junction classification based on flow directions

flow split. Presence of such a ‘pressure spike’ is easier to observe based on Bernoulli equation

analysis; according to which an increase in the flow area (A1 < A2 for ρv1A1 = ρv2A2) ensues in

the deceleration of the flow (v1 > v2) causing an increase in the flowing pressure:

1

2ρv2

1 + ρgz1 + p1 =1

2ρv2

2 + ρgz2 + p2 (3.1)

Figure 3.3: Pressure profiles after single-phase flow split at a branching T-junction

An earlier approach that is worth mentioning is Meier and Gido (1978)’s Method of Characteristics

(MOC) based approach for single-phase flow split at an impacting tee (thus, can be extended

to branching tees easily). Mass and momentum conservation equations are first converted to

characteristic equations. Then, model is built around the idea that pressures at all three ends of a

central tee volume of ‘undefined extent’ (i.e. no real volume) are connected via a pressure drop

equation to a single (nodal) pressure that prevails at the center of the tee. Meier and Gido (1978)

model required definition of 3 separate K-factors (loss coefficients) to account for pressure drops

between the central tee node and each of the nearest nodes of all 3 arms. The K-factors of Bernoulli

29

type mechanical energy equation, are derived from experimental data.

Ki =p− pi12ρivivi

i = 1, 2, 3 (3.2)

Where, i denotes arms of the tee.

When modeling two-phase flow in a network of pipes, it is essential to accurately model the flow at

T-junctions as they can effect fluid distribution (i.e. route preference phenomenon Oranje, 1973)

and associated downstream pressure losses in the network significantly. Typical focus of T-junction

modeling has to do with the associated problem of uneven phase separation at the junction. Liquid

preferential routes determine bottlenecks and impact quality and quantity of gas received at delivery

stations.

3.1 Two-Phase Flow Split

Outlet (downstream run and branch) pressures of a branching tee have major influence on phase

separation, as well as flow split. Consequently, there exists a correspondence, an interdependence

between phase separation and inlet-to-run and inlet-to-branch pressure changes, based on inlet flow

pattern and tee geometry.

Factors affecting two-phase flow split at a T-junction can be summarized as:

• run and branch outlet pressures

• inlet-to-run, inlet-to-branch pressure changes

• inlet stream phasic kinetic energy and/or inertia (momentum in the axial direction)

• inlet flow pattern

• inlet flow mode (turbulent, laminar or transition)

• overall junction geometry (arm inclinations and diameters, split angle, joint corner geometry)

• wall surface properties

• interfacial drag

Ideally, a T-junction model should be consistent and complete in capability; predicting inlet-

to-run and inlet-to-branch pressure changes as well as uneven phase separation at a tee with

reasonable accuracy and consistency. Therefore, there are two components, called ‘sub-models’,

to a T-junction model, that determine overall success and prediction capabilities:

1. Pressure change sub-model

2. Phase separation sub-model

Sub-models are usually tested by substituting experimental data. For instance, experimental phase

distribution data is substituted in the governing equations of pressure change component to match

pressure change information that is typically obtained through extrapolation of observed pressure

profiles for developed flow along the run and branch, thus establishing the pressure change component.

Similarly, pressure change data (often extrapolated from developed flow profiles) is substituted

30

in the governing equations for phase separation component in order to match experimental phase

separation data. However, extrapolation of pressure profiles in order to obtain pressure change data

and separately substituting experimental data in sub-models has the potential danger of leading

to a weak coupling between two components of the model, rendering seemingly separate ‘pressure

change’ and ‘phase separation’ models. Consequently, individual success of sub-models does not

necessarily guarantee consistency for the overall model.

One-dimensional modeling efforts for the analysis of two-phase flow split at T-junctions fall under

three categories (Azzopardi, 1999; Peng and Shoukri, 1997; Lahey, 1986):

1. Empirical correlations; i.e. phase separation sub-model of Seeger et al. (1986).

2. Phenomenological and Mechanistic1 models; i.e. Penmatcha et al. (1996); Ballyk and Shoukri

(1990). Following the discussion of Lahey (1986), approaches such as fluid mechanics based

(hence ‘mechanistic’) models of Penmatcha et al. (1996); Saba and Lahey (1984); Fouda and

Rhodes (1974), are considered under this category.

3. Two-fluid models (control volume based); i.e. TRAC family of thermal-hydraulic codes

(Spore et al., 2001; Steinke, 1996). Although Peng and Shoukri (1997) classifies only two and

three-dimensional models as two-fluid models (i.e. Adechy and Issa, 2004; Ellison et al., 1997),

TRAC extends one-dimensional2 two-fluid model approach over the tee control volume.

Empirical correlations represent the two-phase flow split in a much simplified way that usually

leads to analytical solutions. For instance, Seeger et al. (1986) developed correlations for phase

separation at T-junctions with different branch inclinations. Unfortunately, empirical correlations

are usually valid only in the range of conditions from which they are derived. Besides, accuracy

of predictions typically depends on the amount and quality of experimental data utilized in the

derivation of correlations.

Phenomenological–mechanistic models, on the other hand, represent complicated physics of

flow separation using conservation principles (mass, momentum and/or mechanical energy) and force

balances (i.e. inertial vs. centripetal force balance on fluid particles based on streamline geometry)

that are simplified with assumptions according to physical understanding of the split mechanism. The

simplifications generally lead to models with analytical solutions. Particularly, phenomenological

methods are usually junction type (i.e. diverging, merging or impacting), geometry, and flow

pattern dependent. Also, closure relations or certain parameters such as energy loss coefficients or

momentum correction factors are expected to involve empirical correlations of such dependence.

However, applicability of phenomenological–mechanistic models can be usually extended to wider

range of flow conditions, extrapolated well outside the range on which the model is based.

An intriguing approach is the conformal mapping technique applied to two-phase flow split by

Hatziavramidis et al. (1997). Originally suited for two-phase flows that can be approximated as

potential flow (i.e. irrotational flow of incompressible and inviscid fluids), the model also employs

1Mechanistic T-junction models not to be confused with mechanistic two-phase flow modeling approach2Even for one-dimensional analysis, flow split at a branching tee requires at least two dimensions to be accounted

for; in the directions of (1) run and (2) branch, hence the term ‘one-half’ dimensional (1.5D) model.

31

the concept of the ‘free streamline’ (dividing streamline), defining the boundary of flow split at the

tee (Fig. 3.5).

One-dimensional Two-fluid models are based on the solution of one-dimensional mass and

momentum conservation equations for both phases, in both directions. In essence, two-fluid junction

model is the extension of two-fluid flow model over the junction control volume. As discussed earlier

in Sec. 2.2.2, two-fluid flow model is a rigorous approach for modeling two-phase flow in conduits,

giving rise to the simultaneous solution of coupled conservation equations (mass and momentum

balances) for phases. In fact, what separates two-fluid junction models from mechanistic junction

models is the use of phasic momentum equations in both run and branch directions (hence an

additional, 6th momentum equation) by the two-fluid models (cf. Sec. 4.12).

Success of two-fluid based junction models depends on the description of ‘irreversible’ flow loss

terms; in particular, availability of correlations for appropriate momentum correction k-factors.

Derivation of one-dimensional two-fluid model junction equations are discussed in more detail in

Sec. 4.12

A ‘complete phase separation’ is achieved at ‘critical branch flow split ratio’, when all

of the incoming gas phase is completely extracted through the branch. When complete phase

separation occurs, the T-junction indeed becomes a functional two-phase separator:

λcritical =m3

m1=x1

x3(3.3)

or

m3x3 = m1x1 (3.4)

3.2 Phenomenological and Mechanistic Junction Models

For one-dimensional analysis of steady-state, isothermal two-phase flow through a branching tee,

and assuming that junction geometry (arm inclination and diameters, branching angle, geometry of

joint corners etc.) and wall surface properties (pipe roughness) are known, there are 9 parameters

of primary interest that adequately define (and quantify) flow through the junction, hence the split:

• phasic mass flow rates through each arm (= 2× 3 = 6)

• inlet, run and branch pressures (= 3)

A 10th parameter of particular interest would have been a representative, nodal pressure for the tee.

However, modeling (at least) a two-dimensional pressure distribution with a single pressure node is

a significant challenge (cf. Sec. 4.12 and Appendix B).

For a well posed problem, that is for a unique solution to exist without over specifying the problem,

4 out of 9 parameters can be specified. Then, 5 linearly independent equations are required to

account for remaining 5 unknown parameters.

32

Conventionally, inlet phasic mass flow rates (2) and either run or branch phasic mass flow rates (2),

thus a total of 4 parameters are specified for the analytic solution. In Sec. 4.12 where two-fluid model

based flow split is discussed using a staggered grid arrangement for numerical solution, inlet phasic

mass flow rates and both run and branch outlet pressures are specified as boundary conditions.

Typically, four equations are employed by default when constructing phenomenological–mechanistic

models:

1. mixture mass balance for the tee

2. gas phase mass balance for the tee

3. inlet-to-run mixture momentum balance (or mechanical energy balance)

4. inlet-to-branch mixture mechanical energy balance (or momentum balance)

Mixture (or overall) momentum or mechanical energy balance equations, simplified into inlet-to-run

and inlet-to-branch pressure change equations, indeed constitute the pressure change component of

the model.

A 5th relation, required for completing the equation set, constitutes the phase separation component

of the model determining the degree of separation. In fact, the essential difference between several

T-junction models and what renders them phenomenological or mechanistic is the choice of this 5th

relation and associated constitutive (closure) relations for the calculation of certain parameters (i.e.

two-phase loss multiplier in pressure change equations of homogeneous equilibrium based models

and etc.).

By nature, phenomenological models are expected to be flow pattern dependent. On the other

hand, mechanistic junction models are thought to be flow pattern independent. However, employed

closure relations are frequently flow pattern dependent. Typically, flow pattern specific models

are expected to perform better than independent models. Lahey (1986) provides a list of most

significant branching T-junction studies of the time, categorizing the studies according to analyzed

tee geometry and developed model; i.e. pressure change and/or phase separation model.

3.2.1 Pressure Change Sub-Models

Inlet-to-run and inlet-to-branch pressure changes dominating the flow split and phase separation at

a tee can be calculated using:

1. mechanical energy balances in both directions

2. momentum balances in both directions

3. mechanical energy balance for branch and momentum balance for run direction.

It is well established that (i.e. by reference to Bernoulli equation) when single or two-phase flow

split takes place at a T-junction, flow continuing in the inlet direction through the run experiences

a pressure rise due to deceleration associated with flow area expansion, assuming that main pipe

diameter does not change. On the other hand, typically a pressure drop is experienced by the flow

turning into the branch.

33

Due to limitations of one-dimensional analysis, not all the factors contributing to the pressure change

at a tee can be accounted for explicitly and rigorously in the governing equations. Instead, effects

of factors, such as associated mechanical losses due to change in flow direction or due to secondary

flows, are typically combined into a single parameter (loss coefficient K for mechanical energy

equation and correction factor k for momentum equation) and represented with an additional

term in the pertinent governing equation. In that regard, use of these parameters is similar to the

representation of viscous forces of flow (wall drag) with a friction factor and a friction force for

one-dimensional analysis. Consequently, pressure changes of interest; (1) inlet-to-run pressure rise

and (2) inlet-to-branch pressure change, constitute two parts:

1. Reversible component (∆p)rev: Pressure rise due to Bernoulli effect associated with flow

expansion (or pressure drop in the case of contraction) This is the pressure change naturally

captured by the governing equations without loss coefficients or correction factors.

2. Irreversible component (∆p)irr: Pressure change due to cumulative effects of (a) wall friction,

(b) interfacial drag, (c) recirculation zones (secondary flows, Fig. 3.4) and (d) turning of

flow into the branch. The irreversible component of the total pressure change is completely

accounted for by the loss coefficients in the case of mechanical energy equations. On the other

hand, momentum correction factors predominantly account for the effect of the recirculation

zones and partially account for the effects of turning of flow (momentum transfer), wall and

interfacial frictions; since pertinent loss terms already appear in the governing momentum

equations.

Figure 3.4: Recirculation zones of a branching tee

Overall pressure change depends on relative magnitudes of reversible and irreversible components.

Typically, reversible component along the run direction, the Bernoulli type pressure rise due to

flow expansion, is greater in magnitude compared to pressure drop due to irreversible losses. For

the branch direction, magnitude of the irreversible component, pressure drop associated with flow

turning into branch and other losses, is higher than the pressure rise associated with Bernoulli effect.

34

Governing mechanical energy and momentum equations are usually simplified into pressure change

equations. While pressure equation for inlet-to-run stream is typically based on momentum

conservation, pressure equation for inlet-to-branch stream is based on mechanical energy balance.

Even for single phase flow through a T-junction, it is easier to analyze total pressure change from

inlet-to-branch using mechanical energy conservation because the need to explicitly account for the

directionality of the flow is automatically ruled out. Then, total pressure change can be easily split

into reversible and irreversible components.

∆p13 = (∆p13)rev + (∆p13)irr (3.5)

(∆p13)rev =1

2ρ(v2

3 − v21) (3.6)

(∆p13)irr =1

2K13ρv

21 (3.7)

Where; K13 is the empirical loss coefficient [dimensionless]

On the other hand, for the case of inlet-to-run, Ballyk et al. (1988) base the choice of using a

momentum balance in the run (axial) direction (rather than a mechanical energy balance) to

the better performance of momentum balance predictions in matching experimental data. Then,

inlet-to-run pressure change equation using momentum balance, with momentum correction factor

(k) accounting for an indeterminate axial momentum carried away with the branching flow, is given

as:

∆p12 = k12ρ(v22 − v2

1) (3.8)

However, Buell et al. (1993) reports that breaking the pressure change for inlet-to-run direction

into reversible and irreversible components had the difficulty of generating negative K-factors (loss

coefficients) for certain conditions. Hence, inlet-to-run pressure change can also be expressed, using

mechanical energy balance (with loss coefficient K), without breaking the pressure change into

irreversible and reversible components:

∆p12 =1

2K12ρ(v2

2 − v21) (3.9)

Typically, pressure equations for phenomenological–mechanistic models are extensions of single-

phase mechanical energy or momentum balances. With an analogy to homogeneous equilibrium

model (HEM) of two-phase flow, two-phase flow split through the T-junction is also treated as a

homogeneous mixture. However, two-phase loss multipliers are utilized to account for additional

pressure drop effects (Buell et al., 1993).

Pressure equations for homogeneous flow model can be written in the general form, separately for

(1) inlet-to-run:

∆p12 = k12(ρ2v22 − ρ1v

21) (3.10)

35

and (2) inlet-to-branch:

∆p13 =1

2(ρ∗3v

23 − ρ∗3v2

1) +K131

2

(ρ1v1)2

ρLΦ (3.11)

Where:

ρi Mixture density for arm i = 1, 2, 3 (inlet, run, branch) [kgm−3]

ρ∗3 The ‘homogeneous’ mixture density in branch [kgm−3]

ρL Liquid density (assumed constant through out the tee) [kgm−3]

Φ Two-phase loss multiplier [dimensionless]

Typically, what distinguishes different models are calculation of certain parameters; i.e. mixture

densities, homogeneous mixture density and the loss multiplier (Buell et al., 1993). Follows next is

a brief, chronological review of prominent work on pressure change models.

Fouda and Rhodes (1974) developed a pressure drop model for air–water annular flow in branching

tees, based on separated flow model (SFM) approach. Inlet-to-run pressure change (axial

pressure recovery) is computed by combining phasic momentum equations:

p2 − p1 = ∆p12 = (∆p12)G + (∆p12)L (3.12)

p2 − p1 = ∆p12 = −ksf12 ∆

[m2

AxvL +

m2

A(1− x)vG

](3.13)

p2 − p1 = ∆p12 = −ksf12

(m2

A

)2

[x2

αρG+

(1− x)2

(1− α)ρL

](3.14)

Where, ksf is the separated flow model momentum correction factor [dimensionless].

Inlet-to-branch pressure drop (radial pressure drop) can be computed in a similar fashion but this

time combining phasic mechanical energy balance equations instead of momentum equations because

flow is changing direction (hence Ksf is used):

p3 − p1 = ∆p12 = −Ksf13

2

(m3

A

)2

[x2

αρG+

(1− x)2

(1− α)ρL

](3.15)

Where, Ksf is the separated flow model loss coefficient [dimensionless].

Saba and Lahey (1984) use the homogeneous fluid approach (HEM) with inlet-to-run and inlet-

to-branch mixture mechanical energy balances and splits the pressure change into reversible and

irreversible components. A simplified, ‘homogeneous’ version of the pressure change equation is

then obtained as:

p2 − p1 = ∆p12 = −k12 ∆(ρmixv2mix) (3.16)

Reimann and Seeger (1986) model accounts for a reversible pressure change from inlet-to-throat (a

vena contracta) in the run direction and an additional pressure drop from this throat to a position

36

downstream along the run. Beyond that, the difference between works of Saba and Lahey (1984)

and Reimann and Seeger (1986) is the calculation of two-phase loss multiplier Φ.

Ballyk et al. (1988) use a model similar to Saba and Lahey (1984); homogeneous mixture momentum

equation to determine inlet-to-run pressure drop and homogeneous mixture mechanical energy

balance to determine inlet-to-branch pressure drop for annular flow.

Later, Ballyk and Shoukri (1990) used Fouda and Rhodes (1974) type separated flow momentum

balance for inlet-to-run pressure change and homogeneous flow (mixture) mechanical energy equation

for inlet-to-branch pressure change.

Peng et al. (1993) extend Ballyk et al. (1988) model to annular flow in T-junctions for downwardly

inclined branch arms and develop a pressure drop model based on momentum and mechanical

energy balances.

3.2.2 Phase Separation Sub-Models

Mechanistic junction models employ an additional momentum or mechanical energy equation

as the 5th equation to balance unknowns and this 5th equation determines the phase separation

(Saba and Lahey, 1984; Fouda and Rhodes, 1974).

Fouda and Rhodes (1974) suggest that inlet-to-branch pressure change, in addition to inlet-to-branch

mechanical energy balance (Sec. 3.2.1), could also be estimated using the orifice equation (based on

mechanical energy balance) recognizing that fluid discharge from main line to branch forms a ‘vena

contracta’ (Munson et al., 2009; Bird et al., 2002). If two-phase flow is treated as an homogeneous

mixture then an equation similar to mechanical energy balance (Eqn. 3.15) is obtained. However, a

separated fluid approach (SFM) requires that orifice equation is written individually for phases and

since pressure change should be same for both, model provides an auxiliary relation, a 5th equation

to account for phase separation:

(m3)G = m3 x = CGαA[2ρG(p3 − p1)]12 (3.17)

(m3)L = m3 (1− x) = CL(1− α)A[2ρL(p3 − p1)]12 (3.18)

Where, Ck is the orifice discharge coefficient [dimensionless]

Saba and Lahey (1984) suggest a fluid mechanics based 5th equation to account for phase separation.

However, instead of a control volume based field equation (i.e. momentum or mechanical energy

equation) as in two-fluid models, inlet-to-branch gas phase momentum equation is integrated along

a ‘mean’ (approximate/average) streamline to yield an auxiliary relation for inlet-to-branch pressure

change from which phase separation is determined. The length Ls of the mean streamline is

calculated based on empirical correlations. Gas phase momentum equation is chosen over liquid

phase because gas phase was observed to determine the preferential separation as the ability of gas

to make the turn into branch is the dominant factor.

37

Phenomenological models are typically developed for annular flow pattern in the inlet and

geometrically define proportions of gas and liquid flows extracted from main line into the branch.

There are two fundamental concepts in model development:

1. zones of influence; sections of main line cross-sectional area from which liquid and gas are

diverted into the branch (Fig. 3.6).

2. dividing streamlines that separate incoming (inlet) flow into inlet-to-run and inlet-to-branch

streams for each phase (Fig. 3.5).

Inlet-to-branch streamlines of both phases follow curved paths in order to accomplish a turn into

the branch and typically, phasic streamlines are expected to cross if both phases are to make turn

in the branch direction.

Figure 3.5: Dividing streamlines for two-phase flow in a branching tee

Dividing stream line approach requires geometric analysis of inlet flow pattern to determine ‘zones

of influence’ that affect the outcome of force balance between gas and liquid phases over the

streamline(s) curving into the branch; i.e. competing inertial and centripetal forces. Distance of

streamlines from inlet pipe wall determine zones of influence.

Identifying zones of influence for phases is a semi-empirical step. ‘Zones’, geometric portions of

inlet gas and liquid flows, from which phases extracted to enter the branch are determined by

superimposing cross-sectional phase distribution of the inlet on dividing streamlines. A zone of

influence exists for each phase and dividing streamlines determine both the zone boundary and

phasic (Gas–liquid) interface in the main line. The amount of gas phase extracted to the branch

comes from the area bounded by gas phase dividing streamline and the liquid film boundary.

The concept of ‘zones of influence’ was extended to two-phase flow by Azzopardi and Whalley (1982)

aiming to describe liquid film and gas extraction from the main line for annular flow at a vertical

tee. Based on geometrical considerations only, the ‘zone of influence’ is defined by the angle σ

38

(a) Single-zone model (b) Two-zone model (c) Advanced two-zone model

Figure 3.6: Zones of influence for two-phase flow in a branching tee

covering the sector of material extraction, the liquid film arc of inlet cross-section (Fig. 3.6). The

model constitutes an alternative 5th equation derived from geometrical relations provided by ‘zone

of influence’ concept.

Azzopardi and Whalley (1982) model assumes liquid and gas are extracted from the same zone of

influence. This suggests coincidental dividing streamlines for liquid and gas phases. However, due

to the significant difference in axial momentum fluxes of two phases, the gas and liquid are expected

to have different dividing streamlines.

Recognizing short comings of the single zone of influence used by Azzopardi and Whalley (1982)

and purely geometric approach lead to its definition, Shoham et al. (1987) defined separate zones of

influence for phases, for the analysis of stratified and annular flows. Consequently, phasic zones of

influence are delineated by separate gas and liquid dividing streamlines. Position of liquid streamline

(i.e. distance from pipe wall) is determined by a simplified force balance for competing centripetal

and inertial forces acting on the liquid film, thus establishing the 5th equation for phase separation

component.

Hwang et al. (1988) combined Saba and Lahey (1984) model with zone of influence concept, again

defining a separate zone of influence per phase. Phase separation (gas flow into the branch) is

governed by force balance equations written per phase and derived from a ‘modified version’ of

Euler’s equation of motion (momentum equation for a particle traveling along a streamline). In this

case, an empirical correlation is utilized to determine phasic ‘mean’ streamlines as opposed to the

force balance employed by Shoham et al. (1987). While not lumped into a single equation, whole

phase separation sub-model itself substitutes for the ‘5th equation’ needed for closure.

Both Shoham et al. (1987) and Hwang et al. (1988) models assume that radial (horizontal) distance

of dividing streamlines from pipe wall, defined by (δL, δG) remain constant and independent of

vertical position on the plane of cross-section, in accordance with the original intent of vertical

39

annular flow analysis. However, the major difference between the models is how the positions of

streamlines is determined.

In an attempt to improve the performance of dividing streamlines approach of Hwang et al. (1988)

for annular flow in horizontal, equal-diameter T-junctions, Ballyk and Shoukri (1990) computes

the position of dividing streamlines at varied elevations of the inlet cross-section, introducing the

concept of ‘development length’.

Hart et al. (1991) with Double Stream Model (DSM), and then Ottens et al. (1994)3with Advanced

Double Stream Model (ADSM) further developed the concept of dividing streamlines for separated

flow patterns (i.e. stratified flow) using a purely fluid mechanics based approach. Both models

derive a governing equation for phase separation based on phasic Bernoulli (mechanical energy)

equations and associated loss coefficients. More detailed discussion on Double Stream Model follows

in Sec. 3.2.3.

Peng and Shoukri (1997) further developed and generalized the model of Ballyk and Shoukri (1990)

to account for the effect of gravity, hence branch inclination, on phase separation.

Penmatcha et al. (1996) combined the Hwang et al. (1988) and Shoham et al. (1987) models for

stratified wavy conditions. Separate momentum equations are written for phasic streamlines and

flow areas required for the equations are determined by inlet flow geometry. Marti and Shoham

(1997) extended this model for reduced T-junctions with different branch arm inclinations.

3.2.3 The Double Stream Model

Hart et al. (1991) developed Bernoulli equation based Double Stream Model (DSM) to predict flow

split at T-junctions based on the dividing (separating) streamline concept which assumes phases

flow along their individual channels or streamlines. DSM is essentially a phase separation sub-model

for separated flows. Although Bernoulli equation comes with the inherent simplifying assumption of

constant phasic density all throughout the T-junction volume, it is a reasonable assumption for an

adequately small tee volume.

Bernoulli equations are written for inlet-to-run and inlet-to-branch directions, per phase:

1

2(ρk)1(vk)

21 + (ρk)1gz1 + p1 =

1

2(ρk)2(vk)

22 + (ρk)2gz2 + p2 +

irreversible loss term︷ ︸︸ ︷1

2(ρk)1(vk)

21Kk12 (3.19)

1

2(ρk)1(vk)

21 + (ρk)1gz1 + p1 =

1

2(ρk)3(vk)

23 + (ρk)3gz3 + p3 +

irreversible loss term︷ ︸︸ ︷1

2(ρk)1(vk)

21Kk13 (3.20)

Where:

3Author wishes to acknowledge Dr.Marcel Ottens (Department of Biotechnology, Delft University of Technology,Netherlands) for providing the article, which has not been possible to obtain otherwise

40

(ρk)i = ρk Phase density, constant throughout the tee volume [kgm−3]

Kk Energy loss coefficient for phase k (defined in Tab. 3.1) [dimensionless]

In this arrangement, outgoing phasic velocities (vk)2, (vk)3 are simply controlled by outlet pressures

and irreversible loss terms. Besides the velocities, calculation of mass flow rates going through each

arm, hence actual phase separation, requires void fraction information:

(mk)1 = (ρk)1(vk)1(αk)1A1

(mk)2 = (ρk)2(vk)2(αk)2A2

(mk)3 = (ρk)3(vk)3(αk)3A3

(3.21)

At this point, it is recognized that leaving void fractions (or holdups) as unknowns would require

additional equations to be included in the solution, as is the case with other phenomenological–

mechanistic junction models. This fact gives rise to another simplification via the assumption

(αk)1∼= (αk)2

∼= (αk)3, based on the previous assumption of adequately small tee volume. With the

assumption of constant void fraction through out the tee, DSM establishes that not only outgoing

velocities but also the actual flow split is, in fact, controlled by outlet pressures and irreversible loss

terms.

Four (4) Bernoulli equations (one per phase, per tee arm direction) are re-arranged into a closed

form ‘phase separation equation’ relating gas phase branch mass intake to the liquid phase branch

mass intake:

(a4 − 1)λ2L + 2λL − κ

[(a4 − 1)λ2

G + 2(λG − λ0)]

+1

βL(FrL)13− 2λ0 (3.22)

Where:

a =D2

D3Main line (inlet) to branch diameter ratio [dimensionless]

FrL13 =(vL)2

1

g(D1 −D3)Modified Froude number for liquid phase [dimensionless]

βk Constant depending on inlet phasic flow mode [dimensionless]

βk = 1.00 if flow is turbulent

βk = 1.54 if flow is laminar

κ =ρGv

2G

ρLv2L

Ratio of phasic inlet kinetic energies [dimensionless]

λk =(mk)3

(mk)1Phasic branch mass intake fraction [dimensionless]

λ0 =1

2(1 + (KG)12 − (KG)13) Junction gas phase energy dissipation factor [dimensionless]

Equation 3.22 is further simplified by taking phasic loss coefficients equal (KGi = KLi), based on

41

the assumption that phases are flowing along their separate streamlines (or channels). Nevertheless,

with this assumption DSM became limited to very small holdup values (αL ≤ 0.06) for accurate

phase separation predictions.

The most important point is, because it is based on mechanical energy equations, model had no

true way of recognizing the effect of split angle unless accounted for by other means; in this case by

the irreversible loss terms (and associated loss coefficients). Although Bernoulli equation is applied

on a streamline that curves into the branch, inherently the equation has no means of observing the

directional change in flow. This is to be expected since Bernoulli equation is indeed a non-vectorial

mechanical energy equation derived from a directional momentum balance with certain simplifying

assumptions Please see Munson et al. (2009) for the details of Bernoulli equation derivation from

momentum balance.

To make it more clear, without irreversible loss terms, Bernoulli equation can not account for a

directional change in flow (at least in its usual form and perception) and therefore DSM can not

account for the effect of split angle when irreversible losses are ignored. In a way, Steinke (1996)

also points out to this misconception lead by inadvertent approximation of split angle effects via

irreversible loss terms. However, Steinke (1996) discussion, or Spore et al. (2001) for that matter,

is quite interesting that Bernoulli equation is arranged in a form to resemble a non-conservative

momentum balance accounting for flow direction.

K-factors (loss coefficients) typically depend solely on surface properties (roughness) and geometry

of the conduit. The geometry element is expected to inherently introduce the effect of split angle.

Nevertheless, there is no certainty that a K-factor obtained by matching experimental data for

particular flow conditions, for a particular flow pattern and tee structure, would yield satisfactory

results for significantly different flow conditions and flow pattern if actual physics of the flow is not

properly acknowledged.

Ottens et al. (1994) relieved DSM from the small holdup constraint by deriving appropriate

correlations for liquid phase K-factors (KL) and establishing the Advanced Double Stream Model

(ADSM).

The advantage of DSM (or ADSM thereof), is, while it is certainly necessary to account for the

details of mechanisms controlling the split (i.e. flow pattern geometry), macroscopic balances should

be satisfied at all times and thus a satisfactory model employing simple energy or momentum

balances is preferable to flow pattern dependent methods. Nevertheless, when flow pattern geometry

and fluid distribution in the main line is not accounted for, DSM and similar mechanistic models

are likely to fail if branch diameter is smaller than main line diameter and/or side port orientation

is not horizontal (position of the opening for branch on the main line surface, not the inclination of

branch arm). Especially, if model has no way of recognizing liquid level or elevation of the highest

point the liquid film within the main line reaches it could still predict liquid phase going into the

branch while it is not physically possible. One-dimensional two-fluids model are also subject to

same problem unless a means to account for flow geometry is incorporated in the model.

42

3.2.4 Correction Factor and Loss Coefficient Correlations

For single-phase flow through T-junctions hydraulic loss coefficients (K-factors) and/or momentum

correction factors (k-factors) are dependent on overall junction geometry and branch mass intake

fraction:

λ =m3

m1

Single-phase T-junction parameters (k-factors and K-coefficients) are typically correlated with

experimental data based on polynomial fits:

k =∑

anλn−1

or

K =∑

bnλn−1

Hence, these correlations are specific to experimental conditions and do not account for T-junction

geometry or wall surface properties explicitly. Polynomial coefficients an and/or bn differ between

correlations depending on flow conditions, junction geometry and surface properties for which the

experimental data is obtained. Table 3.1 lists single-phase correlations from three different studies

(Hwang et al., 1988; Ballyk et al., 1988; Buell et al., 1993) for horizontal branching tees with 90o

angle. Buell et al. (1993) and Ballyk et al. (1988) use momentum equations for inlet-to-run direction.

Table 3.1: Correction factor and loss coefficient correlations for a 90o branching tee

inlet-to-run:

Hwang (1988) K12 = 0.3419− 1.223

(m2

m1

)+ 0.8638

(m2

m1

)2

Ballyk (1988) k12 = 0.704− 0.320λ− 0.028λ2

Buell (1993) k12 = 0.57− 0.102λ+ 0.107λ2

inlet-to-branch:

Hwang (1988) K13 = 1− 0.8285λ+ 0.6924λ2

Ballyk (1988) K13 = 1.081− 0.914λ+ 1.050λ2

Buell (1993) K13 = 1− 0.982λ+ 1.843λ2 − 0.717λ3

Gardel (1957), however, derived K-factor correlations for single-phase flow that accounts for changes

43

in the branching angle as well:

K12 = 0.03(1− λ)2 + 0.35λ2 − 0.2λ(1− λ) (3.23)

K13 = 0.95(1− λ)2 + λ2

[1.3 cot

(1

)1− 0.9

√r]

+ 0.8λ(1− λ) cot

(1

)(3.24)

Comparison of single-phase flow K-factor correlations of several authors is given in Fig. 3.7.

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

1.4

0 0.2 0.4 0.6 0.8 1

K13

λG

Gardel (1957)

Ballyk (1988)

Buell (1993)

Figure 3.7: Comparison of K-factors by various correlations

Besides empirical correlations (e.g. Gardel), authors such as Oka and Ito (2005); Bassett et al. (2001)

derive single-phase loss coefficient correlations based on momentum balances written over CVs at the

tee. Pressure change predictions of mechanical energy equations are related to momentum equations

neglecting momentum correction factors and thus analytic expressions for hydraulic (energy) loss

coefficients (K-factors) can be obtained.

However, all the aforementioned correlations are for single-phase flow conditions and for two-phase

conditions these parameters are observed to be dependent on flow pattern and flow mode (i.e.

turbulent or stratified) as well. For instance, El-Shaboury et al. (2007) express phasic loss coefficients

for an impacting tee as functions of Reynolds number.

Similarly, Ottens et al. (1994) derived correlations for liquid phase loss coefficients to complement

Gardel’s single-phase (gas) correlations (for use with ADSM) based on liquid phase Reynolds number

and gas phase K-factors:

(KL)12 =

(11.48 +

1063

1− (ReL)2− 12.67κ

)(KG)12 (3.25)

44

(KL)13 =

(1.247− 69.26

1− (ReL)3− 0.198κ

)(KG)13 (3.26)

Where, κ is the gas phase to liquid phase ratio of inlet kinetic energies and ReL is the liquid phase

Reynolds number.

3.2.5 Film-stop Phenomenon

In annular flow, a substantial part of liquid is flowing as a thin liquid film spread along the conduit

wall. It is, therefore, expected that the liquid film in the main tube segment that is intercepted by

the branch cross-sectional area will be the first component to be extracted into the branch.

During annular flow, as the liquid film approaches the side port of a branching tee, a sudden increase

in the liquid amount going to branch is observed corresponding to a small increase in gas off-take

(Roberts et al., 1995). This ‘step’ increase in liquid amount is attributed to liquid film slowing down

as inlet-to-run pressure spike along the main line increases and ultimately coming to a complete

halt, hence the name ‘film-stop’, at some threshold pressure value, lending itself to easy withdraw

through the side port.

Roberts et al. (1995) extended Azzopardi and Whalley (1982) model to account for the film-stop

effect in annular and semi-annular flows. DSM of Hart et al. (1991) is also applicable to such flow

conditions but does not account for the film-stop effect thus fails to match experimental data beyond

a certain range.

Increasing the branch flow, which corresponds to decreasing the branch downstream pressure, causes

more gas to be extracted through the branch, raising the branch quality. Increasing the branch

flow is associated with a pressure rise through the junction run due to flow expansion. This adverse

pressure rise yields an additional force which tends to drive the flow through the branch. Because

the gas (lighter) phase in the inlet arm has lower axial momentum than the liquid (heavier) phase,

it tends to respond more readily to the decreasing pressure through the branch and the increasing

pressure through the run. Accordingly, increasing the branch flow tends to result in an increase in

the branch flow quality.

For annular flow condition at horizontal tees, inclination of the branch has a significant influence on

the separation of phases because of the non-uniform distribution (on the inlet cross-section) of the

liquid film, attributed to gravity.

3.2.6 Summary

Lahey (1986) concluded that none of the existing models was successfully applicable to a whole

range of flow conditions (different inlet flow patterns and flowing pressures) without first generating

experimental data for loss coefficients or correction factors and recommended a composite model as

an interim solution.

Because Saba and Lahey (1984) model predicted phase separation for high branch mass off-take

45

ratios (0.5 ≤ m3m1≤ 1.0) correctly, Lahey (1986) suggested using Azzopardi and Whalley (1982) model

for the range (0.0 ≤ m3m1≤ 0.05) and a the empirical correlation of Zetzmann4 for the remaining,

uncovered range (0.05 ≤ m3m1≤ 0.5).

Buell et al. (1993) compared experimental results with different phase separation and pressure

change sub-models:

1. Fouda and Rhodes (1974), Separated Flow Model (SFM)

2. Saba and Lahey (1984), Homogeneous Flow Model (HFM)

3. Reimann and Seeger (1986)

4. Hwang et al. (1988)

5. Ballyk et al. (1988)

Recognizing that performance of models varied over branch mass intake fraction, Buell et al. (1993)

also acknowledged the difficulty for a single pressure change model to satisfactorily predict pressure

changes for all flow conditions at a tee, and concluded that ultimately a suite of flow pattern specific

models may be required.

Walters et al. (1998)5 compared performances of 4 pressure change and 4 phase distribution

sub-models for different inlet/branch diameter ratios (reduced tees) (Tab. 3.2).

Table 3.2: Models compared by Walters et al. (1998)

Pressure change: Phase distribution:

Fouda and Rhodes (1974) Shoham et al. (1987)

Saba and Lahey (1984) Azzopardi (1988)

Reimann and Seeger (1986) Hwang et al. (1988)

Hwang et al. (1988) Hart et al. (1991)

Walters et al. (1998) concluded that SFM of Fouda and Rhodes (1974) gave the best overall pressure

drop predictions while DSM of Hart et al. (1991) gave good results for stratified, wavy and annular

flow regimes albeit limited to small holdup values.

4Zetzmann, K. (1982). Phasenseparation und Druckfall in zweiphasen Durchstromten vertikalen Rohrabzweigungen,Doctorate Thesis, University of Hannover, FRG. (Lahey, 1986)

5Azzopardi, B.J., (1988). An additional mechanism in the flow split of high quality gas–liquid flows at a T-junction.UKAEA Report AERE-R 13058. (Walters et al., 1998)

46

Chapter 4

Model Conceptualization

4.1 Conservative Property: FVM vs FDM

Numerical solution to Euler equations can be achieved through mathematical approximation of

governing PDEs using algebraic equations where differential operators are replaced with difference

operators. This leads to the formulation of algebraic Finite Difference Equations (FDEs), also

known as discrete or discretization equations since they require information from discrete points in

the flow domain. Hence, continuous information available in the exact solution of PDEs is replaced

with discrete, approximate information of the FDEs. Approximate values of flow parameters are

then computed for discrete points (nodes) in the domain, which is the basis of the Finite Difference

Method (FDM).

With some physical insight, it is realized that numerical solution could also be achieved by initially

dividing the physical domain of flow into imaginary, discrete (non-overlapping) computation sub-

domains, namely Control Volumes (CVs), that are fixed boundary, open systems over which the

continuum approach and the conservation principles prevail. The algebraic governing equations are

then obtained by applying conservation principles over these CVs. The fluid and flow properties at

the center of each CV typically represent the volumetric average of these values for that CV. This is

the basis of Finite Volume Method (FVM). In fact, writing algebraic equations for a CV can be

seen as a preliminary step in deriving the governing PDEs of flow and thus the algebraic equations

should collapse to the PDEs when a CV of infinitesimal size is considered.

Finite Volume Method equations are also described as integrated forms of the PDEs. However, and

because some type of finite difference (FD) approximation might be required for the definition of

flux values at each CV face, some authors describe the FV approach as an FD method applied to the

integrated form of the flow equations; i.e. Integral (or Integrated) Finite Difference Method (IFDM).

This more general approach to obtain the FVM discretized equations are especially necessary

when complex; irregular geometries are adopted as CVs. Then, governing PDEs are integrated in

a piece wise continuous manner over the CVs and the integral is then evaluated with necessary

approximations.

Governing algebraic equations of the FVM express the conservation principles over finite CVs while

governing PDEs articulate conservation principles over infinitesimal CVs. Assuming infinite machine

precision, FVM retains the ‘conservative property’ at the algebraic equation level as well as at the

47

PDE level. However, because they are derived using mathematically defined differential operators;

algebraic equations of the FDM may not retain the conservative property built in the original PDE

formulation when applied to the finite computational grid. Since FVM algebraic equations possess

the conservative property, conservation of mass, momentum and energy is always satisfied (ignoring

limited computer precision) over any group of CVs and thus over the entire domain regardless of

the size or number of CVs, and not only at the never-attained limit of infinitesimal CV size. This

constitutes the fundamental difference between FDM and FVM. The extensive discussion by Roache

(1998) on the conservative property is highly recommended for the interested party.

In practice, however, what actually matters is to retain the conservative property of the original

PDEs in the final form of the governing algebraic (discretized) equations, which is to retain flux

terms intact and by doing so, ensure that ‘flux matching’ concept prevails at the faces of CVs.

4.2 Single-Phase Flow Equations

All governing equations for fluids in motion are based on the principles of mass conservation,

Newton’s second law, and the first law of thermodynamics applied to a control mass system

(CM). These fundamental physical principles are outlined below. It is important to recognize that

conservation statements are straightforwardly defined for systems of fixed mass (i.e. control mass

systems – Lagrangian view), but they can be readily extended to system of fixed volume, variable

mass (i.e. control volumes – Eulerian view) using the Reynolds Transport Theorem (Munson et al.,

2009; Kleinstreuer, 2003; Todreas and Kazimi, 1990). This discussion is given below.

For a system containing a fixed amount of matter (i.e. the control mass system CM), the three

statements of conservation can be written as:

∂MCM

∂t= 0

i.e. by definition of a control mass system CM, total amount of mass

inside the system is constant (conserved) and does not change with

time. M =∫ρdV in general, and M = ρV for systems of uniform

density.

d ~PCMdt

= ~FCMi.e. the sum of all external forces acting on a CM equals its rate

of change of linear momentum. Linear momentum is defined as

~P =∫ρ~vdV , while ~P = M~v for systems of uniform density.

dECMdt

=(Q− W

)CM

i.e. the rate of change of total energy in the CM equals the amount of

heat energy entering the CM minus the work done by the CM.

48

In this expression;

E =

∫ρedV in general

E = me for systems of uniform density

e = e∗ +v2

2+ g∆z intrinsic energy content per unit mass

e∗ system internal energy

v system velocity

∆z system elevation with respect to a datum z0

Q rate of heat transfer

(positive when added to the system)

W rate of work

(positive when done by the system)

These equations, as stated, apply to systems that maintain fixed amount of mass. In fluid dynamics,

we look at systems of fixed volume stationary in space or ‘control volumes’, rather than systems of

fixed amount of mass. Evidently, in the control volume representation, the amount of fluid contained

inside the CV changes as fluid flows through the CV. The conservation statement that provides

the link between the fixed mass system description and its equivalent control volume description is

called the Reynolds transport theorem, which states that:

dNCM

dt=

rate of change of N

in the CM

=

rate of change of N

in a CV

+

rate of change of flux of N

out of the CV

(4.1)

The Reynolds transport theorem is the statement that relates a CV representation to a CM

representation. dNCMdt refers to the rate of change of any extensive property N within the fixed

mass system representation or CM. The statement above assumes that the CM and CV coincide

at a given time and then tracks how the initial fixed amount of mass leaves the CV at a later

time. The Reynolds transport theorem is a fundamental theorem of fluid dynamics and it is used

in formulating the governing equations of fluid flow for a CV representation or Eulerian frame of

reference. It is important to note that in the theorem above, N represents any extensive property

of the system, such as mass (M), linear momentum (~P ), total energy (E) or entropy (S).

4.2.1 Conservation of Mass

Let us consider the extensive property total mass; M = (ρA∆x), within this finite CV shown in

Fig. 4.1. In a finite volume representation, node point values of properties at x are assumed to be

averages over the entire control volume. The Reynolds transport theorem applied to this CV states:

dMCM

dt=

rate of change of

mass in a CV

+

rate of flux of

mass crossing

the CS

(4.2)

49

Figure 4.1: Finite control volume

Where, CS control surface or surface that defines the CV

Conservation of mass for this CM imposesdMCM

dt= 0, thus the statement above can be re-written

for a discrete period of time ∆t as:

0 =

change of mass

in the CV

during ∆t

+

net flow of mass

leaving the CS

during ∆t

(4.3)

Where:change of mass

in a CV

during∆t

= (ρA∆x)t+∆tx − (ρA∆x)tx

advective mass flux

entering the CS

during∆t

= (ρ~vA)t′

x−∆x2

∆t

advective mass flux

leaving the CS

during∆t

= (ρ~vA)t′

x+ ∆x2

∆t

In above expressions, t′ represents any suitable time reference within the time interval t′ = t and

t′ = t+ ∆t. This selection can make the resulting algebraic numerical scheme to be fully implicit

(for the case t′ = t+ ∆t), fully explicit (for the case t′ = t) or a hybrid of these two. Substituting

these expressions into Eqn. 4.3, one obtains:

[(ρA∆x)t+∆t

x − (ρA∆x)tx]

+[(ρ~vA)t

x+ ∆x2

∆t− (ρ~vA)t′

x−∆x2

∆t]

= 0 (4.4)

50

or equivalently, by dividing Eqn. 4.4 by ∆x and ∆t:

(ρA)t+∆tx − (ρA)tx

∆t+

(ρ~vA)t′

x+ ∆x2

− (ρ~vA)t′

x−∆x2

∆x= 0 (4.5)

Equation 4.5 represents the 1D Finite Volume representation of conservation of mass for advective

flow implemented in this study. Please note that if one takes the limit ∆t,∆x→ 0 then the PDE

form of conservation of mass equation is readily obtained.

∂(ρA)

∂t+∂(ρ~vA)

∂x= 0 (4.6)

which, for a system of constant cross-sectional area, becomes:

∂ρ

∂t+

∂x(ρv) = 0 (4.7)

An alternate way of obtaining the FVM equation is to start with the corresponding PDE represen-

tation (Eqn. 4.6) and carry out its integration over time (t to t+ ∆t) and discrete space (x− ∆x2 to

x+ ∆x2 ) to obtain:

x+ ∆x2∫

x−∆x2

t+∆t∫t

∂(ρA)

∂tdt

dx+

t+∆t∫t

x+ ∆x

2∫x−∆x

2

∂x(ρ~vA)dx

dt = 0 (4.8)

Because the sequence of integration is irrelevant, the most convenient sequence that groups the

relevant differentials is elected. Thus;

x+ ∆x2∫

x−∆x2

[(ρA)t+∆t − (ρA)t

]dx+

t+∆t∫t

[(ρ~vA)x+ ∆x

2− (ρ~vA)x−∆x

2

]dt = 0 (4.9)

In a discretized domain, node point values are assumed to be averages over the domain x− ∆x2 <

x < x+ ∆x2 , and the first integral is easily evaluated. For the second integral a decision must be

made regarding whether fluxes (ρ~v) will be considered constant and equal to their values at time

level t (explicit scheme), or t + ∆t (implicit scheme) or some combination (Crank-Nicholson or

Semi-Implicit). By using a generic time reference t′, one obtains:

[(ρA)t+∆t

x − (ρA)tx] x+ ∆x

2∫x−∆x

2

dx+[(ρ~vA)t

x+ ∆x2

− (ρ~vA)t′

x−∆x2

] t+∆t∫t

dt = 0 (4.10)

51

or

(ρA)t+∆tx − (ρA)tx

∆t+

(ρ~vA)t+∆tx+ ∆x

2

− (ρ~vA)t+∆tx−∆x

2

∆x= 0 (4.11)

Eqn. 4.11 is identical to Eqn. 4.5. In either expression the time dependency of the flux terms has

not been explicitly defined. This decision would render the resulting numerical scheme implicit,

explicit or Crank-Nicholson or semi-implicit.

4.2.2 Conservation of Momentum

Figure 4.2: Control volume

Let us now consider the case of conservation of linear momentum. For the CV depicted in Fig. 4.2,

one can define:

Extensive property = Linear momentum = ~P = M~v = (ρA∆x)~v

Intensive property = Momentum per unit mass = ~v

Flux of

Extensive property = mass flux × intensive property = (ρ~v)~v

by advection

For the case of the momentum balance equation, Reynolds transport theorem states:

d ~PCMdt

=

time rate increase

of momentum

in CV

+

net rate of flux of

momentum out

the CS

(4.12)

where d ~PCMdt = ~FCM , as defined by the second law of Newton applied to the CM. The application of

the Reynolds transport theorem assumes that the CM and CV coincide at the same initial time,

which allows one to write ~FCMs = ~FCV . The conservation statement defined by Eqn. 4.12 can thus

52

be written for a discrete period of time ∆t as:

~FCV ∆t =

increase of momentum

in CV

during ∆t

+

net flow of momentum

crossing the CS

during ∆t

(4.13)

Where:change of momentum

in a CV

during ∆t

= (ρA∆x~v)t+∆tx − (ρA∆x~v)tx

momentum flux

entering the CV surface at LHS

during ∆t

= (ρ~v~vA)t′

x−∆x2

∆t

momentum flux

leaving the CV surface at RHS

during ∆t

= (ρ~v~vA)t′

x+ ∆x2

∆t

Thus Eqn. 4.14 is obtained;

~FCV ∆t =[(ρ~vA)t+∆t

x ∆x− (ρ~vA)tx∆x]

+[(ρ~v~vA)t

x+ ∆x2

∆t− (ρ~v~vA)t′

x−∆x2

∆t]

(4.14)

Equation 4.14 states that the sum of all forces (surface and body forces) acting on the finite CV over

the finite period of time ∆t equals the accumulation of linear momentum in the CV plus the net

flux of momentum though the CV surface that takes place over during the time period ∆t. Forces

(surface and body forces) acting on the CV of fluid of interest are shown in Fig. 4.3. Body Forces

( ~FB) act on the entire system (i.e. gravitational forces) while Surface Forces ( ~Fs) are transmitted

through an area or surface. The net force acting on the CV is hence given by:

Figure 4.3: Force Balance in an Elementary Pipe CV

53

~FCV = ~FS + ~FB (4.15)

Surface forces can be classified as:

1. Tangential forces: Those tangential to flow surface of CV (i.e. friction or shear force).

2. Normal forces: Those perpendicular to the flow surface of CV (i.e. pressure). Pressure force

is isotropic (i.e. it has the same magnitude in all directions) and it is always considered to be

applied in compression (not tension).

From the free body diagram in Fig. 4.3, one can readily derive:

~FCV =

Net force

in the direction

of flow

= −[(pA)x+ ∆x

2− (pA)x−∆x

2

]− [τws]− [Mg sin θ] (4.16)

Where:

θ Angle of inclination [rad]

s CV surface in contact with pipe or total wetted area [m2]

τw Shear stress [Nm−2]

Shear stress (τw) in 1D flow is typically evaluated as a function of the dimensionless friction factor

coefficient. By definition, friction factor is a measure of the ratio of shear stress to inertial force per

unit area. Darcy-Weisbach friction factor is defined as:

f = 4τw

12ρv

2(4.17)

Substituting the force balance in Eqn. 4.16 into the conservation statement in Eqn. 4.14, one can

obtain:

[(ρ~vA)t+∆t

x ∆x− (ρ~vA)tx∆x]

+[(ρ~v~vA)t

x+ ∆x2

∆t− (ρ~v~vA)t′

x−∆x2

∆t]

= (pA)t′

x−∆x2

− (pA)t′

x+ ∆x2

+ [−τws− ρA∆xg sin θ]t′

x ∆t (4.18)

Equation 4.18 can now be re-written per unit volume and time as shown below:

(ρ~vA)t+∆tx − (ρ~vA)tx

∆t+

(ρ~v~vA)t′

x+ ∆x2

− (ρ~v~vA)t′

x−∆x2

∆x

= −(pA)t

x+ ∆x2

− (pA)t′

x−∆x2

∆x+ [−τwLw − ρgA sin θ]t

x (4.19)

Equation 4.19 represents the 1D linear momentum conservation equation for a finite volume element.

54

In this expression, Lw is wetted perimeter (Lw = πD). When one takes the limit as ∆t,∆x→ 0,

the PDE form of conservation of momentum is obtained:

∂(ρ~vA)

∂t+∂(ρ~v~vA)

∂x= −∂(pA)

∂x− τwLw − ρgA sin θ (4.20)

One can also derive the FVM expression in Eqn. 4.19 by direct integration of the PDE in Eqn. 4.20,

as shown in the previous section for the case of mass conservation.

4.2.3 Conservation of Energy

Figure 4.4: Energy Balance in an Elementary Pipe CV

For the case of energy conservation in a CV (Fig. 4.4), one defines:

Extensive property = Energy = E = Me = (ρA∆x)e

Intensive property = Energy per unit mass = e = e∗ + v2

2 + g∆zel

Flux of

extensive property = mass flux × intensive property = (ρ~v)e

by advection

Where:

∆zel System elevation with respect to a datum z0 Since x represents the distance

from center of CV to the entrance of pipe: ∆zel = x sin θ

θ Pipe inclination

For the case of energy conservation over the control volume in Fig. 4.4, the Reynolds transport

theorem states:

d ~ECMdt

=

time rate of change of

energy in CV

+

net rate of

energy leaving

through the CS

(4.21)

55

Where;∂Esys

∂t =(Q− W

)sys

, as required by the first law of thermodynamics and(Q− W

)sys

=(Q− W

)CV

because CM and CV are assumed to coincide at the initial time.

For a finite period of time ∆t Eqn. 4.21 becomes:

(Q− W

)CM

∆t =

change of energy

in CV

during ∆t

+

net flow of energy

out of CS

during ∆t

(4.22)

Where:change of energy

in CV

during ∆t

= (ρA∆xe)t+∆tx − (ρA∆xe)tx

energy flux

entering the CV surface at LHS

during ∆t

= (ρ~veA)t′

x−∆x2

∆t

energy flux

leaving the CV surface at RHS

during ∆t

= (ρ~ve)t′

x+ ∆x2

A∆t

change in energy

in CV due to heat and work

during ∆t

=(Q− W

)CV

∆t

Thus;(Q− W

)CV

∆t = +[(ρeA)t+∆t

x ∆x− (ρeA)tx∆x]

+[(ρ~veA)t

x+ ∆x2

∆t− (ρ~veA)t′

x−∆x2

∆t]

(4.23)

Heat transfer with the surroundings (Q) in pipe flow is typically calculated as a function of an

overall heat transfer coefficient (cU ). The overall heat transfer coefficient (cU ) relates the total heat

transfer (Q) to the surface area available for heat transfer (s) and temperature difference (∆T ), as

shown below:

Q = cUs∆T (4.24)

Where:

s = πD∆x

∆T = (T∞ − T )

Q is positive when heat enters the CV (T∞ > T ) and negative if heat is leaving the CV (T∞ < T ). cU

is a required design parameter in the calculation of energy balances of comprehensive pipeline models.

It characterizes the heat interaction between the flowing fluid and the environment (i.e. materials

surrounding the pipeline). In non-insulated pipes, pipe walls are considered infinitely conductive

56

because they do not pose any significant heat flow resistance due to the typically large conductivity

of pipe materials and usually small wall thicknesses. In such situations, cU basically equals the

effective convective coefficient between external side of the pipe and the environment. Values of cU

ultimately depend on the environmental conditions around the pipe and for example, whether the

pipe is buried or not. For a buried pipe, typical values of cU are 0.1 to 2 BTU/ft2 − hr−o F , while

it could vary between 2 and 20 BTU/ft2 − hr −o F when the pipeline is directly exposed to the

atmosphere.

For buried pipes, cU ends up being more or less equal to the conductivity of the ground or soil. This

statement assumes that soil temperature is prescribed at a distance L = 1ft from the pipe external

diameter, since cU = k/L for approximately constant heat flow areas. Thermal conductivity of

earth materials is found around k = 1.4BTU/ft2 − hr −o F , with values changing with respect

to soil type, soil density, moisture content, salt concentration, among others. While soil thermal

science literature provide several correlations for conductivity as a function of these properties,

reasonable values of cU can be determined from known values used for old lines placed in similar

soil conditions. Most authors, however, would assume the average value of 1 BTU/ft2 − hr −o Ffor buried pipes for preliminary calculations. When a pipeline is directly exposed to the atmosphere,

cU becomes practically equal to the convection heat transfer coefficient between the pipe external

surface and the air surrounding it. A pipe can be exposed to either natural (free) convection or

forced convection in windy environments. For reliable estimations of cU for pipes under external

forced convection, one can use correlations of ‘convective coefficient’ for flow across cylinders found

in standard heat transfer books. Typical values of heat transfer coefficients for forced gas convection

can be around 2 to 50 BTU/ft2 − hr −o F . When pipeline goes through a water body (i.e. river),

this value can be greater than 20 because heat transfer coefficients for forced convections in liquids

can be found within the range of 2 to 2000 BTU/ft2 − hr −o F .

In the absence of rotating shafts, work can be done on a CV by forces or agents that can influence

fluid expansion/compression or cause the deformation of the CV surface. The pressure force, for

instance, causes the fluid to enter and leave the control volume and does work, which is typically

called flow energy or flow work in the thermodynamic analysis of open systems. Flow work

is done by the inlet pressure to push the fluid into the control volume, and work is done by the

control volume to push the fluid out of the system. The presence of terms accounting for flow

work in an energy balance gives rise to enthalpy terms, when combined with internal energy values.

The friction force, on the other hand, can not do work because it is a stationary force unable to

cause deformation of the CV boundaries or any fluid expansion/compression. The friction force

is a stationary force applied at the boundary, and stationary forces do not do work. Wall shear

or frictional effects are often associated with heat generation inside the CV, but internal heat

generation due to friction is largely neglected in this type of formulations.

In a flow system, the rate at which work is done by a force is equal to the product of that force

times the velocity component in the direction of the force:

W = ~F · |~v| (4.25)

57

Where, velocities are evaluated at the fluid point where the force is being applied. Therefore, the

work done by pressure force can be described as:

WCV = −[(pA~v)t

x+ ∆x2

− (pA~v)t′

x−∆x2

](4.26)

The convention positive work (+W ) for forces in the direction of flow (work done by the fluid) and

negative work (−W ) for forces acting against the flow (work done on the fluid) is maintained. It is

important to point out that, other than pressure, none of the forces present in the free body diagram

of the CV (Fig. 4.3) can be considered to do work on the fluid. They do not cause deformation

of the CV surface or fluid expansion/compression. Such is the case for both friction and gravity

forces. The contribution of gravity (or the gravity force) to the energy balance is, however, fully

captured by the definition of total intrinsic energy (e), where the displacement of the fluid through

the gravitational field comes associated with changes in fluid’s potential energy.

By introducing the definitions of heat transfer and work into the energy conservation statement of

Eqn. 4.23, one obtains:

[(ρeA)t+∆t

x ∆x− (ρeA)tx∆x]

+[(ρ~veA)t

x+ ∆x2

∆t− (ρ~veA)t′

x−∆x2

∆t]

− [cULw∆x∆T∆t]x −[(pA~v)t

x+ ∆x2

− (pA~v)t′

x−∆x2

]∆t = 0 (4.27)

Or equivalently, dividing the equation Eqn. 4.27 by ∆x and ∆t;

[(ρeA)t+∆t

x − (ρeA)tx]

∆t+

[(ρ~veA)t

x+ ∆x2

− (ρ~veA)t′

x−∆x2

]∆x

+

[(pA~v)t

x+ ∆x2

− (pA~v)t′

x−∆x2

]∆x

− [cULw∆T ]x = 0 (4.28)

Or, in terms of enthalpy (h) flux:

[(ρeA)t+∆t

x − (ρeA)tx]

∆t+

[(ρ~vhA)t

x+ ∆x2

− (ρ~vhA)t′

x−∆x2

]∆x

− [cULw∆T ]x = 0 (4.29)

Equation 4.28 represents the 1D statement of conservation of energy for a finite control volume

implemented in this study. Again, this expression collapses to the PDE form as shown in Eqn. 4.30

after taking the limits ∆t,∆x→ 0:

∂(ρeA)

∂t+∂(ρ~veA)

∂x+∂(pA~v)

∂x− [cULw∆T ] = 0 (4.30)

58

Or;∂(ρeA)

∂t+∂(ρ~vhA)

∂x− [cULw∆T ] = 0 (4.31)

Where; Lw = πD

4.3 Two-Fluid Model Equations

In this study, a one-dimensional, single-pressure two-fluid model is developed with the assumption

of thermodynamic equilibrium; i.e. mechanical, thermal and chemical equilibrium. Mechanical

equilibrium suggests that phasic pressures are same at any point over the cross-section of the

conduit. Similarly, local thermal equilibrium means that phases are at the same temperature over

the cross-section; hence, it is sufficient to use a single energy equation ensuing in the 5 equation

single-pressure two-fluid model.

The two-fluid model treats each phase in a multiphase flow environment as a separate fluid with

its own set of governing equations. In the most detailed case each phase is considered to have its

own velocity, temperature and pressure. In this study, however, the time lag associated with the

attainment of thermal and mechanical equilibrium between the phases is considered insignificant

in comparison to the characteristic time it would take for pressure and temperature conditions to

change during flow along the axis of pipe. Therefore, pressure and temperature non-equilibrium is

neglected and pressure and temperature are assumed to be same for both phases.

It is important to note that some sort of an averaging process (e.g. volume) is almost always

involved in the derivation of governing two-fluid model PDEs (and FDEs thereof), regardless of

the flow dimensions involved; i.e. for 3D and 2D models with Navier-Stokes equations as well.

The need for an averaging process essentially stems from the fact that the two phases are indeed

interpenetrating and there are moving boundaries between the phases to account for; at macroscopic

as well as microscopic (local) scale. A ‘local instant’ formulation is possible without resorting to

an averaging and results in the ‘direct numerical simulation’ approach. However, besides being

computationally too demanding, for most practical applications only ‘average’ flow properties are

of interest. Governing equations in this text are derived using the ‘macroscopic balance’ approach

because it is easier to visualize, has a physical basis and thus simpler to comprehend. Nevertheless,

it can mask the averaging process if one is not careful. Hence, it should be remembered at all times

that flow properties (e.g. ρ, v, α) represented in the equations of this text are average properties

both at microscopic (local) and macroscopic (control volume) scale. Extensive detail on various

averaging processes can be found in the texts by Prosperetti and Tryggvason (2007); Kolev (2005);

Ishii and Hibiki (2003); Todreas and Kazimi (1990).

4.3.1 Conservation of Mass

One might write separate conservation statements for each of the flowing phases found in the CV in

Fig. 4.5. In the case of mass conservation for the liquid, one considers:

59

Figure 4.5: Control Volume for Stratified Two-Phase Flow

Extensive property = Liquid mass = ρLVL = ρL(αLA∆x)

Superficial liquid mass flux =Liquid mass flow

Total area=ρL ~vLAL

A= ρL ~vLαL

Intensive property = liquid mass per unit mass = 1 (one)

The statement of conservation of liquid mass, based on Reynolds Transport theorem is written as:

dMCM

dt= 0 =

increase of liquid mass

in CV

over period ∆t

+

net efflux of liquid

out of CS

during period∆t

(4.32)

Where:increase of liquid mass

in CV

over period ∆t

=[(ρLαLA)t+∆t

x ∆x− (ρLαLA)tx∆x]

net efflux of liquid

out of CS

during period ∆t

=

−(ρL ~vLαLA)t′

x+ ∆x2

∆t

mass leaving CS via advection

+(ρL ~vLαLA)t′

x−∆x2

∆t

mass entering CS via advection

+(ΓLA∆x)t′x∆t

mass crossing CS due to mass transfer

Here; t′ is any suitable time, either t′ = t or t′ = t+ ∆t, which renders the numerical scheme either

explicit or implicit. Therefore, the 1D Finite volume statement of mass conservation for the liquid

phase can be written as:

(ρLαLA)t+∆tx − (ρLαLA)tx

∆t+

(ρL ~vLαLA)t′

x+ ∆x2

− (ρL ~vLαLA)t′

x−∆x2

∆x= ΓLA = −ΓGA (4.33)

60

Equivalently for the gas phase we can write:

(ρGαGA)t+∆tx − (ρGαGA)tx

∆t+

(ρG ~vGαGA)t′

x+ ∆x2

− (ρG ~vGαGA)t′

x−∆x2

∆x= ΓGA = −ΓLA (4.34)

The PDE form of the continuity equation for the two-fluid model can be obtained as one takes the

limit ∆t,∆x→ 0 in these expressions. For a generic phase k we have:

∂t(ρkαkA) +

∂x(ρk ~vkαkA) = ΓkA (4.35)

Where; αG + αL = 1 and ΓG + ΓL = 0.

4.3.2 Conservation of Momentum

Figure 4.6: Control Volume for Stratified Two-Phase Flow

In the case of momentum conservation for the liquid phase, one identifies:

Extensive property = Linear momentum = ρLVL ~vL = ρL(αLA∆x) ~vL

Intensive property = Momentum per unit mass = ~vL

Flux of = Liquid mass flux × intensive property = (ρL ~vLαL) ~vL

extensive property

The momentum conservation statement for the CV in Fig. 4.6 is thus, for the liquid phase:

~FCV ∆t =

increase of momentum

in CV

during ∆t

+

net efflux of momentum

leaving the CS

during ∆t

(4.36)

Where:

61increase of momentum

in CV

during ∆t

= (ρL ~vLαLA∆x)t+∆tx − (ρL ~vLαLA∆x)tx

net momentum efflux

out of CS

during ∆t

= −

momentum leaving CS

via advection

+

momentum entering CS

via advection

+

momentum gained by CV

due to mass transfer

Momentum leaving CS via advection −(ρL ~vL ~vLαLA)t′

x+ ∆x2

∆t

Momentum entering CS via advection +(ρL ~vL ~vLαLA)t′

x−∆x2

∆t

Momentum gained by CV

as a result of mass transfer+(ΓLA∆x~vc)

t′x∆t

Here; ~vc = ~vL if mass transfer is from gas-to-liquid (ΓL > 0) and ~vc = ~vG if mass transfer is from

liquid-to-gas (ΓL < 0) as discussed by Shoham (2006); Todreas and Kazimi (1990).

The net force acting on the liquid CV is given by:

~FCV ∆t =[(pαLA)t

x−∆x2

− (pαLA)t′

x+ ∆x2

]+ [τisi − τwLswL − ρLαLA∆xg sin θ]t

x ∆t (4.37)

Although Todreas and Kazimi (1990) leaves the derivation at this stage, it is important to note that

as a consequence of the aforementioned averaging process, which is required for the derivation of

two-phase flow equations (see introduction of Sec. 4.3), holdup (αL, or void fraction αG for that

matter) is assumed to be constant all throughout the control volume.

(αL)t′x = (αL)t

x−∆x2

= (αL)t′

x+ ∆x2

Hence; pressure gradient accelerating the liquid volume is defined as the fraction of overall pressure

gradient available for the liquid phase (across the CV) and it is customary to write the holdup

outside the parenthesis to avoid later misconceptions. This matter is further elaborated after

staggered grid is introduced in Sec. 4.7 and then following the discussion of boundary conditions in

Sec. 4.8.

~FCV ∆t = (αL)t′x

[(pA)x−∆x

2− (pA)t

x+ ∆x2

]+ [τisi − τwLswL − ρLαLA∆xg sin θ]t

x ∆t (4.38)

62

Then, 1D Finite Volume equation for liquid phase momentum balance is:

(ρL ~vLαLA)t+∆tx − (ρL ~vLαLA)tx

∆t+

(ρL ~vL ~vLαLA)t′

x+ ∆x2

− (ρL ~vL ~vLαLA)t′

x−∆x2

∆x

= (ΓLA~vc)t′x − (αL)t

′x

(pA)t′

x+ ∆x2

− (pA)t′

x−∆x2

∆x+ [τiLi − τwLLwL − ρLαLAg sin θ]t

′x (4.39)

The PDE counterpart of this equation, for a generic phase k, obtained at the limit ∆t,∆x→ 0 as:

∂t(ρk ~vkαkA) +

∂x(ρk ~vk ~vkαkA) + αk

∂x(pA) = δiτiLi − τwkLwk − ρkαkAg sin θ + ΓkA~vc (4.40)

Where; δi = 1 for k = L and δi = −1 for k = G since τi is positive for the usual case ~vG > ~vL and

τi is negative for the case ~vG < ~vL, by the interfacial friction factor definition given in Eqn. 4.42.

Hence, interfacial force is ‘adding’ to the momentum of liquid phase by accelerating it, i.e. by

dragging it with the faster gas phase velocity.

Phase wall shear stresses are calculated using single phase friction factors calculated for the

corresponding phase velocity:

fk =

[τw

12ρ|~v|~v

]k

(4.41)

Interfacial shear stress is calculated as a function of interfacial friction factors, which are defined

using an inertia term that employs the slip velocity between the phases and the density of the

continuous phase (assumed gas in this study):

fi = 4τi

12ρG| ~vG − ~vL|( ~vG − ~vL)

(4.42)

It is important to note again that a consequence of one-dimensional modeling is the loss of certain

details regarding the flow; i.e. information pertinent to the ‘eddies’ in the case of turbulent flow.

Nevertheless, associated hydraulic losses along the axial direction of the conduit are captured

through the use forcing functions (friction forces) with one-dimensional Euler equations. Friction

force terms are calculated using friction factors (Eqns. 4.50 and 4.51) based on Reynolds number of

the flow which is an indicator of the flow mode; i.e. laminar, transition or turbulent (cf. Sec. 4.4).

4.3.3 Conservation of Energy

Based on the thermal equilibrium assumption, a single, overall (mixture) energy equation is adequate

for the two-fluid model approach.

63

Figure 4.7: Control Volume for Stratified Two-Phase Flow

Extensive property = Energy = Ek = Mkek = (ρkAk∆x)ek

Intensive property = Energy per unit mass = ek = e∗k +v2k2 + g∆zel

Flux of

extensive property = mass flux × intensive property = (ρk ~vkαk)ek

by advection

Where;

∆zel System elevation with respect to a datum z0 Since x represents the distance

from center of CV to the entrance of pipe: ∆zel = x sin θ

θ Pipe inclination

The conservation statement for the liquid CV becomes:

(Q− W

)CV

∆t =

increase of Total energy

in CV

during ∆t

+

net flux of energy

leaving CS

during ∆t

(4.43)

Where:increase of total energy

in CV

during ∆t

= [ρGαGA∆teG + ρLαLA∆teL]t+∆tx − [ρGαGA∆teG + ρLαLA∆teL]tx

net flux of energy

leaving CS

during ∆t

=+ (ρk ~vkαkAek)

t′

x+ ∆x2

∆t

energy leaving CS via advection

+ (ρk ~vkαkAek)t′

x−∆x2

∆t

energy entering CS via advection

Q = [cUsw∆t]t′

x

W = (pAG ~vG + pAL ~vL)x−∆x2− (pAG ~vG + pAL ~vL)x+ ∆x

2

64

Please note that work done against the liquid CV incorporates the pressure work contribution and

that of the interfacial friction force. However, since overall energy equation incorporates work done

on both phases, interfacial friction force terms cancel each other. Based on these considerations,

Eqn. 4.43 becomes:

[(ρGαGeG + ρLαLeL)A]t+∆tx − [(ρGαGeG + ρLαLeL)A]tx

∆t

+[(ρG ~vGαGeG + ρL ~vLαLeL)A]t

x+ ∆x2

− [(ρG ~vGαGeG + ρL ~vLαLeL)A]t′

x−∆x2

∆x

= −(pAG ~vG)t

x+ ∆x2

− (pAG ~vG)t′

x−∆x2

∆x−

(pAL ~vL)t′

x+ ∆x2

− (pAL ~vL)t′

x−∆x2

∆x+ [cULw∆t]t

x (4.44)

Or, in terms of enthalpy flux:

[(ρGαGeG + ρLαLeL)A]t+∆tx − [(ρGαGeG + ρLαLeL)A]tx

∆t

+[(ρG ~vGαGhG + ρL ~vLαLhL)A]t

x+ ∆x2

− [(ρG ~vGαGhG + ρL ~vLαLhL)A]t′

x−∆x2

∆x

+ [cULw∆t]t′

x (4.45)

As discussed previously, a two-fluid model for steady one-dimensional flow can be composed of

up to six conservation equations. Such sort of models would include two equations accounting for

continuity, two equations accounting for momentum balance and two equations for energy balance

in the liquid and gas phases. However, use a combined (mixture) energy equation for the system

(by lumping the liquid and gas phase balances together) leads to the five equation version of the

two-fluid model. The use of combined energy equation is predicated upon the assumption of zero

interfacial temperature gradient (i.e. thermal equilibrium) and it could also be obtained by summing

up individual, phasic energy conservation equations. The PDE version of the combined energy

statement can be obtained as one takes the limits ∆t,∆x→ 0, as shown:

∂t[(ρLαLeL + ρGαGeG)A] +

∂t[(ρL ~vLαLhL + ρG ~vGαGhG)A] = cULw∆t (4.46)

It is important to note that heat transfer coefficient cU should indeed become a function of flow

pattern, gas and liquid phase conductivities since liquid and gas phases typically have different heat

conductivities. However, in this study cU is assumed constant for practical purposes based on the

assumption of low liquid loading conditions.

4.4 Closure Relationships

The use of a one-dimensional two-fluid model requires describing the interaction between phases

themselves and the pipe wall, which depends on the nature of the flow pattern describing the

gas-liquid distribution over the pipe cross-section. One of the important phase interactions that

65

needs to be properly described is the interfacial mass transfer term. Interfacial mass transfer between

the phases can be calculated in different ways which depend on the nature of flowing fluids and

version of the two-fluid model. Two immiscible phases, for example, would not exchange mass and

this calculation is omitted. TRAC, for example, calculates mass transfer between the phases as a

function of heat input and latent heat required to vaporize the liquid phase (Spore et al., 2001).

For the case of natural gas condensation in a pipeline, mass transfer from gas phase to liquid phase

within two neighboring cells is computed based on thermodynamic equilibrium gas mass fractions of

neighboring cells (Mucharam et al., 1990):

ΓL =[(ρGvGαG + ρLvLαL)A](fmgi+1 − fmgi)

Adx(4.47)

Where; fmgi is the (thermodynamic) equilibrium gas mass fraction for the flow stream (hence the

composition) crossing the cell face j and evaluated at p− T conditions of upstream cell i; while,

fmgi+1 is the equilibrium gas mass fraction for the same flow stream evaluated at p− T conditions

of downstream cell i+ 1. Equilibrium gas mass fractions are computed via an EOS based on p− Tconditions of pertinent cell.

It is important to recognize that this mass transfer model assumes thermodynamic equilibrium and

ignores kinetic aspect of the mass transfer process by considering conclusion of mass transfer within

the time it takes for flow stream to traverse the CV. While this may not be the best solution for

studies involving small sized CVs where significant or sharp changes in pressure and temperature

between neighboring cells is expected, it becomes a reasonable assumption for studies with large

CVs where the time it takes for flow to enter and leave the CV is long enough for mass transfer to

be complete or where change in pressure and temperature between neighboring CVs is insignificant.

Momentum transfer accompanying the mass transfer is defined by the momentum transfer term

present in Eqn. 4.39:

Momentum source term for phase k = ΓkA~vc (4.48)

Where; ~vc = ~vL if mass transfer is from gas-to-liquid (ΓL > 0) and ~vc = ~vG if mass transfer is from

liquid-to-gas (ΓL < 0) as discussed by Shoham (2006); Todreas and Kazimi (1990).

Another important parameter, which accounts for the momentum exchange between the phases, is

the interfacial force (or interfacial stress) as defined in Sec. 4.3.2. Relative motion of the phases

is usually described as the difference between the gas phase velocity and that of the interface.

Interface velocity is typically approximated as being equal to the liquid phase velocity and surface

area between the phases is defined as the ratio of the surface area of contact between the phases

and the total volume of the fluid. Calculations of such interfacial parameters as friction factors,

surface areas as well as contact surfaces between each phase and the pipe wall and associated wall

frictions are carried out based on the flow regime.

These flow regime dependent closure relationships simply involve calculations based on flow geometry

and can be found in detail in several text books (i.e. Shoham, 2006) or in the text by Ayala (2001),

66

hence will not be repeated here extensively. However, calculation of pertinent parameters for two of

the utilized flow patterns is shown for the sake of completeness.

Wall shear force calculation:

Fwk = Awkfwk4

ρk|vk|vk2

(4.49)

For laminar flows wall friction factor (Darcy-Weisbach) is calculated as:

f =64

Rek(4.50)

For turbulent flows wall friction factor (Darcy-Weisbach) is calculated using Chen (1979) formula:

1√f

= −2 log10

3.7065Dk− 5.0452

Reklog10

(1

2.8257

Dk

)1.1098

+5.8506

Re0.8981k

)](4.51)

Where:

ε Pipe roughness [m]

Dk Phasic hydraulic diameter [m]

Rek Phasic Reynolds number [dimensionless]

Phasic hydraulic diameter is defined as:

Dk = 4Volume of phase k

Area wetted by phase k(4.52)

Interfacial friction force calculation:

Fi = Aifi4

ρG|vG − vL|(vG − vL)

2(4.53)

Flow pattern dependent phasic hydraulic diameter, wetted area and interfacial area calculations are

given in Tab. 4.1 for mist and stratified flow patterns, where rdrop is mean droplet radius.

Table 4.1: Flow pattern dependent parameters

Flow pattern: DG = DL = AwG = AwL = Ai =

Stratified Flow

(παG

π − θ + sin θ

)D

(παLθ

)D

(π − θπ

)4

D

π

)4

D

(sin θ

π

)4

D

Mist Flow D D αG4

DαL

4

D

3(1− αG)

rdrop

67

4.5 Primary Unknowns

Single-phase flow in a conduit can be described by three physical parameters, namely; pressure

(p), velocity (v) and temperature (T ), where temperature is no longer an unknown in the case

of isothermal flow. Governing algebraic equations are solved for these three primary parameters,

hereafter referred to as primary unknowns or simply the ‘primaries’. The choice of primary unknowns

is based on the fact that they are all fundamental and directly measurable properties of the flow

and fluid. Density and fluid energetic content (i.e. internal energy or enthalpy) are derived from

p− T information by applying a suitable Equation of State (EOS). In this work, the Peng-Robinson

EOS is implemented for fluid property predictions (Peng and Robinson, 1976). Detailed information

about the implementation of Equations of State and associated fluid property prediction is found in

Appendix C.

For the case of two-phase flow, two additional primary unknowns are added to the system, liquid

phase velocity (vL) and void fraction (α). A list of primary unknowns and related secondary

parameters are given in Tab. 4.2 and

Table 4.2: Primary unknowns

Primary unknown Secondary parameter

Pressure (p) Phasic density (ρk, through EOS)

Gas phase velocity (vG) —

Liquid phase velocity (vL) —

Void fraction (αG) —

Temperature (T ) Phasic density (ρk, through EOS)

4.6 Advective Property: Donor Cell Scheme

Looking at the algebraic equations developed for FVM, it is recognized that values of advected

properties, particularly mass (or density), momentum (or momentum per unit mass – velocity)

and energy (or energy per unit mass – specific energy), are required both at the center (as the CV

average) and at the edges of CVs. Unfortunately, not all of these values can be left as unknowns in

the system of governing equations. Otherwise, number of equations and number of unknowns are

not balanced. Therefore, in order to balance number of unknowns and the equations, typically edge

values are written in terms of cell center values through various approximation schemes. Here, it

should be noted again that cell center values of parameters in a CV also represent the volumetric

average of these values for that CV.

Focusing on the continuity equation and assuming that all required velocity values are available (i.e.

velocity field is known a priori and not a concern); approximation of density values at cell edges is

needed. A linear interpolation between the cell center densities of two neighbor CVs carries the

68

potential danger of yielding unrealistic results, especially for transient analysis, as evidenced by the

donor cell exercises provided by Patankar (1980).

One remedy to this problem is the donor cell scheme, also known as the upwind method or upstream

weighting. Donor cell is, in essence, based on the foundation that fluid passing through a point

carries the information from its upstream but has little or no information about the downstream.

Therefore, upstream weighted averaging (100% upstream in the context of this text) is favored

over linear averaging (which corresponds to central difference scheme) in the case of advective flow

modeling, specifically for the case of transient analysis.

4.7 Staggered Grid

When deciding for the nodes where primary unknowns will be evaluated in FVM equations, following

the donor cell discussion, pressure and temperature are solved for at the cell centers since density is

calculated based on p− T information in the case of compressible flow. The intuitive approach is

to place all primary unknowns (p, v and T ) at the same node, cell center, as a convenient choice.

This approach is typically known as the ‘co-located’ or coincidental grid, where all unknowns are

calculated at the center of the same grid block and typically is the default setting for FDM based

models. A typical coincidental grid is shown in Fig. 4.8. In this arrangement, mass, energy, and

momentum CVs coincide and there is a single computational grid for the placement of unknown

variables.

Figure 4.8: Coincidental Grid

Placing pressures and velocities at the same node has the potential danger of producing wavy

(checkerboard) patterns for pressure and velocity values, particularly in the case of incompressible

flow. This potential danger associated with the use of coincidental grid is discussed in detail by

Patankar (1980). In this type of grids, calculation of the pressure forces driving the flow necessarily

employ an average of pressure values of neighboring cells centers, causing the actual CV pressure

value to be leapfrogged in the calculation. Checkerboard pressures can not be effectively dissipated

in these co-located numerical grids, since pressure information does not appear explicitly in the

governing equations but rather appears in the form of a momentum source (pressure drop) or

through fluid densities. In the case of incompressible flow, continuity equation can neither help to

deduce the pressure information (since density is not a function of pressure in this case) nor be

used to determine the velocity field (both pressure and velocity are unknowns). The difficulty of

wavy pressure patterns is likely to be eliminated in the case of compressible flow where density is

an explicit function of pressure. In those cases, pressure values would be constrained by density

values (or vice a versa) through the use of an appropriate EOS and the physical connection between

velocities and pressures are then established by the FVM equations.

69

As presented by Patankar (1980), the most traditional approach to avoid the checkerboard pressure

problem is the implementation of staggered grids in the computational domain. A staggered grid

implies the definition of non co-located, staggered location for the primary unknowns. Typically,

cell pressures are evaluated at the cell centers but fluid velocities are defined at the cell edges.

This implies the definition of two, displaced numerical grids, as shown in Fig. 4.9, one used for the

implementation of mass and energy FV equations (p-cells) and the other for the implementation of

the momentum FV equations (v-cells). The basis of this approach is that neither the nodes where

primary unknowns are calculated nor the CVs for the conservation equations have to coincide in

order to obtain a numerical solution.

Figure 4.9: Staggered Grid

Besides preventing checkerboard type pressure fields, the use of a staggered grid arrangement brings

additional advantages. The apparent, direct advantage is that velocities are now evaluated between

pressure couples that are actually driving the flow and thus a solid physical link is established.

In fact, paying closer attention to the FVM equations and their derivation, it is observed that

the natural location for velocities is the cell edges. This is because edge velocities are frequently

needed in the governing equations for the calculation of fluxes at cell edges. The staggered approach

will also facilitate the formulation of conservation equations for T-junctions, which is discussed in

Sec. 4.12.

The staggered grid approach is also applied in both OLGA (Bendiksen et al., 1989) and TRAC

(Spore et al., 2001). While details of the application of this computational grid were not available

for OLGA, the staggered grid approach utilized in this study is slightly different than the form

applied in TRAC. TRAC employs the non-conservative form of the momentum equations meaning

that momentum equations are written specifically at the edges of cells. The non-conservative form

of the momentum equations are derived from the conservative FVM equations and one advantage

of using the non-conservative form is that it eliminates the need for extensive use of cell center

velocities, which have to be approximated using edge values, in momentum equations. Nevertheless,

in this study, the conservative form of the momentum equations is employed in order to assure and

reinforce rigorous conservation.

It is important to point out that Rhie and Chow (1983) have presented a pressure treatment that

allows the formulation of coincidental grids that do not suffer from the checkerboard pressure problem.

In their coincidental grid application, all primary unknowns are evaluated at cell centers and CVs

of the conservation equations coincide (Fig. 4.8). Their method requires the implementation of an

70

additional pressure equation that decouples implicit pressure-velocity information of the governing

equations and allows the formulation of an explicit link for pressure information. Co-located methods

are particularly useful with complex, irregular geometries.

A low accuracy, simple alternative to staggered grid is using forward differences for momentum flux

terms and backward differences for pressure gradient term (Wesselling, 2001). However, in this

study staggered grid is applied in order to extend governing Euler equations over the interfaces of

tee control volume.

As a consequence of the staggered grid arrangement, momentum equations are now written for a

displaced momentum CV overlapping two neighboring mass CVs. This requires that the average

holdup (or void fraction) for the momentum CV (multiplying the pressure gradient term as discussed

when two-phase momentum PDE is derived in Sec. 4.3.2) described as a function of neighboring

mass CV holdups. In this study, a volumetric average of neighboring mass CV holdups (reduces to

a linear average when diameter is same for both mass CVs) is employed (Frepoli et al., 2003).

4.8 Boundary Conditions

A potential inconvenience of the staggered grid approach is the handling of boundary conditions

(BC). Because of the displaced nature of the staggered grid for mass and momentum equations

(see Fig. 4.9), it can be seen that first momentum cell is not directly connected to pipe inlet

information while center of last momentum CV exactly coincides with pipe outlet. This last cell,

moreover, has half of its volume defined outside the actual pipe physical domain. As a consequence

of this staggered arrangement, none of the momentum equations written for the system explicitly

incorporate information defined at the pipe inlet (i.e. inlet pressure). Consequently pressure is

specified at pipe outlet while velocity (or mass flow rate) and temperature are specified at pipe

inlet. Then, the flow problem is stated as:

• Flow rate qin is injected at temperature Tin at point ‘inlet’ to be delivered at pressure pout, at

point ‘outlet’.

In fact, this arrangement of boundary conditions is quite useful for the analysis of branching

T-junction problem as outlet pressures have significant influence on the flow split.

Nevertheless, it should be noted that for flow specification in the reverse direction, pressure becomes

specified at the inlet while flow rate is now being specified at the outlet. Furthermore, by adding an

additional momentum CV at the beginning of the grid, thus increasing number of momentum CVs

one more than mass CVs, it could be possible to link inlet pressure specification to the pipe. Of

course, this would require removing the outlet pressure specification.

Prosperetti and Tryggvason (2007) discuss that flow at the pipe outlet should essentially mimic an

adequately long flow domain where outflow has no effect on the solution; i.e. as if there is no real

outlet but outlet location is a point within an infinitely long conduit where fully developed flow

is taking place and ‘boundary layer’ effects are insignificant. This is typically accommodated by

assuming that flow profile is relatively smooth around the outlet and flow continues beyond the

71

outlet without being disturbed (i.e. no jumps in the pressure and velocity profiles). Such a smooth

outlet condition is typically imposed on the velocity profile since pressure is already specified as

BC. Prosperetti and Tryggvason (2007) lists four general outflow boundary conditions that allow a

smooth exit of the flow; (1) convective, (2) parabolic, (3) zero gradient and (4) zero second gradient

(zero curvature).

In this study, phasic velocities at the edge of the last momentum CV, temperature and void fraction

values for the imaginary boundary mass CV (i.e. ghost or phantom cell) are all calculated based on

the zero second gradient condition (zero curvature, i.e. free flow). In other words, these properties

are directly calculated as a linear extrapolation of the values of previous two CVs.

Inlet flow rate specification is converted to inlet phasic velocities based on void fraction specification

and phasic density calculation at inlet p− T conditions. For isothermal studies, inlet pressure does

not require updating. However, for thermal studies inlet pressure is also updated using the zero

curvature method. Otherwise, profiles of flow parameters display jumps, disturbances over the first

couple of cells; associated with the sudden expansion of gas phase into the pipe and Joule-Thompson

cooling.

On the other hand, there are three options for inlet void fraction specification; (1) fixed (2) calculated

internally based on no-slip condition, which also sets inlet phasic velocities equal (3) updated based

on zero curvature. While first and second options cause disturbances in flow profiles near the

inlet, third option adheres completely with the smooth, fully developed inlet flow condition, hence

preferred for steady-state analysis.

Developed model also has the option of specifying outlet void fractions as boundary conditions,

solely for checking purposes. However, an outlet void fraction (or holdup) specification is quite

a counter-intuitive attempt since; typically, there is no true way of controlling or knowing it in

advance, unlike pressure or flow rate. This is more clear when a branching tee is considered; as

phase separation at the junction is not known in advance, outlet void fraction specification(s) would

perhaps impose a certain phase separation ratio on the solution. Besides, specification of void

fraction at the pipe outlet has been observed to cause no significant effect on the profiles of flow

parameters for single pipe study, at least according to the two-fluid model formulation employed in

this thesis.

Singh (2009) specifies outlet void fractions (both of them for the case of a branching tee). However,

through the course of this study, it has been observed that the only time void fraction specification

effects flow profiles of a single pipe is when (and if) the void fraction term is included within the

pressure gradient term (Eqn. 4.39). That is to say, two different void fractions (i.e. void fractions of

neighboring mass CVs) are utilized when calculating the pressure force accelerating the flow stream

within the momentum CV. In fact, only then it is possible to match single pipe, two-phase results

from Singh (2009) with outlet void fraction specifications. This brings the discussion back to the

average holdup issue discussed in Sec. 4.3.2.

72

4.9 Order of Accuracy

From FDM point of view, numerical solutions are approximate solutions of the governing PDEs

where differential operators are replaced with finite difference operators. A finite difference operator

is in essence a truncated Taylor series and, by the formal definition, order of accuracy is the order

of highest order term in this truncated Taylor series. Hence, order of accuracy is a measure of

the accuracy of the numerical solution (i.e. how many digits after the decimal point matches the

analytical solution). How to determine theoretical order of accuracy using Taylor series analysis is

available in great detail in the text by Roache (1998) therefore it will not be repeated here.

From practical stand point, the order of accuracy indicates how much the numerical solution can

improve when one increases (e.g. doubles) the number of steps or CVs (or decreases the step or

CV size) of the solution. Evidently, numerical solution gets closer to the analytical solution with

increasing number of steps (i.e. finer grid). Practical (actual) order of accuracy of a numerical

method can be estimated through Richardson extrapolation (Roache, 1998). It is particularly

important to check the order of accuracy with this technique since practical order of accuracy

is usually lower than the theoretical value determined by Taylor series analysis. Furthermore,

enhancements done in the numerical method with best of intentions may end up lowering the

practical order of accuracy.

4.9.1 Higher Order Approximations

The accuracy of a numerical solution improves with increasing number of CVs, at the expense

of computational cost and time. Typically, there is an optimum number of cells beyond which

improvements are no longer significant

‘First-order accurate’ implies that numerical error associated with the discretization of PDE(s)

decreases linearly with with step size; may it be time or space. For instance, with a discretization

method ‘second-order accurate’ in space, discretization error is expected to decrease quadratically

as CV (step) size is halved.

One important objective in numerical simulation is to decrease the number of optimum cells

necessary to get a reliable solution, and this can be accomplished by implementing higher-order

approximations. Higher-order approximations include more information from the neighboring CVs

(or nodes) when calculating the solution at a particular cell (or node). This can be observed from the

Taylor series perspective of replacing differential operators. However, a more physical way of looking

at higher order approximations is realizing that these are indeed polynomial approximations of the

changes in variables through each cell. On the other hand, a relatively lower order approximation

(i.e. first-order) is merely a linear approximation of these changes. One good example of higher

resolution (higher-order) schemes is the QUICK (Quadratic Upstream Interpolation for Convective

Kinematics) scheme by Leonard (1979) where edge densities needed for the governing FVM equations

are evaluated as:

ρi+1/2 =3

8ρi+1 +

3

4ρi −

1

8ρi−1 (4.54)

73

ρi−1/2 =3

8ρi +

3

4ρi−1 −

1

8ρi−2 (4.55)

Utilization of higher resolution schemes ensues in estimating the change in a variable through a

CV with polynomials. This frequently introduces the wiggles (cf. Sec. 4.10) around shock fronts

and even stability problems. For instance, the study of Jelinek (2005) shows that stability issues

surface for the case of two-phase flow in nuclear systems suggesting that higher-order schemes are

not always suitable. Therefore, in this study, ‘theoretically’ second-order accurate upwind scheme

(the donor cell) is utilized. The donor cell scheme is known to be stable and robust albeit at the

expense of significant numerical diffusion (Frepoli et al., 2003; Spore et al., 2001), which is not a

major concern in the current context of steady-state two-phase flow split at T-junctions.

4.10 Transient Analysis

Transient analysis is concerned with the propagation of a perturbation, an induced change in

flow conditions, traveling through the conduit as a shock front, or as a slower change in the flow

conditions; i.e. a change in the injected fluid density or composition. Sudden changes in the flow

(i.e. change in boundary conditions such as the injection rate or temperature) cause shock fronts

(discontinuities); while slow, gradual changes cause smoother waves (slow transients) in the profiles

of flow parameters. Both types of perturbations are transmitted within the conduit, analogous to

waves. For instance, the information of the change in pressure will travel at the speed of sound waves

in a fluid while a change in the inlet temperature, retaining the wave nature of the perturbation,

travels at the speed of flow when conduction is ignored. Evidently, capturing the propagation of

perturbations which move at different speeds is particularly important when modeling compressible,

thermal flow in long pipes.

The wave equation conveys the information that perturbation is taking place in a continuous medium

and a part of the fluid moves as a whole (i.e. a particle in the flow). In the case of Euler equations,

the information of continuous medium is expressed by the equation of continuity (derived from

conservation of mass) while the information that a part of the fluid moves as whole is supplied

by the equation of motion (derived from Newton’s second law, conservation of linear momentum

Lindsay, 1931).

The conservative form of Euler equations tend to smooth such jumps in the numerical solutions

obtained for flow profiles, despite the inviscid nature of advective Euler equations. This difficulty is

known as numerical diffusion or numerical viscosity1 and typically due to averaging of fluxes or flow

parameters to obtain second order accurate approximations. In fact, from Taylor series point of

view, the second order terms that end up being included in the solution are the terms that resemble

viscosity/diffusivity terms.

Utilization of higher resolution schemes (i.e. QUICK, cf. Sec. 4.9.1) allows higher order approxima-

tions and decrease the effects of numerical diffusion, thus capture shock fronts better. However,

such schemes cause non-physical wiggles (oscillations, overshoots and undershoots) in the shock

1Not to be confused with artificial viscosity

74

front profiles of parameters. These wiggles are not due to stability issues but they are merely

mathematical results of the polynomials utilized in the numerical solutions by the higher resolution

schemes (Roache, 1998).

The non-physical wiggles in the solution imply that fluxes of conserved properties are not physical.

Hence, one remedy of preventing these wiggles is to limit computed fluxes to realistic values (values

that can be realized) and make the solution Total Variation Diminishing (TVD – Harten, 1983).

However, typical applications of flux limiters cause practical order of accuracy of higher resolution

schemes to decrease around the shock fronts. Furthermore, it can cause convergence problems.

Another remedy is to introduce an artificial viscosity term in the solution (that will ultimately limit

the flux). However, the adjustment required for such an artificial parameter renders the method

problem specific and prevents it from being a generalized solution (Roache, 1998).

Another method of uniformly retaining higher order accuracy while eliminating the wiggles is

the Essentially Non-Oscillatory (ENO) Schemes that use an adaptive stencil compared to the

fixed stencils of typical higher order schemes. In other words, an ENO scheme decides how many

additional points and in which direction (upstream or downstream) to include in the polynomial

approximation of the variable depending on the position of the shock front, hence prevents possible

oscillations (Harten et al., 1987).

Another option to maintain higher order accuracy is using an Adaptive Mesh Refinement (AMR)

algorithm. The AMR algorithm checks the coarse grid solution of a domain for certain criteria,

such as order of accuracy based on Richardson analysis. The algorithm, then spots the cells over

which the criterion is not met and establishes a finer grid within those cells and repeats the solution.

Therefore, the regions that require smaller CVs for precision are automatically picked and processed

accordingly.

According to the utilized time integration scheme (i.e. explicit, semi-implicit or implicit), stability

of the numerical solution depends on CFL condition (Courant number) which limits the maximum

allowable time-step duration based on grid spacing, flow velocity and velocity of the fastest traveling

perturbation (speed of sound –the pressure wave):

v∆t

∆x+ c0≤ 1 (4.56)

or

∆t ≤ ∆x+ c0

v(4.57)

Where; c0 represents the speed of sound.

Using an implicit scheme removes, in theory, this limit which can be quite constraining. However,

due to the difficulty of capturing wave propagation of different speeds, i.e. flow (void fraction,

density) and pressure wave accurately, it is typically recommended that a material CFL condition is

applied at all times:v∆t

∆x≤ 1 (4.58)

75

or

∆t ≤ ∆x

v(4.59)

4.11 Numerical Treatment

Volume averaged two-fluid model equations (the governing equations), after proper discretization,

could be solved numerically with any method suitable for non-linear system of equations, such

as Newton-Raphson. Due to computational intensity of this direct approach, different classes of

algorithms were developed (Prosperetti and Tryggvason, 2007):

1. Segregated algorithms: Governing equations are solved sequentially, such as SIMPLE

(Semi-Implicit Method for Pressure-Linked Equations) similar to IMPES (Implicit Pressure

Explicit Saturation) algotihm in reservoir simulation. These algorithms are typically suitable

for relatively slow transients and for flows that are not highly coupled and thus not requiring

a simultaneous solution. Typically, continuity (mass balance) and momentum equations are

arranged to form the pressure Poisson equation. During iterations, first the pressure equation

is solved to obtain pressures at iteration level k+1, then pressures at k+1 are utilized to

calculate the velocity field at k+1. Successive pressure corrections are utilized to adjust

densities and velocities to satisfy required conservation at convergence.

2. Partially coupled algorithm: Sequential solution of governing equations using segregated

algorithms causes accumulation of errors if time and spatial step sizes are adequately large.

Forcing iteration process to limit such accumulations with higher degree of convergence

typically ensues in increased computational times. Instead, simultaneous solution of the

pressure Poisson equation with coupled momentum equations yields velocities and pressures.

An automatic time-step control is needed to prevent volume fraction errors from growing

beyond an acceptable limit.

3. Coupled algorithms: Simultaneous solution of governing non-linear equations. Typically

suited for flows where interaction between phases is very strong or problems with short time

scales (fast transients).

4.11.1 Newton-Raphson Method

The set of coupled, non-linear algebraic (discretized) equations that govern the fluid flow are solved

simultaneously using a simple, generalized Newton-Raphson iterative procedure (coupled algorithm).

A brief description of the Newton-Raphson scheme is outlined next.

Let us assume that x∗1 and x∗2 form the actual solution set for equations f1(x1, x2) = 0 and

f2(x1, x2) = 0 and xk1, xk2 are initial approximations (guesses) to this solution. Newton-Raphson

procedure utilizes a turncated Taylor series expansion (Eqn. 4.60 and Eqn. 4.61) in order to come

up with the necessary increments ∆x1 and ∆x2 that can improve the current guesses closer to the

76

actual solution of the system of equations (Eqn. 4.63).

f1(xk1 + ∆xk1, xk2 + ∆xk2) = f1(xk1, x

k2) + ∆xk1

∂f1

∂x1+ ∆xk2

∂f1

∂x2+

(∆xk1)2

2!

∂2f1

∂x21

+(∆xk2)2

2!

∂2f1

∂x22

+ · · ·

f2(xk1 + ∆xk1, xk2 + ∆xk2) = f2(xk1, x

k2) + ∆xk1

∂f2

∂x1+ ∆xk2

∂f2

∂x2+

(∆xk1)2

2!

∂2f2

∂x21

+(∆xk2)2

2!

∂2f2

∂x22

+ · · ·

(4.60)

∆xk1∂f1

∂x1+ ∆xk2

∂f1

∂x2

∼= −f1(xk1, xk2)

∆xk1∂f2

∂x1+ ∆xk2

∂f2

∂x2

∼= −f2(xk1, xk2)

(4.61)

The system of equations in the form of Eqn. 4.61 can be expressed in a matrix formation and solved

for the increments ∆xk1 and ∆xk2 (Eqn. 4.62). Finally, ∆xk1 and ∆xk2 are utilized to get improved

approximations to the solution.

∂f1

∂x1

∂f1

∂x2

∂f1

∂x3· · · ∂f1

∂xn

∂f2

∂x1

∂f2

∂x2

∂f2

∂x3· · · ∂f2

∂xn

∂f3

∂x1

∂f3

∂x2

∂f3

∂x3· · · ∂f3

∂xn

......

.... . .

...

∂fn∂x1

∂fn∂x2

∂fn∂x3

· · · ∂fn∂xn

k

∆x1

∆x2

∆x3

...

∆xn

k

= −

f1

f2

f3

...

fn

k

(4.62)

x∗1 = xk1 + ∆xk1

x∗2 = xk2 + ∆xk2(4.63)

Similarly, algebraic forms of governing equations are written in the ‘residual’ form (Eqn. 4.64) and

then their derivatives with respect to the primary unknowns are numerically computed (Eqn. 4.65)

to construct the Jacobian (coefficient matrix) for the matrix solution (Eqn. 4.66).

Rmssk = 0

Rmomk = 0

Renek = 0

(4.64)

∂Ri∂p∼=Ri(p+ δp, T, v)−Ri(p, T, v)

δp(4.65)

77

∂Rmss1

∂p1

∂Rmss1

∂v1

∂Rmss1

∂p2· · · ∂Rmss1

∂vn

∂Rmom1

∂p1

∂Rmom1

∂v1

∂Rmom1

∂p2· · · ∂Rmom1

∂vn

∂Rmss2

∂p1

∂Rmss2

∂v1

∂Rmss2

∂p2· · · ∂Rmss2

∂vn

......

.... . .

...

∂Rmom1

∂p1

∂Rmom1

∂v1

∂Rmom1

∂p2· · · ∂Rmom1

∂vn

k

∆p1

∆v1

∆p2

...

∆vn

k

= −

Rmss1

Rmom1

Rmss2

...

Rmomn

k

(4.66)

The perturbation amount required to numerically evaluate derivatives should be a function of (1)

the actual variable value and (2) significant digits that can be recognized as a perturbation by the

residual. Besides, perturbation should be small enough to get a good estimate of the analytical

derivative (Frepoli et al., 2003).

δxn = max[σ, (σ × xn)] (4.67)

Where; typically σ = 10−6 but different multiplier (σ) values could be used for different primary

unknowns based on the order of magnitude of the pertinent residual and the primary unknown.

More elaborate discussion on the subject can be found in the text by Dennis and Schnabel (1996).

A fixed multiplier value of σ = 10−8 has proved to be adequate for the purposes of this study.

Convergence is achieved when maximum absolute values of both residuals and the improvements

are below certain limits:

1. |R|max < εresidual = 1e−8

2. |∆x|max < εimprovement = 1e−8

4.11.2 Phase Appearance-Disappearance

A typical problem encountered when conducting multiphase flow studies is the handling of phase

appearance-disappearance throughout the system.

In the case of two-fluid models, a phase can ‘disappear’ for two reasons:

1. Phase can be left behind, i.e. ‘hold-up’, as a result of the hydrodynamic requirements; i.e.

due to low interfacial drag, high wall friction and inertia.

This is a major concern for transient analysis since phases can not be completely ‘hold-up’ at

steady-state conditions due to conservation of mass. This is easier to observe when a system

is considered without mass transfer (i.e. air-water); where all the injected phases has to reach

to the outlet of the conduit for steady-state to be achieved.

2. Phase can disappear as a result of the mass transfer, which is a concern for both steady-state

78

and transient analysis.

Furthermore, because void fraction is a primary unknown and none of the governing equations

mathematically bound its value within the physical limits of [0, 1], void fraction can assume values

beyond the physical limits. This is simply a consequence of the N-R process when iterations swing

around the actual solution; typically when actual solution is located very close to the physical

limits of [0, 1]. Nevertheless, true convergence can only be achieved when conservation is physically

sound. That is to say, if void fraction is persistently going above the upper limit of one (1) or going

below the lower limit zero (0), then this means that hydrodynamic conditions require single-phase

conditions in that particular CV.

When void fraction is equal to zero (or near zero; αG ∼= 0, i.e. practically all liquid) or one (or near

one; αG ∼= 1, i.e. practically all gas); mass and momentum equations of the disappearing phase can

no longer provide flow information, and current equation set becomes ‘singular’ as balance equations

for the disappearing phase no longer respond to perturbations. Consequently, numerical derivatives

of the missing phase residuals can not be calculated. At this point, solution of the jacobian is

difficult if not impossible as the matrix itself also becomes singular; i.e. diagonal contains zero (or

near zero elements when void fraction is very close to the limits), causing difficulties with the matrix

inversion process.

One of the options for preventing governing equation set from becoming singular is to prevent the

disappearing phase from ‘truly’ disappearing due to hydrodynamics; by artificially increasing the

magnitude of momentum forces (i.e. drag force) as the void fraction approaches the physical limits.

In that regard, this option resembles the presence of capillary forces in reservoir engineering that

prevent phases from being completely drained from a reservoir block during the simulation; that is,

unless phases disappear due to mass transfer. Similar approach can also be taken to limit the mass

transfer in order to leave behind a trace amount of the disappearing phase and setting secondary

phase fluid properties equal to those of the primary phase.

In this study, phase appearance-disappearance problem is handled by switching balance equations of

the disappearing phase with ‘dummy’ equations as described by Frepoli et al. (2003). For instance,

upon disappearance of the liquid phase, liquid mass balance equation (which no longer provides any

useful information) is replaced by a dummy void fraction equation:

RmssL = 1− αG = 0 (4.68)

This dummy equation serves two purposes:

1. Fixes the value of αG = 1, which has become a ‘bogus’ parameter in the case of single-phase

flow; hence, preventing it from assuming unphysical values as a result of the iteration process;

which, otherwise would cause conservation problems for the gas phase mass and momentum

balances.

2. Ensures that one (1) is registered in the diagonal of the jacobian as a result of the numerical

79

evaluation of the derivative of dummy residual with respect to void fraction. Therefore,

prevents jacobian from becoming a singular matrix.

Fixing the value of αG = 1 renders the liquid phase momentum equation invalid (0 = 0). Since there

is no actual use for liquid phase velocity for single-phase gas flow conditions, a similar treatment; i.e.

a simple dummy equation with the sole purpose of registering 1 at the jacobian diagonal, could be

applied. However, Frepoli et al. (2003) suggest a better treatment for the liquid phase momentum

equation; by replacing void fraction in the original liquid momentum equation with a constant upper

limit value (i.e. αmax = 0.999999) above which flow is practically single-phase from numerical point

of view. Then, liquid velocity is calculated using this new, ‘dummy’ liquid momentum equation for

a droplet velocity that would have prevailed should liquid phase re-appeared as a dispersed (mist)

flow pattern. Frepoli et al. (2003) concludes that this approach prevents the need for ‘accelerating’

(or decelerating) the liquid droplets from an entirely irrelevant ‘bogus’ liquid velocity to the actual

droplet velocity when liquid phase re-appears. Therefore, the method provides the Newton-Raphson

iteration with a reasonable initial guess. Frepoli et al. (2003) method is adopted in this study

in order to provide a good estimate of the liquid velocity profile for the case of single-phase to

two-phase transition in steady-state analysis.

Because this study is focused on steady-state analysis, primary concern is to address phase

appearance-disappearance due to mass transfer. Phasic mass transfer term, defined by Eqn. 4.47

in Sec. 4.4, is calculated based on thermodynamic equilibrium assumption, ignoring kinetics of

the transfer mechanism and assuming that the time it takes flow stream to traverse the CV is

enough to establish thermodynamic equilibrium, and complete the mass transfer. Consequently,

steady-state analysis requires secondary phase to be present (or absent) in the CV, regardless of the

hydrodynamic conditions, once thermodynamic conditions predict its appearance (or disappearance).

For instance, between two neighboring CVs, if p− T couple falls outside of the two-phase region

(Fig. 1.2) in the downstream cell, calculated mass transfer term should remove all liquid phase and

add it to the gas phase. This is simply a requirement of (and complies with) both the steady-state

assumption and the thermodynamic equilibrium assumption. In this case, dummy void fraction

equation (Eqn. 4.68) serves a third purpose; it drives the value of αG to 1 (along with the mass

transfer term) upon predicting the disappearance of the liquid phase based on the thermodynamics.

Equation switching is applied as follows:

1. At the beginning of an iteration, control volumes are marked (flagged) as single-phase or

two-phase based on cell p− T conditions. This may ensue in neighboring single-phase and

two-phase cells, which poses no problem since discretized governing equations are able to

handle such discontinuities.

2. Iteration (construction of the jacobian) is completed based on phase flag assigned at the

beginning of the iteration; i.e. equation set employed for a CV corresponds to its phase flag.

No equation switching is allowed when numerically evaluating the residual derivative; i.e.

perturbed thermodynamic conditions may indicate a phase change but equation set is not

switched.

80

3. Following the update of primary variables, cells are re-evaluated for the phase flag and the

equation switch now takes place if new thermodynamic conditions require and calculated gas

mass fraction (fmg) for the CV is beyond the practical limits [10−6, 0.999999].

4. Strictly for steady-state purposes, a cell is forced to use the single-phase or two-phase equation

set solely based on its flag which is assigned according the new thermodynamic conditions.

It is important to note that, aforementioned equation switching algorithm is valid only for steady-

state conditions where void fraction could only become 1 (or liquid holdup become 0, for that matter)

due to thermodynamic mass transfer between the phases. For steady-state conditions, second phase

(holdup or void fraction thereof) simply cannot disappear completely due to conservation of mass,

unless one of the phases is completely removed by the mass transfer.

Equation switching process becomes more detailed and complicated in the case of transient analysis

where cells are flagged based on an intermediate hydrodynamic void fraction prediction before the

iteration commences (Frepoli et al., 2003; Spore et al., 2001).

4.11.3 Flow Pattern Transition

Everything else being the same, i.e. p, vG, α, vL, T ; magnitudes of calculated wall (fwk) and

interfacial friction forces (fi) can differ significantly for different flow patterns. Such differences in

the magnitudes could be large enough to cause discontinuities in the momentum residuals should

there be a change in the flow pattern. These jumps can eventually cause problems when evaluating

derivatives of momentum residuals if a flow pattern change is predicted after the perturbation.

Unless pattern transition is smoothened, problems can surface, ultimately delaying or preventing

N-R iteration from converging.

Pattern transition could be handled in a manner similar to single-phase to two-phase transition.

That is to say, no flow pattern transition is permitted during an iteration step but only after

the iteration is completed following the successful construction of jacobian. However, this would

probably slow down the convergence rate. Another approach could be to smoothen friction force

calculations using some sort of an averaging technique. For instance, Spore et al. (2001) uses

weighted averages of the friction forces calculated individually for the transition flow patterns:

(Fw)equation = (1− σ)(Fw)mist + σ(Fw)stratified (4.69)

Where; σ is a weighing factor.

4.11.4 Alignment of Residuals and Primary Unknowns

The proper order, proper arrangement of residuals (rows) and primary unknowns (columns) when

constructing the jacobian matrix is an important aspect of the problem in order not to produce

zero (or near zero) elements in the jacobian diagonal. Along those lines, considering single-phase

incompressible flow, it is apparent that pressure should be aligned with primary phase momentum

81

equation while velocity should be aligned with the mass balance since mass balance would not

recognize pressure perturbations in the case of incompressible flow.

Extending the approach for two-phase flow conditions, keeping primary phase arrangement same as

before, secondary phase velocity is aligned with the secondary phase momentum equation while

void fraction is aligned with the secondary phase mass balance since, in the absence of secondary

phase, its mass balance is replaced with a dummy void fraction equation. Here, it is important to

recognize that which phase being primary or secondary is not the main concern in the sense that

momentum equation of the disappearing phase will be replaced with a momentum equation using a

constant, ‘limit’ void fraction (i.e. α = 0.999999 or α = 10−6); hence, it will still be responsive to

pressure and velocity perturbations.

An important observation to note is; instead of primary phase mass balance, initially a combined

mass balance (arithmetic summation of phasic mass balances) was utilized along with the secondary

phase mass balance. Idea was to simplify the alignment of residuals and primary unknowns if

primary phase were to disappear instead of the secondary phase, since combined mass balance is

always (i.e. single-phase gas, single-phase liquid or two-phase conditions) a valid equation. For

single pipe studies this approach successfully generated the same results as were with individual

phasic mass balances. However, later it has been observed to create mass balance problems at the

tee CV. While no additional analysis was performed to pinpoint the exact reason of this problem,

this approach is simply left out as an option.

4.12 Two-Fluid Junction Model

The FVM based formulation, along with the staggered grid arrangement, allows extension of

governing 1D Euler equations over the T-junction volume regardless of the junction geometry (i.e.

irregular shape). This would allow a seamless algorithm for the network analysis. Therefore, as

shown in Fig. 4.10, tee structure is divided into control volumes.

Figure 4.10: Control Volumes for a T-Junction

82

An important point to underline is that tee cell has 3 interfaces; 2 with neighboring cells of the

main line (inlet and run arms) and one with the side line (branch). Conservation equations for this

CV are written regarding all three interfaces. Flow leaving the tee CV through its arms (or the

split itself for that matter) is to be determined by momentum balances that regard directionality of

the flow.

Change in momentum along each axial flow direction at the tee should be associated with the

presence of forces in these directions. If a coincidental mesh had been used (i.e. velocities at cell

centers along with pressures), then defining effective pressure forces along any of the two primary

directions of flow could have been complicated. This is because, it would then have been necessary

to superpose projection of pressure force in one direction on the other direction which is specifically

a problem for the case of a merging tee structure. Moreover, it would have been necessary to place

a second velocity within the tee CV (perhaps at the cell center as well) in order to account for the

momentum flux in the branch direction. Here, the convenience of the staggered grid is obvious, by

evaluating the velocities at the faces of the tee CV all these difficulties are worked out.

Nevertheless, there are three aspects of the 1D staggered formulation that complicates what, perhaps

otherwise, would have been a straight-forward application:

1. Accurate conservation of momentum in the tee CV.

Among the 3 principle conserved quantities (i.e. mass, momentum and energy), momentum is

the only one with vector attribute. Since it is a vector quantity, like force, momentum must

be conserved in all 3 primary directions of the flow. However, 1D momentum equations within

the Euler set of equations prevails only in one direction. This is not a problem for the case of

a pipe because only concern is the flow in the ‘single’ axial direction of the pipe. On the other

hand, in the case of a tee, tee CV to be more specific, there are two axial directions to account

for. Therefore, momentum conservation has to be accounted for both along the run (x) and

along the branch (x′) axial directions in order to account properly for momentum conservation

at the tee CV. Furthermore, writing the second momentum equation is in fact a necessity in

order to balance the number of equations of the system to the number of unknowns.

2. Proper representation and approximation of 3D pressure and velocity fields (distributions)

within the tee CV, in 1D.

Flow ‘loss’ terms included in 1D momentum equations are expected to make up for the essential

details lost by approximating 3D pressure and velocity distributions in 1D, such as irregular

geometry, flow area changes (diameters of the lines are not necessarily equal), secondary flows

(recirculation zones) and change in flow direction due to branching of flow.

3. Proper definition of momentum fluxes.

Due to staggered nature of momentum CVs, three momentum cells will be enclosing three faces

(edges) of the tee CV (mass CV). While faces of inlet and run momentum cells are parallel

to each other branch momentum cell’s influx face will be making an angle with the faces of

83

inlet and run momentum cells. Besides, incoming momentum flux is to be split among the

influx faces of run and branch momentum cells. Therefore, proper definition of (1) momentum

flux across all three of the ‘momentum faces’ and (2) branch momentum flux contribution for

conservation of momentum along the run direction are essential. Contribution from this third

interface could be observed as a sink/source term as well.

For practical purposes, and with the assumption of a reasonably small tee CV, density and other fluid

properties can perhaps be safely approximated as constants through out the tee cell. Assumption of

a small enough CV could provide the basis for the assumption of a same void fraction for all outgoing

streams. But this is also a consequence of the finite volume concept; that is, phasic mass fluxes

leaving the tee CV typically have the same fluid properties (i.e. density) and same void fraction

since flow is originating from the same ‘pool’. However, if void fractions of outgoing streams are

assumed to be same then, necessary adjustments for outgoing flow rates could only be established

through phasic velocities as previously pointed out with the DSM discussion in Sec. 3.2.3.

From FVM modeling perspective, a tee is merely a 3 pipe system. Both branch and run can be

visualized as separate pipes with outlet pressure specifications. Then, solution for an outgoing arm

solely depends on the flow going through it and the outlet pressure specification. This is because,

in the staggered grid arrangement inlet pressure and void fraction specifications have no significant

influence on the solution; particularly for isothermal, steady-state analysis. However, what really

matters is the phasic flow rates going through arms, where flow rates are supposed to be determined

by void fraction and phasic velocities of the tee CV.

Therefore, once flow split (and phase separation) is determined for the tee CV, that is phasic flow

rates for outgoing arms are known, pressure profiles of outgoing arms is solely based on specified

outlet pressures. Nevertheless, momentum equations at outgoing tee interfaces should link the

pressures obtained for the first cells of outgoing arms (next to tee CV) to the tee CV pressure which

becomes the outlet pressure specification for the inlet arm when considered a separate, third pipe.

Hence, an accurate solution can be obtained in the main line (inlet arm) as well.

If this pressure link is not correct; that is, it might be possible to obtain an incorrect tee cell pressure

with a seemingly correct flow split unless proper momentum equations are utilized, connecting all 3

arms of the tee, solution in the inlet arm would be incorrect as well.

Governing mass and momentum equations for the tee cell and associated momentum CVs are

obtained as follows for the tee structure given in Fig. 4.10:

Mass equation for the T-cell:

Vp3(ρn+1p3 − ρnp3)

∆t= [ρp2vv2Av2 − ρp3vv3Av3 − ρp3vv6Av6]n+1 (4.70)

84

Momentum equation for cell v2:

Vv2

[(ρv2vv2vv2)n+1 − (ρv2vv2vv2)n

]∆t

=

ρp1vv1vv1Av1 − ρp2vv2vv2Av2

+ (p2Ap2 − p3Ap3)

−ρv2Av2∆xv2g sin θ

− (τws)v2

n+1

(4.71)

Momentum equation for cell v3:

Vv3[(ρv3vv3vv3)n+1 − (ρv3vv3vv3)n]

∆t=

(ρp2vv2Av2 − ρp3vv6Av6) vv2 − ρp3vv3vv3Av3

+ (p3Ap3 − p4Ap4)

−ρv3Av3∆xv3g sin θ

− (τws)v3

+irr. loss

n+1

(4.72)

Where; the term [−ρp3vv6Av6], representing mass loss to the branch, is used for the calculation of

actual momentum flux going straight into the run.

Momentum equation for cell v6:

Vv6[(ρv6vv6vv6)n+1 − (ρv6vv6vv6)n]

∆t=

(ρp3vv6A

∗v6) vv2 cos(180− β)− ρp3vv6vv6Av6

+(p3A

∗p3 − p6Ap6

)−ρp3Av6∆xv6g sin θ

− (τws)v6

+irr. loss

n+1

(4.73)

Where; A∗p3 is the cross-sectional area of the mass flux within the T-cell, along the branch direction.

There could be different approaches on the derivation of momentum equations; for instance if

momentum influx to CV v3 is calculated based on the mass flux at edge v3, then there would be no

need to account for branch contribution. Nevertheless, correct form of the momentum equations

should work effectively without the loss terms. One way of testing momentum equations, without

the loss terms, is to see if they predict a pressure spike (pressure rise) for the first cell of run and if

pressure drops would be equal for both outgoing arms when there is no split angle; i.e. branch is

parallel.

Using momentum equations at the tee CV, instead of Bernoulli equations, allows accounting for

(to a degree) branching angle effects without requiring any loss terms. Therefore, inlet-to-branch

pressure drop predicted by momentum equations (without loss terms) should represent the reversible

part of pressure drop.

Calculation of momentum loss terms is, in fact, another challenge since, to the best of author’s

knowledge, and unlike Bernoulli equation energy loss terms, there is no general, explicit description

for these terms. Besides, calculation of these loss terms should be dependent on the derivation of

85

momentum equations and calculation of momentum fluxes at the faces of tee momentum cells. Based

on mechanistic junction models and Bernoulli equation loss terms (Sec. 3.2.1) typically momentum

loss terms are expected to be in the form:

loss12 = k12 ρtee v21

loss13 = k13 ρtee v21

(4.74)

The k-factors utilized with this definition of momentum loss terms are perhaps slightly different than

the k-factors used with the inlet-to-run pressure drop equation (Eqn. 3.8) in Sec. 3.2.1. Moreover,

because inlet-to-branch k-factor correlations are not available in the literature, they should be

obtained before this approach is pursued any further.

Steinke (1996) discusses that momentum correction k-factors can be written in terms of energy

loss coefficient K-factors by arranging non-conservative form of the momentum equations (used in

TRAC; Spore et al., 2001) in the form of Bernoulli equations. For single-phase problems, appropriate

momentum k-factors can be obtained this way. However, extending the approach for two-phase

conditions may not be straight-forward:

• Two-phase K-factors (energy loss coefficients) are not widely available in the literature. It is

typically assumed that single-phase K-factors should suffice in capturing two-phase pressure

drops; i.e. via separated flow model or two-phase loss multipliers used with homogeneous

mixture approach (cf. Sec. 3.2.1).

• Transformation of Euler type momentum equations into Bernoulli type mechanical energy

equations requires certain assumptions; (1) Steady flow, (2) incompressible flow, (3) flow along

a streamline or (4) irrotational flow (cf. Munson et al., 2009 pages 279 - 284).

Theory of an alternative FVM formulation, one that accounts for different void fractions for outgoing

streams, is presented in Appendix B. The theory is an extension of Steinke (1996) discussion based

on the need for using two separate pressures within the tee CV in order to better represent 3D

pressure distribution in the tee CV for combining tees. Steinke’s approach is adopted for branching

tees with the ‘buffer zone’ concept that introduces certain simplifications in the governing model

equations.

Another approach is to use Bernoulli equations (mechanical energy balance) at the tee CV, based

on phasic streamline approach, instead of momentum equations. Then, phasic Bernoulli equation

for inlet-to-run stream ([p1 − p2]):

1

2ρkv

2k1 + ρkgzk1 + p1 =

1

2ρkv

2k2 + ρkgzk2 + p2 +

1

2Kk12ρkv

2k1 (4.75)

And, phasic Bernoulli equation for inlet-to-branch stream ([p1 − p3]):

1

2ρkv

2k1 + ρkgzk1 + p1 =

1

2ρkv

2k3 + ρkgzk3 + p3 +

1

2Kk13ρkv

2k1 (4.76)

86

Where;

vk1 =qk1

αk1A1(4.77)

vk2 =qk2

αk2A2(4.78)

vk3 =qk3

αk3A3(4.79)

From aforementioned equations, it seems as if Bernoulli equations account for exit (outgoing) void

fractions at a tee, not directly but through velocities. In practice, velocities and void fractions are

distinct primary unknowns of the FVM numerical solution. After an iteration, updated values of

the primaries are directly substituted in the governing equations. However, it is recognized that

these Bernoulli equations are not really dependent on void fractions if values of velocities are to be

substituted directly (or vice a versa). Therefore, these equations (at least in their current form)

do not carry information on void fractions of outgoing stream. Consequently it is assumed that

outgoing void fractions are same as the tee cell itself.

Phasic Bernoulli equations could replace phasic momentum equations derived previously (Eqn. 4.71

to Eqn. 4.73) and then could be solved along with mass balances for the tee CV; in order to account

for exit velocities of the tee cell. This is, indeed, the basis of Double Stream Model of Hart et al.

(1991) introduced in Sec. 3.2.3. However, observations regarding this transformation have shown that

the assumption KG = KL, which led to the derivation of final, closed form DSM equation (Eqn. 3.22),

renders simultaneous solution of phasic Bernoulli equations and mass balances inconsistent. Because

DSM is essentially a phase separation sub-model, its implementation for the tee CV; connecting

DSM pressures and velocities to the FV momentum grid and prediction of pressure drop across

the faces of tee has to be worked out. Consequently, DSM is incorporated in the FVM solution

by solving only gas phase Bernoulli equations (for both run and branch directions). Gas phase

Bernoulli equations link run and branch inlet pressures to the tee cell pressure which becomes the

outlet pressure specification for inlet arm. Liquid phase momentum equations are replaced with

dummy equations that set outgoing liquid velocities to the velocities determined by closed form

DSM equation (Eqn. 3.22) which is solved separately after completing each iteration in order to

determine phase separation. Solving the DSM equation separately, in fact, slows the convergence

rate (i.e. increases number of iterations); yet, has been observed to be stable. Ultimately, two-phase

flow separation problem is addressed by incorporating an already verified, mechanistic junction

model.

It is important to repeat once more that due to limitations of one-dimensional analysis, not all the

factors contributing to the pressure change at a tee can be accounted for explicitly and rigorously in

the governing equations. Instead, for instance, mechanical losses due to change in flow direction

or effect of secondary flows (recirculation zones, Fig. 3.4), are typically combined into a single

parameter (loss coefficient K for mechanical energy equation and correction factor k for

momentum equation) and represented with an additional irreversible loss term in the pertinent

governing equation. In that regard, use of these parameters is similar to the representation of viscous

forces of flow (i.e. wall drag) with a friction factor and associated friction force in one-dimensional

87

analysis.

4.13 Fluid Composition After Split

For the practical purpose of steady-state analysis, stream composition is assumed to be constant

along each pipe element and along each arm of a tee during an iteration. In other words, all

the CVs enclosed by a straight section of the network and the tee cells themselves have particular

compositions assigned at the beginning of an iteration and these compositions remain constant

during all calculations until the iteration is completed. Compositions are updated before new

iteration commences; according to phase split and/or flow merge at tee cells.

For a particular pipe, composition is assigned to all of it’s CVs based on it’s inlet stream, may it

be a boundary condition prescribed in the input file (if pipe is connected to a boundary node) or

outlet composition of an upstream network element (in this case another pipe or a tee) connected

to the inlet of this pipe.

For a T-junction; first, overall fluid composition arriving at the tee CV is calculated according to

incoming streams. Once the overall composition at the tee cell is determined, compositions for

outgoing arms are obtained based on the ‘separator approach’; i.e. using outgoing phasic velocities

(Martinez and Adewumi, 1997).

Following the completion of each iteration and after updating primary unknowns (and correcting

boundary conditions as necessary), pipe, tee arm and tee CV (tee cell) compositions are updated.

Then, secondary parameters; i.e. phase mass fraction (fmg), density, viscosity etc. (properties based

on composition) are computed according to newly assigned composition. The ‘update’ procedure

for the compositions is described next.

First, total composition arriving at the tee cell is computed by adding up number of moles of each

component flowing into the cell with each incoming stream. Total number of moles flowing-in

through each arm’s interface can be easily calculated by dividing total mass flow (combination of

both phases) at that face by the mixture molecular weight of the upstream cell (due to donor cell

scheme).

Total mass entering the tee cell is a summation of mass entering through each incoming arm:

min =

N (=3)∑j=1

mj (4.80)

Where, N is total number of incoming arms and for the particular case of branching tee, N = 1.

Total mass entering through arm j is given by:

mj = [mG + mL]j (4.81)

88

And, mass carried in by each phase is:

[mG = ρG vG αA]j

[mL = ρL vL (1− α)A]j(4.82)

Mass entering through each arm could also be expressed as a function of number of moles:

mj = nj

[∑ciMWi

]j[

mG = nG∑

xiMWi

]j[

mL = nL∑

yiMWi

]j

(4.83)

Mixture molecular weight for arm j:[MWmix =

∑ciMWi

]j

(4.84)

Gas phase molecular weight for arm j:[MWG =

∑xiMWi

]j

(4.85)

Liquid phase molecular weight for arm j:[MWL =

∑yiMWi

]j

(4.86)

Therefore, number of moles entering the tee cell through arm j can be written as:

nj =mj∑ciMWi[

nG =mG∑xiMWi

]j[

nL =mL∑yiMWi

]j

(4.87)

Or, substituting definition of mj from Eqns. 4.81 and 4.82:[n =

(ρg vg αg + ρg vg (1− αg))A∑ciMWi

]j[

nG =ρg vg αg A∑xiMWi

]j[

nL =ρl vl (1− αg)A∑

yiMWi

]j

(4.88)

89

Number of moles entering through arm j could also be given as:

nj = [nG + nL]j (4.89)

Total number of moles entering the tee cell:

nin =N∑j=1

nj (4.90)

Where, N is the number of arms with incoming flow to the tee, for the particular case of branching

tee, N = 1.

Number of moles of component i coming into tee cell through arm j:

[ni = n ci]j (4.91)

Total moles of component i entering tee cell:

(ni)in =

N∑j=1

[ni]j =

N∑j=1

[nci]j (4.92)

Now that total moles of all the components arriving at the tee cell are known, overall composition

at the tee cell is known and it is ‘flashed2’ at tee cell p− T conditions to determine outgoing phase

composition and fluid properties.

Overall compositions of outgoing arms are then determined according to the amount of each phase

that goes through an arm, hence depends on the phasic velocities and void fractions (please see

‘double channel’ development in Appendix B for further discussion on the theory of separate void

fractions for each outgoing arm) associated with each outgoing face of the tee cell since density and

composition of each phase is assumed to be constant through out the tee CV.

Mass of gas phase leaving the tee cell through arm j:

[mG = ρG vG αGA]j (4.93)

Mass of liquid phase leaving tee CV through arm j:

[mL = ρL vL (1− αG)A]j (4.94)

2Equation of state based determination of second phase formation and associated computation of chemicalcomponent distribution among gas and liquid phases is called ‘flash’ calculations.

90

Total mass leaving the tee cell through arm j:

[m = (ρG vG αG + ρL vL (1− αG))A]j (4.95)

Number of moles leaving the tee cell through arm j within gas phase:

[nG]j =[ρGvGαGA]j

[∑xiMWi]tee

(4.96)

Number of moles leaving tee cell through arm j within liquid phase:

[nL]j =[ρL vL (1− αG)A]j

[∑yiMWi]tee

(4.97)

Number of moles of component i leaving the tee cell through arm j within gas phase:

[(nG)i]j = [nG]j [xi]tee =[ρG vG αA]j

[∑xiMWi]tee

[xi]tee (4.98)

Number of moles of component i leaving the tee cell through arm j within liquid phase:

[(nL)i]j = [nL]j [yi]tee =[ρL vL (1− α)A]j

[∑yiMWi]tee

[yi]tee (4.99)

Total number of moles of component i leaving the tee cell through arm j:

[ni = (nG)i + (nL)i]j (4.100)

Total number of moles leaving the tee cell through arm j:

[n = nG + nL]j (4.101)

Now that number of moles of all the components going through arm j is known, composition for

arm j can be easily obtained as: [ci =

nin

]j

(4.102)

91

Chapter 5

Results and Discussion

5.1 Single-Phase Compressible Flow

Pressure profiles generated by the numerical model for an isothermal, single-phase compressible

study are compared with the analytical solution provided by Tian-Adewumi design equation:

dA2MW

f m2Z RT(p2

2 − p21)− d

fln

(p2

2

p21

)+ L = 0 (5.1)

Tian-Adewumi design equation differs from the steady-state gas flow equation (Eqn. 2.19) introduced

in Sec. 2.1.1 in the sense that:

1. Tian-Adewumi equation (Eqn. 5.1) expresses the flow in terms of mass flow rate while

steady-state gas flow equation (Eqn. 2.19) expresses the flow in terms of volumetric flow rate.

2. Kinetic energy terms ignored during the derivation of steady-state gas flow equation (Eqn. 2.19)

are included in the derivation of Tian-Adewumi equation (Eqn. 5.1).

Tian-Adewumi equation is derived from the very same Euler mass and momentum equations without

any simplifications (Tian and Adewumi, 1992). Thus, model and the analytical results should be in

very good agreement. Specifics of the study; pipe and fluid properties and boundary conditions

are given in Tab. 5.1. Pipe properties are taken from Tables 6.4 and 6.5 of Ayala (2001). However,

injected fluid is chosen as air with a constant gas compressibility (z-factor) of Z = 0.9 both for the

Tian-Adewumi equation and for the numerical model in order to have a fair comparison. Otherwise,

the design equation requires externally computed z-factors to be provided over each increment.

Comparison of pressure profiles given in Fig. 5.1 is in good agreement for the over 300 km long

pipeline. Numerical solution seems to deviate from the analytical solution towards the inlet of the

pipe where inlet pressure prediction by Tian-Adewumi equation is ∼= 15103 kPa and numerical

model predicts an inlet pressure of ∼= 15008 kPa with 20 CVs and ∼= 15054 kPa with 40 CVs. Error

in pressure prediction with 20 CVs is ∼ 0.628% at the outlet (where difference is the most significant)

and decreases to ∼ 0.321% when 40 CVs is employed by the model. The fact that error is at most

halved by doubling the number of cells suggests practical order of accuracy of the numerical model

is indeed lower than the theoretical value; as expected. Figure 5.2 shows the percentage error change

across the pipe length suggesting that numerical error increases with pipe length. An important

92

Table 5.1: Single-phase compressible flow study data

Pipe length 321.869 km

Pipe diameter 0.70485 m

Pipe roughness 0.24384× 10−3 m

Gas flow rate 170.550 kg/s

Gas MW (Air) 29 g/mole

Gas compressibility 0.9

Flow temperature 388.705 K

Ambient temperature 299.816 K

Heat transfer coefficient 5.6783 W/m2/K

Outlet pressure 9643.697 kPa

observation is that based on the shape of the curves (Figure 5.2) error does not grow beyond a

certain value however increasing towards the pipe inlet.

15,05415,103

15,008

9,000

10,000

11,000

12,000

13,000

14,000

15,000

16,000

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Pre

ssu

re [

kPa]

x [km]

Tian-Adewumi

20 Cell

40 Cell

Figure 5.1: Single-phase compressible flow study pressure profiles

93

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Pre

ssu

re E

rro

r [%

]

x [km]

20 Cell

40 Cell

Figure 5.2: Single-phase compressible flow study numerical error profile

5.2 Two-Phase Flow

For two-phase analysis, the gas condensate fluid described by Ayala and Adewumi (2003) is

implemented in this study. The hydrocarbon mixture is composed of three ‘lumped’ components

where molar composition, component properties, binary interaction coefficients for EOS calculations

and parachor parameters for interfacial tension calculations are given in Tabs. D.1, D.2, D.3 and

D.3 respectively, in Appendix D. The mixture is injected at the inlet of the pipeline described in

Tab. 5.2.

Table 5.2: Two-Phase flow study data

Pipe: SI Units US Units

Length 16093.44 m 52,800 ft (10 miles)

Diameter 0.3048 m 12 in

Roughness 3.048e-4 m 0.001 ft

Heat Trans. Coef. 5.6783 J m2 s−1K−1 1 BTU ft−2 hr−1 oF−1

Ambient temperature 26.67 oC (299.8167 K) 80 oF (539.67 R)

Gas injection rate 65.5483 m3s−1 200 MMscfD

Inlet pressure 13.65MPa 1980 psia

Inlet temperature 60.183 oC (333.3 K) 140 oF (600 R)

94

Two-phase flow performance of the FVM based two-fluid model (NR - with Newton-Raphson

solution technique) has been assessed for mist flow in the entire pipe and compared against a

Runge-Kutta based marching two-fluid model (RK). The pipe is discretized anywhere in between 2

CVs (dx = 26,400 ft) to 52,800 CVs (dx = 1 ft) in order to study the robustness of the RK and NR

methods for two-phase flow predictions. The operational conditions in this pipeline are such that

the gas is being injected to the pipeline very close to dew point conditions. Therefore, retrograde

hydrocarbon condensation is triggered upon pressure depletion, and multiphase conditions are found

throughout the entire pipe. The relatively small quantities of liquids formed in this case and the

significant amounts of turbulence in the gas phase impose mist flow (dispersed liquid) conditions

through out the pipe.

Figures 5.3 to 5.7 show the agreement between NR and RK methods when RK method uses very

small step size (i.e. dx = 1 ft) however NR model has significantly large spatial discretization (dx

= 528 ft); and how the RK solution deviates from the actual profile if computational grid is any

coarser (i.e. dx > 2 ft). Further details on the two-phase flow performance assessment of the FVM

model and comparison against RK based marching algorithm can be found in the text by Ayala

and Alp (2008).

1500

1550

1600

1650

1700

1750

1800

1850

1900

1950

2000

0 10000 20000 30000 40000 50000 60000

Pre

ssu

re (

psi

)

Distance (ft)

NR (dx = 528 ft)

RK (dx = auto)

RK (dx = 1 ft)

RK (dx = 2.5 ft)

RK (dx = 5 ft)

RK (dx = 10 ft)

Figure 5.3: Two-phase RK and NR pressure profiles

95

575

580

585

590

595

600

0 10000 20000 30000 40000 50000 60000

Tem

pe

ratu

re (

R)

Distance (ft)

NR (dx = 528 ft)

RK (dx = auto)

RK (dx = 1 ft)

RK (dx = 2.5 ft)

RK (dx = 5 ft)

RK (dx = 10 ft)

Figure 5.4: Two-phase RK and NR temperature profiles

18.0

18.5

19.0

19.5

20.0

20.5

21.0

21.5

22.0

22.5

0 10000 20000 30000 40000 50000 60000

Ve

loci

ty (

ft/s

)

Distance (ft)

NR (dx = 528 ft)

RK (dx = auto)

RK (dx = 1 ft)

RK (dx = 2.5 ft)

RK (dx = 5 ft)

RK (dx = 10 ft)

Figure 5.5: Two-phase RK and NR gas velocity profiles

96

18.0

18.5

19.0

19.5

20.0

20.5

21.0

21.5

0 10000 20000 30000 40000 50000 60000

Ve

loci

ty (

ft/s

)

Distance (ft)

NR (dx = 528 ft)

RK (dx = auto)

RK (dx = 1 ft)

RK (dx = 2.5 ft)

RK (dx = 5 ft)

RK (dx = 10 ft)

Figure 5.6: Two-phase RK and NR liquid velocity profiles

0.000

0.005

0.010

0.015

0.020

0.025

0 10000 20000 30000 40000 50000 60000

Ho

ldu

p (

Frac

tio

n)

Distance (ft)

NR (dx = 528 ft)

RK (dx = auto)

RK (dx = 1 ft)

RK (dx = 2.5 ft)

RK (dx = 5 ft)

RK (dx = 10 ft)

Figure 5.7: Two-phase RK and NR holdup profiles

97

5.3 Two-Phase Flow and Secondary Phase Appearance

Two-phase flow capability of the model is demonstrated again for mist flow pattern along with the

handling of phase appearance-disappearance. For this purpose, the same hydrocarbon mixture with

lumped components (Appendix D) is injected at the inlet of the pipeline described in Tab. 5.1.

Figures 5.8 and 5.9 show pressure and temperature profiles obtained under different thermal

conditions (HTC: heat transfer coefficient cU , [W K−1m−2]) and for isothermal conditions. The

reason of increased pressure drop along the pipeline is more clear when void fraction profiles in

Fig. 5.10 is observed; as the amount of secondary phase (liquid hydrocarbon) increases, it induces

additional pressure drop along the line. Pressure and temperature profiles obtained for HTC

= 5.6783 [W K−1m−2] is in good agreement with Figure 6.8 of Ayala (2001).

Void fraction plots in Fig. 5.10 show how two-phase conditions prevail at different distances from

the inlet, based on p− T conditions along the pipeline. As expected, with increased heat loss to the

environment (increasing HTC and greater temperature drop) liquid hydrocarbons released earlier in

the pipe and model captures this phenomenon successfully.

9,000

10,000

11,000

12,000

13,000

14,000

15,000

16,000

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

pre

ssu

re [

kPa]

x [km]

HTC = 5.6783

HTC = 1.0000

HTC = 0.0000

Isothermal

Figure 5.8: Phase appearance p profiles - 20 CVs

Phasic velocity profiles given in Figs. 5.11 and 5.12 indicate that gas velocity and dummy liquid

velocity values are almost identical in the single-phase region. This is to be expected since in the

absence of liquid phase, liquid phase properties such as density and viscosity are simply set equal

to the gas phase properties by the EOS module (otherwise EOS module does not provide any

information on the absent phase).

98

280

300

320

340

360

380

400

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Tem

pe

ratu

re [

K]

x [km]

HTC = 5.6783

HTC = 1.0000

HTC = 0.0000

Isothermal

Figure 5.9: Phase appearance T profiles - 20 CVs

0.9550

0.9600

0.9650

0.9700

0.9750

0.9800

0.9850

0.9900

0.9950

1.0000

1.0050

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Vo

id F

ract

ion

[-]

x [km]

HTC = 5.6783

HTC = 1.0000

HTC = 0.0000

Isothermal

Figure 5.10: Phase appearance αG profiles - 20 CVs

99

The staggered nature of the momentum CVs along with the Frepoli et al. (2003) type dummy

momentum equation causes a jump in liquid velocity (not plotted in the figures) right before the

momentum CV where liquid phase appears the first time (∼ 50 km in Fig. 5.11). This jump is

not physical and only a consequence of the current formulation; i.e. densities substituted in the

staggered momentum equations is a weighted average of neighboring mass CVs. In the case of

dummy liquid momentum equation written for the inlet face of a mass CV where liquid phase

appears the first time; the upstream liquid density is simply set equal to the gas density (by choice)

and downstream density is the actual liquid determined by the EOS. This arrangement, along with

the dummy momentum equation is the source of jump and indicates that current implementation of

the dummy momentum equation, while not essential for steady-state analysis, needs improvement

since its purpose is to accelerate the iteration process by preventing such jumps in the profiles that

could ensue in convergence or stability issues.

Another important thing to notice is the effect of steep temperature drop during the first 100 km of

the pipe; the density increase accompanying the temperature drop (Fig. 5.13) is slowing down the

phases, hence the initial descend in the velocity profiles of Fig. 5.11 is not observed in Fig. 5.12 where

temperature drop is not that sharp due to smaller heat transfer coefficient (cU = 1 [W K−1m−2],

Fig. 5.14).

2.5

2.7

2.9

3.1

3.3

3.5

3.7

0.955

0.960

0.965

0.970

0.975

0.980

0.985

0.990

0.995

1.000

1.005

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Ve

loci

ty [

m/s

]

Vo

id F

ract

ion

[-]

x [km]

Void fraction [-]

Liquid velocity [m/s]

Gas velocity [m/s]

Figure 5.11: Phase appearance αG vs vk profiles - 20 CVs, HTC ∼= 5.6783W K−1m−2

Figures 5.15 and 5.16 compare p−T and void fraction vs cumulative mass transfer profiles obtained

with 20 and 40 CVs. No significant difference is observed in the profiles besides the slight change in

the location of secondary phase appearance; within [40− 50] km with 20 CVs and [45− 55] km with

40 CVs, as a result of finer griding. It is also observed that jump in the liquid velocity profile is still

100

3.1

3.3

3.5

3.7

3.9

4.1

4.3

0.980

0.982

0.984

0.986

0.988

0.990

0.992

0.994

0.996

0.998

1.000

1.002

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Ve

loci

ty [

m/s

]

Vo

id F

ract

ion

[-]

x [km]

Void fraction [-]

Liquid velocity [m/s]

Gas velocity [m/s]

Figure 5.12: Phase appearance αG vs vk profiles - 20 CVs, HTC ∼= 1.0W K−1m−2

2.5

2.7

2.9

3.1

3.3

3.5

3.7

0

50

100

150

200

250

300

350

400

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Ve

loci

ty [

m/s

]

De

nsi

ty [

kg/m

3]

x [km]

Gas density [kg/m3]

Liquid density [kg/m3]

Liquid velocity [m/s]

Gas velocity [m/s]

Figure 5.13: Phase appearance density and velocity profiles - 20 CVs, HTC ∼= 5.6783W K−1m−2

101

3.1

3.3

3.5

3.7

3.9

4.1

4.3

90

140

190

240

290

340

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Ve

loci

ty [

m/s

]

De

nsi

ty [

kg/m

3]

x [km]

Gas density [kg/m3]

Liquid density [kg/m3]

Gas velocity [m/s]

Liquid velocity [m/s]

Figure 5.14: Phase appearance density and velocity profiles - 20 CVs, HTC ∼= 1.0W K−1m−2

of the same magnitude when velocity profiles are compared (Fig. 5.17).

With steep temperature drops (i.e. due to heat loss to the environment) stability problems have

been observed and required providing better initial conditions to the system. In fact, mass transfer,

associated momentum transfer and heat loss terms; all evaluated at current iteration level (k+1),

increase the non-linearity of the problem and cause difficulties in convergence. Tricking the way

around by lagging the calculations of transfer and heat loss terms one iteration behind could help

attain convergence; yet would definitely increase number of iterations.

Another important observation is; using small CVs (fine grid) also ensues in (1) smaller transfer

terms because values of gas mass fractions (fmg) of neighboring cells are closer to each other (cf.

Sec. 4.11.2) and (2) smaller heat loss terms because pipe surface area is lesser. While fluxes entering

and leaving the CVs remain the same, as these terms get smaller in magnitude, it is easier for the

system to digest associated non-linearity, albeit at the expense of increased computational load due

to increased number of CVs. Comparison of mass transfer terms per CV; given in Fig. 5.18, clearly

indicate the difference in the magnitude of calculated mass transfer terms however cumulative mass

transfer is same (Fig. 5.16).

Also, looking at Fig. 5.11 and Fig. 5.16 the effect of hydrodynamic holdup is clearly observed; as

gas-to-liquid mass transfer diminishes towards the last 100 km of the pipeline, void fraction begins

to ascend (or holdup begins to decrease for that matter) indicating that liquid is slowly being left

behind not being able to catch up with the gas flow.

102

250

270

290

310

330

350

370

390

410

9,000

10,000

11,000

12,000

13,000

14,000

15,000

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Tem

pe

ratu

rer

[K]

Pre

ssu

re [

kPa]

x [km]

P 20 cell [kPa]

P 40 cell [kPa]

T 20 cell [K]

T 40 cell [K]

single phase

Figure 5.15: Hydrocarbon mixture p− T profiles - 20 and 40 CVs, HTC ∼= 5.6783 [W K−1m−2]

0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

0.955

0.960

0.965

0.970

0.975

0.980

0.985

0.990

0.995

1.000

1.005

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Cu

mu

lati

ve M

ass

Tras

nfe

r [k

g/s

]

Vo

id F

ract

ion

[-]

x [km]

Void frac. 20 cell [-]

Void frac. 40 cell [-]

Mass Trans. 20 cell [kg/s]

Mass Trans. 40 cell [kg/s]

Figure 5.16: Phase appearance αG vs cum. mass transfer profiles - 20 and 40 CVs, HTC ∼=5.6783W K−1m−2]

103

2.5

2.7

2.9

3.1

3.3

3.5

3.7

3.9

0.00 50.00 100.00 150.00 200.00 250.00 300.00 350.00

Ve

loci

ty [

m/s

]

x [km]

velG 20 cell [m/s]

velL 20 cell [m/s]

velG 40 cell [m/s]

velL 40 cell [m/s]

single phase

Figure 5.17: Phase appearance velocity profiles - 20 and 40 CVs, HTC ∼= 5.6783W K−1m−2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

0.955

0.960

0.965

0.970

0.975

0.980

0.985

0.990

0.995

1.000

1.005

0.0 50.0 100.0 150.0 200.0 250.0 300.0 350.0

Mas

s Tr

ansf

er

pe

r C

V [

kg/s

]

Vo

id F

ract

ion

[-]

x [km]

Void frac. 20 cell [-]

Void frac. 40 cell [-]

Mass Trans. 20 cell [kg/s]

Mass Trans. 40 cell [kg/s]

Figure 5.18: Phase appearance αG and mass transfer per CV profiles - 20 and 40 CVs, HTC∼= 5.6783W K−1m−2

104

5.4 Flow Split and Phase Separation

We begin flow split and phase separation analysis using air-water mixture. Data of the tee structure

(from Hart et al., 1991 and Singh, 2009) is given in Tab. 5.3. Flow is assumed to occur isothermally

and no mass exchange is considered; i.e. no water vapor or air dissolution in the water (mass

transfer = 0). Stratified flow pattern is assumed to prevail in the junction in accordance with DSM.

Besides, typically phase separation is equal (i.e. λG = λL) for dispersed flow patterns such as mist

flow since dispersed phase is entrained in the continuous phase. While gas flow rate is kept constant,

simulation is repeated with different liquid flow rates in order to observe the effect of increased

liquid presence on uneven phase separation. For a given water flow rate branch and run outlet

pressures are separately increased (from the initial 101.325 kPa) to observe the effects of outlet

pressure on the solution.

Table 5.3: Air-water flow split data

Arm length L 15 m

Number of CVs nCV 5

Arm diameter D 0.051 m

Pipe roughness ε 0.005 m

Gas flow rate mG 1.973× 10−2 kg/s

Air MW MWair 29 g/mole

Gas compressibility Z 0.9

Water density ρw 998 kg/m3

Flow temperature T 288.71 K

Run outlet pressure prun 101.325 kPa

Branch outlet pressure pbranch 101.325 kPa

Typical pressure and void fraction profiles for the tee structure are given for water flow rate of

0.0057 kg/s in Figs. 5.19 and 5.20 respectively. Model captures pressure rise in the run entrance as

expected (Fig. 5.19). And Fig. 5.20 shows how significantly void fraction can change (in the run)

following an uneven phase separation at the junction when most of the gas phase is diverted into

the branch (for λL = 0.52 and λG = 0.90).

Looking at Fig. 5.21, it is observed that there is always a pressure increase at the run entrance and

the pressure difference between run and branch entrances is always positive; i.e. (prun > pbranch)tee.

More liquid tends to turn into branch direction as this pressure differential increases. Also, with

increasing liquid flow rate (as the liquid momentum increases) it is harder to divert liquid phase

into the branch. This is to be expected since same amount of pressure drop now has to overcome

greater liquid inertia in order to draw liquid phase into the branch.

105

101.2

101.4

101.6

101.8

102.0

102.2

102.4

102.6

102.8

103.0

0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0

Pre

ssu

re [

kPa]

x (m)

inlet

run

branch

Figure 5.19: Air-water flow pressure profile at T-junction for mL = 0.0057 kg/s

0.91

0.92

0.93

0.94

0.95

0.96

0.97

0.98

0.99

1.00

0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0

Vo

id f

ract

ion

[-]

x (m)

inlet

run

branch

Figure 5.20: Air-water flow αG profile at T-junction for mL = 0.0057 kg/s and prun = 102.125 kPa

106

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.000 0.010 0.020 0.030 0.040 0.050 0.060 0.070

λ L

(prun- pbranch)tee [kPa]

mL=1.427e-4 kg/s

mL=1.814e-3 kg/s

mL=3.140e-3 kg/s

mL=5.700e-3 kg/s

Figure 5.21: Air-water flow (∆p)tee vs λL plots for different water flow rates

Plotting a similar graph for outlet pressures tells more on the uneven phase separation. Looking at

Fig. 5.22, the effect of liquid flow rate is observed in a different manner; with increasing water flow

rate less liquid tends to flow into branch when both outlet pressures are equal (prun = pbranch or

prun − pbranch = 0). When run outlet pressure is increased (to the right of the y axis in Fig. 5.22)

more of the both phases are diverted into the branch as expected since fluids prefer flowing towards

a lower potential; in this case, into the branch. Similarly, when branch outlet pressure is increased

(to the left of the y axis in Fig. 5.22) lesser amounts of both phases go into the branch. It is easier

to observe the flip-flop effect when water flow rate curves of mL = 1.427e−4 and mL = 1.814e−3

[kg/s] are considered; slight change in the outlet pressure differential causes water turn into branch

entirely. This signifies how sensitive uneven phase separation can be in response to downstream

(outlet) pressures.

When λG vs λL is plotted (Fig. 5.23) curves with characteristics to Fig. 5.22 is observed. Suggesting

that primary (continuous) phase split is not significantly effected. Plotting (∆p)out vs λG (Fig. 5.24)

verifies the fact that for small holdup values primary phase split is simply governed by outlet

pressure differential as is the case with single-phase flow split.

107

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

-1.000 -0.800 -0.600 -0.400 -0.200 0.000 0.200 0.400 0.600 0.800 1.000

λ L

(prun - pbranch)out [kPa]

mL=1.427e-4 kg/s

mL=1.814e-3 kg/s

mL=3.140e-3 kg/s

mL=5.700e-3 kg/s

prun > pbranch

prun < pbranch

Figure 5.22: Air-water flow (∆p)out vs λL plots for different water flow rates

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

λ L

λG

mL=1.427e-4 kg/s

mL=1.814e-3 kg/s

mL=3.140e-3 kg/s

mL=5.700e-3 kg/s

Figure 5.23: Air-water flow λG vs λL plots for different water flow rates

108

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

-1.000 -0.800 -0.600 -0.400 -0.200 0.000 0.200 0.400 0.600 0.800 1.000

λ G

(prun - pbranch)out [kPa]

mL=1.427e-4 kg/s

mL=1.814e-3 kg/s

mL=3.140e-3 kg/s

mL=5.700e-3 kg/s

prun > pbranch

prun < pbranch

Figure 5.24: Air-water flow (∆p)out vs λG plots for different water flow rates

Figure 5.25 is another widely used representation of phase separation; the ‘complete phase separation’

curve. This curve shows at what total mass flow rate fraction all the liquid will be diverted into

the branch. It is important to note that this representation differs slightly from the discussion in

Sec. 3.1 which was aimed for total gas phase separation. Consequently, mathematical description of

the liquid phase separation needs to be arranged as follows:

m1(1− x1) = Inlet liquid mass rate

m3(1− x3) = Branch liquid mass rate(5.2)

5.4.1 Hydrocarbon Mixture Uneven Phase Separation

In this section, the hydrocarbon mixture (Appendix D) used for the analysis of two-phase flow in

Sec. 5.3 is utilized again for uneven phase separation analysis. Table 5.4 describes the tee structure.

In order to ensure that liquid hydrocarbon is present in the flow stream throughout arms of the

whole tee structure, average operating p− T conditions are chosen to fall within the phase envelope

(i.e. p = [12536 − 12696] kPa) and, given the length of arms (150m), flow is assumed to occur

isothermally at T = 299.82 K and without mass transfer.

Based on average operating pressure different amounts of liquid is formed within the inlet arm and

consequently different phase separation curves are obtained as shown in Fig. 5.26. As an outcome of

the retrograde condensation (keeping the operational temperature constant) more liquid is formed

109

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00

(1-x3)/(1-x1)

m3/m1

mL=1.427e-4 kg/s

mL=1.814e-3 kg/s

Complete Separation

Figure 5.25: Air-water complete phase separation curves

Table 5.4: Hydrocarbon mixture flow split data

Arm length L 150 m

Number of CVs nCV 5

Arm diameter D 0.70485 m

Pipe roughness ε 0.24384× 10−4 m

Gas flow rate qG 178.29 m3/s

Flow temperature T 299.82 K

110

in the line with decreasing average operating pressure; i.e. inlet void fraction values lie in the range

[0.98− 0.95]. Hence, slopes of phase separation curves decrease with decreasing operational pressure

(more liquid), as observed previously for air-water mixture (Fig. 5.23).

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

λ L

λG

p = 12536.81 kPa

p = 12636.81 kPa

p = 12696.81 kPa

Figure 5.26: Hydrocarbon mixture flow λG vs λL plots for different average flow pressures

Table 5.5 lists overall composition of hydrocarbon streams going through each of the arms of the

tee when all the liquid phase is diverted into the branch. Consequently, mole fractions of heavier

components (i.e. Lumped-3 and Lumped-2) are greater in the branch stream; compared to inlet

and run streams, since heavier components condense within the liquid phase; while, fraction of

the lighter component (i.e. Lumped-1) has decreased in the branch stream. However, the changes

in the compositions are not significant and points to the conclusion that for small holdup values

compositional change can be ignored for the purposes of network analysis.

Table 5.5: Hydrocarbon mixture flow change in overall composition after flow split at the T-junction;pavg = 12636.81 kPa, λG = 0.55 and λL = 1

Mole fraction:

Component: Inlet Run Branch

Lumped-1 0.7658 0.7679 0.7641

Lumped-2 0.2240 0.2226 0.2251

Lumped-3 0.0102 0.0095 0.0108

111

5.5 Open Network Application

The significance of uneven phase separation is further demonstrated for air-water mixture flowing

through a relatively short, open network with two 90o (horizontal) branches joining a main line.

Network data is given in Tab. 5.6.

Table 5.6: Air-water flow open network data

Main line L 60 m

Main line num. of CVs nCV 20

Branch (1 & 2) L 15 m

Branch num. of CVs nCV 5

Arm diameter D 0.051 m

Pipe roughness ε 0.005 m

Inlet gas flow rate mG 1.973× 10−2 kg/s

Air MW MWair 29 g/mole

Gas compressibility Z 0.9

Inlet water flow rate mL 2.23× 10−2 kg/s

Water density ρw 998 kg/m3

Flow temperature T 288.71 K

Run outlet pressure prun 109.39 kPa

Branch outlet pressure (1st) pbranch1 110.925 kPa

Branch outlet pressure (2nd) pbranch2 109.39 kPa

Pressure distribution in the network is given in Fig. 5.27. Important observation to note is the

pressure rise (the spike) in the main line right after each split (around 20m and around 52m). Void

fraction (αG) distribution of the network is given in Fig. 5.28; demonstrating how significantly main

line void fraction can change as a result of uneven phase separation.

112

109.0

109.5

110.0

110.5

111.0

111.5

112.0

112.5

0 10 20 30 40 50 60 70

Pre

ssu

re [

kPa]

x [m]

Inlet

Run

Branch

Branch 2

110.93

110.94

110.95

110.96

110.97

15 17 19 21 23 25

109.50

109.52

109.54

109.56

109.58

109.60

47 49 51 53 55

Figure 5.27: Air-water open network pressure profile

0.945

0.95

0.955

0.96

0.965

0.97

0.975

0.98

0 10 20 30 40 50 60 70

Pre

ssu

re [

kPa]

x [m]

Inlet

Run

Branch

Branch 2

Figure 5.28: Air-water open network αG profile

113

5.6 Single-Phase Closed-Loop Network Application

A simple, single phase, steady-state network presented in Fig. 5.29, is solved using the model

proposed in this study. All required data is shown in Fig. 5.29 (i.e. supply/demand and delivery

pressures) and given in Tab. 5.7 (i.e. fluid and pipe properties). Results are compared to the

traditional gas network solution that uses generalized analytical gas flow equations. It is important

to note that upstream and downstream nodes are assigned randomly to describe the necessary

pipe interconnectivity and to provide an initial guess for flow directions. Actual flow directions

are ultimately determined by the models and both methods expected to be in agreement. Given

the inlet flow rate and exit pressures, the network is first solved using the FVM based model to

predict exit flow rates and then these exit flow rates are fed back in the conventional model (based

on Kirchhoff’s laws and analytic equations) in order to compare pressure distribution predictions

for the network. Pipe properties are kept same for the whole network.

Adopting the FVM model of this network required some effort since the nodes, where both a

supply/demand has been assigned and 3 pipes meet, need to be replaced with two consecutive tee

structures. However, the cells of these structures should be small enough not to significantly change

the network topology. This configuration; consecutive tee junctions, is also expected to be a better

representation of actual field conditions where pipes are typically connected using 90o tees and it

is not practical to connect more than 3 pipes to the same pipe junction. The FVM arrangement

built for the network is presented in Fig. 5.30. The shaded cells (gray and light green) are each

1m long while all the other cells are 2414/3 = 804.67m long. In fact, Fig. 5.30 demonstrates how

similar in essence the FVM representation of a pipeline network to the typical single layer reservoir

representations in simulation studies.

Exit flow rates obtained by the FVM model, as well as the input rate, are supplied to the analytical

model. Pressure specification at only one node (in this case at node 3) is sufficient for the analytical

solution. Pressure distribution obtained by the FVM code is compared to the analytical solution

in Tab. 5.8. Pressures seem to be in good agreement with the analytical results. Up to this point,

the proposed FVM model has been successfully applied to the study of single-phase networks and

multiphase flow conditions in single pipes and tees.

114

Figure 5.29: Single-phase closed-loop network

Table 5.7: Single-phase network data

Parameter SI Units US Units

Pipe length 2414 m 7920 ft

Pipe diameter 0.1016 m 4.00 in

Pipe roughness 1.8288E-04 m 0.0072 in

Gas gravity (γG) 0.5524

z-factor (Z) 0.90

Operating Temperature 23.89 oC 75.00 oF

115

Figure 5.30: FVM representation of the single-phase closed-loop network

Table 5.8: Comparison of pressure results of of the single-phase network

Node Pressure (Analytic) Cell Pressure (FVM) Deviation

# kPa psia # kPa psia %

N1 1974.452 286.37 1 1939.340 281.28 1.7783

N2 1283.459 186.15 37 1279.124 185.52 0.3378

N3 1103.161 160.00 23 1103.204 160.01 -0.0039

N4 1159.078 168.11 7 1162.302 168.58 -0.2782

N5 1116.054 161.87 17 1116.336 161.91 -0.0252

N6 1089.303 157.99 29 1090.359 158.14 -0.0970

N7 1011.047 146.64 11 1014.038 147.07 -0.2958

N8 1010.771 146.60 54 1013.140 146.94 -0.2343

N9 1008.358 146.25 33 1010.805 146.60 -0.2426

116

Chapter 6

Concluding Remarks

Current state-of-the-art modeling software available to oil and gas industry do not address uneven

phase separation (route preference) issue and typically do not provide a distinct T-junction component

for modeling purposes. This study includes a comprehensive review of branching T-junction and

phase separation models available in the open literature and classification of modeling efforts. In that

regard, a FVM based one-dimensional, two-fluid model has been developed for steady-state analysis

of low-liquid loading two-phase flow split at branching T-junctions with the purpose of uneven phase

separation and route preference analysis in gas condensate networks. The model uses steady-state,

one-dimensional Euler equations and outlet pressure specifications. Uneven phase separation at

the junction has been captured through incorporating a fluid mechanics based phase separation

sub-model at the junction; double stream model (DSM) of Hart et al. (1991), where Bernoulli type

mechanical energy equation (gas phase) ultimately replaced phasic momentum equations at the

junction CV and closed form DSM equation is solved separately in order to determine amount of

phase separation.

Developed model captures the expected pressure rise in the run direction following the flow split

at the junction and predicts different amounts of liquid phase separation with increasing liquid

flow rate and different outlet pressure specifications, in agreement with experimental data and

trends available in the literature. With increasing liquid flow rate, lesser amounts of the liquid

phase goes into the branch as a result of higher inertia and axial momentum; while, decreasing

branch outlet pressure (or increasing run outlet pressure for that matter) has been observed to

divert more of the liquid phase into the branch by creating a greater centripetal force on the liquid

phase. Depending on the phase separation, significantly different void fractions can ensue in the

branch or run directions compared to inlet void fraction (Fig. 5.28). Compositional changes in the

outgoing hydrocarbon mixtures following the split have been observed to be insignificant for the

particular case of low-liquid loading conditions (i.e. αL < 0.06, Tab. 5.5)

6.1 Conclusions

Conclusions based on the most prominent challenges of this study and respective observations

follows below:

• No apparent difficulties have been observed related to the assignment of gas phase properties as

liquid properties when liquid phase is absent (for hydrocarbon mixtures). Typically, following

117

a pressure (or temperature) perturbation, when constructing the jacobian matrix, severe

discontinuities are anticipated to surface when EOS module decides formation of the liquid

phase and computes actual liquid properties in order to replace bogus liquid phase properties

that were equivalent of gas phase. Despite the fact that system continues to use the same

equation set associated with the phase flag assigned to a CV at the beginning of an iteration,

significant changes in the phasic properties (i.e. density and viscosity) could cause problems.

• While single-phase to mist flow pattern (two-phase) transition had no problems, convergence

problems have been detected when a direct transition from single-phase to stratified flow

pattern is artificially induced. This is due to larger differences in gas phase and liquid phase

velocities in stratified flow and associated magnitudes of wall and interfacial friction forces.

stability and convergence problems when trying to switch from single-phase to stratified flow

pattern directly.

• The dummy momentum equation, along with the use of gas phase properties for the absent

liquid phase and averaging of neighboring densities to be substituted in the liquid momentum

equation, has been observed to cause an artificial jump in the liquid phase profile.

• Without the irreversible losses, utilizing phasic momentum equations at the T-junction CV has

been observed to predict a fixed phase separation (i.e. λL/λG = constant) for different flow

rates. However, with the inclusion of loss terms (Ottens et al., 1994) convergence problems

has been noted.

• Replacing phasic momentum equations with Bernoulli equations and taking phasic loss

coefficients to be equivalent (the inherent assumption of DSM) or computing liquid phase loss

coefficients based on Ottens et al. (1994) correlations (basis of advanced DSM) have been

found out to be inconsistent. Instead, closed form DSM equation is solved separately and

phasic momentum equations are replaced with gas phase Bernoulli equation.

• The DSM benefits from the fact that macroscopic mechanical energy balances should be

satisfied at all times. Nevertheless, when flow pattern geometry and phasic distribution in the

main line is not accounted for, DSM and similar mechanistic approaches (i.e. one-dimensional

momentum equations) are likely to fail if branch diameter is significantly smaller than main

line diameter and/or side port orientation1 is not horizontal. Especially, if model has no way

of recognizing the highest elevation the liquid film (or the liquid surface) within the main

line, it could still predict liquid phase going into the branch while it is not actually physically

possible; i.e. liquid level lower than the entrance of side port opening.

• The distances from tee center to CV faces have been observed to have an effect on convergence;

as these distances get smaller system required initial conditions closer to the actual solution;

else, either convergence rate has slowed or failed completely.

• Phases are not necessarily at thermodynamic equilibrium initially (i.e. upon entering the

CV); especially when a secondary stream merges with a primary stream at a combining

1Location of the branch entrance on the main line surface, not the inclination of branch arm

118

T-junction. There is a time factor involved for the establishment of chemical equilibrium

among the merging streams within a CV. Hence, chemical equilibrium may not be established

before phasic streams leave the current CV and move to the next one at new p-T conditions.

For sufficiently long CVs however; chemical equilibrium assumption is reasonable based on

volumetric averages utilized in FVM. Furthermore, over a fine grid where change in p − Tis insignificant between neighboring CVs; assumption of thermodynamic equilibrium is not

expected to cause any divergence from the actual physics unless sharp changes are observed in

p− T ; i.e. due to sudden enlargement at a T-junction. This is the only source of anticipated

sharp changes in p− T parameters in steady-state analysis, apart from merging of streams at

a T-junction. Regardless, a kinetic model for mass transfer, accounting for the time it takes

fluid streams to traverse the CV, should establish the correct amount of mass transfer without

the chemical equilibrium assumption.

• Heat loss to the surroundings is calculated based on a simplified model that inherently assumes

both phases to be in thermal equilibrium which could introduce errors in the prediction of

overall heat transfer to the environment when using finer grids following the argument in the

previous paragraph.

6.2 Recommendations

Based on aforementioned conclusions following recommendations are in order for future work:

• Although providing better initial guesses (initial conditions) has almost always alleviated the

convergence problems that occurred during the course of this study, the need for a globally

convergent Newton-Raphson algorithm; one that introduces obtained improvements in fractions

whenever iterations begin to diverge from the locality of actual solution, is apparent in order

to assure convergence when initial conditions are outside the the radius of convergence.

• For steady-state conditions, use of following equation in place of the missing phase momentum

equation should eliminate the jump observed in the absent phase velocity profile, when

single-to-two phase transition occurs:

0 = vG − vL (6.1)

• A ‘front tracking’ algorithm, similar in theory to the ‘level tracking’ algorithm discussed by

Aktas (2003), could be useful in predicting the actual location of phase appearance within a

long CV.

• Impact of (1) near T-junction griding and (2) distance of CV faces from the tee center; on the

pressure rise in the run direction and convergence issues is to be further inspected.

• Inclusion of advanced DSM to relieve the current model from the small holdup constraint

and accounting for the liquid level in the main line would improve the model applicability

significantly.

119

• It could be possible to obtain better correlations for phasic momentum correction factors (or

phasic energy loss coefficients) for two-phase flow at a T-junction through a separate 3D CFD

study.

• Proper modeling of two-phase flow through counter-combining and impacting T-junctions,

in addition to branching T-junctions, is a must for a full scale, closed-loop gas condensate

network analysis.

• Discussion in the last two paragraphs of Sec. 6.1 suggests that a grid sensitivity analysis is

essential for T-junctions along with the sensitivity analyses for changes in mixture composition,

void fraction, outlet pressures and a wider range of gas-liquid flow rates.

120

References

Adechy, D. and Issa, R. I. (2004). Modelling of Annular Flow Through Pipes And T-Junctions.

Computers and Fluids, 33(2):289–313. (Cited on pages viii, 6, 7, and 30.)

Aktas, B. (2003). Level Tracking in Thermal-Hydraulic Simulations of Nuclear Reactors. PhD

thesis, The Pennsylvania State University. (Cited on page 118.)

Al-Wazzan, A. (2000). The Split of Horizontal Two-Phase Flow at a T-Junction – CFD Study. In

PHOENICS User Conference. (Cited on page 6.)

Ayala, L. F. (2001). A Unified Two-fluid Model for Multiphase Flow in Natural Gas Pipelines.

Master’s thesis, The Pennsylvania State University. (Cited on pages 65, 91, 97, and 156.)

Ayala, L. F. and Adewumi, M. A. (2003). Low-liquid Loading Multiphase Flow in Natural Gas

Pipelines. Journal of Energy Resources Technology, 125(4):284–293. Transactions of the ASME.

(Cited on pages 4, 24, and 93.)

Ayala, L. F. and Alp, D. (2008). Evaluation of “Marching Algorithms” in the Analysis of Multiphase

Flow in Natural Gas Pipelines. Journal of Energy Resources Technology, 130(4):04303–10. (Cited

on pages 4, 24, and 94.)

Azzopardi, B. J. (1999). Phase Separation At T Junctions. Multiphase Science and Technology,

11(4):223–329. (Cited on page 30.)

Azzopardi, B. J. and Rea, S. (2000). Phase Separation Using a Simple T-junction. In SPE Annual

Technical Conference and Exhibition, Dallas, TX. Society of Petroleum Engineers. (Cited on

page 2.)

Azzopardi, B. J. and Whalley, P. B. (1982). The Effect of Flow Patterns on Two-phase Flow in a

T-Junction. International Journal of Multiphase Flow, 8(5):491–507. (Cited on pages 6, 37, 38,

44, and 45.)

Ballyk, J. D. and Shoukri, M. (1990). On the Development of a Model for Predicting Phase Separation

Phenomena in Dividing Two-phase Flow. Nuclear Engineering and Design, 123(1):67–75. (Cited

on pages 5, 27, 30, 36, 39, and 137.)

Ballyk, J. D., Shoukri, M., and Chan, A. M. C. (1988). Steam-Water Annular Flow in a Horizontal

Dividing T-Junnction. International Journal of Multiphase Flow, 14(3):265–285. (Cited on

pages 34, 36, 42, and 45.)

121

Bassett, M. D., Winterbone, D. E., and Pearson, R. J. (2001). Calculation of Steady Flow Pressure

Loss Coefficients for Pipe Junctions. In Proceedings of the Institution of Mechanical Engineers

(IMechE), volume 215, pages 861–881. (Cited on page 43.)

Beggs, D. H. and Brill, J. P. (1973). A Study of Two-Phase Flow in Inclined Pipes. SPE Journal of

Petroleum Technology, 25(5):607–617. (Cited on page 26.)

Bendiksen, K. H., Malnes, D., Moe, R., and Nuland, S. (1989). Dynamic Two-fluid Model OLGA:

Theory and Application. Society of Petroleum Engineers of AIME, pages 171–180. (Cited on

pages 22 and 69.)

Berger, E. L., Kolthoff, K. W., Schrodt, J. L. G., Long, S. L., and Pauley, J. C. (1997). The

SpliTigatorTM : A Device for the Mitigation of Phase Splitting. In International Thermal

Operations and Heavy Oil Symposium, Bakersfield, California. Society of Petroleum Engineers.

(Cited on page 2.)

Bird, R. B., Stewart, W. E., and Lightfoot, E. N. (2002). Transport Phenomena. Wiley, 2nd edition.

(Cited on page 36.)

Buell, J. R., Soliman, H. M., and Sims, G. E. (1993). Two-Phase Pressure Drop and Phase

Distribution of a Horizontal Tee Junction. ASME Publications Fluid Engineering Division (FED),

Gas-Liquid Flows, 165:25–37. (Cited on pages 34, 35, 42, and 45.)

Chen, N. H. (1979). An Explicit Equation for Friction Factor in Pipe. Industrial and Engineering

Chemistry, Fundamentals, 18(3):296–297. (Cited on page 66.)

Chisholm, D. (1967). A Theoretical Basis for the Lockhart-Martinelli Correlation for Two-Phase

Flow. International Journal of Heat and Mass Transfer, 10(12):1767–1778. (Cited on page 21.)

Coulter, D. M. (1979). New Equation Accurately Predicts Flowing Gas Temperatures. Pipe Line

Industry, 50(5):71–73. (Cited on page 15.)

Crowe, C. (2006). Multiphase Flow Handbook. CRC Press. (Cited on pages 20 and 21.)

Dennis, J. E. and Schnabel, R. B. (1996). Numerical Methods for Unconstrained Optimization and

Nonlinear Equations. Classics in Applied Mathematics. SIAM. (Cited on page 77.)

El-Shaboury, A., Soliman, H., and Sims, G. (2007). Two-Phase Flow in A Horizontal Equal-sided

Impacting Tee Junction. International Journal of Multiphase Flow, 33:411–431. (Cited on

page 43.)

Ellison, T. K., Hatziavramidis, D. T., Sun, B., and Gidaspow, D. (1997). Computational Fluid

Dynamics (CFD) Model for Phase Separation at Branching Tee Junctions. In SPE Western

Regional Meeting, Long Beach, California. Society of Petroleum Engineers. (Cited on pages 6

and 30.)

Fouda, A. E. and Rhodes, E. (1974). Two-Phase Annular Flow Stream Division in a Simple Tee.

Trans. Instn. Chem. Eng., 52:354–360. (Cited on pages 2, 30, 35, 36, and 45.)

122

Frepoli, C., Mahaffy, J. H., and Ohkawa, K. (2003). Notes on the Implementation of a Fully-Implicit

Numerical Scheme for a Two-Phase Three-Field Flow Model. Nuclear Engineering and Design,

225(2-3):191–217. (Cited on pages 22, 70, 73, 77, 78, 79, 80, and 99.)

Hart, J., Hamersma, P. J., and Fortuin, J. M. H. (1991). Phase Distribution During Gas–liquid

Flow Through Horizontal Dividing Junctions. Nuclear Engineering and Design, 126(3):293–312.

(Cited on pages iii, 7, 8, 27, 39, 44, 45, 86, 104, and 116.)

Harten, A. (1983). High Resolution Schemes for Hyperbolic Conservation Laws. Journal of

Computational Physics, 49(3):357–393. (Cited on page 74.)

Harten, A. M. I., Engquist, B., Osher, S., Chakravarthy, S., and California Univ, L. A. (1987). Uni-

formly High Order Accurate Essentially Non-Oscillatory Schemes. III. Journal of Computational

Physics, 71:231–303. (Cited on page 74.)

Hatziavramidis, D., Sun, B., and Gidaspow, D. (1997). Gas-liquid Flow Through Horizontal

Tees of Branching and Impacting Type. American Institute of Chemical Engineers Journal,

43(7):1675–1683. (Cited on pages 6 and 30.)

Hong, K. C. (1978). Two-Phase Flow Splitting at a Pipe Tee. Journal of Petroleum Technology,

February:290–296. (Cited on page 6.)

Hwang, S. T., Soliman, H. M., and Lahey, R. T. (1988). Phase Separation in Dividing Two-Phase

Flows. International Journal of Multiphase Flow, 14(4):439–458. (Cited on pages 38, 39, 42,

and 45.)

Hwang, S. T., Soliman, H. M., and Lahey, R. T. (1989). Phase Separation in Impacting Wyes and

Tees. International Journal of Multiphase Flow, 15(6):965–975. (Cited on page 5.)

Ishii, M. and Hibiki, T. (2003). Thermo-Fluid Dynamics of Two-phase Flow. Springer Sci-

ence+Business Media, New York, NY. (Cited on pages 20, 22, and 58.)

Issa, R. I. and Oliveira, P. J. (1994). Numerical Prediction of Phase Separation in Two-Phase Flow

through T-Junctions. Computers Fluids, 23(2):347–372. (Cited on page 6.)

Jelinek, M. (2005). Study of Higher Order Numerical Schemes Applied to Full Two-Phase Flow.

Master’s thesis, The Pennsylvania State University. (Cited on page 73.)

Jones, J. and Williams, R. L. (1993). A Two-Phase Flow-Splitting Device That Works. SPE

Production and Facilities, 8(3):197–202. (Cited on page 2.)

Kalkach-Navarro, S., Lee, S. J., R. T. Lahey, J., and Drew, D. A. (1990). The Prediction of

Phase Separation in a Branching Conduit Using a Three Dimensional Two-Fluid Model, chapter

Multiphase Transport Formulation, pages 3–24. Multiphase Transport and Particulate Phenomena:

Volume 1. (Cited on page 6.)

Kappos, L. and Economides, M. J. (2004). Definitive Method of Flow Analysis of Surface Pipeline

Networks. In SPE Annual Technical Conference and Exhibition, pages 83–98. (Cited on page 135.)

123

Ke, S. L. and Ti, H. C. (2000). Transient Analysis of Isothermal Gas Flow in Pipeline Network.

Chemical Engineering Journal, 76(2):169–177. (Cited on page 135.)

Kleinstreuer, C. (2003). Two-Phase Flow: Theory and Applications. Taylor & Francis. (Cited on

pages 19, 22, and 47.)

Kolev, N. I. (2005). Multiphase Flow Dynamics, volume 1 Fundamentals. Springer. (Cited on

pages 58 and 127.)

Kumar, S. (1987). Gas Production Engineering, volume 4 of Contributions in Petroleum Geology &

Engineering. Gulf Pub. Co., Houston. (Cited on page 14.)

Lahey, R. T. (1986). Current Understanding of Phase Separation Mechanisms in Branching Conduits.

Nuclear Engineering and Design, 95:145–161. (Cited on pages 2, 30, 32, 44, and 45.)

Lahey, R. T. (1990). The Analysis of Phase Separation and Phase Distribution Phenomena Using

Two-fluid Models. Nuclear Engineering and Design, 122:17–40. (Cited on pages viii, 2, 6, 7,

and 27.)

Larock, B. E., Jeppson, R. W., and Watters, G. Z. (2000). Hydraulics of Pipeline Systems. CRC

Press, Boca Raton, FL. (Cited on pages 12 and 135.)

Leonard, B. P. (1979). A Stable and Accurate Convective Modelling Procedure BasedoOn Quadratic

Upstream Interpolation. Computer Methods in Applied Mechanics and Engineering, 19(1):59–98.

(Cited on page 72.)

Lindsay, R. B. (1931). Wave-Motion and the Equation of Continuity. Proceedings of the National

Academy of Sciences of the United States of America, 17:420–426. (Cited on page 73.)

Margaris, D. P. (2007). T-Junction Separation Modelling in Gas-liquid Two-Phase Flow. Chemical

Engineering and Processing, 46(2):150–158. (Cited on page 2.)

Marti, S. and Shoham, O. (1997). A Unified Model for Stratified-Wavy Two-Phase Flow Splitting

at a Reduced T-Junction with an Inclined Branch Arm. International Journal of Multiphase

Flow, 23(4):725–748. (Cited on page 39.)

Martinez, F. F. A. and Adewumi, M. A. (1997). Two-Phase Gas-Condensate Flow in Pipeline

Open-Network Systems. SPE Production and Facilities, 12(4):218–224. (Cited on pages 7, 25,

and 87.)

Meier, R. W. and Gido, R. G. (1978). One-Dimensional Calculation of Flow Branching Using

the Method of Characteristics. Technical Report NUREG/CR-0064 LA-7306-MS, Los Alamos

National Laboratory, Los Alamos, NM. (Cited on page 28.)

Mucharam, L. and Adewumi, M. A. (1991). A Study of the Effect of Condensate Formation on

Performance Predictions in a Gas Pipeline Network. In SPE Production Operations Symposium,

Oklahoma City, Oklahoma. Society of Petroleum Engineers. (Cited on page 26.)

124

Mucharam, L., Adewumi, M. A., and Watson, R. W. (1990). Study of Gas Condensation in

Transmission Pipelines with a Hydrodynamic Model. SPE Production Engineering, 5(3):236–247.

(Cited on page 65.)

Munson, B. R., Young, D. F., Okiishi, T. H., and Huebsch, W. W. (2009). Fundamentals of Fluid

Mechanics. Wiley, 6th edition. (Cited on pages 36, 41, 47, and 85.)

Oka, K. and Ito, H. (2005). Energy Losses at Tees With Large Area Ratios. Transactions of the

ASME, 127:110–116. (Cited on page 43.)

Oranje, L. (1973). Condensate Behaviour in Gas Pipelines is Predictable. Oil and Gas Journal,

(71):39–44. (Cited on pages 2, 6, and 29.)

Ottens, M., Hoefsloot, H. C. J., and Hamersma, P. J. (2001). Transient Gas-liquid Flow in Horizontal

T-Junctions. Chemical Engineering Science, 56(1):43–55. (Cited on pages 7 and 8.)

Ottens, M., Swart, A. d., Hoefsloot, H. C., and Hamersma, P. J. (1994). Gas-Liquid Flow Splitting

in Regular, Reduced and Impacting T-Junctions. In Conference on Multiphase Flow Dynamics in

Industrial Plants, Ancona, Italy. (Cited on pages 39, 41, 43, and 117.)

Passut, C. A. and Danner, R. P. (1972). Correlation of Ideal Gas Enthalpy, Heat Capacity, and

Entropy. Industrial and Engineering Chemistry, Process Design and Development, 11(4):543–546.

(Cited on pages vii, 154, and 157.)

Patankar, S. V. (1980). Numerical Heat Transfer and Fluid Flow. Series in computational methods

in mechanics and thermal sciences. McGraw-Hill, Newyork. (Cited on pages 68 and 69.)

Pauchon, C. L., Fabre, J., Dhulesia, H., and Cirlot, G. B. (1994). TACITE: A Transient Tool for

Multiphase Pipeline and Well Simulation. In SPE Annual Technical Conference and Exhibition,

volume Pi, pages 311–326. (Cited on page 21.)

Peng, D.-Y. and Robinson, D. B. (1976). A New Two Constant Equation of State. Industrial and

Engineering Chemistry, Fundamentals, 15(1):59–64. (Cited on page 67.)

Peng, F. and Shoukri, M. (1997). Modelling of Phase Redistribution of Horizontal Annular Flow

Divided in T-Junctions. Canadian Journal of Chemical Engineering, 75(1):264–271. (Cited on

pages 30 and 39.)

Peng, F., Shoukri, M., and Chan, A. M. C. (1993). Annular Flow in T-Junctions with Horizontal

Inlet and Downwardly Inclined Branches. ASME Publications Fluid Engineering Division (FED),

Gas-Liquid Flows, 165:13–24. (Cited on page 36.)

Penmatcha, V. R., Ashton, P. J., and Shoham, O. (1996). Two-Phase Stratified Flow Splitting at

a T-junction. International Journal of Multiphase Flow, 22(6):1105–1122. (Cited on pages 30

and 39.)

Prosperetti, A. and Tryggvason, G. (2007). Computational Methods for Multiphase Flow. Cambridge

University Press. (Cited on pages 58, 70, 71, and 75.)

125

Reimann, J. and Seeger, W. (1986). Two-Phase Flow in a T-Junction with a Horizontal Inlet. Part

II: Pressure Differences. International Journal of Multiphase Flow, 12(4):587–608. (Cited on

pages 35, 36, and 45.)

Rhie, C. M. and Chow, W. L. (1983). Numerical Study of the Turbulent Flow Past an Airfoil

with Trailing Edge Separation. American Institute of Aeronautics and Astronautics Journal,

21:1525–1532. (Cited on page 69.)

Roache, P. J. (1998). Fundamentals of Computational Fluid Dynamics. Hermosa Publishers,

Albuquerque, NM. (Cited on pages 47, 72, and 74.)

Roberts, P. A., Azzopardi, B. J., and Hibberd, S. (1995). The split of horizontal semi-annular flow

at a large diameter T-junction. International Journal of Multiphase Flow, 21(3):455–466. (Cited

on page 44.)

Saba, N. and Lahey, R. T. (1984). The Analysis of Phase Separation Phenomena in Branching

Conduits. International Journal of Multiphase Flow, 10(1):1–20. (Cited on pages 30, 35, 36, 38,

44, and 45.)

Seeger, W., Reimann, J., and Muller, U. (1986). Two-Phase Flow in a T-Junction with a Horizontal

Inlet. Part I: Phase Separation. International Journal of Multiphase Flow, 12(4):575–585. (Cited

on page 30.)

Shoham, O. (2006). Mechanistic Modeling of Gas-Liquid Two-Phase Flow in Pipes. Society of

Petroleum Engineers, Richardson, TX. (Cited on pages viii, 2, 4, 17, 19, 20, 21, 23, 25, 26, 61,

and 65.)

Shoham, O., Brill, J. P., and Taitel, Y. (1987). Two-Phase Flow Splitting in a Tee Junction

Experiment and Modelling. Chemical Engineerieg Science, 42(11):2667–2676. (Cited on pages 5,

38, 39, and 45.)

Singh, G. (2009). Finite-Volume Analysis of Two-Phase Split at T-Junctions. Master’s thesis, The

Pennsylvania State University. (Cited on pages 8, 71, and 104.)

Smoglie, C., Reimann, J., and Muller, U. (1987). Two Phase Flow through Small Breaks in a

Horizontal Pipe with Stratified Flow. Nuclear Engineering and Design, 99:117–130. (Cited on

page 2.)

Spore, J. W., Mahaffy, J. H., and Murray, C. (2001). TRAC-M/FORTRAN 90 (version 3.0) Theory

Manual. Office of Nuclear Regulatory Research, U.S. Nuclear Regulatory Commission. (Cited on

pages 8, 22, 26, 30, 41, 65, 69, 73, 80, and 85.)

Steinke, R. G. (1996). TRAC-P and Bernoulli-Equation Momentum-Convection Terms for the Tee-

Component Internal-Junction Cell Interfaces. Technical report, Los Alamos National Laboratory,

Los Alamos, NM. (Cited on pages 30, 41, 85, and 139.)

Sung, W., Huh, D., Lee, J., and Kwon, O. (1998). Optimization of Pipeline Networks with a

126

Hybrid MCST-CD Networking Model. SPE Production and Facilities, 13(3):213–219. (Cited on

page 135.)

Szilas, A. P. (1985). Production and Transport of Oil and Gas. Developments In Petroleum Science.

Elsevier, New York, 2nd completely rev. edition. (Cited on pages 127, 130, and 132.)

Tae, S.-J. and Cho, K. (2006). Two-Phase Split of Refrigerants at a T-junction. International

Journal of Refrigeration, 29(7):1128–1137. (Cited on page 2.)

Tian, S. and Adewumi, M. A. (1992). Development Of Analytical Design Equation For Gas Pipelines.

volume Pi, pages 989–995. Society of Petroleum Engineers of AIME. (Cited on page 91.)

Todreas, N. E. and Kazimi, M. S. (1990). Nuclear systems. Hemisphere Pub. Corp., New York.

(Cited on pages 2, 20, 47, 58, 61, and 65.)

Walters, L., Soliman, H., and Sims, G. (1998). Two-Phase Pressure Drop and Phase Distribution

at Reduced Tee Junctions. International Journal of Multiphase Flow, 24:775–792. (Cited on

pages vii and 45.)

Wesselling, P. (2001). Principles of Computational Fluid Dynamics. Spinger. (Cited on page 70.)

127

Appendix A

Single-Phase Flow in Pipe Networks

Pipeline systems (networks) exist in several engineering applications such as water distribution

systems in civil engineering, oil and gas pipelines in petroleum engineering and reactor cooling

systems in nuclear engineering.

Single-phase, compressible or incompressible flow in networks is typically modeled using one-

dimensional, steady-state, analytical expressions which are closed form equations derived from

fundamental principles of conservation of mass and energy applied over the entire pipeline length.

By analogy to electric circuits, these flow equations are coupled with Kirchhoff laws to obtain

steady-state solutions. Kirchhoff laws stand for conservation of mass at nodes (junctions of pipes)

and conservation of energy around loops.

From very simple to very complex pipeline systems, every network is composed of two basic

components (Szilas, 1985):

1. Node: Nodes are junctions where two or more pipes meet and supply/demand is introduced

to the system (i.e. wellheads, delivery stations etc.). Kolev (2005) uses the term ‘knots’ to

distinguish junction nodes from supply/demand nodes.

2. Node Connecting Element (NCEs): All the elements laying between nodes are considered as

NCE, such as legs (straight pipe sections), compressors, chokes and etc.

There are three distinct connection types for NCEs (Fig. A.1):

1. Series connection; outlet of one NCE is connected only to the inlet of another NCE

2. Parallel connection; inlet of one NCE is connected to the inlet of another NCE and their

outlets are connected to each other as well

3. Loop connection; three or more NCEs are connected in series to form a geometrically closed

structure (outlet of the last NCE is connected to the inlet of first NCE in the loop). However,

there is no restriction on the number of connections that the nodes of the loop can have.

The two kinds of pipeline systems can be built with these connection types; loopless systems and

looped systems. Loopless systems are better known as spanning trees or open networks where the

NCEs of the system form no closed loops Fig. A.2.

128

Figure A.1: Connection Types

Figure A.2: An Open Network

Looped systems can be further classified as single loop systems in which there is only one loop

formed by the NCEs and multiple loop systems in which several loops exist (Fig. A.3).

In a network which has a single or no input source, number of loops (around which independent

energy equations can be written) is equal to the number of NCEs minus number of nodes plus one

(Eqn. A.1).

Num. of loops = Num. of NCE−Num. of nodes + 1 (A.1)

Figure A.3: Closed Networks

129

If the network has more than one input source, number of loops is simply equal to the number of

NCEs minus number of nodes (Eqn. A.2).

Num. of loops = Num. of NCE−Num. of nodes (A.2)

A.1 Kirchhoff Analysis for Flow in Networks

Conventionally, before modeling a pipeline system a mathematical hydraulic model for every NCE in

the pipeline system is defined; usually in the form of a pressure drop versus flow rate (throughput)

relationship. Then, by recognizing an analogy between pipeline systems and electric circuits,

Kirchhoff’s laws are applied to the pipeline system.

The hydraulic model for steady-state, compressible, single-phase gas flow under isothermal conditions,

shown in Eqn. A.3, is the closed form equation derived from principle conservation equations as

previously presented in Sec. 2.1.1 (Eqn. 2.19).

qG = C

(Tscpsc

)√1

f

√p2in − p2

out es

γGTavgZavgLeD

52 (A.3)

This equation relates pressure drop across the pipe to the gas flow rate, pipe length, diameter

and roughness. This hydraulic model for single-phase compressible flow through straight pipes is

typically arranged in the form of Eqns. A.4 and A.5 with an analogy to Ohm’s law for conductors

in an electric circuit.

∆(p2) = q2GR (Resistance) ⇔ V (Voltage) = I (Current)R (Resistance) (A.4)

∆(p2) = p2in − p2

out = q2G (

pscTsc

)2 f

C2

γGTavgZavgLeesd5︸ ︷︷ ︸

Resistance R

(A.5)

where: qG is the flow rate through the leg (straight pipe section), ∆(p2) is the difference of squared

inlet and outlet pressures of the pipe and represents the potential for flow, R is the ’resistance’ term

in units of p2/q2G with an analogy to circuitry. The ‘resistance’ term accounts for the cumulative

effect of pipe properties and flow variables such as diameter, friction factor, fluid properties etc.

Kirchhoff’s first law, also known as the ‘node law’ is indeed an application of conservation of

mass at a node where algebraic sum of mass flow coming into and going out of a node is zero

Eqn. A.6. It is also possible to observe the first law as a node/junction continuity equation

nconn∑i

(qG)i = 0 (A.6)

Where; (qG)i is the flow in the i’th NCE connected to the node and nconn is the number of NCEs

connected to the node.

130

Kirchhoff’s second law or the ‘loop law’ is the application of conservation of energy around a

loop where algebraic sum of potential change (i.e. pressure drop) along each NCE in the loop, with

a consistent sign convention, is zero. Hence it is also known as the energy loop equation (Eqn. A.7).

With an analogy to circuitry, this second law could be interpreted as the movement of a particle

around the loop; a particle begins the movement at a certain node and always follows a certain

direction (clockwise or anti-clockwise). One loop is completed when the particle arrives at the

starting node. Then, (potential) energy of the particle should not change since no work has been

done or since particle arrived at the same reference point of same potential.

nLoop∑i

∆(p2)i = 0 (A.7)

Where; ∆(p2)i is the potential change in the i’th NCE of the loop and nLoop is the number of NCEs

in the loop.

A.2 Steady-State Analysis of Pipeline Systems

Flow in pipeline systems is typically assumed to be at steady-state in order to simplify the modeling

task. In this section, the most traditional approaches for steady-state analysis of pipeline network

are presented.

A.2.1 Open Network Analysis

In a ‘loopless’ system with n nodes and m = n− 1 pipes, if parameters to compute NCE resistances

(R) are present as well as necessary initial and boundary conditions, pressure at any node of the

network could be computed in a sequential manner. The necessary initial and boundary conditions

include a proper set of flow rates in to and out of the nodes and any of the terminal pressures (any

one of the network inlet or outlet pressures). Then, Beginning with the NCE connecting the node of

known terminal pressure to the network; hydraulic equations for each NCE are solved successively

along with the application of Kirchhoff’s first law.

A.2.2 Closed Network Analysis

Cross Method

For looped system analysis, one of the early methods is again an iterative procedure, proposed by

Cross for the analysis of a low pressure looped system and later extended by Hain for high pressure

systems (Szilas, 1985).

Let us consider the analysis of a single loop system where flow rate and pressure at the inlet and

input/output rates at the nodes are known but pressures at the nodes and flow rates in the pipes

are to be determined. This analysis requires the application of the Kirchhoff’s second law, in which

a consistent sign convention for flow direction is needed. Conventionally, clockwise direction is

131

positive. Initially; a flow rate is assumed for one of the lines emerging from the inlet node and then

node continuity equation is applied at every successive node till one loop is completed in a certain

direction. Therefore, flow rates in pipes are computed and node pressures can be easily obtained.

If calculated flow rates for the lines differ from the actual rates by an ∆q amount due to initially

assumed flow rate in the first pipe, then Kirchhoff’s second law is applied for the loop and initially

assumed inlet flow rate is corrected by an ∆q amount as expressed in Eqn. A.8:

(qG)actualinlet = (qG)computedinlet + ∆q (A.8)

Kirchhoff’s second law for a loop is arranged as shown in Eqn. A.9 by substituting the analytical

flow equation in the resistance form as given in Eqn. A.7.

nLoop∑i

∆(p2)i =

nLoop∑i

(q2G)iRi (A.9)

Equation A.9 is further modified to accommodate the computing procedure as shown in Eqn. A.10

and can be re-written again for the correction ∆q as presented in Eqn. A.11.

nLoop∑i

(qG)i|qG|iRi =

nLoop∑i

δi(q2G)iRi = 0 (A.10)

nLoop∑i

(qG)i + ∆q |(qG)i + ∆q|Ri =

nLoop∑i

δi (qG)i + ∆q2Ri (A.11)

Where, δi is a sign function to account for adopted sign convention correctly.

Subtracting Eqn. A.10 from Eqn. A.11 yields Eqn. A.12:

nLoop∑i

δi

(q2G)i + 2(qG)i∆q + ∆q2

Ri = 0 (A.12)

For |∆q| << | qG|, ∆q2 ≈ 0 then Eqn. A.12 simplifies to Eqn. A.13:

nLoop∑i

δi

(q2G)i + 2(qG)i∆q

Ri = 0 (A.13)

Finally required correction for the flow rates can be obtained from Eqn. A.14:

∆q =

nLoop∑i

δi(q2G)iRi

−2

nLoop∑i

δi(qG)iRi

(A.14)

132

In a multiple loop system, initially; flow rates are assumed for individual pipes in loops. Correction

amount ∆q is individually computed for loops and pipe flow rates are corrected loop after loop.

Flow rates in pipes shared by two (or more) loops are corrected using the ∆q determined for every

loop containing the pipe. In other words; flow rate in this shared pipe is corrected each time a

correction done in any of the loops it belongs to.

Renouard Method

The Cross method converges rather slowly for a multiple loop system. One reason of the slow

convergence is the individual corrections applied for each shared pipe. Because loops are addressed

independently; an improvement of the flow rate in a shared pipe by one of the loops can be attenuated

by another loop sharing the same pipe. In order to overcome this convergence difficulty Renouard

(Szilas, 1985) offered a variant of the Cross method where interaction of the loops sharing common

elements are taken into account. In this method, flow correction for a pipe is made using the sum

of individual corrections from the loops sharing the same element and this requires simultaneous

solution of the loop equations.

Stoner Method

Stoner introduced a method capable of treating looped systems with different types of NCEs and

the system response to any parameter of these NCEs while Cross method was limited to flow rate

and node pressure analysis. The basis of the analysis is the simultaneous solution of all network

governing equations using a multivariable Newton-Raphson iterative procedure (Szilas, 1985).

A pipeline system of n nodes and m NCEs yields a system of n equations with (2n+m) variables

when node continuity equations (Kirchhoff’s second law) are written at each node. While the

(2n+m) variables are; n node pressures, n supply/demand rate at nodes and either m NCE flow

resistance or m NCE flow rate; this system of n equations can be solved for any set of n variables

provided that other (n+m) variables are known.

As described before; continuity equation at any node can be expressed as shown in Eqn. A.15:

Fj = (qG)s +

nconn∑i

(qG)i = 0 (A.15)

Where:

(qG)s Flow in to or out of node j (supply/demand)

(qG)i Flow in the i‘th NCE connected to node j

nconn Number of pipes connected to node j

Fj Node continuity equation at node j

Flow in any NCE that is connected to a node can be further defined using the mathematical

hydraulic model for the NCE. Hence, node continuity equation itself becomes a function of all the

133

parameters controlling the flows in the NCEs reaching the node and can be presented as shown in

Eqn. A.16.

Fj = Fj(x1, x2, x3, ..., xn) = 0 (A.16)

Similar equations written for every node in the system constitutes a non-linear system of equations

since hydraulic model expressions involve powers and multiplications of unknowns. Nevertheless,

system can be solved with various iterative techniques (such as Newton-Raphson) as long as involved

equations are linearly independent.

A.3 Transient Analysis of Pipeline Networks

Flow in gas pipeline systems is frequently at transient (unsteady-state) and though steady-state

modeling may be adequate for several instances; the aim of transient modeling is the analysis of

planned modifications to the system as well as unplanned situations such as a leak causing a sudden

pressure drop in the system.

Transient analysis of a pipeline system is carried out in two parts; first, transient flow in NCEs

must be modeled then node continuity can be applied. The hydraulic model for straight pipes –

Eqn. 2.19 derived in Sec. 2.1.1 - no longer describes the flow accurately since the simplifications

lead to its derivation are based on the assumption of steady-state conditions. When modeling

transient flow in a single pipe, it is not possible to simplify the governing conservation equations.

Hence, simultaneous solution of the two fundamental equations; mass and momentum conservation

equations, is required. For the case of non-isothermal flow; a third equation, energy conservation

equation, is needed as well.

In typical network transient analysis, Euler equations (mass and momentum conservation) are

applied over the pipes while junctions are still treated solely as nodes at which only the conservation

of mass (Kirchhoff’s first law) is applied. This implies the assumption of steady-state conditions

at the nodes and is reasonable for the practical purposes of studying oil and gas networks where

volumes of junctions are insignificant compared to the very long pipes.

The fundamental equations, whether in conservative form or not, constitute a system of non-

linear partial differential equations (NL-PDE) which may be solved in several different ways.

Most frequently adopted techniques include Finite Difference Method (FDM) and Method of

Characteristics (MOC) and its variations.

The node continuity equations applied at every node will also need to be solved simultaneously. The

fact that each qi in every node continuity equation will be computed using one of the aforementioned

methods makes the computation procedure very laborious. Conventionally, assuming an up front

insignificant loss in accuracy, some simplifications are done on the equations defining the transient

flow in the pipe. These assumptions include isothermal conditions, neglecting elevation changes and

removing some of the terms in momentum equation which are usually smaller than other terms by

an order of magnitude.

134

For the pipes connecting to the same node, simplified form of mass conservation equations are

written at the nodes. These equations are summed together and all the flux terms ( ∂∂x(ρ v)) are

collected on one side. After rearranging, the transient form of node continuity equation, Eqn. A.17,

is arrived:

Vj kdpjdt

= (qG)s,j +

nconn∑i

(qG)i,j (A.17)

Where:

(qG)s,j Flow in to or out of node j (supply/demand)

(qG)i,j Flow in the ith NCE connected to node j

nconn Number of pipes connected to node j

Vj Pseudo node volume which is equal to

the sum of half the volumes of every pipe connected to node j

pj Pressure at node j

k Represents the remaining terms

Simplified and arranged form of momentum equation is substituted in place of flow rates in this new

form of the node continuity equation (Eqn. A.17). A new system of NL-PDEs is obtained following

the same steps at every node.

A.3.1 Finite Difference Method

Governing NL-PDEs of the network are first transformed into algebraic equations using spatial

and temporal discretization schemes. Later on, algebraic equations can be solved for the values

of variables at certain points in the flow domain. FDMs are classified according to temporal

discretization scheme utilized.

In an Explicit Method; NL-PDEs are transformed into algebraic equations where unknown values of

variables at time t+ ∆t depend on the values at time t which are known. Hence algebraic form of

conservation equations can be solved individually in a straight forward manner.

In an Implicit Method; NL-PDEs are transformed into a system of algebraic equations where values

of unknown variables at time t + ∆t depend on the values of variables at neighboring points at

t+ ∆t. Therefore; simultaneous solution of the equations are needed.

Crank - Nicholson Method may be considered as a combination of implicit and explicit methods

where values of unknown variables at a time t + ∆t depend on both the values of variables at

neighboring points at time t+ ∆t and the values of variables at time t. So, simultaneous solution of

the equations is still needed.

A.3.2 Method of Characteristics

In the method of characteristics, by finding paths (characteristic lines or characteristics) in the plane

of natural coordinates (independent variables) of the equation system, the NL-PDEs are reduced to

135

ordinary differential equations (ODE) which can then be solved easier numerically (Larock et al.,

2000).

A.4 Description of the Network Model for Numerical Solution

Natural Gas transmission networks can be represented as a directed graph which provides all

the necessary connection information. Connections between nodes are easily described using a

connection matrix and loops are defined by a loop matrix. Loop matrix can be extracted from the

connection matrix using a connected graph which is a ‘loopless’ graph. This connected graph is

called the matching matrix and in essence contains two individual matrices derived from connection

matrix. One of these matrices (the matrix of tree branches) contains the information of all NCEs

except the ones creating loops in the system. NCEs causing loops in the system are collected in

another matrix (the matrix of chord branches) that is added to the end of tree matrix to form the

matching matrix. Conventionally NCEs form the columns and nodes form the rows in all of these

matrices (Eqn. A.18).

A =

a1,1 a1,2 a1,3 . . . a1,m

a2,1. . .

.... . .

an,1 . . . an,m

︸ ︷︷ ︸

NCEs(m)

nodes(n) (A.18)

Any element ai,j of the matrix A will be assigned:

• +1 if edge j emerges from node i

• −1 if edge j ends in node i

• 0 if edge j is not connected to node i

Once the network is defined by the connection and loop matrices, node continuity equation and

energy loop equation can be expressed as matrix operations to be carried out in the computer.

Kappos and Economides (2004) provide a simple, iterative algorithm for the analysis of pipeline

networks. The algorithm neither requires prior knowledge of split node location nor good initial

estimates of the solution. Instead, code successively computes node pressures, pipe flow rates and

flow directions.

Sung et al. (1998) developed a network optimization tool based on a minimum cost spanning tree

approach (in which occurrence of loops are not allowed in the network) for the optimum path and

constrained derivative approach for the optimum diameter. Their solution algorithm for the network

flow includes the conventional Newton-Raphson method.

Ke and Ti (2000) extended the conventional analogy with electrical networks to analyze transients

in an isothermal gas pipeline network. In this approach, other circuitry concepts like capacitance

136

and inductance are accounted for. A Major benefit of this approach is that governing equations are

transformed in to a set of first order ODE instead of dealing with second order (hyperbolic) PDE,

hence computational costs are greatly reduced while reasonably accurate results were obtained.

While it is possible to reduce governing equations to a single, closed form relation (the hydraulic

model) for single-phase steady-state flows it has not been yet shown that a similar approach can be

taken without making any simplifying assumptions which eventually lead to loss of an adequate

mathematical description of the system and the two-phase flow phenomenon. However, without

the closed form hydraulic models for pipes or NCEs, it is not convenient to form an analogy

to the circuitry leaving Kirchhoff technique as an unfeasible option for the network analysis of

two-phase flow. Instead, a proper solution technique needs to accommodate simultaneous solution

of conservation equations for pipes and/or NCEs.

137

Appendix B

Double Channel T-Junction Model

Presence of a pressure spike, a pressure increase in the main line, following the flow split at a

T-junction is recognized through simple analysis of Bernoulli equation. The decrease in mass flow

rate along the main line of the tee (after the split) has an effect similar to flow area expansion along

a single pipe. Therefore, Bernoulli equation analysis indicates a pressure increase in the run. This

is most clear when irreversible losses are neglected:

1

2ρ1v

21 + ρ1gz1 + p1 =

1

2ρ2v

22 + ρ2gz2 + p2 +

1

2ρ1v

21k12 (B.1)

For;

2A1 = A2

ρ1 = ρ2

z1 = z2

12ρ1v

21k12

∼= 0

Then, by conservation of mass:

ρ1v1A1 = ρ2v2A2

ρ1v1A1 = 2ρ1v2A1

(B.2)

v2 =v1

2(B.3)

Solving Bernoulli equation for p2 requires that p2 > p1:

12ρ1v

21 + p1 = 1

2ρ1(v12 )2 + p2

12ρ1v

21 + p1 = 1

8ρ1v21 + p2

p2 = 38ρ1v

21 + p1

p2 > p1

(B.4)

Proper treatment and capture of flow split at a T-junction requires acknowledging that outgoing

arms of the tee do not necessarily share the same single pressure and same single void fraction

although their extended inlets originate from within the same control volume; i.e. tee node or the

knot (the tee cell itself in FVM grid).

The fact that arms of the tee should ‘feel’ different pressures can simply be observed from experimental

data by linear extrapolation of pressure profiles of run and branch arms towards the tee (Ballyk

138

and Shoukri, 1990).

Flow going out through each arm (or the split itself for that matter) is to be determined by

momentum balances that regard directionality of the flow. From FVM point of view, once the split

is determined, pressure profiles of outgoing arms is solely based on specified outlet pressures for the

arms. Nevertheless, utilized set of equations should tie pressures of outgoing arms, as well as the

flow rates, to the main line so that an accurate solution can be obtained in the main line as well.

Depending on the equations that establish the connection between tee cell inlet and tee cell

outlets, both the split and thereof pressure profiles along all arms of the tee would be different.

In other words, without proper accounting for the actual physics, different solutions would be

possible. However, purpose of the momentum equations written specifically for the tee (not

the regular momentum equations written for faces) is to predict correct split amounts so that

both outlet pressure specifications and momentum conservation within the tee cell are satisfied

properly/adequately.

From FVM perspective, the tee component itself is merely a 3 pipe system where flow coming out

of main pipe is split between branch and run. Both branch and run can be visualized as separate

pipes with outlet pressure specifications. Then, solution for an outgoing arm merely depends on the

flow going through it and it’s outlet pressure specification. This is because, in the staggered grid

approach inlet pressure specification has no significant influence on the solution (particularly for

isothermal analysis). However, void fraction (along with velocities) has control over the flows going

through arms.

From this point of view, it seems as if outgoing arms can be considered to feel a single pressure at

the tee and therefore only purpose of momentum equations written for the tee is to predict outgoing

flow rates (or velocities and void fractions for that matter). Perhaps this would have been the case

if any possible effect tee pressure might have on outgoing flow rates was to be ignored.

Nevertheless, even if a single tee cell pressure had no effect on outgoing flow rates, the connection

between pressures in the first cells of outgoing arms and tee cell has to be properly established

simply because, as a separate pipe, solution in the main pipe depends on the tee cell pressure which

becomes it’s outlet pressure specification.

In other words, it might be possible to obtain an incorrect tee cell pressure with a seemingly correct

flow split unless proper momentum equations are utilized to connect all 3 arms of the tee, thus

leading a different solution in the main line. In essence, this means not only the velocities of outgoing

faces but also outgoing pressures have to be properly linked to inlet pressure and velocities through

specialized equations for the tee.

To repeat it once more, even if the effect of a singular tee cell pressure on flow split (outgoing

velocities) could be neglected, the correct solution for the main pipe still depends on the link that

would be established between pressures felt by outgoing arms and the tee cell.

Another important point is that, in the case of two-phase flow, outgoing arms may as well have

139

separate/different void fractions, despite the fact that flow originates from the same pool - tee

CV itself. Unless two-phase pattern in the tee cell is mist (or bubbly) flow, behaving like an

homogeneous fluid, assumption of a single void fraction for both branch and run arms could be

misleading/erroneous Perhaps, it would have been a valid assumption for an homogeneous two-phase

flow model though. Instead, whether void fractions of outgoing streams are same or not should be

determined through application of conservation principles.

The whole point of this discussion is to emphasize that derivation of proper equations for flow split

should begin by recognizing outgoing pressures, void fractions and velocities as unknowns. These

unknowns are to be solved for via specialized equations written for the tee and not by regular

edge/face momentum equations. However, it is shown later in the following sections that equations

derived for tee could collapse to a specialized set of edge/face momentum equations for outgoing

faces/edges of tee cell.

At this stage of the analysis, there are a total of 10 unknowns to account for at a tee (Fig. B.1):

p31, p32, p33, α31, α32, α33, vG3, vL3, vG6, vL6.

B.1 Double Channel Model

For the derivation of proper tee split equations and to generate adequate number of equations

accounting for all unknowns, a variation of Steinke (1996) approach is adopted (shall be referred to

as Double Channel Model, DCM). Based on author’s original derivation for combining T-junctions,

(1) main pipe to run and (2) main pipe to branch are flows considered to occur through separate

channels;by dividing the tee CV into two smaller CVs, similar to the separating (or dividing)

streamlines approach adopted by mechanistic junction models. Doing so, Steinke recognizes that an

additional pressure at the tee (pT ), as well as portion of main line cross-sectional area (A313) should

be accounted for.

Figure B.1: The double channel model control volumes

140

Flow turning into branch can be treated as (contracting) bent pipe flow. Hence, presence of

additional pressure pT is simply a requirement of the bent pipe flow. This can be observed by

considering momentum balance over a bent pipe of uniform cross-sectional area. Ignoring irreversible

losses due to wall friction or bent of flow, momentum conservation in the direction of run (x-direction)

requires that momentum of the incoming fluid be balanced by a pressure increase at about where

flow deflects in the bent. In other words, a pressure force in the opposite direction is required

to slow down the incoming momentum in the run direction, before diverting the flow. Without

acknowledging such a pressure increase in the bent around the deflection point, momentum of

incoming stream in the run direction (x-direction) could only be balanced by irreversible losses. This

would mean imitation of actual physics of flow curvature by blending its effect within irreversible

losses.

Despite generating two additional unknowns (pT and A313), imagining of the flow channels has

other advantages that is worth the effort:

1. Could help facilitate a better understanding of flow mechanics at the tee.

2. Separate momentum equations can be written for each channel, providing additional equations

for the solution, as opposed to writing momentum balances over the whole tee itself (which,

in theory, should yield similar results).

3. There is no need to account for the contribution to/from branch flow when writing momentum

balances over individual flow channels, while it would have been necessary to do so in the case

of writing momentum balances for the whole tee cell or outgoing edges.

B.2 The Buffer Zone

For practical purposes and as a requirement of the momentum balance around the inlet face of the

tee, it is reasonable to assume that tee densities (one per phase, ρG3 and ρL3) and the tee void

fraction α31 prevail in the tee cell way before the split occurs. By creating an imaginary ‘buffer zone’

before flow splits into channels within the tee, it is possible to convert incoming mass flow based on

upstream phase densities and velocities to mass flow based on phase densities and velocities of tee

cell. Further more, it could also be assumed that all the mass transfer (and momentum transfer

thereof) taking place at the tee occurs within this buffer zone. Nevertheless, this convenience comes

at the price of recognizing a new set of velocities (vG31 and vL31) at the entry point of channels.

Extending flow channels to the inlet face of the tee and ignoring such an imaginary buffer zone

might have helped avoid creating new velocities as unknowns. Yet, it might have as well been

difficult to account for tee densities, void fraction and the momentum balance over the inlet face.

Also, assuming that incoming momentum based on upstream properties would be equal to the

momentum at the end of buffer zone (based on tee densities, and velocity) before the split would

not be correct as can be seen from a single phase compressible flow analysis:

141

mass balance over the buffer zone:

ρ2v2A2 = ρ31v31A31 (B.5)

where A31 = A2;

v31 =ρ2v2

ρ31(B.6)

If ρ2 = 2ρ31 then 2v2 = v31

Hence;

(ρ2v2A2)v2 6= (ρ31v31A2)v31 (B.7)

At this stage, there are a total of 14 unknowns with the inclusion of (vG31 and vL31).

B.3 Simplifying Assumptions

In order not to further complicate the problem, a series of simplifying assumptions are introduced

before commencing the derivation of governing equations:

1. Assuming that imaginary flow channels are adequately short (or the tee volume itself is

adequately small), density and void fraction changes as well as mass transfer (and momentum

transfer thereof), are neglected within the channels (or entirely within the tee).

This implies, due to conservation of mass, change of velocity over a channel is caused by

a change in flow area. Constant void fraction assumption along a channel is more of a

convenience, because of which the bent pipe treatment for branch channel and calculation of

pressure forces in both channels are simplified.

Also, ignoring mass transfer in the channels requires that all the mass transfer supposed to

occur in the tee takes place in the buffer zone, unless completely ignored for the sake of

simplicity.

2. Pressure gradient for the run channel is approximated as p31− p32∼= p31− p4, thus eliminating

p32 as an unknown.

3. Pressure gradient for the branch channel beyond the deflection point is approximated as

pT − p33∼= pT − p6, thus eliminating p33 as an unknown.

Approximating pressure gradients with downstream cell pressures provides the basis for extending

momentum equations written for channels to further include the volume of edge momentum cells

of the tee, over which regular edge momentum balances would have been written. This helps

establishing an improved physical link (1) between outgoing velocities at tee faces and the pressure

gradients by placing velocities actually inside the pressure gradient, and (2) between pressure profiles

of outgoing arms and the tee cell pressures.

142

B.4 Governing Equations

At this point, for the sake of completeness, governing equations are developed without any additional

simplifications to the model with separate void fractions accounted for the arms of the tee. Additional

simplifications are to be introduced should there be a need; either to balance number of available

equations or for the purposes of easier implementation.

First step is to acknowledge all 13 unknowns at a tee: p31, pT , α31, α32, α33, vG3, vL3, vG6, vL6,

vG31, vL31, A312 and A313. T31 of the cell is ignored since it is associated with energy balance by

default.

B.4.1 Mass Balances

Linearly independent mass balance equations that can be written for the system are:

1) Gas phase mass balance for branch

mssG23 −mssG6 = 0

ρG3vG31α33A313 − ρG3vG6α33A6 = 0(B.8)

2) Liquid phase mass balance for branch

mssL23 −mssL6 = 0

ρL3vL31(1− α33)A313 − ρL3vL6(1− α33)A6 = 0(B.9)

3) Gas phase mass balance for run

mssG22 −mssG3 = 0

ρG3vG31α32A312 − ρG3vG3α32A3 = 0(B.10)

4) Liquid phase mass balance for run

mssL22 −mssL3 = 0

ρL3vL31(1− α32)A312 − ρL3vL3(1− α32)A3 = 0(B.11)

5) Gas phase mass balance for buffer zone

mssG2 −mssG31 − Γ3 = 0

ρG2vG2α2A2 − ρG3vG31α31A2 − Γ3 = 0(B.12)

6) Liquid phase mass balance for buffer zone

mssL22 −mssL3 + Γ3 = 0

ρL2vL2(1− α2)A2 − ρL3vL31(1− α31)A2 + Γ3 = 0(B.13)

Please recall that in Sec. B.3 flow channels are assumed to be adequately short enough to ignore

143

mass transfer within: Γchannel ∼= 0. Similarly, transfer within the buffer zone could be ignored

(Γ3∼= 0) for convenience

Important auxiliary relations to recognize as independent equations are:

7)

A2 = A31 = A312 +A313 (B.14)

8)

α31A3 = α32A312 + α33A313 (B.15)

Equation B.15 is a requirement of phasic mass balance in the buffer zone, before the split: 8)

ρG3vG31α31A2 = ρG3vG31α32A312 + ρG3vG31α33A313 (B.16)

Or;

8)

ρL3vL31(1− α31A2) = ρL3vL31(1− α32)A312 + ρL3vL31(1− α33)A313 (B.17)

It is very important to recognize that, despite being written for separate phases, Eqns. B.16 and

B.17 do not constitute linearly independent equations. This is quite clear as both equations collapse

to same Eqn. B.15 should mass flux terms be canceled in both equations.

It should be noted that phasic mass balances over the whole tee do not constitute linearly independent

equations because they can be easily obtained by adding mass balances of the same phase from above

list. Thus, they would not provide additional information to the system. However, for convenience

and purposes of easier implementation, any phase mass balance from above list could be replaced

by a mass balance of the the same phase over the whole tee:

7) Gas phase mass balance over whole tee (Eqns. B.12 + B.16):

ρG2vG2α2A2 − ρG3vG3α32A3 − ρG3vG6α33A6 − Γ3 = 0 (B.18)

8) Liquid phase mass balance over whole tee (Eqns. B.13 + B.17):

ρL2vL2(1− α2)A2 − ρL3vL3(1− α32)A3 − ρL3vL6(1− α33)A6 + Γ3 = 0 (B.19)

B.4.2 Momentum Balances

It is possible to write 6 linearly independent momentum equations. Assuming that length of branch

channel along the run driection is short enough to neglect friction and additional irreversible losses

along this section, friction and any additional irreversible losses are only attributed to the section

of channel along the branch direction. Slightly complicated alternatives include: (1) partitioning

of friction and irreverible losses according to length and (2) accounting for separate friction and

irreverible losses for the run and branch sections of the branch channel. Also, gravitational forces

144

are ignored along both channels.

1) Gas phase momentum balance for branch, in run direction

momG33 −momG6 cos θ + (p31 − pT )α∗33A∗313 + (pT − p6)α∗33A

∗6 cos θ

−fwG33 cos θ − fi33 cos θ − 12k13mom

∗G33 cos θ = 0

(B.20)

2) Liquid phase momentum balance for branch, in run direction

momL33 −momL6 cos θ + (p31 − pT )(1− α∗33)A∗313 + (pT − p6)(1− α∗33)A∗6 cos θ

−fwL33 cos θ + fi33 cos θ − 12k13mom

∗L33 cos θ = 0

(B.21)

3) Gas phase momentum balance for branch, perpendicular to run direction

−momG6 sin θ + (pT − p6)α∗33A∗6 sin θ − fwG33 sin θ − fi33 sin θ

−12k13mom

∗G33 sin θ = 0

(B.22)

4) Liquid phase momentum balance for branch, perpendicular to run direction

−momL6 sin θ + (pT − p6)(1− α∗33)A∗6 sin θ − fwL33 sin θ + fi33 sin θ

−12k13mom

∗L33 sin θ = 0

(B.23)

5) Gas phase momentum balance for run

momG32 −momG4 + (p31 − p4)α∗32A∗3 − fwG32 − fi32 − 1

2k12mom∗G32 = 0 (B.24)

6) Liquid phase momentum balance for run

momL32 −momL4 + (p31 − p4)(1− α∗32)A∗3 − fwL32 + fi32 − 12k12mom

∗L32 = 0 (B.25)

Where:

momG4 = (ρG3vG3α32A3)vG3

momG6 = (ρG3vG6α33A6)vG6

momL4 = (ρL3vL3(1− α32)A3)vL3

momL6 = (ρL3vL6(1− α33)A6)vL6

A∗312∼= A312

A∗313∼= A313

A∗3∼= A3

A∗6∼= A6

α∗32 is a weighted (volumetric) average of α32 and α4 and

α∗33 is a weighted (volumetric) average of α33 and α6 as is the case for a typical momentum balance

equation around an edge.

145

Due to staggered grid and donor cell arrangements momG3 = (ρG2vG2α2A2)vG2. Hence, momG3 has

to be partitioned based on flow areas and void fractions in order to obtain momG32 and momG33

per channel. This is recognized by analysis of mass and momentum partition right before the split

(Eqn. B.16):

mssG31 = mssG32 +mssG33

ρG3vG31α31A2 = ρG3vG31α32A312 + ρG3vG31α33A313

ρG3vG31α31A2 = ρG3vG31α31A2α32A312

α31A2+ ρG3vG31α31A2

α33A313

α31A2

mssG31 = mssG31α32A312

α31A2+mssG31

α33A313

α31A2

Where;

mssG32 = mssG31α32A312

α31A2

mssG33 = mssG31α33A313

α31A2

Similarly,

momG3 = momG32 +momG33

momG3 = momG3α32A312

α31A2+momG3

α33A313

α31A2

Then,

momG32 = momG3α32A312

α31A2(B.26)

momG33 = momG3α33A313

α31A2(B.27)

Where momG3 = (ρG2vG2α2A2)vG2.

Instead of writing momentum balances along orthogonal axes (x-y) of run (primary) direction,

it is equivalent to write momentum balances along orthogonal axes (x′-y′) of branch (secondary)

direction. Then, a secondary equation set written for orthogonal axes of branch could replace the

current set. However, it should be remembered that equation sets are not linearly independent from

one other.

Second set of alternative momentum balances:

1) Gas phase momentum balance for branch, in branch direction

momG33 cos θ −momG6 + (p31 − pT )α∗33A∗313 cos θ + (pT − p6)α∗33A

∗6

−fwG33 − fi33 − 12k13mom

∗G33 = 0

(B.28)

146

2) Liquid phase momentum balance for branch, in branch direction

momL33 cos θ −momL6 + (p31 − pT )(1− α∗33)A∗313 cos θ + (pT − p6)(1− α∗33)A∗6−fwL33 + fi33 − 1

2k13mom∗L33 = 0

(B.29)

3) Gas phase momentum balance for branch, perpendicular to branch direction

momG33 sin θ + (p31 − pT )α∗33A∗313 sin θ = 0 (B.30)

4) Liquid phase momentum balance for branch, perpendicular to branch direction

momL33 sin θ + (p31 − pT )(1− α∗33)A∗313 sin θ = 0 (B.31)

Please note that phasic momentum balances for the run channel could be decomposed into x′

and y′ directions (that is unless x′ = x, depending on the split angle) by taking cos and sin of

all the components in the equations. Yet, in essence, the equations simply would not be linearly

independent, hence not repeated in the second set.

Typically, as it is the case with mass balances, writing momentum balances for the whole tee volume

would not bring any new information to the system because same information is already included

within run and branch momentum balances. For instance, the gas phase momentum balance in the

run direction for the whole tee volume is simply equivalent to mathematical summation of (1) gas

phase momentum balance in run direction for the run and (2) gas phase momentum balance in run

direction for the branch.

It is easier to write overall momentum balances instead of phasic balances for the tee, considering

the complications associated with void fractions multiplying pressure gradients (that is if pressure

forces calculations are to be consistent with the rest of the moodel), alternative momentum balances

can be written for the system are:

1) overall momentum balance for whole tee, in run direction:

summation of Eqns. B.20 + B.21 + B.24 + B.25

momG3 +momL3 − (momG4 +momL4)− (momG6 +momL6) cos θ

+(p31 − pT )A∗313 + (p31 − p4)A∗3 + (pT − p6)A∗6 cos θ

−fwG32 − fwG33 cos θ − 12k12mom

∗G32 −

12k13mom

∗G33 cos θ

−fwL32 − fwL33 cos θ − 12k12mom

∗L32 −

12k13mom

∗L33 cos θ = 0

(B.32)

2) overall momentum balance for whole tee, perpendicular to run direction

147

summation of Eqns. B.22 + B.23

−(momG6 +momL6) sin θ + (pT − p6)A∗6 sin θ − fwG33 sin θ − fwL33 sin θ

−12k13mom

∗G32 sin θ − 1

2k13mom∗L32 sin θ = 0

(B.33)

A second set of alternative momentum balances are as follows:

1) overall momentum balance for whole tee, in branch direction

summation of Eqns. B.28 + B.29 + (B.25 + B.25) cos θ

(momG3 +momL3) cos θ − (momG4 +momL4) cos θ − (momG6 +momL6)

+(p31 − pT )A∗313 cos θ + (p31 − p4)A∗3 cos θ + (pT − p6)A∗6−fwG32 cos θ − fwG33 − 1

2k12mom∗G32 cos θ − 1

2k13mom∗G33

−fwL32 cos θ − fwL33 − 12k12mom

∗L32 cos θ − 1

2k13mom∗L33 = 0

(B.34)

2) overall momentum balance for whole tee, perpendicular to branch direction

summation of Eqns. B.30 + B.31 + (B.25 + B.25) sin θ

(momG3 +momL3) sin θ − (momG4 +momL4) sin θ

+(p31 − pT )A∗313 sin θ + (p31 − p4)A∗3 sin θ

−fwG32 sin θ − 12k12mom

∗G32 sin θ

−fwL32 sin θ − 12k12mom

∗L32 sin θ = 0

(B.35)

Where:

momG3 = (ρG2vG2α2A2)vG2

momL3 = (ρL2vL2(1− α2)A2)vL2

It is also important to note that, because momentum is a vectoral quantity, mathematical summation

of equations in both directions are not necessarily equivalent to vectoral summation of the equations.

In other words, for instance, only vectoral summation of momentum balance in run direction and

momentum balance perpendicular to run direction is mathematically equivalent to momentum

balance in branch direction.

For branch channel, it is a convenient alternative to use momentum equations in the branch and

run directions instead of branch (or run) direction and with its associated perpendicular direction

momentum balance.

So far it has been possible to write more than 13 linearly independent equations to balance number

of unknowns.

148

B.4.3 Parallel Split

When branch is parallel to run (split angle is 0o), momentum conservation orthogonal to run (or

momentum conservation orthogonal to branch if second set of momentum balances are considered)

does not provide any useful information because none of the momentums or forces would have

orthogonal components (sin(0o) = 0), thus equation becomes singular. Also, there is no longer an

actual pT to account for; when branch is parallel to run pT = p31 indeed and pT becomes a dummy

variable.

From the perspective of numerical solution, if branch is parallel to the run, while constructing the

jacobian (coefficient) matrix which hosts numerical derivatives of the residuals (governing equations),

these perpendicular momentum equations for branch are likely to cause jacobian to be singular as

well .

In order to prevent this singularity and retain dummy equations in the system when split angle is

0o, orthogonal momentum balance of a phase could be substituted with a mathematical summation

of momentum balances of the same phase in both directions.

However, when branch is parallel to run, the fact that pT = p31 may pose other requirements for the

dummy equation replacing singular equations. As pT becomes a dummy variable when split angle is

0o, it is necessary to fix its values to p31. This is typically established via a dummy equation that

explicitly sets variable to the desired value and, for the purposes of jacobian matrix solution, that

does not generate a zero when its derivative is taken with respect to the variable:

pT − p31 = 0 (B.36)

B.4.4 Reduction of Unknowns

For isothermal analysis of regular, straight pipe; there are 2 equations solved per CV (phasic mass

balances) and 2 equations solved per face (phasic momentum balances). Therefore, as long as

governing equations permit, it is convenient to express several of the 13 unknowns with explicit

equations involving remaining unknowns of the tee, which shall become the primary unknowns.

This is to reduce the number of equations solved for a tee and match the general scheme of 2 mass

equations per CV and 2 momentum equations per downstream face.

First, using Eqn. B.14:

A312 = A2 −A313 (B.37)

Then, vG31 can be defined separately using flow channel gas mass balances (Eqns. B.8 and B.10):

vG31 =vG3A3

(A2 −A313)

vG31 =vG6A6

A313

149

Then A313 is obtained by setting two expressions of vG31 equal:

A313 = A2vG6A6

vG3A3 + vG6A6(B.38)

Then;

A312 = A2 −A2vG6A6

vG3A3 + vG6A6= A2

vG3A3

vG3A3 + vG6A6(B.39)

And final expressions for vG31 is:

vG31 =vG3A3 + vG6A6

A2(B.40)

Because of constant void fraction (α) assumption, phasic mass balances for the flow channels alone

do not provide information to express α32 and α33 explicitly. Therefore, it is necessary to utilize

momentum balances to determine α32 and α33.

Nevertheless, liquid mass balance equations (Eqns. B.9 and B.11) can be employed to obtain either

vL3 or vL6 explicitly with new definitions of vL31 and A313 at hand.

A better option to express vL31 is to use Eqn. B.9 and the new definition of A313 (Eqn. B.38):

vL31 =vL6

vG6

vG3A3 + vG6A6

A2= vL6

vG31

vG6(B.41)

Using Eqn. B.11 and the definition of A312:

vL31 =vL3

vG3

vG3A3 + vG6A6

A2= vL3

vG31

vG3(B.42)

It is possible to get vL6 by setting both expressions of vL31 equal:

vL6 = vL3vG6

vG3(B.43)

Also, from Eqn. B.15:

α31 =α32A312 + α33A313

A2(B.44)

or

α31 = α32 − (α32 − α33)vG6A6

vG3A3 + vG6A6(B.45)

At this point, 6 unknowns has been defined in terms of other unknowns using 5 (out of 7) mass

balances and 1 auxiliary relation (Eqn. B.14).

Using Eqn. B.33 pT can be defined as:

pT = p6 +momG6 +momL6 + fwG33 + fwL33 + 1

2k13mom∗G32 + 1

2k13mom∗L32

A∗6(B.46)

150

Using Eqn. B.33 p31 can be defined as:

pT = p6 +momG6 +momL6 + fwG33 + fwL33 + 1

2k13mom∗G32 + 1

2k13mom∗L32

A∗6(B.47)

151

Appendix C

Fluid Properties

Nomenclature

Roman letters:

a Attraction parameter [Pam6mol−2orPam6mol−2]

b Co-volume [m3mol−1]

c Component mole fraction [dimensionless]

kij Binary interacting coefficient [dimensionless]

mi A function of accentric factor ω [dimensionless]

p Pressure [kgm−1 s−2]

A A function of (aα) [dimensionless]

B A function of co-volume (b) [dimensionless]

MWβ Molecular weight of phase β [gmol−1]

R Gas constant [8.31447 J K−1mol−1]

T Temperature [K]

V Molar volume [m3mol−1]

V TCβ Volume translation coefficient of phase β [gm3 kg−1mol−1]

Z Compressibility factor [dimensionless]

Greek letters:

α Dimensionless parameter [dimensionless]

ε Pipe roughness [m]

ω Accentric factor [dimensionless]

φβi Fugacity coefficient for component i in phase β [dimensionless]

ρβ Density of phase β [kgm−3]

Ωai Component and EOS specific constant [dimensionless]

Ωbi Component and EOS specific constant [dimensionless]

Subscript:

c Critical property

152

i Individual component

m Mixture

r Reduced property

β Phase β

C.1 Equation of State

An equation of state (EOS) relating the pressure, volume and temperature (PVT) properties of a

mixture is required for the purpose of numerical modeling of the flow. Because we are only concerned

with hydrocarbon (HC) components, the Peng-Robinson EOS is utilized for the computation of

needed state functions in the model.

p =RT

V − bm− (aα)mV 2 + 2bmV − b2m

(C.1)

Attraction parameter (a) is defined as shown:

ai = ΩaiR2T 2

ci

Pci(C.2)

Dimensionless parameter alpha (α) is defined as:

αi = [1 +mi(1− T 0.5ri )]2 (C.3)

Function miis defined as:

mi =

0.374640 + 1.542260ωi − 0.269920ω2

i ωi ≤ 0.49

0.379642 + 1.485030ωi − 0.164423ω2i + 0.016666ω3

i ωi > 0.49(C.4)

Hence;

(aα)i = ΩaiR2T 2

ci

pci[1 +mi(1− T 0.5

ri )]2 (C.5)

For a mixture, (aα)m is computed through the application of quadratic mixing rule:

(aα)m =

n∑i

n∑j

cicj(1− kij)((aα)i(aα)j)0.5 (C.6)

Co-volume parameter (b) is defined as:

bi = ΩbiRTcipci

(C.7)

153

For a mixture, bm is computed through the application of quadratic mixing rule:

bm =n∑1

cibi (C.8)

Law of corresponding states enables the definition of real gas equation with the introduction of Z

factor (compressibility factor):

pV = ZRT (C.9)

The EOS in Eqn. C.1 can be re-arranged in the cubic form by utilizing z-factor (Z) concept of real

gas equation and then solved for the unknown z-factor.

Z3 − (1−B)Z2 + (A− 2B − 3B2)Z − (AB −B2 −B3) = 0 (C.10)

where: A is given by:

A = (aα)p

(RT )2(C.11)

and B is defined as:

B = bp

RT(C.12)

C.2 Density calculation

The phase densities are calculated using an arranged form of the real gas law:

ρβ =pMWβ

ZβRT(C.13)

The density values computed using the z-factors from PR-EOS needs to be corrected utilizing the

volume translation (volume shift) concept:

ρcorrectedβ =ρβMWβ

MWβ − V TCβρβ(C.14)

VTC of a phase is calculated as shown:

V TCm =

n∑1

ciV TCi (C.15)

C.3 Fugacity calculation

Fugacity, a measure of chemical potential, is utilized as an indication of chemical equilibrium. Phases

are in chemical equilibrium when fugacity of each component is same in both phases. Fugacity of

154

component i in phase β is calculated as:

fβi = φβicβip (C.16)

Fugacity coefficient is calculated as a function of parameters of the cubic EOS:

lnφβi =BβiBβ

(Zβ − 1)− ln(Zβ −Bβ)

+Aβ

2√

2Bβ

BβiBβ− 2

N∑j=1

cβj(1− kij)√AβiAβj

ln

(Zβ + (1 +

√2)Bβ

Zβ − (1−√

2)Bβ

)(C.17)

C.4 Enthalpy calculation

Mixture enthalpy is calculated by adding ideal enthalpy to the enthalpy of departure:

hm = hidealm + hdeparturem (C.18)

Departure enthalpy is obtained by:

hdeparturem =1000

MWm

[RT (Z − 1) +

T d(aα)mdT − (aα)m

2√

2bmln

(Z + (

√2 + 1)Bm

Z − (√

2− 1)Bm

)](C.19)

The derivative termd(aα)mdT of Eqn. C.19 is obtained as shown:

d(aα)mdT

=−R2√T

n∑i

n∑j

cicj(1− kij)

[mi

√ΩaiTciPci

(aα)0.5j +mj

√ΩajTcjPcj

(aα)0.5i

](C.20)

Where:

hdeparturem [Jkg−1]d(aα)mdT

[Pa2m6mol−2K−1]

Component ideal enthalpy is evaluated by using the Passut and Danner (1972) correlation:

hideali = cf(APD +BPDT + CPDT

2 +DPDT3 + EPDT

4 + EPDT5)

(C.21)

Where; APD, BPD, CPD, DPD, EPD, FPD are coefficients provided by Passut and Danner (1972)

and:

hideali Component ideal enthalpy [Jkg−1]

cf Conversion factor [2.326000000768900103Jkg−1 −BTUlbm−1]

T Temperature [R]

155

And, finally, mixture ideal enthalpy is obtained as:

hidealm =n∑1

ciMWi

MWmhideali (C.22)

156

Appendix D

Hydrocarbon Data

Lumped hydrocarbon mixture data is converted to SI units from Ayala (2001). Please refer to

Appendix C for nomenclature of this chapter.

Table D.1: Hydrocarbon mixture composition and critical properties

Molar MW pc Tc Vc

Component fraction g/mole kPa K m3/kg

Lumped-1 0.7658 16.20 4588.25414 189.89 6.1286E-03

Lumped-2 0.2240 39.37 4548.67823 350.44 4.6571E-03

Lumped-3 0.0102 130.00 1585.79418 722.22 4.1721E-03

Table D.2: Hydrocarbon mixture Peng-Robinson EOS specific parameters

Component ω Ωa Ωb Parachor

Lumped-1 1.075E-2 4.5724E-1 7.7796E-2 7.6520E+1

Lumped-2 1.342E-1 4.5724E-1 7.7796E-2 1.3424E+2

Lumped-3 1.550E-1 4.5724E-1 7.7796E-2 3.6276E+2

Table D.3: Hydrocarbon mixture binary interacting coefficients

Lumped-1 Lumped-1 Lumped-1

Lumped-1 0.0000 0.0280 0.0280

Lumped-2 0.0280 0.0000 0.0280

Lumped-3 0.0280 0.0280 0.0000

157

Table D.4: Hydrocarbon mixture Passut and Danner (1972) coefficients

Component A B C D E F

Lumped-1 -5.5083 5.6072E-1 2.7943E-4 4.1205E-7 -1.5057E-10 1.9329E-14

Lumped-2 2.6694 2.1301E-1 1.6983E-4 2.3631E-7 -1.2312E-10 1.8461E-14

Lumped-3 30.263 -2.1654E-1 4.5852E-4 -9.7970E-8 1.0465E-11 -3.1320E-16

Vita

Doruk Alp graduated from Yukselis Koleji highschool (Ankara) in 1997. Based on his score in the

nationwide university exam, taken by over 600 thousand students that year, he was placed in the

Petroleum and Natural Gas Engineering program at Middle East Technical University (Ankara).

During undergraduate years, Doruk had internships with Turkish Petroleum Corporation (1999)

and NV Turkse Perenco (2000), briefly held treasurer position of SPE METU student chapter

and worked on ‘Gas Pipeline Deliverability Analysis’ for his graduation project with Dr.Mahmut

Parlaktuna. Doruk received the BS degree with honors in Petroleum and Natural Gas Engineering

in June 2001.

Following his graduation, Mr.Alp was admitted to the MS program in Petroleum and Natural Gas

Engineering at METU. During the course of his study, Doruk was TA for more than 5 different

course (with associated lab and recitation hours) including; fluid mechanics, well logging, natural gas

engineering, rock & reservoir properties, reservoir engineering and engineering graphics. He worked

on numerical modeling of ‘Gas Production from Hydrate Reservoirs’ with Dr.Mahmut Parlaktuna

(METU) and Dr.George Moridis (Lawrence Berkley National Lab. - LBNL) as his advisors and

received the MS degree in Petroleum and Natural Gas Engineering in July 2005.

Mr.Alp was admitted to the PhD program in Petroleum and Natural Gas Engineering at Penn

State in August 2005. He became a graduate assistant to Dr.Luis F. Ayala for the production

process engineering lab and class. Doruk had a reservoir engineering internship with Chevron ETC

(Houston) in 2008.

At the time of writing, Doruk has accepted a reservoir engineering position with Chevron ETC,

pending his graduation.


Recommended