You are on page 1of 193

A PRACTICAL MODEL TO PREDICT GAS HYDRATE FORMATION,

DISSOCIATION AND TRANSPORTABILITY IN OIL AND GAS

FLOWLINES

by

Luis E. Zerpa

c Copyright by Luis E. Zerpa, 2013

All Rights Reserved


A thesis submitted to the Faculty and the Board of Trustees of the Colorado School of Mines

in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Petroleum Engi-

neering).

Golden, Colorado

Date

Signed:
Luis E. Zerpa

Signed:
Dr. Hossein Kazemi
Thesis Advisor

Signed:
Dr. Carolyn A. Koh
Thesis Advisor

Golden, Colorado

Date

Signed:
Dr. William Fleckenstein
Professor and Head
Department of Petroleum Engineering

ii
ABSTRACT

The oil and gas industry is facing very challenging production issues with offshore explorations

in deeper and colder waters. Longer subsea tiebacks will be required to transport hydrocarbon fluids

from the wellhead to production and processing platforms. The formation of solid deposits, such

as gas hydrates, waxes, asphaltenes and scale, may plug the flowlines, preventing production and

generating a safety hazard. The flow assurance of the produced hydrocarbon stream is a technical

discipline that focuses on the design of facilities and procedures for the uninterrupted transport of

reservoir fluids from the reservoir to the point of sale. The rapid formation of gas hydrates, which is

promoted by typical high pressure/low temperature operation conditions in deep subsea facilities,

is considered one of the most challenging flow assurance problems. A transient gas hydrate model,

that predicts when and where hydrate plugs will form in flowlines, will have significant utility for the

flow assurance engineers in the oil and gas industry. The Colorado School of Mines Hydrate Kinetics

model (CSMHyK) was specially designed to predict hydrate formation in oil-dominated systems.

The objective of this research work is to develop a comprehensive hydrate model, extending and

improving the CSMHyK model for the prediction of hydrate formation and transportability in oil,

water and gas-dominated systems. The mechanisms of hydrate formation and transportability in

pipelines is studied through the analysis of experimental data obtained at the Center for Hydrate

Research laboratory of the Colorado School of Mines and two large scale flow loops (ExxonMobil

and Tulsa University flow loops). A set of conceptual pictures is developed to explain the physical

phenomena of gas hydrate formation in flowlines. The mathematical models developed in this work

represent a significant advancement for the prediction of hydrate plugging risk in the pipelines of

oil and gas transport facilities, and can be used as a tool to design flow assurance strategies. These

models improve our capability to predict hydrate formation, by considering dynamic aggregation

phenomena in oil-dominated systems, flow regime transition in high water cut systems, and hydrate

film growth in gas-dominated systems.

iii
TABLE OF CONTENTS

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

NOMENCLATURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xviii

LIST OF ABBREVIATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxiv

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxvi

DEDICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxviii

CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Flow Assurance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.1.1 Fluid characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.2 Steady-state thermal-hydraulic calculations . . . . . . . . . . . . . . . . . . . . 9

1.1.3 Transient flow thermal-hydraulic calculations . . . . . . . . . . . . . . . . . . 13

1.1.4 Final system design and operation procedures for flow assurance . . . . . . . 14

1.2 Colorado School of Mines Hydrate Kinetics Model (CSMHyK) . . . . . . . . . . . . 15

1.2.1 Kinetics model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.2.2 Transport model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

1.2.3 Cold flow model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.3 Thesis objectives and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

1.4 Structure of this thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

CHAPTER 2 THE ROLE OF FLUID-FLUID AND FLUID-SOLID INTERFACE ON


HYDRATE PLUG FORMATION . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.1 Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

iv
2.1.1 Effects of surfactants in gas hydrate formation . . . . . . . . . . . . . . . . . 31

2.1.2 Effect of surfactants on contact angle, wettability, and cohesion/adhesion


forces of hydrate particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.2 Emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.2.1 Hydrate formation in emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . 41

2.2.2 Effects of hydrates on emulsion stability . . . . . . . . . . . . . . . . . . . . . 42

2.2.3 Agglomeration mechanism of hydrate particles . . . . . . . . . . . . . . . . . 43

2.2.4 Formulation dependence of hydrate antiagglomeration agents . . . . . . . . . 45

2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

CHAPTER 3 MULTIPHASE FLOW IN PIPELINES AND HYDRATE FORMATION . . . 48

3.1 A review of single-phase flow concepts . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.2 Multiphase flow concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.3 Multiphase flow modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.3.1 Homogeneous fluid model approach . . . . . . . . . . . . . . . . . . . . . . . . 54

3.3.2 Separated-flow model approach . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.3.3 Transient multiphase flow in pipe modeling approach . . . . . . . . . . . . . . 56

3.4 Effect of hydrate formation on multiphase flow . . . . . . . . . . . . . . . . . . . . . 62

3.4.1 Hydrodynamic slug flow model . . . . . . . . . . . . . . . . . . . . . . . . . . 63

3.4.2 Extension of a simple hydrodynamic slug flow model for transient hydrate
kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.4.2.1 Transient hydrate kinetics model . . . . . . . . . . . . . . . . . . . . 69

3.4.2.2 Mixture energy conservation equation . . . . . . . . . . . . . . . . . 70

3.4.3 Example of application of the three-phase hydrodynamic slug flow model . . 71

3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

CHAPTER 4 FLOW LOOP EXPERIMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . 77

v
4.1 ExxonMobil flow loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.2 Tulsa University flow loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

CHAPTER 5 MATHEMATICAL MODEL FOR HYDRATE FORMATION . . . . . . . . . 84

5.1 Model for hydrate formation in oil-dominated systems . . . . . . . . . . . . . . . . . 84

5.1.1 Description of dynamic particle cohesion force model . . . . . . . . . . . . . . 86

5.1.2 Implementation of the model into a transient multiphase flow simulator . . . 89

5.1.3 Validation of the model against flow loop data . . . . . . . . . . . . . . . . . 91

5.2 Model for hydrate formation in water-dominated systems . . . . . . . . . . . . . . . 97

5.2.1 Description of the model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.2.2 Implementation of the model into a transient multiphase flow simulator . . . 106

5.2.3 Validation of the model against flow loop data . . . . . . . . . . . . . . . . . 107

5.3 Model for hydrate formation in gas-dominated systems . . . . . . . . . . . . . . . . . 112

5.3.1 Description of the model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

5.3.2 Implementation of the model into a transient multiphase flow simulator . . . 116

5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

CHAPTER 6 APPLICATION OF THE HYDRATE MATHEMATICAL MODEL . . . . . . 119

6.1 Estimating hydrate plugging risk in oil-dominated systems . . . . . . . . . . . . . . . 119

6.1.1 Offshore well hydrate risk assessment . . . . . . . . . . . . . . . . . . . . . . . 120

6.1.2 Well restart procedure optimization . . . . . . . . . . . . . . . . . . . . . . . 130

6.2 Estimating hydrate plugging risk in water-dominated systems . . . . . . . . . . . . . 134

6.2.1 Uninsulated case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

6.2.2 Thermal insulated case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.3 Hydrate film growth in a gas-dominated horizontal pipeline . . . . . . . . . . . . . . 139

vi
6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

CHAPTER 7 SUMMARY AND CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . 143

CHAPTER 8 SUGGESTIONS FOR FUTURE RESEARCH . . . . . . . . . . . . . . . . . . 146

REFERENCES CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

APPENDIX - DERIVATION OF THE MIXTURE ENERGY CONSERVATION


EQUATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

vii
LIST OF FIGURES

Figure 1.1 Wax deposit on a pipe wall narrowing conduit diameter and modifying wall
roughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

Figure 1.2 Example of asphaltene molecular structures . . . . . . . . . . . . . . . . . . . . . 5

Figure 1.3 Common gas hydrate structures; adapted from Sloan (2011, p. 3). . . . . . . . . 8

Figure 1.4 Pressure versus temperature diagram for a hydrocarbon–water system;


adapted from Sloan and Koh (2008a, p. 198). . . . . . . . . . . . . . . . . . . . . 9

Figure 1.5 Pressure versus temperature diagram showing the WAT curve (green dashed
curve), upper and lower asphaltene precipitation loci (red dash-dot curves)
and hydrate equilibrium curve (blue continuous curve) from the
characterization of reservoir fluids; adapted from Ayers (2003). . . . . . . . . . 10

Figure 1.6 Pressure versus temperature diagram showing a typical subsea flowline
operation conditions curve, the WAT curve, the lower asphaltene precipitation
locus and hydrate equilibrium curves; adapted from Ayers (2003). . . . . . . . . 11

Figure 1.7 Pressure versus temperature diagram showing a typical subsea flowline
operation conditions curve and hydrate equilibrium curves at different
methanol concentrations; adapted from Sloan and Koh (2008c, p. 646). Points
on the flowline operation conditions curve indicate distance from wellhead in
kilometers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Figure 1.8 Conceptual model for hydrate formation in oil-dominated multiphase flow
systems consisting of gas, oil, and water; adapted from Turner (2005), with
input from J. Abrahamson (U. Canterbury, Christchurch, NZ). . . . . . . . . . 15

Figure 1.9 Bishnoi’s laboratory data used to regress the intrinsic kinetic rate constants . . 17

Figure 1.10 Universal curve of turbulent flow sub-ranges in terms of Weber and Reynolds
dimensionless numbers that affect mean droplet size in an emulsion, and
experimental data obtained in an autoclave mixing cell and a large scale flow
loop, using different crude oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

Figure 1.11 Illustration of a water droplet with a surrounding hydrate shell showing the
boundary layer around the spherical particle. . . . . . . . . . . . . . . . . . . . 20

Figure 1.12 Conceptual picture of the BP-Statoil-SINTEF stabilized flow concept


(redrawn from Davies, 2009). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

viii
Figure 1.13 Illustration of a hydrate particle with a surrounding water film showing the
boundary layer around the spherical particle. . . . . . . . . . . . . . . . . . . . 23

Figure 2.1 Conceptual model for hydrate formation in a multiphase flow systems
consisting of water, oil, and gas; adapted from Turner (2005). . . . . . . . . . . 27

Figure 2.2 Schematic of an anionic surfactant molecule, identifying the negatively


charged hydrophilic group and the lipophilic group. . . . . . . . . . . . . . . . . 27

Figure 2.3 Adsorption of a surfactant molecule at: (a) gas/liquid interface, (b)
liquid/liquid interface, (c) liquid/nonpolar-solid interface, (d)
liquid/polar-solid interface, and (e) liquid/polar-solid interface with increased
cationic surfactant concentration. . . . . . . . . . . . . . . . . . . . . . . . . . . 29

Figure 2.4 Variation of surface tension and electrical conductivity as a function of


surfactant concentration, indicating the discontinuity at the CMC. . . . . . . . 30

Figure 2.5 Association of anionic surfactant molecules forming normal micelles in an


aqueous phase and inverse micelles in an oleic (oil) phase. . . . . . . . . . . . . 31

Figure 2.6 Bicontinuous structure of microemulsions formed by associating both types of


micelles and resulting in a zero curvature surfactant film that separates oil
and water zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

Figure 2.7 Balance of interfacial tensions among the three phases at contact point,
defining contact angle of the liquid with the solid surface (left: solid is oil-wet;
right: solid is water-wet). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

Figure 2.8 Summary of cohesion forces between cyclopentane hydrates particles in the
presence of Caratinga crude oil, measured by Dieker et al. (2009). . . . . . . . . 35

Figure 2.9 Mechanisms of emulsion stability (adapted from Salager, 2000a). . . . . . . . . 36

Figure 2.10 Schematic representation of phase behavior described by Winsor for different
values of the ratio of net interaction energy of the surfactant with the oil
phase to the net interaction energy of the surfactant to the aqueous phase . . . 38

Figure 2.11 Effect of formulation variables on type and properties of emulsions . . . . . . . 40

Figure 2.12 Diagram of generalized formulation-composition bidimensional map . . . . . . . 41

Figure 2.13 PVM images of 68 vol% of water dispersed in crude oil: a) initial W/O
emulsion, b) hydrate formation around water droplets, c) large hydrate
agglomerates during hydrate dissociation, c) w/O/W emulsion after hydrate
dissociation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

Figure 3.1 Flow regimes for gas/liquid flow in a horizontal pipe p. 167]Brennen:2005. . . . 52

ix
Figure 3.2 Flow regimes for gas/liquid flow in a vertical pipe p. 170]Brennen:2005. . . . . 52

Figure 3.3 Flow regimes for slurry (solid particles of sand dispersed in water) flow in a
horizontal pipe p. 169]Brennen:2005. . . . . . . . . . . . . . . . . . . . . . . . . 53

Figure 3.4 Illustration of basic mass fields in a transient multiphase flow model, including
gas, oil and water as continuous phases, and dispersed phases within
continuous phases (i.e., oil and water droplets dispersed in continuous gas,
water droplets and gas bubbles dispersed in continuous oil, and oil droplets
and gas bubbles dispersed in continuous water). . . . . . . . . . . . . . . . . . . 61

Figure 3.5 Pipe segment divided into equally spaced sections, showing node numbering
convention. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

Figure 3.6 Liquid holdup distribution along the pipeline length at four different times:
(a) t = 0.04 s, (b) t = 100 s, (c) t = 500 s, (d) t = 700 s. . . . . . . . . . . . . . 66

Figure 3.7 Conceptual picture used to define the phases and multiphase flow concepts. . . 67

Figure 3.8 Cross section of pipe with stratified smooth flow regime. . . . . . . . . . . . . . 70

Figure 3.9 Hydrate equilibrium curve for pure methane (V) and liquid water (L)
calculated with CSMGem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

Figure 3.10 Distribution of water (continuous blue line) and hydrate (segmented green
line) holdup along the pipeline length at four different times, (a) 40 s, (b)
3230 s, (c) 4000 s, (d) 6000 s, showing the transition from stratified to slug
flow upon hydrate formation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

Figure 3.11 Comparison of predicted results and experimental data of fluid and ambient
temperature as function of time. . . . . . . . . . . . . . . . . . . . . . . . . . . 75

Figure 4.1 Schematic of ExxonMobil flow loop. . . . . . . . . . . . . . . . . . . . . . . . . 78

Figure 4.2 Schematic of University of Tulsa hydrate flow loop. . . . . . . . . . . . . . . . . 82

Figure 5.1 Conceptual model for hydrate formation in oil-dominated multiphase flow
systems consisting of gas, oil, and water; adapted from Turner (2005), with
input from J. Abrahamson (U. Canterbury, Christchurch, NZ). . . . . . . . . . 85

Figure 5.2 Illustration of the capillary liquid bridge between two hydrate particles . . . . . 87

Figure 5.3 Comparison of experimental measurements (points) and model prediction


(solid line) of cyclopentane hydrate cohesive forces as function of contact time . 89

x
Figure 5.4 Illustration of mass fields considered in the oil-dominated hydrate formation
model coupled with the OLGA R
transient multiphase flow simulator,
including gas and oil as continuous phases, water and hydrate is assumed to
be dispersed in the oil phase, and oil and water may be dispersed as droplets
in the continuous gas phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

Figure 5.5 Plot of hydrate volume fraction as function of time, comparing experimental
data and oil-dominated model prediction for an ExxonMobil flow loop
constant pressure experiment with Conroe oil, 70% liquid loading, 37% water
cut, and 2.3 m/s fluid mixture velocity (Test # XoM-2004-2). . . . . . . . . . . 92

Figure 5.6 Plot of cumulative gas injection into the loop as function of time, comparing
experimental data and oil-dominated model prediction for a University of
Tulsa flow loop constant pressure experiment with Caratinga oil, 50% liquid
loading, 25% water cut, and 1.2 m/s fluid mixture velocity (Test # TU-2004-3). 92

Figure 5.7 Plot of flow loop pressure as function of time, comparing experimental data
and oil-dominated model prediction for an University of Tulsa flow loop
constant volume experiment with Troika oil, 50% liquid loading, 25% water
cut, and 1.2 m/s fluid mixture velocity (Test # TU-2004-4). . . . . . . . . . . . 93

Figure 5.8 Plot of fluid and ambient temperature as function of time, comparing
experimental data and oil-dominated model prediction for an ExxonMobil
flow loop experiment with Conroe oil, 50% liquid loading, 15% water cut, and
1 m/s fluid mixture velocity (Test # XoM-2011-9). . . . . . . . . . . . . . . . . 94

Figure 5.9 Plot of phases volume fraction as function of time, for an ExxonMobil flow
loop experiment with Conroe oil, 50% liquid loading, 15% water cut, and 1
m/s fluid mixture velocity. The estimated experimental hydrate volume
fraction is also shown for comparison with the model prediction (Test #
XoM-2011-9). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

Figure 5.10 Plot of fluid and ambient temperature as function of time, for an ExxonMobil
flow loop experiment with Conroe oil, 75% liquid loading, 15% water cut, and
1 m/s fluid mixture velocity (Test # XoM-2011-14). . . . . . . . . . . . . . . . 96

Figure 5.11 Plot of phases volume fraction as function of time, for an ExxonMobil flow
loop experiment with Conroe oil, 75% liquid loading, 15% water cut, and 1
m/s fluid mixture velocity. The estimated experimental hydrate volume
fraction is also shown for comparison with the model prediction (solid lines)
(Test # XoM-2011-14). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Figure 5.12 Plot of hydrate volume fraction as function of time, comparing experimental
data with predictions using the oil-dominated hydrate formation model for an
ExxonMobil flow loop experiment with King Ranch condensate, 50% liquid
loading, 25% water cut, and 1 m/s fluid mixture velocity (Test #
XoM-2011-25). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

xi
Figure 5.13 Plot of hydrate volume fraction as function of time, comparing experimental
data and oil-dominated model prediction for an ExxonMobil flow loop
experiment with Conroe oil, 50% liquid loading, 75% water cut, and 1 m/s
fluid mixture velocity (Test # XoM-2011-23). . . . . . . . . . . . . . . . . . . . 98

Figure 5.14 Conceptual picture for hydrate formation in water-dominated systems


consisting of gas and water phases. . . . . . . . . . . . . . . . . . . . . . . . . . 98

Figure 5.15 Schematic illustration of gas-liquid flow regimes in horizontal pipe; adapted
from Brennen (2005). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

Figure 5.16 Schematic illustration of slurry flow (solids dispersion in liquid phase) regimes;
adapted from Brennen (2005) and Kitanovski and Poredos (2002). . . . . . . . 100

Figure 5.17 Typical behavior of measured pressure drop observed after hydrate formation
onset in high water cut flow loop experiments; adapted from Joshi et al.
(2011a). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

Figure 5.18 Conceptual model for hydrate formation mechanism in water-dominated


systems consisting of an aqueous phase, a gaseous phase and solid hydrate
particles, showing transition of flow regimes with increasing hydrate volume
fraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

Figure 5.19 Correlation of transition hydrate concentration (Φtransition ) as function of


Reynolds number (Re) and liquid loading (LL) in vol.%, developed from
experimental data obtained at the ExxonMobil and University of Tulsa flow
loops at constant mixture velocities; presented in the August 2012 CSM
hydrate consortium meeting by Giovanny Grasso. . . . . . . . . . . . . . . . . . 102

Figure 5.20 Cross section of pipe with stratified smooth flow regime. . . . . . . . . . . . . . 103

Figure 5.21 Illustration of gas bubbles entrainment in the liquid slug, due to energy
dissipation in the front of the liquid slug; adapted from Andreussi and
Bendiksen (1989). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

Figure 5.22 Illustration of mass fields considered in the water-dominated hydrate


formation model coupled with the OLGA R
transient multiphase flow
simulator, including gas and water as continuous phases, and hydrate and gas
as dispersed phases in the water continuous phase. . . . . . . . . . . . . . . . . 106

Figure 5.23 Plot of fluid and ambient temperature as function of time, comparing
experimental data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 1 m/s fluid mixture velocity (Test #
XoM-2011-1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

xii
Figure 5.24 Plot of gas injected into flow loop as function of time, comparing experimental
data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 1 m/s fluid mixture velocity (Test #
XoM-2011-1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

Figure 5.25 Plot of hydrate volume fraction as function of time, comparing experimental
data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 1 m/s fluid mixture velocity (Test #
XoM-2011-1), also showing the calculated Φtransition . . . . . . . . . . . . . . . . 110

Figure 5.26 Plot of fluid and ambient temperature as function of time, comparing
experimental data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 2.5 m/s fluid mixture velocity (Test #
XoM-2010-2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

Figure 5.27 Plot of gas injected into flow loop as function of time, comparing experimental
data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 2.5 m/s fluid mixture velocity (Test #
XoM-2010-2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

Figure 5.28 Plot of hydrate volume fraction as function of time, comparing experimental
data with predictions using the model for hydrate formation in
water-dominated systems for an ExxonMobil flow loop experiment with 50%
liquid loading, 100% water cut, and 2.5 m/s fluid mixture velocity (Test #
XoM-2010-2), also showing the calculated Φtransition . . . . . . . . . . . . . . . . 111

Figure 5.29 Comparison of experimental data with predictions using the model for
hydrate formation in water-dominated systems for Test XoM-2011-21 with
Conroe oil, 50% liquid loading, 90% water cut, and 1.75 m/s fluid mixture
velocity: (a) ambient and fluid temperature as function of time, (b) hydrate
volume fraction as function of time, including the calculated Φtransition . . . . . 112

Figure 5.30 Comparison of experimental data with predictions using the model for
hydrate formation in water-dominated systems for Test XoM-2011-24 with
Conroe oil, 50% liquid loading, 75% water cut, and 2.5 m/s fluid mixture
velocity: (a) ambient and fluid temperature as function of time, (b) hydrate
volume fraction as function of time, including the calculated Φtransition . . . . . 113

Figure 5.31 Conceptual picture for hydrate formation in gas-dominated systems consisting
of gas, some condensate and water; adapted from Lingelem et al. (1994). . . . . 115

xiii
Figure 5.32 Illustration of the mass fields considered in the gas-dominated hydrate
formation model coupled with the OLGA R
transient multiphase flow
simulator, including a gas phase saturated with water and hydrate film
deposited on the pipe wall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

Figure 6.1 Geometry of well, flowline, and riser, based on typical geometries from the
Caratinga field located at the Campos Basin in Brazil. . . . . . . . . . . . . . . 120

Figure 6.2 Diagram of pressure versus temperature showing flowline operation conditions
during steady-state flow and hydrate equilibrium curves at different ethanol
concentrations. Points on flowline operation conditions curve indicate distance
from wellhead. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

Figure 6.3 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature corresponding to the
steady-state flow predicted by the oil-dominated gas hydrate formation model
for Case 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

Figure 6.4 Fluid distribution in terms of volume fraction along the pipeline length,
corresponding to the steady-state flow predicted by the oil-dominated gas
hydrate formation model for Case 1. . . . . . . . . . . . . . . . . . . . . . . . . 124

Figure 6.5 Hydrate slurry relative viscosity along the pipeline length, corresponding to
the steady-state flow predicted by the oil-dominated gas hydrate formation
model for Case 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

Figure 6.6 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature corresponding to the
steady-state flow predicted by the oil-dominated gas hydrate formation model
for Case 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

Figure 6.7 Fluid distribution in terms of volume fraction along the pipeline length,
corresponding to the steady-state flow predicted by the oil-dominated gas
hydrate formation model for Case 2. . . . . . . . . . . . . . . . . . . . . . . . . 126

Figure 6.8 Hydrate slurry relative viscosity along the pipeline length, corresponding to
the steady-state flow predicted by the oil-dominated gas hydrate formation
model for Case 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

Figure 6.9 Performance measures (∆Pf lowline , Φhyd , and µr ) used to establish the
hydrate plugging risk as a function of ethanol concentration, predicted by the
oil-dominated gas hydrate formation model for Case 2. . . . . . . . . . . . . . . 127

Figure 6.10 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature corresponding to the
steady-state flow predicted by the oil-dominated gas hydrate formation model
for Case 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

xiv
Figure 6.11 Fluid distribution in terms of volume fraction along the pipeline length,
corresponding to the steady-state flow predicted by the oil-dominated gas
hydrate formation model for Case 3. . . . . . . . . . . . . . . . . . . . . . . . . 128

Figure 6.12 Hydrate slurry relative viscosity along the pipeline length, corresponding to
the steady-state flow predicted by the oil-dominated gas hydrate formation
model for Case 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

Figure 6.13 Performance measures (∆Pf lowline , Φhyd , and µr ) used to establish the
hydrate plugging risk as a function of ethanol concentration, predicted by the
oil-dominated gas hydrate formation model for Case 3. . . . . . . . . . . . . . . 130

Figure 6.14 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature, after 50 hours of shut-in,
predicted by the oil-dominated gas hydrate formation model for Case 2
without ethanol. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

Figure 6.15 Fluids distribution in terms of volume fraction along the pipeline length, after
50 hours of shut-in, predicted by the oil-dominated gas hydrate formation
model for Case 2 without ethanol. . . . . . . . . . . . . . . . . . . . . . . . . . . 131

Figure 6.16 Valve opening schedule used during the restart of production. . . . . . . . . . . 132

Figure 6.17 Evolution of the production restart in terms of pressure drop along the
pipeline (left plot) and hydrate volume fraction (right plot) at a point in the
middle of the pipeline with respect to time, predicted by the oil-dominated
hydrate formation model for Case 2 with 20 wt% ethanol injected before the
shut-in and during the restart. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Figure 6.18 Performance measures — max {∆Pf lowline (t)}; max {Φhyd (t)}; and
max {µr (t)} — used to evaluate the hydrate plugging risk as a function of
ethanol concentration in transient restart operations, predicted by the
oil-dominated hydrate formation model for Case 2. . . . . . . . . . . . . . . . . 134

Figure 6.19 Diagram of pressure versus temperature showing flowline operation conditions
during steady-state flow, and hydrate equilibrium curve, predicted by the
water-dominated hydrate formation model for the uninsulated case. . . . . . . . 135

Figure 6.20 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature corresponding to the
steady-state flow, predicted by the water-dominated hydrate formation model
for the uninsulated case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

Figure 6.21 Fluid distribution in terms of volume fraction along the pipeline length,
corresponding to the steady-state flow, predicted by the water-dominated
hydrate formation model for the uninsulated case. . . . . . . . . . . . . . . . . . 136

xv
Figure 6.22 Hydrate volume fraction along the pipeline length, showing the calculated
Φtransition as plugging criterion, corresponding to the steady-state flow
predicted by the water-dominated hydrate formation model for the
uninsulated case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

Figure 6.23 Temperature distribution along the pipeline length, showing system
temperature and hydrate equilibrium temperature corresponding to the
steady-state flow, predicted by the water-dominated hydrate formation model
for the thermal insulated case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

Figure 6.24 Fluid distribution in terms of volume fraction along the pipeline length,
corresponding to the steady-state flow, predicted by the water-dominated
hydrate formation model for the thermal insulated case. . . . . . . . . . . . . . 138

Figure 6.25 Hydrate volume fraction along the pipeline length, showing the calculated
Φtransition as plugging criterion, corresponding to the steady-state flow
predicted by the water-dominated hydrate formation model for the thermal
insulated case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

Figure 6.26 Plot of gas hydrate deposit thickness along the pipeline length at 30, 60 and
100 days, showing an increase in deposit thickness with time, predicted by the
hydrate formation model for gas-dominated systems. . . . . . . . . . . . . . . . 140

Figure 6.27 Plot of system pressure along pipeline length, showing an increase in inlet
pressure with time required to maintain a constant mass flow rate at the inlet
of the pipeline, predicted by the hydrate formation model for gas-dominated
systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

Figure 6.28 Plot of hydrate deposit thickness along the pipeline length at 100 days, for
different levels of water saturation of the gas phase; 150, 200 and 300 ppm of
water in the gas phase, predicted by the hydrate formation model for
gas-dominated systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

xvi
LIST OF TABLES

Table 3.1 Classification of empirical correlations for pressure drop prediction in


multiphase flow based on the homogeneous fluid model . . . . . . . . . . . . . . . 55

Table 3.2 Input parameters for hydrodynamic slug flow model. . . . . . . . . . . . . . . . . 66

Table 3.3 Input parameters for three-phase (gas/liquid water/hydrate) hydrodynamic slug
flow model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

Table 4.1 Experiments performed in the ExxonMobil flow loop in Fall 2010. . . . . . . . . . 79

Table 4.2 Experiments performed in the ExxonMobil flow loop in Fall 2011. . . . . . . . . . 80

Table 4.3 Additional experimental data from the ExxonMobil flow loop . . . . . . . . . . . . 80

Table 4.4 Available experimental data from the University of Tulsa hydrate flow loop . . . . 83

Table 5.1 Experiments from the ExxonMobil flow loop selected for predictions using the
water-dominated hydrate formation model. . . . . . . . . . . . . . . . . . . . . . . 108

Table 5.2 Summary of prediction error of hydrate volume fraction at the hydrate
transition concentration (Φtransition ), for the ten high water cut cases studied. . . 113

Table 6.1 Properties and composition of Caratinga crude oil . . . . . . . . . . . . . . . . . . 121

xvii
NOMENCLATURE

ACO . . . . . . . . . . . . . . . . . . . net interaction energy of the surfactant with the oil phase

ACW . . . . . . . . . . . . . . . . net interaction energy of the surfactant with the aqueous phase

Ai . . . . . . . . . . . . . . . . . . . . . . . . . . cross sectional area of solid hydrate bridge (m2 )

Ak . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe cross sectional area occupied by phase k

Ap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe cross sectional area (m2 )

As . . . . . . . . . . . . . . . . . . surface area between water and hydrocarbon phases (m2 /m3 )

aT . . . . . . . . . . . . . . . . . . . . . . temperature coefficient of optimum salinity (ln(S)/K)

Cbulk . . . . . . . . . . . . . . . . . . . . . hydrate guest concentration in the bulk phase (kg/m3 )

Ceq . . . . . . . . hydrate guest concentration in water phase in the presence of hydrate (kg/m3 )

Cp,k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . heat capacity of phase k (J/kg/K)

Csol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . gas solubility in water (molar fraction)

Csol−hyd . . . . . . . . . gas concentration in water in equilibrium with hydrates (molar fraction)

d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . immersion depth of liquid bridge (m)

d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe internal diameter (m)

dA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate agglomerate diameter (m)

DA . . . . . . . . . . . . . . . . . . . diffusivity of gas molecules through the hydrate shell (m2 /s)

d¯D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . mean droplet diameter (m)

db,32 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sauter mean diameter of gas bubble (m)

db,max . . . . . . . . . . . . . . . . . . . . . . . . . . . maximum stable gas bubble diameter (m)

Dgas−water . . . . . . . . . . . . . . . . . . . . . . . . diffusivity coefficient of gas in water (m2 /s)

dP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate particle diameter(m)

xviii
ek . . . . . . . . . . . . . . . . . . . . . . . . . . internal energy per unit mass of phase k (J/kg)

f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . dimensionless friction factor

f . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . fractal dimension

Fa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . inter-particle cohesion force (mN)

f (A) . . . . . . . . . . . . . . . . . . . . . . . . . . . . function of alcohol type and concentration

fi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . interfacial friction factor

FIP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . inter-particle cohesion force (mN)

FS . . . . . . . . . . . . . . . . . . . . . . . . inter-particle cohesion force due to sintering (mN)

fslv . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . generic three phase free energy term

g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . acceleration of gravity (m/s2 )

h . . . . . . . . . . . . . . . . . . . . . . . . . . . convective heat transfer coefficient (J/m2 /K/s)

h . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . elevation (m)

H . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . total height of liquid bridge (m)

Hk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . holdup of phase k

hk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . enthalpy of phase k (J/kg)

HLs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . liquid holdup in the slug

HT Chyd deposit . . . . . . . . . . . . . . . heat transfer coefficient of hydrate deposit (W/m2 /K)

K . . . . . . . . . . . . . . slope of the logarithm of the optimum salinity as a function of ACN

k1 . . . . . . . . . . . . . . . . . . . . . . . . . . . intrinsic kinetic rate constant 1 (kg/m2 /s/K)

k2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . intrinsic kinetic rate constant 2 (K)

kcomp . . . . . . . . . . . . . . . . . . . . . thermal conductivity of the hydrate shell (J/s/m/K)

khyd . . . . . . . . . . . . . . . . . . . . . thermal conductivity of the hydrate phase (J/s/m/K)

kmass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . mass transfer coefficient (m/s)

kwater . . . . . . . . . . . . . . . . . . . . . . thermal conductivity of the water phase (J/s/m/K)

xix
LL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . liquid loading (vol.%)

mgas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . mass of gas (kg)

Nbubbles . . . . . . . . . . . . . . . . . . . . . . . number of gas bubbles per unit volume (1/m3 )

nH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydration number (kg of water/ kg of gas)

NSC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Schmidt number

p . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pressure (Pa)

Pi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . wetted perimeter of the interphase (m)

Pk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . wetted perimeter of phase k (m)

q . . . . . . . . . . . . . . . . . . . . . . heat transfer through pipe wall per unit volume (J/m3 )

q̇ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe wall heat flux (J/m2 /s)

Qk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . volumetric flow rate of phase k (m3 /s)

r . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe radius (m)

R . . . . . . . . . Winsor ratio of net interaction energies of surfactant to aqueous and oil phases

R∗ . . . . . . . . . . . . . . . . . . . . . . . . . . . . harmonic mean radius of a particle pair (m)

Re . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reynolds number

ReLs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . liquid slug Reynolds number

rp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate particle radius (m)

Rs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . gas mass fraction

rw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . radius of water core (m)

S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . salinity (gr NaCl/100 cm3 of aqueous phase)

SE,hyd . . . . . . . . . . . . . energy source term related to exothermic hydrate reaction (J/kg/s)

sk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . entropy of phase k (J/kg)

Sk . . . . . . . . . . . . . . . . . . . . . . . . . . external source/sink term of phase k (kg/s/m3 )

Sw/s . . . . . . . . . . . . . . . . spreading coefficient of a water drop on a solid surface (mN/m)

xx
t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . time (s)

T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . temperature (K)

tcontact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate particle contact time (s)

Thyd eq . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate equilibrium temperature (K)

Tsystem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . system temperature (K)

u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . kinetic rate correction factor

v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . velocity (m/s)

V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . volume (m3 )

v0D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . fall velocity of droplets (m/s)

vk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . linear velocity of phase k (m/s)

vL∗ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . characteristic liquid phase velocity (m/s)

vm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . mixture velocity (m/s)

vo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . drift-flux velocity (m/s)

vs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . slip velocity (m/s)

vSk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . superficial velocity of phase k (m/s)

W e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Weber number

W ecrit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . critical Weber number

x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . axial coordinate along the pipe length (m)

X . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lockhart-Martinelli parameter

Xgas hyd . . . . . . . . . . . . . . . . . . . . molar fraction of hydrate guest in the hydrate phase

Greek Letters

α. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . liquid bridge embracing angle (degrees)

α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . wetted angle (rad)

αs . . . . . . . . . . . . . . . . . . . . . . . . . . characteristic parameter of nonionic surfactant

xxi
β . . . . . . . . . . . . . . mixture Joule-Thomson coefficient multiplied by mixture Cp (m3 /kg)

γ̇ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . fluid flow shear rate (1/s)

∆hhyd . . . . . . . . . . . . . . . . . . . . . . . . . enthalpy change of hydrate formation (J/kg)

∆Tsub . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . subcooling (K)

δ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate shell thickness (m)

δhyd deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate deposit thickness (m)

ε . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . pipe surface roughness (m)

ε . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . porosity of the hydrate shell

ηk . . . . . . . . . . . . . . . . . . . . . . . . . . . Joule-Thomson coefficient of phase k (K/Pa)

θ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . contact angle (degrees)

θ . . . . . . . . . . . . . . . . . . . . pipe angle of inclination with respect to horizontal (degrees)

θP . . . . . . . . . . . . . . . . . . . wetting angle of water on hydrate particle surface (degrees)

µk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viscosity of phase k (Pa.s)

µ◦o . . . . . . . . . . . . . . . . . . Standard chemical potential of the surfactant in the oil phase

µr . . . . . . . relative viscosity of the hydrate slurry to the viscosity of the continuous oil phase

µ◦w . . . . . . . . . . . . . . . . Standard chemical potential of the surfactant in the water phase

ρ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . density (kg/m3 )

ρk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . density of phase k (kg/m3 )

σ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . interfacial tension (mN/m)

σs . . . . . . . . . . . . . . . . . . . . . . . . . . . . characteristic parameter of ionic surfactant

τi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . interfacial shear stress (Pa)

τt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate tensile strength (MPa)

τw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . wall shear stress (Pa)

τw,k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . wall shear stress of phase k (Pa)

xxii
υk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . specific volume of phase k (m3 /kg)

Φ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . hydrate particle volume fraction

Φef f . . . . . . . . . . . . . . . . effective volume fraction of the agglomerated hydrate particles

φ2GO . . . . . . . . . . . . . . . . two-phase correction factor for gas-only frictional pressure drop

φhyd deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . porosity of hydrate deposit

φ2LO . . . . . . . . . . . . . . . two-phase correction factor for liquid-only frictional pressure drop

Φmax . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . maximum particle packing fraction

Φtransition . . . . . . . . hydrate transition concentration for water-dominated systems (fraction)

χ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . radius of contact (m)

ψd . . . . . . . . . . . . . . . . . . . . liquid droplet deposition rate into liquid phase (kg/m3 /s)

ψe . . . . . . . . . . . . . . . . . . . . . liquid droplet entrainment rate into gas phase (kg/m3 /s)

ψg . . . . . . . . . . . . . . . . . . . mass transfer rate between gas and liquid phases (kg/m3 /s)

xxiii
LIST OF ABBREVIATIONS

AA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Anti-Agglomerant

CAPEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Capital Expenditures

CMC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Critical Micelle Concentration

CSM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Colorado School of Mines

CSMGem . . . . . . . . . . . . . . Colorado School of Mines Gibbs Energy Minimization model

CSMHyK . . . . . . . . . . . . . . . . . . . . Colorado School of Mines Hydrate Kinetics model

DSC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Differential Scanning Calorimetry

EACN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Equivalent Alkane Carbon Number

FBRM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Focus Beam Reflectance Method

HLB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydrophilic-Lipophilic Balance

HPU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydraulic Power Unit

HTGC . . . . . . . . . . . . . . . . . . . . . . . . . . . . High Temperature Gas Chromatography

KHI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kinetic Hydrate Inhibitor

OPEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Operational Expenditures

O/W . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oil-in-Water emulsion

o/W/O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oil-in-water-in-oil multiple emulsion

PVM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Particle Video Microscope

PVT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pressure-Volume-Temperature analysis

SAD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surfactant Affinity Difference

SARA . . . . . . . . . . . . . . . . . . . . . . . . Saturates, Aromatics, Resins, Asphaltenes test

SDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Sodium Dodecyl Sulfate

xxiv
THI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Thermodynamic Hydrate Inhibitor

WAT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wax Appearance Temperature

WOR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Water-Oil Ratio

W/O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Water-in-Oil emulsion

w/O/W . . . . . . . . . . . . . . . . . . . . . . . . . . . . Water-in-oil-in-water multiple emulsion

xxv
ACKNOWLEDGMENTS

I would like to express my sincere appreciation to Dr. Dendy Sloan, who persuaded me to

work on my PhD degree at the Colorado School of Mines, and for introducing me to the world of

gas hydrates and flow assurance. I would like to thank Dr. Sloan for providing support from the

beginning, and for his excellent advice and guidance during my research work at Mines.

I would like to thank Dr. Carolyn Koh and Dr. Amadeu Sum for all their support, advice and

guidance during my research work at the Center for Hydrate Research, and for their comments,

suggestions and corrections to many technical communications.

I would like to acknowledge Dr. Hossein Kazemi for giving me the opportunity to work with

him, and for all his support, advice and guidance during the writing of this dissertation.

I would like to thank my thesis committee members, Dr. Mike Batzle, Dr. Yu-Shu Wu and

Dr. Xiaolong Yin, for their time spent helping me shape my research project and reading this

dissertation.

I acknowledge the financial support from DeepStar and the CSM Hydrate Consortium, which

has been sponsored by BP, Champion Technologies, Chevron, ConocoPhillips, Eni, ExxonMo-

bil, Halliburton, MultiChem, Nalco, Petrobras, Schlumberger, Shell, SPT Group, Statoil, and

Total. Special thanks for the following individuals, who provided direct input to this research

project: Jeff Creek, Douglas Estanga, and Siva Subramanian from Chevron; Tom Danielson from

ConocoPhillips; Larry Talley, Mike Eaton, Doug Turner, and Jason Lachance from ExxonMobil;

Alexander de Freitas formerly from Petrobras; and Inge Wold from SPT Group.

I would like to acknowledge ExxonMobil (Larry Talley, Mike Eaton, Doug Turner, and Jason

Lachance) and University of Tulsa (Mike Volk and Emmanual DelleCase) for access to flow loops,

and for guidance in experimental observations.

I would like to acknowledge the Viola Vestal Coulter Foundation for granted me a fellowship

in the first year of my career at Mines, and the Roberto Rocca Education Program for granted me

the Roberto Rocca Fellowship during the last two years of my PhD at Mines.

Very special thanks goes to the present and past graduate and undergraduate students that form

part of the Hydrate Busters, for all the discussions in and out of the hydrate office and specially for

xxvi
your friendship. I would like to particularly thank the following Busters for their direct input to

my work at the Center for Hydrate Research: Zach Aman, John Boxall, Piyush Chaudhari, Simon

Davies, Giovanny Grasso, Eric Grzelak, Sanjeev Joshi, Patrick Lafond, Ishan Rao, Kyle Springer,

Matt Walsh, and Eric Webb.

I would like to thank my family and friends, who have given me unconditional support all these

years. To my parents, for teaching me the value of perseverance, discipline and honesty, for being

my sources of inspiration and motivation, and for all the support and love they have given me;

without this foundation I would not have achieved this goal.

Finally, I would like to thank Patricia Cuba, my wife, for being with me through this journey,

for sharing a common goal of seeking new knowledge, for cheering me during tough moments, for

making me a better person, and for loving me so much.

xxvii
This work is dedicated to the loving memory of Solange Martinez, mi Abuela.

xxviii
CHAPTER 1

INTRODUCTION

The offshore oil and gas industry has experienced a continuous growth since its inception in the

1940’s (Sandrea and Sandrea, 2007). Currently, offshore production represents 30% of global oil

and gas production (Serbutoviez, 2012). With the implementation of new exploration and drilling

technologies, it is expected for the offshore industry to continue its growth towards deep (500 - 1500

m) and ultra-deep (> 1500 m) waters. From market predictions, the offshore industry is expected

to invest $210 billion for new developments in five years (2011 - 2015), a 60% increase compared

to the previous period (2006 - 2010), with pipelines and flow systems representing the 38% of this

budget (Davis, 2011).

The offshore industry will face more challenging scenarios with production from deeper and

colder waters. Subsea facilities will require longer subsea tiebacks in satellite fields to transport

hydrocarbon fluids from the wellhead to existing production and processing platforms, and may

require the transportation of processed gas and condensate streams to export facilities through

subsea pipelines. The formation of solid deposits, such as gas hydrates, waxes, asphaltenes and

mineral scale, may plug the flowlines, preventing production and generating a safety hazard. The

flow assurance of the produced hydrocarbon stream has become a technical discipline that focuses

on the design of safe and secure operation techniques for the uninterrupted transport of reservoir

fluids from the reservoir to the point of sale (Kaczmarski and Lorimer, 2001). The rapid formation of

gas hydrates, which is promoted by typical high pressure and low temperature operation conditions

in deep subsea facilities, is considered one of the most challenging flow assurance problems (Sloan,

2005).

It had been recognized that predicting hydrate formation conditions using thermodynamic cal-

culations (with excellent accuracy) is not sufficient to estimate hydrate plugging risk (Sloan, 2005).

Instead, this can be used to design hydrate avoidance methods (e.g., chemical injection, thermal

insulation, active heating, and pressure reduction), keeping the systems out of the hydrate sta-

bility region (Creek et al., 2011). The offshore oil and gas industry has progressed toward using

longer tiebacks to connect subsea wells with platforms (Ronalds, 2005), rendering hydrate avoid-

1
ance methods economically unfeasible. The alternative would be to consider hydrate management

approaches, where hydrates are allowed to form, but the plugging risk is low (Creek et al., 2011;

Sloan, 2005). Models for hydrate formation kinetics coupled with transportability models may be

helpful during the design and assessment of hydrate management approaches, providing estimates

of the amount of hydrates that could form and its transportability in specific scenarios, to estimate

hydrate plugging risk.

A gas hydrate model that estimates hydrate plugging risk in pipelines is of significant interest

in design and optimization of flowlines and operation procedures. Such a model should account

for transient mechanisms present in different systems of oil/gas production facilities such as: oil-

dominated systems (oil as hydrate carrier phase), water-dominated systems (water as hydrate

carrier phase), and gas-dominated systems (hydrate film growth on pipe walls). By predicting gas

hydrates formation and transportability, the comprehensive model can be applied in the design and

optimization of oil/gas transport facilities, focusing on prevention, management or remediation of

gas hydrates in flowlines.

An intense research effort at the Center for Hydrate Research of Colorado School of Mines

(CSM) led to the development of the CSM Hydrate Kinetics model (CSMHyK), a gas hydrate

model specially designed for oil-dominated systems, which has been incorporated as a plug-in

module in a transient multiphase flow simulator (Bendiksen et al., 1991; Boxall et al., 2009; Turner

et al., 2005). The objective of this thesis is to extend and significantly advance the gas hydrate

model (CSMHyK) towards a comprehensive model that includes transient mechanisms present in

oil, water and gas-dominated systems, with the goal of accounting for the principal scenarios of

hydrate formation found in oil/gas subsea flowlines.

This introductory chapter presents a general overview of flow assurance focusing on solid de-

posits (waxes, asphaltenes, mineral scale and gas hydrates), a brief description of the current version

of CSMHyK for oil-dominated systems, the thesis objectives and scope, and structure of the thesis.

1.1 Flow Assurance

The flow assurance of produced fluids from offshore wells has become a technical discipline,

critical for the economic success of offshore production projects (Kaczmarski and Lorimer, 2001).

Flow assurance involves the design of facilities and operation procedures for the uninterrupted

2
transport of reservoir fluids from the reservoir to the point of sale, over the life of a project.

The flow assurance contributions to subsea facilities design involve specification of pipeline size,

production facility materials, and pressure boosting requirements (Bai and Bai, 2005). The pipe di-

ameter should guarantee an optimum hydraulic performance of the system, in terms of the pressure

drop versus production flow rate relationship, over a range of operating conditions (i.e., minimum

and maximum allowable flow rates). The selection of materials should consider compatibility with

the fluids to be transported, corrosion issues and thermal insulation requirements. Pressure boost-

ing equipment may be required if the wellhead pressure is not enough to transport fluids up the

riser to the processing platform.

The flow assurance design process also involves fluid properties characterization and thermal-

hydraulic calculations, for the development of operation strategies that prevent problems such as

solid blockages, severe slugging, sand erosion and corrosion. Four issues have been identified in the

flow assurance design process (Bai and Bai, 2005):

1. Fluid characterization.

2. Steady-state thermal-hydraulic calculations.

3. Transient flow thermal-hydraulic calculations.

4. Final system design and operation procedures for flow assurance.

Each issue is discussed below in this section.

1.1.1 Fluid characterization

The first stage consists of the characterization of hydrocarbon fluids and produced water, to

obtain the fluid properties required for the thermo-hydraulic calculations and to assess potential

problems with solid deposits. For hydrocarbon fluids the characterization involves pressure-volume-

temperature (PVT) and compositional analysis. The PVT analysis provides properties such as

density, viscosity, oil and gas compressibility, saturation pressure, and gas solubility in oil. The

compositional analyses are typically performed through components with at least C36+ carbon

number, and should include density and molecular weight of the petroleum fractions from C7 and

above components (Guo et al., 2005).

3
The characterization of hydrocarbon fluids also includes special measurements for the flow

assurance design, to assess potential problems with deposition of waxes or asphaltenes. Wax is

the name given to heavy components of crude oil that consist of linear n-paraffins and branched

iso-paraffins with a carbon number from C20 to C70 (Sarica and Panacharoensawad, 2012). To

determine the amount of wax in a crude oil, a compositional analysis is required that detects

heavier components in the oil, typically using high temperature gas chromatography (HTGC), which

determines paraffin composition as high as C100 (Alboudwarej et al., 2006). A key parameter used

to determine problems of wax deposition is the crude cloud point or wax appearance temperature

(WAT), defined as the temperature below which wax crystals start to form. Another wax parameter

of interest for flow assurance is the pour point, which is related to the oil phase rheology. The pour

point is the temperature at which the oil phase solidifies because of formation of a gel network of

precipitated wax crystals; this could prevent restart of fluid flow during transient shut-down/restart

operations.

The deposition of wax on the pipe walls narrows the flow conduit diameter and changes the wall

roughness increasing pressure drop and reducing production flow rate, and can eventually plug the

pipeline (Figure 1.1). Wax deposition is a slow mechanism mainly dominated by molecular diffusion

of heavy paraffins (Guo et al., 2005). The molecular diffusion process is driven by a concentration

gradient, which is a function of the radial temperature gradient inside the flowing pipeline. The

solubility of heavy paraffins is greater at the higher temperature at the center of the pipe, while

the solubility is lower at the low temperature at the pipe wall. If the temperature at the wall is

lower than the WAT point at a given pressure, then wax will precipitate and deposit on the wall.

Figure 1.1: Wax deposit on a pipe wall narrowing conduit diameter and modifying wall roughness
(Mokhatab and Towler, 2009).

4
Essentially asphaltenes are large polar molecules primarily formed by carbon and hydrogen,

with one to three heteroatoms (nitrogen, oxygen or sulfur) per molecule (McCain, 1990). The het-

eroatoms can be part of aromatic rings clusters or can be in the links between the rings (Figure 1.2).

The polar groups of an asphaltene molecule provides a driving force for interfacial adsoption, due

to the ionization of the polar group at the oil-water interface. For example, the carboxylic group

(COOH), shown to the left of the asphaltene molecule in Figure 1.2, can dissociate in water result-

ing in a carboxylate anion (COO- ), which has a high affinity for the water, and a positively charge

hydrogen ion (Verruto and Kilpatrick, 2008).

Figure 1.2: Example of asphaltene molecular structures (Coelho et al., 2012).

Asphaltenes are petroleum fractions that solubilize in aromatic solvents (e.g., toluene and ben-

zene), but are insoluble in light aliphatic solvents (e.g., pentane and heptane) (Mitchell and Speight,

1973). Asphaltenes are insoluble in the bulk oil phase, but are stabilized as dispersed colloids by

the adsorption of resins. Resins are formed by two groups, a polar group that contains heteroatoms

(e.g., oxygen, sulfur and nitrogen) and a nonpolar paraffinic group. The polar group of the resins is

adsorbed to the surface of asphaltene micelles by hydrogen bonding between the heteroatoms and

dipole-dipole interactions generated from the polarities of the resin and asphaltene. The nonpolar

paraffinic group of the resins acts a transition to the bulk oil phase.

The stability of asphaltenes changes with pressure, temperature and composition. Generally,

asphaltene precipitation occurs near the oil bubble point because of the vaporization of light hydro-

carbons; i.e., a change in composition of the liquid hydrocarbon phase (Hammami and Ratulowski,

5
2007). Near the oil bubble point, the different compressibility between lighter and heavier compo-

nents translates into an increase in volume fraction of lighter components, affecting the equilibrium

that exists in the nonpolar portion of the crude oil. This cause desorption of resins from the sur-

face of asphaltene micelles, which reduces the stability of the colloidal dispersion, promoting the

agglomeration of asphaltene micelles and eventual precipitation.

The amount of asphaltenes in an oil sample can be measured with a SARA analysis, which

determines the amount of saturates, aromatics, resins and asphaltenes by a combination of solvent

induced precipitation and column chromatography. Asphaltene stability in dead oils can be mea-

sured using titration tests, where an aliphatic solvent is added and a light transmittance probe

is used to detect the asphaltene flocculation onset, which is marked by a maximum in the light

transmittance signal (Guo et al., 2005). A simple method to determine the asphaltene precipita-

tion tendency of a crude oil is the de Boer plot (de Boer et al., 1995), which identifies three regions

(severe, slight and no problem) based on the difference of reservoir pressure and bubble point pres-

sure as function of oil density at in-situ conditions; however predictions from this method could

be overly pessimistic (Wang et al., 2006). Hammami and Ratulowski (2007) presented a review of

special methods to determine the upper and lower asphaltene precipitation loci in terms of pressure

and temperature for live oil samples.

The precipitation of asphaltenes can cause plugging of wells and pipelines by deposition on

tubular walls (similar to wax deposition showed in Figure 1.1). Additionally, asphaltenes can cause

oil/water separation problems by emulsion stability (Sjöblom et al., 2003) and fouling of surface

equipment (Dickakian and Seay, 1988).

The characterization of produced water allows assessment of potential problems with inorganic

precipitates and provides information to estimate the hydrate equilibrium curve. Inorganic pre-

cipitates or scales are mineral solids that precipitate from the water phase, caused due to changes

in pressure and temperature. The most common scales encountered in oil and gas pipelines are

calcium carbonate (calcite, CaCO3 ), barium sulfate (BaSO4 ), strontium sulfate (SrSO4 ), and cal-

cium sulfate (anhydrate, CaSO4 ) (Guo et al., 2005). Carbonate scales form from formation water

with reduced pressure, increased temperature and increased pH. Sulfate scales form with mixing

of different waters (e.g., formation water mixed with sea water). Problems with scale deposition

increase with increasing water cut. Scale deposits can also plug a flowline by reducing the conduit

6
diameter.

Gas hydrates are crystalline compounds formed by cages of hydrogen-bonded water molecules

(host) stabilized by encapsulation of small gas molecules (guest), like methane, ethane, propane

or carbon dioxide (Sloan and Koh, 2008a). The cages of hydrogen-bonded water molecules are

typically formed by pentagonal and hexagonal faces, and sometimes by square faces depending on

the crystal structure, as shown in Figure 1.3 for the most common gas hydrate structures. The

gas hydrate structure I (sI) unit cell consists of two small cages, each formed by twelve pentagonal

faces (512 ) and six large cages, each formed by twelve pentagonal and two hexagonal faces (512 62 ).

The gas hydrate structure II (sII) unit cell consists of sixteen small 512 cages and eight large cages,

each formed by twelve pentagonal and four hexagonal faces (512 64 ). The gas hydrate structure H

(sH) unit cell consists of three small 512 cages, two small cages, each formed by three square, six

pentagonal and three hexagonal faces (43 56 63 ) and one large cage formed by twelve pentagonal and

eight hexagonal faces (512 68 ). The stability of a specific hydrate structure depends on the fit of the

guest molecule inside the host cage. Small hydrocarbon molecules, such as methane and ethane

can stabilize gas hydrate sI. However, sII is the typical gas hydrate structure encountered in oil and

gas pipelines because of the presence of larger hydrocarbon molecules, like propane and iso-butane

that fit in the large cage of sII, but do not fit in the large cage of sI (Sloan and Koh, 2008a).

The pressure versus temperature diagram, shown in Figure 1.4, illustrates a typical phase

behavior of hydrocarbon-water systems, such as, ethane–water, propane–water, and isobutane–

water (Sloan and Koh, 2008b). In this diagram the symbols I, LW , H, V, and LHC represent ice,

liquid water, hydrate, vapor, and liquid hydrocarbon, respectively. The quadruple points represent

the coexistence of four phases, Q1 for I–LW –H–V, and Q2 for LW –H–V–LHC . The continuous blue

line formed by line LW –H–LHC at conditions above Q2, line LW –H–V between Q1 and Q2, and line

I–H–V at conditions below Q1, represents the hydrate equilibrium boundary. To the right of this

curve the hydrate phase is not stable (i.e., no hydrates are possible), and at lower temperatures

and higher pressures to the left side of the hydrate equilibrium curve the hydrate phase is stable

(i.e., hydrates can form). The red segmented line in the upper portion of the diagram, represents

the vapor-liquid hydrocarbon equilibrium curve with liquid water to the right of Q2 and hydrate

to the left of Q2. Below the V–LHC equilibrium curve the hydrocarbon will be in the vapor state,

and above the V–LHC equilibrium curve the hydrocarbon will be in the liquid state. The dash-dot

7
Figure 1.3: Common gas hydrate structures; adapted from Sloan (2011, p. 3).

green line is the I–LW equilibrium curve with a vapor phase below Q1 and a hydrate phase above

Q1. To the left of the I–LW curve, water will be in the solid state (ice or hydrate), and to the right

of the I–LW curve, water will be in liquid state or in the hydrate phase. The complexity of the

phase behavior diagram changes with the composition of the system; more details about hydrate

phase diagrams for water-hydrocarbon systems can be found in Sloan and Koh (2008b).

The gas hydrate phase envelope, defined by equilibrium pressure and temperature conditions,

can be determined by thermodynamic flash calculations based on the van der Waals and Plat-

teeuw (1959) model with excellent engineering accuracy (Ballard and Sloan, 2004b). Several com-

mercial thermodynamic programs (e.g., Multiflash


R
from Infochem Computer Services, Ltd., and

PVTSim
R
from Calsep) implement a modified version of this model for the prediction of gas hy-

drate properties (Sloan and Koh, 2008a). Another program developed at the Colorado School of

Mines, based on the Gibbs energy minimization method (CSMGem), can also be used to estimate

the gas hydrate equilibrium conditions (Ballard and Sloan, 2004a). The composition of hydrocar-

bon fluids and produced water, obtained during the fluids characterization, are used as input to

these programs to obtain the hydrate equilibrium curve.

8
HC
L −H−L
Pressure (log scale)
− L HC

w
Lw − V
H−V−L
HC

Q2

V
I−L −H


H

w

w
L
−V Q1
I−H
I − Lw − V

Temperature

Figure 1.4: Pressure versus temperature diagram for a hydrocarbon–water system; adapted from
Sloan and Koh (2008a, p. 198).

Figure 1.5 shows a pressure versus temperature diagram that includes the WAT curve, the upper

and lower asphaltene precipitation loci and the hydrate equilibrium curve from the characterization

of a Gulf of Mexico reservoir fluid. This kind of diagram allows one to identify conditions corre-

sponding to potential formation of solid deposits. For the wax and hydrate curves, the right side

of the curves represents the regions with no precipitation/formation of solids, and to the left of the

curves are the regions of wax crystals precipitation and hydrate formation, respectively. In the case

of asphaltenes, there are two precipitation curves, an upper and a lower precipitation curve. The

asphaltene precipitation region is found between the upper and lower precipitation curves, while

there are two asphaltene stable regions, one above the upper precipitation curve and other below

the lower precipitation curve.

1.1.2 Steady-state thermal-hydraulic calculations

The second stage of the flow assurance design process consists of steady-state thermal-hydraulic

calculations using multiphase flow simulators (e.g., PipeSim


R
from Schlumberger, PipePhase
R

from Invensys, and PipeFlow


R
from Daxesoft). One of the purposes of this stage is to establish

9
70
Wax No Wax
60 Asphaltene
stable region

Pressure (MPa)
50 Asphaltene
precipitation region
40
Asphaltene
30 stable region

20 Hydrates

10 No
Hydrates

0
260 280 300 320 340 360
Temperature (K)
Figure 1.5: Pressure versus temperature diagram showing the WAT curve (green dashed curve),
upper and lower asphaltene precipitation loci (red dash-dot curves) and hydrate equilibrium curve
(blue continuous curve) from the characterization of reservoir fluids; adapted from Ayers (2003).

pipeline diameter, considering the geometry of the flowline from the wellhead to the platform

in terms of horizontal distance and elevation. The pressure drop versus production flow rate

relationship is studied through a parametric study modifying initial and boundary conditions. The

pipe diameter selected should guarantee proper hydraulic performance of the system over a range

of operating conditions during the life of the project.

Potential problems with solid deposits can be determined by comparing the flowline pressure

and temperature conditions obtained from the steady-state thermal-hydraulic calculations with

the conditions of solid precipitation/formation obtained during the fluid characterization studies.

Wherever the flowline operating conditions curve lies to the left of the WAT, asphaltene precip-

itation or hydrate equilibrium curves, the system is considered to potentially have solid deposits

problems. Figure 1.6 shows a pressure versus temperature diagram where a typical subsea flowline

operation conditions curve is compared to the WAT curve, upper and lower asphaltene precipita-

tion loci and hydrate equilibrium curve. The reservoir fluids flowing out of the wellhead through

the horizontal portion of the flowline, start transferring heat to the cold environment (typical sea

floor temperature is 277 K) through the pipe wall, represented by a considerable decrease in tem-

perature and a slight decrease in pressure in the flowline. There is a major pressure decrease when

10
the fluids flow up the riser with an increase in temperature. A Joule-Thomson cooling effect at a

valve near the platform is responsible for an additional decrease in temperature. In this example

the flowline operation conditions intercept the wax precipitation boundary and then the hydrate

equilibrium curve during fluid flow in the horizontal portion of the pipeline, indicating that this

system can have a potential problem with wax and hydrate solid deposits. The operation condi-

tions of the pipeline are below the lower asphaltene precipitation locus (i.e., inside the asphaltene

stable region), indicating that there is no risk of asphaltene precipitation in this flowline. Note that

if the operation pressure is high enough to shift the flowline conditions to inside the asphaltene

precipitation region, then asphaltene precipitation would be an additional potential problem. At

this stage, it is possible to study different flow assurance strategies, such as thermal insulation of

the pipeline to prevent the system from entering in the solid precipitation/formation regions.

60
Wax No Wax

50 Asphaltene
Pressure (MPa)

precipitation region

40
Asphaltene
stable region
30
No
Hydrates Hydrates
20
Wellhead
Flowline operation
10 conditions
Platform

0
260 280 300 320 340 360
Temperature (K)
Figure 1.6: Pressure versus temperature diagram showing a typical subsea flowline operation
conditions curve, the WAT curve, the lower asphaltene precipitation locus and hydrate equilibrium
curves; adapted from Ayers (2003).

A flow assurance strategy typically used to avoid hydrate formation is the injection of ther-

modynamic hydrate inhibitors (THI), such as methanol (MeOH) or monoethylene glycol (MEG).

11
Thermodynamic hydrate inhibitors compete for water molecules through hydrogen-bonding, pre-

venting the formation of hydrate cages, thus inhibiting the hydrate formation. This results in the

hydrate equilibrium curve shifting to the left by decreasing the hydrate equilibrium temperature

and increasing the hydrate equilibrium pressure. Figure 1.7 shows a pressure versus temperature

diagram where a typical subsea flowline operation conditions curve is compared to hydrate equi-

librium curves at different methanol concentrations. A significant portion of the flowline is inside

the hydrate forming region when no methanol is injected. A methanol concentration greater than

20% is required to completely inhibit hydrate formation in this system.

Hydrate forming region


20 0%
30% 20% 10% MeOH
Pressure (MPa)

15 11 km
24 16
32

40
10 Wellhead
48

5 56 Hydrate
Platform 64 72 free
region
80
0
270 280 290 300
Temperature (K)
Figure 1.7: Pressure versus temperature diagram showing a typical subsea flowline operation
conditions curve and hydrate equilibrium curves at different methanol concentrations; adapted
from Sloan and Koh (2008c, p. 646). Points on the flowline operation conditions curve indicate
distance from wellhead in kilometers.

The gas and oil industry is currently in a transition from using hydrate avoidance methods (such

as injection of thermodynamic hydrate inhibitors) to risk management methods (Sloan, 2005). The

use of inhibitors that act at the interfaces of hydrate particles, such as kinetic hydrate inhibitors

(KHI), have been proven to reduce the required concentrations of chemical additives in a more

12
economical and efficient flow assurance strategy. KHIs are low molecular weight polymers that

dissolve in the water phase, such as poly-N-vinylcaprolactam (PVCap) and poly-N-vinylpyrrolidone

(PVP). The purpose of KHIs is to delay the hydrate nucleation and crystal growth for a period of

time by adsorbing at the surface of the hydrate crystal (Sloan and Koh, 2008c).

1.1.3 Transient flow thermal-hydraulic calculations

The third stage of the flow assurance design process consists of transient-flow thermal-hydraulic

calculations using transient multiphase flow simulators. This stage involves the study of transient

operations such as planned or unplanned shut-down of production and subsequent restart. As-

suming that the system is originally outside of the hydrate formation region, during a production

shut-down the fluids temperature will decrease, reaching equilibrium with the cold external ambient

temperature, entering into the hydrate formation region. The formation of gas hydrates during a

production shut-down represents a major concern because of its rapid formation at high subcool-

ing (i.e., difference between hydrate equilibrium temperature and fluids temperature). Generally,

wax deposition does not represent a problem during production shut-down because the molecular

diffusion that dominates the deposition mechanism is a slow process requiring a great volume of

fluid flow to create a deposit.

The transient thermal-hydraulic calculations complement the design of thermal insulation by

estimating the cool-down time, which is the time required for the system to reach the hydrate

formation conditions from the steady-state flow conditions after a production shut-down. The

cool-down time should be long enough to accommodate a no-touch time and an operation time.

During the no-touch time, no means of hydrate mitigation are implemented and the operator should

try to determine the problem that caused the shut-down. The operation time is the time required

to perform a hydrate mitigation strategy during the shut-down, such as depressurization or dead

oil displacement. If the operator is able to find a solution before the no-touch time, then the system

can be restarted without the implementation of any hydrate mitigation strategy. Generally, the

cool-down time is between 12 to 24 hours (Bai and Bai, 2005).

During planned production shut-down operations longer that the no-touch time, it is possible

to perform chemical inhibitor injection previous to shut-down to prevent hydrate formation. On

the other hand, during unplanned production shut-down operations the system will not count with

13
hydrate inhibition, so the process must include other hydrate remediation strategies if the no-

touch time is surpassed. Typically, after the no-touch time hydrate inhibitors are injected in the

system to prevent hydrate formation in the wellbore, wellhead, jumpers and manifolds, followed

by depressurization or dead oil displacement. The depressurization of the pipeline is performed by

blow-down of the pipeline from the top sides, lowering the pressure of the system outside of the

hydrate formation region.

The transient thermal-hydraulic calculations also serve to estimate the pipeline liquids inventory

after a production shut-down, which consist of estimating the amount and distribution of gas, oil

and water inside the pipeline (Hagesaether et al., 2007). The amount of hydrate inhibitor required

can be determined from the amount of water in the system. Also, the transient thermal-hydraulic

calculations can be used to study potential problems with severe slugging. Liquids can accumulate

in local low spots in the pipeline, which can result in a liquid surge coming to the platform during

restart, or generate severe slugging at low restart flow rates. If the upstream pressure is low and

the flow rate is reduced, the gas and liquid velocities might not be enough to continuously carry the

fluids up the riser, resulting in intermittent fluid flow to the top sides (i.e., processing equipment).

Liquid accumulates in the base of the riser until pressure behind the liquid is high enough to push

the slug up the riser resulting in a huge amount of liquids flowing into the separator.

1.1.4 Final system design and operation procedures for flow assurance

The fourth and final stage of the flow assurance design process consists of finalizing the system

design and establishing the operating philosophy for flow assurance issues, by integrating the results

of the previous three stages. The final system design will depend strongly on the operation strategies

adopted to avoid problems with solid blockages. Pigging (passing a cylinder down the line to clean

it) is an alternative to remove solid deposits of wax, asphaltene or mineral scale, by mechanically

removing the deposit from the wall. If pigging is selected as an operation strategy, the system should

include round-trip pigging capabilities in the facilities design. Another alternative to manage solid

deposits is chemical inhibitor injection, which requires the inclusion of chemical inhibitor delivery

lines and platform chemicals storage tanks in the facilities design.

The flow assurance design process of an offshore field faces an additional challenge; to balance

capital cost (CAPEX) and operational cost (OPEX) minimizing the overall economic expenses of

14
a project (Guo et al., 2005). The selection of materials (thermal insulation, high grade materials)

and subsea equipment along with the design of solid deposits mitigation systems are related to

CAPEX. Operation procedures like chemical inhibitor injection, pigging, flow monitoring are related

to OPEX.

1.2 Colorado School of Mines Hydrate Kinetics Model (CSMHyK)

An intense research effort at the Center for Hydrate Research of the Colorado School of Mines

led to the development of the Colorado School of Mines Hydrate Kinetics model (CSMHyK), a

gas hydrate model specially designed for oil-dominated systems. CSMHyK was developed for oil-

dominated systems based on the conceptual model presented in Figure 1.8, which represents an

approximation to the mechanism of hydrate plug formation and is divided into four main stages:

1. Water entrainment: The water droplets are dispersed in the continuous oil phase by flow

shear, as a water-in-oil emulsion.

2. Hydrate growth: A hydrate shell forms at the interface between the water droplets and the

surrounding oil phase.

3. Agglomeration: The hydrate-encrusted water droplets can agglomerate, increasing into larger

hydrate masses or agglomerates.

4. Plugging: The agglomeration of hydrate particles leads to an increase in the slurry viscosity,

which can eventually form a plug.

Figure 1.8: Conceptual model for hydrate formation in oil-dominated multiphase flow systems
consisting of gas, oil, and water; adapted from Turner (2005), with input from J. Abrahamson (U.
Canterbury, Christchurch, NZ).

15
CSMHyK has been incorporated as a plug-in module in a transient multiphase flow simulator

(Bendiksen et al., 1991; Turner et al., 2005). The transient multiphase flow model considers four

phases: gas, oil, water, and hydrate. The governing equations for one-dimensional multiphase flow

in a pipeline consist of the continuity equation, the momentum conservation equation and the energy

equation. These governing equations are constrained by geometrical data; initial and boundary

conditions; fluid properties; and inter-field mass, momentum and heat transfer. The model solves

the system of equations to obtain pressure, phase velocities, phase masses, and temperature. A

detailed description of the governing equations of the transient multiphase flow model is presented

in Section 3.3.3 of Chapter 3, which presents a review of multiphase flow concepts and modeling

approaches.

CSMHyK is used to calculate additional contributions to the conservation equations. The hy-

drate kinetic model (presented in Section 1.2.1) calculates the hydrate formation rate, which is

used to calculate mass transfer terms between the gas, water and hydrate phases. The hydrate

equilibrium curve is used as a thermodynamic constraint to determine the conditions when the hy-

drate phase start to form. A hydrate particle agglomeration model is used to calculate the effective

viscosity of the hydrate carrier phase, which affects the momentum conservation equations. The

modified viscosity will induce modifications in the friction factors. The hydrate heat of formation

acts as an energy source inside the pipeline, affecting the energy equation and the resulting system

temperature.

The current version of the CSMHyK model comprises three sub-models: the kinetics model,

the transport model and the cold flow model.

1.2.1 Kinetics model

The kinetics model assumes that all the water is dispersed into the continuous oil phase as

water droplets of fixed mean diameter. The surface area between water and hydrocarbon phases is

calculated using the Hinze (1955) correlation. Hydrate nucleation is assumed to occur immediately

after a specified subcooling (∆Tsub = Thyd eq − Tsystem ), and the hydrate growth is calculated using

the following intrinsic kinetics equation of first order with an adjustable rate constant (Boxall,

16
2009),
 
dmgas k2
= −uk1 exp As ∆Tsub (1.1)
dt Tsystem
The intrinsic kinetics equation gives the rate of gas consumption during hydrate formation as a

function of the intrinsic kinetic rate constants (k1 and k2 ), the surface area (As ) between water and

hydrocarbon phases, and the subcooling as the thermal driving force. Vysniauskas and Bishnoi

(1983) and Englezos et al. (1987) measured the gas consumption rate during the formation of

mixed methane and ethane hydrate at different temperatures and subcooling, in the absence of

either mass or heat transfer resistances. Their experimental data, shown in Figure 1.9, were used

to regressed the intrinsic kinetic rate constants, k1 = 7.3548 × 1017 and k2 = −13600 K. These

constants correspond to the kinetic reaction of hydrate structure I, which represents a limitation

of the model since hydrate structure II is the expected structure in oil and gas flowlines. In order

to match flow loop data, a scaling factor u = 1/500 was introduced, indicating the need to include

mass and heat transfer resistances to the model (Boxall et al., 2009). The need for this large

correction factor (1/500) indicates that our initial model was a fit to the data, rather than a true

physical representation of the phenomena.

Figure 1.9: Bishnoi’s laboratory data used to regress the intrinsic kinetic rate constants (Englezos
et al., 1987; Vysniauskas and Bishnoi, 1983).

17
Once the hydrate particle has formed, the size of the hydrate particle agglomerates is calculated

using a model presented by Camargo and Palermo (2002), based on a steady-state balance between

the inter-particle cohesion force and the fluid flow shear forces. Assuming a constant cohesion force

(Fa ), the following non-linear equation is solved for the hydrate agglomerate diameter (dA ),
"  #2
dA 3−f

Φ
4−f Fa 1 −

dA Φmax dP
− "  3−f # = 0 (1.2)
dP dA
d2P µ0 γ̇ 1 − Φ
dP
where dA is the hydrate agglomerate diameter; dP is the hydrate particle diameter; Φ is the

hydrate particle volume fraction; Φmax is the maximum volume fraction to which particles can

pack (assumed to be equal to 4/7, the packing concentration of randomly packed spheres of same

diameter); f is the fractal dimension that accounts for the porosity of the aggregates, assumed to

be equal to 2.5 (a value inside the reported range of fractal dimensions of aggregates under shear

conditions, Hoekstra et al., 1992); Fa is the inter-particle cohesion force, assumed to be equal to 50

mN/m (a value that corresponds to capillary attraction between hydrate particles by a liquid bridge,

obtained from theoretical calculations by Camargo and Palermo, 2002); µ0 is the oil viscosity; and

γ̇ is the shear rate.

Due to the fractal structure of the aggregates, the effective volume fraction of the agglomerated

hydrate particles is calculated as


 3−f
dA
Φef f = Φ (1.3)
dP
Then, the relative viscosity of the hydrate slurry to the viscosity of the continuous oil phase

is calculated using a modification of the Mills (1985) equation that considers the effective volume

fraction of the hydrate particle aggregates:

1 − Φef f
µr =  (1.4)
Φef f 2

1 − Φmax

Camargo and Palermo (2002) presented the derivation and description of this model for the

hydrate slurry relative viscosity and compared the predictive performance of the model against

experimental data, showing good agreement between calculation and experimental data.

18
A hydrate plug is identified by a large viscosity increase, which causes an unacceptably large

pressure drop in the line, prohibiting flow.

1.2.2 Transport model

The subsequent transport model, similarly, assumes that all the water is dispersed by the flow

shear forces into the continuous oil phase as water droplets. The transport model includes a revised

correlation to calculate the water mean droplet size (Boxall, 2009) as a function of the viscosity of

the oil phase and water-oil interfacial tension, which is an improvement from the Hinze correlation.

Boxall (2009) developed a universal curve in terms of two dimensionless numbers, the Weber

number defined as W e = ρo v 2 d/σ, and the Reynolds number defined as Re = ρo vd/µo , that

identified two separate regions in turbulent flow that affect mean droplet size in an emulsion. The

first region is the inertial sub-range, where the droplet size is determined by a balance between

the interfacial stress that holds the drop together, and the turbulent inertial stress that attempts

to break-up the droplet. In this region the droplet sizes are independent of viscosity. The second

region is the viscous sub-range, where the droplet size is determined by a balance between the

interfacial stress and the sub-eddy viscous stress. Figure 1.10 shows this universal curve, which

uses experimental data from an autoclave mixing cell with a Particle Video Microscope probe and

a large scale flow loop, involving different crude oils. The transition from the inertial to the viscous

sub-ranges occurs when W e > 0.0674Re5/4 . A mean droplet size correlation was developed for each

region using the measured mean droplet diameter. For the inertial subrange the mean droplet size

correlation is,

d¯D = 0.063dW e−3/5 (1.5)

and the droplet size correlation for the viscous sub-range is,

d¯D = 0.016dRe1/2 W e−1 (1.6)

It is important to note that the use of the Weber number is a combination of shear (Re) and

surface properties (σ); this combination is our first quantitative recognition that flow properties

and chemistry affect water droplet size, as opposed to more usual flow properties alone (e.g., Re).

Instantaneous hydrate nucleation occurs at a specified subcooling forming an initial hydrate

shell at the water droplets interface. For this model, it is assumed that further hydrate growth

19
Figure 1.10: Universal curve of turbulent flow sub-ranges in terms of Weber and Reynolds dimen-
sionless numbers that affect mean droplet size in an emulsion, and experimental data obtained in
an autoclave mixing cell and a large scale flow loop, using different crude oils (Boxall, 2009).

occurs on the interior of the shell (i.e., a shrinking core model), and that hydrate growth is limited

by slowest mass or heat transfer through the boundary layer surrounding the hydrate particle or

through the hydrate shell, as illustrated in Figure 1.11.

Figure 1.11: Illustration of a water droplet with a surrounding hydrate shell showing the boundary
layer around the spherical particle.

The hydrate growth rate from external mass and heat transfer through the boundary layer (i.e.,

external resistances) is calculated as,


 
dmgas
= kmass As (Cbulk − Ceq ) (1.7)
dt ext,mass

and

   
dmgas hAs ∆Tsub Xgas hyd M Wgas
= (1.8)
dt ext,heat ∆hhyd M Whyd

20
where kmass is the mass transfer coefficient calculated using correlations for the inertial and viscous

sub-ranges (Davies, 2009), As is the surface area of the hydrate particle, Cbulk is the hydrate guest

concentration in the bulk phase, Ceq is the hydrate guest concentration in the water phase in the

presence of hydrate, h is the convective heat transfer coefficient calculated using correlations for the

inertial and viscous sub-ranges (Davies, 2009), Xgas hyd is the guest concentration in the hydrate

phase, M Wgas is the molecular weight of the gas, M Whyd is the molecular weight of the hydrate

phase, and ∆hhyd is the hydrate heat of formation.

The hydrate growth rate from internal mass transfer across the hydrate shell is based on the

diffusion of gas molecules through the hydrate shell, and is calculated using the following equation:
 
dmgas Dgas As
=   (Cbulk − Ceq ) (1.9)
dt int,mass δ rrPw

Here Dgas is the diffusivity of gas molecules through the hydrate shell, δ is the shell thickness,

rP is the particle radius, and rw is the radius of the water core.

The hydrate growth rate from internal heat transfer across the hydrate shell is calculated using

the following equation,


   
dmgas kcomp As ∆Tsub Xgas hyd M Wgas
=   (1.10)
dt int,heat δ rrPw ∆hhyd M Whyd

Here kcomp is the thermal conductivity of the hydrate shell with porosity ε, assuming that the

pores are filled with water

kcomp = εkwater + (1 − ε) khyd (1.11)

The agglomeration of hydrate particles is calculated in the same form as in the kinetics model,

but considering the inter-particle cohesion force as a function of temperature using a correlation

developed at CSM from micro-mechanical force measurements (Dieker et al., 2008),

dP
Fa = [0.0017 (7.7 − ∆Tsub ) + 0.0007] (1.12)
2
This correlation captures the increase in the value of inter-particle cohesion forces with increas-

ing temperature, which is an improvement to the previous assumption of a fixed value for the

inter-particle cohesion force.

In a similar way to the kinetics model, a hydrate plug is identified by a large viscosity increase,

resulting in a large pressure drop in the flowline, prohibiting flow.

21
1.2.3 Cold flow model

The transport model has been adapted to simulate the BP-Statoil-SINTEF stabilized cold flow

concept (Gudmundsson, 2002; Wolden et al., 2005), in which dry hydrate flowable slurries are

generated. In this stabilized flow concept, warm production fluids are contacted with a stream of

dry hydrate particles in a mixing zone. The water from the production fluids coats the dry hydrate

particles and rapidly converts to dry hydrate in the reaction zone. Part of the dry hydrate slurry

can then be transported to the downstream processing facilities and the rest of the slurry is pumped

around a subsea recycle loop, as shown in Figure 1.12.

Figure 1.12: Conceptual picture of the BP-Statoil-SINTEF stabilized flow concept (redrawn from
Davies, 2009).

The cold flow model was specially designed to simulate the hydrate formation in the reaction

zone, where the fundamental difference to the transport model is that dry hydrate particles are

pre-formed, and then coated with a water film. For this model it is assumed that further hydrate

growth occurs on the interior of the water film with a growing hydrate core, and that hydrate growth

is limited by slowest mass or heat transfer through the boundary layer surrounding the hydrate

particle or through the water film, as illustrated in Figure 1.13. A similar calculation procedure to

the one implemented in the transport model is used in the cold flow model to estimate the hydrate

growth rate.

The lower mass transfer resistance across the water film is responsible for rapid conversion of

water into hydrate. This minimizes particles agglomeration, resulting in a transportable hydrate

slurry.

22
Figure 1.13: Illustration of a hydrate particle with a surrounding water film showing the boundary
layer around the spherical particle.

In the cold flow model the number of particles is conserved throughout the simulation. The

particle size is determined from the particle concentration and the volume fraction of water and

hydrate, by dividing the hydrate and water volumes between the particles.

1.3 Thesis objectives and scope

The previous version of CSMHyK was developed for oil-dominated systems, where the oil is

the continuous phase. However, the increasing water cut in the production stream during the life

of a reservoir will eventually lead to a water-dominated system, where water is the hydrate carrier

phase. Gas-dominated systems are also important scenarios for hydrate formation that includes

deposition of hydrates on the pipe wall, sloughing and jamming to form a plug.

The objective of this thesis was to develop a comprehensive model to predict gas hydrate

formation and transportability in oil and gas flowlines. The model includes transient mechanisms

present in oil, water, and gas-dominated systems, to account for the principal scenarios of hydrate

formation found in oil and gas subsea flowlines. As a result, the CSMHyK model was upgraded.

The accomplished items are:

• Developed conceptual models to explain the mechanisms of hydrate formation in different

scenarios of oil/gas production facilities, from observations and data of flow loop experiments.

• Built mathematical models, based on experimental observations and data, of the mechanisms

of hydrate formation and transportability in different scenarios of oil/gas production facilities.

• Included the effect of naturally inhibited oils and the injection of anti-agglomerant surfactants.

• Generated examples of applications of the new gas hydrate models.

23
• Presented a workflow for hydrate flow assurance in the design of subsea transport flowlines

using the gas hydrate model.

1.4 Structure of this thesis

A review of surface chemistry concepts is presented in Chapter 2, with the principal objective to

identify interfacial phenomena and surface chemistry interactions involved in gas hydrate formation

and agglomeration in oil and gas pipelines. There are five types of interfaces where gas hydrates may

form and aggregate: gas/liquid, liquid/liquid, gas/solid, liquid/solid, and solid/solid; where gas is

hydrocarbon gas, liquid is either oil, water or condensate, and solid is either gas hydrate or pipe

wall surface. A review of fundamental interfacial concepts can help create a better understanding

of phenomena at these interfaces. Two areas of surface chemistry have been selected to illustrate

the concepts and mechanisms associated with these systems: surfactants and emulsions. Examples

from the literature pertaining to gas hydrates are presented for each system.

The fluid flow in a pipeline can be classified in single-phase, multiphase and slurry flow (which

includes solid particles and dispersions), by considering the number of phases present. A review

of fundamental concepts of fluid flow in pipes is presented in Chapter 3, along with a review of

multiphase flow modeling approaches. The effect of gas hydrate formation on multiphase flow is

also illustrated by extending a simple hydrodynamic slug flow model.

The mechanisms of hydrate formation and transportability in pipelines was studied through

the analysis of experimental data obtained at the Center for Hydrate Research laboratory of the

Colorado School of Mines and two large scale flow loops (i.e., Tulsa University flow loop and

ExxonMobil flow loop). Chapter 4 presents a description of the flow loop facilities available during

this study, and the experimental procedure followed in a flow loop experiment. Chapter 5 presents a

set of conceptual pictures developed to explain the physical phenomena of gas hydrate formation in

flowlines. Different hydrate formation mechanisms are identified for different systems encountered

in oil/gas flowlines. For oil-dominated systems, dispersed water droplets are converted into hydrate

particles, and the agglomeration of these hydrate particles increase the slurry viscosity leading to

a plug. Experimental observations in flow loop experiments of water-dominated systems, showed

that pressure drop in the pipe depends on a flow regime transition involving multiple phases and

solid dispersion in the liquid phase. For gas-dominated systems, hydrates deposit and grow on pipe

24
walls, narrowing the flow diameter and increasing surface roughness.

A description of the mathematical models for hydrate formation in oil, water and gas-dominated

systems is presented in Chapter 5, including the models’ contributions to the transient conservation

equations. This work presents improvements to the hydrate aggregation module used for oil-

dominated systems, based on experimental data, which account for temperature, particle-particle

contact time, excess water, and the presence of surface active compounds. The model for water-

dominated systems consists of a mass transfer-based hydrate growth model and hydrate plugging

criterion, based on fluid velocity and liquid holdup. The model for gas-dominated systems consists

of a combined heat and mass transfer model for hydrate film growth on pipe walls. Examples of

applications of these models applied to a typical subsea well/flowline/riser geometry are presented

in Chapter 6.

Chapter 7 summarizes the main conclusions of this thesis, and Chapter 8 presents suggestions

for future research.

25
CHAPTER 2

THE ROLE OF FLUID-FLUID AND FLUID-SOLID INTERFACE ON HYDRATE PLUG

FORMATION

Gas hydrates are crystalline compounds formed by hydrogen-bonded water molecules in a lat-

tice structure that is stabilized by encapsulating a small guest molecule, like methane or ethane

(Sloan and Koh, 2008a). Hydrates form in the presence of appropriate quantities of gas and wa-

ter molecules, typically at high pressures and low temperatures. The hydrocarbon-water interface

represents an ideal location for the formation of gas hydrates in production systems.

Figure 2.1 illustrates a proposed model for hydrate formation in a multiphase flow system

containing water, oil, and gas, where hydrates form at the interface of water droplets entrained in

the oil phase, and/or from gas bubbles entrained in the water phase (Turner, 2005). Frequently

as little as 4 vol% of the water is formed into hydrate, when a plug occurs (Austvik, 1992). In

the oil phase, these hydrate-encrusted water droplets can agglomerate into larger hydrate masses,

leading to an increase in the slurry viscosity, which can eventually form a plug. The surface of water

droplets in water-in-oil (W/O) emulsions is a critical location for the formation and agglomeration

of hydrates. Another aspect that calls the attention of industry and academia are the so-called

naturally inhibited oils, which are characterized by having natural surfactants that seem to prevent

the agglomeration of hydrate particles and allow the transport of hydrate slurries without the

formation of plugs (Leporcher et al., 1998).

This chapter presents a review of surface chemistry concepts, with the principal objective to

identify interfacial phenomena and surface chemistry interactions involved in gas hydrate formation

and agglomeration in oil and gas pipelines. There are five types of interfaces where gas hydrates may

form and aggregate: gas/liquid, liquid/liquid, gas/solid, liquid/solid, and solid/solid; where gas is

hydrocarbon gas, liquid is either oil, water or condensate, and solid is either gas hydrate or pipe

wall surface. A review of fundamental interfacial concepts can help create a better understanding of

the phenomena at these interfaces. Two areas of surface chemistry have been selected to illustrate

the concepts and mechanisms associated with these systems: surfactants and emulsions. Examples

from the literature pertaining to gas hydrates are presented for each system.

26
Figure 2.1: Conceptual model for hydrate formation in a multiphase flow systems consisting of
water, oil, and gas; adapted from Turner (2005).

2.1 Surfactants

The presence of natural surfactants (acids/salts, alcohols, ethers, esters, and other organic com-

pounds) in oil systems and synthetic surfactants (corrosion inhibitors, hydrate anti-agglomerants,

hydrate kinetic inhibitors, demulsifiers, antifoaming agents, etc.) introduced into the well produc-

tion stream motivated the consideration of surfactants in this review. Surfactants, are amphiphilic

molecules which exhibit a double affinity for polar and non-polar substances. In its simplest con-

figuration, a surfactant molecule has two types of functional groups, a hydrophilic or polar group

(water-soluble), and a lipophilic or non-polar group (oil-soluble) which is generally a hydrocarbon

chain with at least 12 carbon atoms; one example of a common surfactant is sodium dodecyl sulfate

(SDS) and a schematic of a typical molecular structure is illustrated in Figure 2.2.

Figure 2.2: Schematic of an anionic surfactant molecule, identifying the negatively charged hy-
drophilic group and the lipophilic group.

Surfactants can be classified, based on the dissociation of its hydrophilic group in water, in four

general types: anionic, cationic, nonionic and amphoteric (or zwitterionic). Anionic surfactants

have a hydrophilic group negatively charge that is associated to an inorganic metal (a cation, typi-

27
cally sodium). In water, the anionic surfactant molecule is divided in two free ions, the counter-ion

with positive charge and the anionic surfactant molecule with negative charge. Cationic surfac-

tants have a hydrophilic group positively charge associated to an inorganic anion (a counter-ion

with negative charge). The cationic surfactant is ionized in water by the separation of the cationic

surfactant monomer with positive charge and the counter-ion with negative charge. Cationic surfac-

tants are highly adsorbed in negative surfaces. Non-ionic surfactants have no counter-ion, resulting

in no electrical charges. Amphoteric surfactants exhibit two oppositely charged groups in their

hydrophilic part.

The balance of effects produced by the type, size and strength of the hydrophilic and lipophilic

groups will determine the properties of the surfactant molecule. The surfactant molecule will tend

to be water-soluble if the hydrophilic group is more important, in particular if the surfactant is

ionized and the lipophilic hydrocarbon chain is relatively short (e.g., <12 carbon atoms in length).

On the other hand, a long hydrocarbon chain (e.g., >16 carbon atoms) will make the surfactant

molecule more oil soluble. The surfactant chemical structure and other parameters depending on

the different fluids, temperature and pressure can alter the affinity of surfactants with the water

and oil phases, as will be discussed later.

Surfactant molecules present two fundamental properties: interfacial adsorption and self asso-

ciation, which are the essential processes leading to surfactants forming structures and enhancing

solubilization. The adsorption of surfactant molecules to interfaces is driven by its double affinity

to polar and non-polar substances, where a minimum in free energy of the system is attained.

The surfactant can diffuse from a bulk phase to an interface, such as gas/liquid (Figure 2.3(a)), liq-

uid/liquid (Figure 2.3(b)) and liquid/solid (Figure 2.3(c), d and e) interfaces, decreasing the surface

or interfacial tension, modifying the contact angle between the phases, modifying the wettability

of solid surfaces, and changing surface charge and surface viscosity (Shah, 1977). The surfactant

adsorption onto solid surfaces depends on the original wettability of the solid. The intrinsic solid

wettability depends on the nature of the solid and liquid substances, particularly their polarity.

The orientation of surfactant molecules depend on the system, in particular with respect to the

electric charges on a solid surface (Durbut, 1999). If the solid is oil-wet, the surfactant adsorbs

from the lipophilic tail, regardless of the surfactant type, as shown in Figure 2.3(c). If the solid

is water-wet (i.e., a polar surface), the surfactant molecules can interact with the polar surface

28
through their hydrophilic part, as shown in Figure 2.3(d). In Figure 2.3(d), a negatively charged

hydrophilic solid is shown with cationic surfactants adsorbed to the surface from the hydrophilic

head, attracted by electrostatic forces. This is a typical phenomenon associated with cationic sur-

factant such as, fatty amine salts and quaternary ammonium salts, used in applications of surface

hydrophobation, corrosion inhibition and solid particles dispersant. An increase in the cationic

surfactant concentration promotes the formation of surface bilayers or admicelles (Figure 2.3(e)),

through interactions between the lipophilic part of the surfactant molecules (Esumi, 2006).

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

Figure 2.3: Adsorption of a surfactant molecule at: (a) gas/liquid interface, (b) liquid/liquid
interface, (c) liquid/nonpolar-solid interface, (d) liquid/polar-solid interface, and (e) liquid/polar-
solid interface with increased cationic surfactant concentration.

Initially, when water soluble surfactants are added to an aqueous solution phase, surfactant

molecules adsorb to any available interface until the interface is saturated. With increasing sur-

factant concentration, surfactant molecules start to associate, forming spherical aggregates called

micelles. The surfactant association is driven by the hydrophobicity of its lipophilic group which

avoids mixing with water. The surfactant concentration at which the first micelle is formed is known

as the critical micelle concentration (CMC), and may be detected from a discontinuity in the change

of several variables such as surface tension, osmotic pressure, viscosity, electrical conductivity and

density (Preston, 1948). Figure 2.4 shows a typical plot of surface tension and electrical conduc-

tivity as a function of surfactant concentration, indicating the discontinuity of these properties at

the CMC.

The CMC depends upon the surfactant molecular structure and physicochemical conditions (i.e.,

temperature, pH, ion concentration). Above the CMC, additional surfactant molecules incorporate

into micelles, while the concentration of un-associated single surfactant molecules (monomers)

29
Figure 2.4: Variation of surface tension and electrical conductivity as a function of surfactant
concentration, indicating the discontinuity at the CMC.

remains nearly constant in a dynamic equilibrium between monomers and micelles, with constant

swapping of molecules between micelles and the bulk phase. A further increase in surfactant

concentration can promote the formation of more complex aggregates in the form of cylindrical,

hexagonal-packed, lamellar structures, worm like micelles, vesicles, liposomes, and other structures

(Shah, 1977). Inverse micelles are formed in oil-soluble surfactants when the hydrophilic part

aggregates to avoid mixing with the oil phase. In the case of inverse micelles of ionic surfactants, a

minimum number of hydrated water molecules is needed that reduces the repulsive forces between

the head groups (Lee, 2008). Figure 2.5 shows a schematic of normal and inverse micelles.

Another important property of surfactant solutions is the solubilization of non-polar molecules

(oil) inside normal micelles in the aqueous phase or water inside inverse micelles in the oleic phase.

Above the CMC, the solubilization increases by incorporation of a non-soluble phase inside the

micelles. When the formulation of the solution is close to the hydrophilic-lipophilic balance of the

surfactant with both oil and water phases, a microemulsion phase can be formed. Bicontinuous

microemulsions are typically defined as containing normal and inverse swollen micelles (filled with

oil and water), which are combined to result in a single phase structure, as shown schematically in

Figure 2.6, where a close to zero curvature of a surfactant layer is limiting the oil and water zones

30
Figure 2.5: Association of anionic surfactant molecules forming normal micelles in an aqueous
phase and inverse micelles in an oleic (oil) phase.

(Clausse et al., 1981; Salager et al., 2005; Scriven, 1976).

Figure 2.6: Bicontinuous structure of microemulsions formed by associating both types of micelles
and resulting in a zero curvature surfactant film that separates oil and water zones (Salager et al.,
2005).

2.1.1 Effects of surfactants in gas hydrate formation

The association of surfactant molecules in the form of micelles at concentrations above the CMC

increases the gas hydrate formation rate and reduces the induction time in quiescent systems, which

is of special interest for the application of gas hydrates in the storage and transportation of natural

gas (Kalogerakis et al., 1993). Zhong and Rogers (2000) studied the effect of SDS on the hydrate

formation mechanism using ethane and natural gas as guest molecules. They suggested that micelles

act as a nucleation point by increasing the solubility of hydrocarbon gas in the aqueous phase and

by inducing the formation of hydrate crystals around the micelle in the bulk water phase below

31
the gas/water interface. Consequently, the hydrate formation rate was observed to increase by

more than 700 times and the induction time for nucleation decreased compared to systems without

surfactant. Other studies had obtained similar effects with different surfactants, but invoked a

different mechanistic explanation, questioning the CMC formation requirement (Lee et al., 2010;

Zhang et al., 2004).

The morphology of the hydrate film that forms at hydrocarbon/water interfaces is affected

by the adsorption of surfactant molecules at the interface. Luo et al. (2007) studied gas hydrate

formation in a methane bubble column without surfactants, observing the formation of a hydrate

shell around gas bubbles, which hindered further formation of gas hydrates. The effect of SDS,

below the CMC, on hydrate formation on gas bubbles using a static mixer was studied by Tajima

et al. (2009), who reported an increase in hydrate formation rate and a change in the morphology

of the hydrate film with the addition of SDS. The adsorption of surfactant at the bubble interface

promoted the formation of a rougher hydrate film with a weaker structure that easily collapsed;

the increase in hydrate formation rate was attributed to higher mass transfer through the hydrate

film and surface renewal by the film collapse.

In another study, Kalogerakis et al. (1993) compared the effect of anionic and nonionic sur-

factants in a stirred cell, reporting a greater increase in methane hydrate formation rate with an

anionic surfactant (SDS). From the experiments with anionic surfactant, aggregates of hydrate par-

ticles suspended in the liquid phase were observed, increasing the slurry viscosity. Hydrate growth

on the reactor wall was attributed to a more water-wet wall due to the anionic surfactant solution.

The effect of anionic surfactants seems to be detrimental for the transportability of hydrate slurries

and should be avoided in pipelines, while nonionic surfactants seem to prevent the agglomeration

of hydrate particles.

These studies exemplify the changes in the interfacial properties between the gas and water

phases caused by the presence of surfactants. Surfactants tend to promote the formation of hydrates

at the interface by lowering the interfacial tension at the water/gas interface, thus facilitating the

transport of gas in contact with water.

32
2.1.2 Effect of surfactants on contact angle, wettability, and cohesion/adhesion forces
of hydrate particles

The wettability of a solid surface by a liquid is determined by the contact angle, θ, which a

liquid drop makes with the surface at the three-phase contact point (see Figure 2.7). Young’s

equation (eqn. 2.1) represents the relationship of the balance of interfacial tensions between the

three phases at the contact point with the contact angle,


σso − σsw
cos (θ) = (2.1)
σwo
This equation shows the dependence of the contact angle on the nature of the liquids and the

solid, i.e., interfacial tension of the liquid phases, σwo , and the solid free energies with the liquids,

σsw and σso . A solid is water-wet if the water drop contact angle is lower than 90o , and it is oil-wet

if the contact angle is greater than 120o . For intermediate values of the contact angle, there is no

preference for the liquid phase to wet the solid.

Figure 2.7: Balance of interfacial tensions among the three phases at contact point, defining contact
angle of the liquid with the solid surface (left: solid is oil-wet; right: solid is water-wet).

Young’s equation is useful to determine the wetting equilibrium of a solid. However, the wetta-

bility can be determined thermodynamically using the spreading coefficient defined in terms of the

interfacial tensions (Kanicky et al., 2001),

Sw/s = σso − σwo − σsw (2.2)

Spontaneous spreading of the water drop occurs when Sw/s > 0, i.e., decreasing the interfacial

tension between water-oil and solid-water. The addition of a surfactant that decreases the interfacial

tension between water-oil and solid-water can make the solid more water-wet. However, the actual

wettability of the solid will depend on the nature of the surfactant.

33
It is often difficult to determine the wettability of hydrate surfaces, as a flat and smooth surface

is required for measurements of the contact angle. Høiland et al. (2005) qualitatively studied the

wettability of freon hydrates in water/oil emulsions by comparing the emulsion inversion point

of different crude oils with and without hydrate particles, demonstrating that the wettability of

hydrate particles depends on the oil composition. Oil-wet hydrate particles are attributed to the

adsorption of natural surface active components of crude oils on the hydrate surface, which are

correlated to have a low hydrate plugging risk in pipelines. Intermediate and water-wet hydrate

particles are believed to contain less or none of these oil-wetting components, and are correlated

to have a high hydrate plugging risk. Natural oil components that interact with hydrate surfaces

are characterized to have polar functional groups and to be surface active, such as petroleum acids

and bases (Borgund et al., 2009).

Aspenes et al. (2010b) studied the influence of oil composition on the wettability of pipeline

surfaces, suggesting that the same natural components of oil that interact with hydrate surfaces

also influence the pipeline surfaces. It was suggested that oils with high acids content compared to

bases are more likely to form oil-wetting surfaces, resulting in lower hydrate plugging tendency.

The non-plugging hydrate tendency of systems with oil-wet hydrate particles and pipeline sur-

faces is supported by the work of Dieker et al. (2009) where micromechanical force measurements

were used to quantify the cohesion forces between cyclopentane hydrate particles. An increase

in the cohesion forces between hydrate particles was measured when the surface active compo-

nents were removed from the crude oil. Figure 2.8 is a summary of the cohesion force measured

by Dieker et al. (2009) for the Caratinga crude oil, which is known to be a naturally-inhibited,

hydrate-transportable oil. From a SARA analysis the Caratinga oil has an asphaltene content of

6.2%, saturates 39.8%, aromatics 39.8% and resins 14.3% (Sjöblom et al., 2010). Note the increase

in cohesion force when the acidic components where extracted from the Caratinga oil (through a

“pH-14 extraction” process), and a further increase in cohesion forces when the asphaltenes were

extracted. The observed increase in cohesion forces correlates well with an increase in interfacial

tension from its original value of 23 mN/m to 31 mN/m after the removal of acids, which implies

that the absence of surface active compounds on the hydrate surface contribute to the increment

on cohesion forces (Sjöblom et al., 2010).

34
Figure 2.8: Summary of cohesion forces between cyclopentane hydrates particles in the presence
of Caratinga crude oil, measured by Dieker et al. (2009).

Additional evidence that associates oil-wet surfaces with low hydrate plugging tendency is

provided by the work of Aspenes et al. (2010a) that studied adhesion forces between cyclopentane

hydrate particles and different solid surface materials. It was concluded that adhesion forces are

higher for water-wet pipeline surfaces, which may be part of the mechanism for hydrate plug

formation, and that addition of petroleum acids to the oil phase drastically reduce adhesion forces.

At this point the importance of the role of surfactants can be seen, with the surfactant funda-

mental property of interfacial adsorption, in the formation of gas hydrates at hydrocarbon/water

interfaces and defining the wettability of hydrate particle and solid surfaces.

2.2 Emulsions

In this section, a brief discussion of fundamental concepts and mechanisms of emulsions behavior

will set the framework to understand the effect of emulsions on hydrate formation/dissociation and

plugging formation tendency in pipelines.

In the petroleum industry, two types of emulsions are frequently encountered: water-in-oil

(W/O) and oil-in-water (O/W) emulsions. Emulsions are defined as thermodynamically unstable

mixtures of two immiscible liquids (one continuous and one dispersed), which means that they will

separate over time. The separation of the immiscible phases will depend on the rate of coalescence

of the dispersed phase. The stability of emulsions can be extended by kinetics means, making the

separation time longer through complex interfacial mechanisms that delay the coalescence of the

35
dispersed phase.

The formation of an emulsion requires at least three components: two immiscible phases (water

and oil) and an emulsifying agent, which may be a natural component of one phase or an emulsifying

agent may be added to one of the phases. The emulsifying agent usually is a surfactant that adsorbs

at the interface between the immiscible phases; some natural surface-active components in crude oils

include asphaltenes, resins with organic acids and bases, naphthenic acids, and carboxylic acids,

among many others (McLean and Kilpatrick, 1997; Sjöblom et al., 2003, 2007). Solid particles

can also act as stabilizing agents, as in examples of sand, clays, mineral incrustations, corrosion

products, paraffin wax, and precipitated asphaltenes (Sjöblom et al., 2003, 2007; Sullivan and

Kilpatrick, 2002).

The rupture mechanism of an emulsion involves three basic steps: (i) long distance approach of

dispersed drops by sedimentation and creaming, (ii) drainage of the thin film between drops, and

(iii) coalescence of drops to become a single larger drop. The kinetic stability of an emulsion is

attained by mechanisms that influence the thin film drainage. Four mechanisms have been listed as

the principal mechanisms acting to stabilize emulsions: (1) electrostatic repulsion between charged

interfaces, (2) steric repulsion, (3) Marangoni-Gibbs effect, and (4) formation of a rigid cross-linked

network adsorbed at the droplet interface (McLean and Spiecker, 1998; Sullivan and Kilpatrick,

2002). A schematic representation of the emulsion stability mechanisms is presented in Figure 2.9.

Figure 2.9: Mechanisms of emulsion stability (adapted from Salager, 2000a).

When two droplets come into contact, the charged surfaces repel each other preventing ag-

glomeration and coalescence (Becher, 2001; Schramm, 2005); this mechanism plays a significant

role in stabilizing O/W emulsions (McLean and Spiecker, 1998; Schramm, 2005). Steric repulsion

prevents approaching droplets from coalescing due to physical barriers of adsorbed material such as

polymer or surfactant chains (Schramm, 2005). The Marangoni-Gibbs effect describes the process

36
whereby an expanding surface shows a subsequent decrease in surfactant concentration and increase

in interfacial tension; the interfacial tension gradient generated provides the driving force for the

flow of surfactant to areas of high tension, opposing film thinning (McLean and Spiecker, 1998;

Schramm, 2005). The adsorption characteristics of the surfactant has a direct effect on the kinetics

phenomena (Salager, 2000a). The formation of rigid cross-linked networks at droplet interfaces is

associated with asphaltenes (McLean and Spiecker, 1998). Even low amount of resins in the oil

phase can improve the stability of emulsions by breaking up large asphaltene aggregates, which

are then more mobile and hence able to adhere at water-oil interfaces forming a strong interfacial

network (Spiecker et al., 2003).

Emulsion properties, such as emulsion type, droplet size distribution, stability and rheology

depend mostly on the following formulation variables: nature of the oil (equivalent alkane carbon

number, EACN), salinity of the aqueous phase (type and concentration of electrolyte), nature and

concentration of surfactant, and temperature. A general relationship may be established between

the generalized formulation and the trend variation, i.e. the increase or decrease of the different

emulsion properties, whose intensity is influenced by the previous mechanisms (Salager, 2000b).

Other variables which influence emulsion properties are pressure, order of mixing, and energy

input.

With the studies of surfactant-water-oil systems, different formulation approaches have been

developed that can be used to predict or estimate the emulsions type and properties (Solans et al.,

1996). Developed from empirical observations, the Bancroft (1912) rule states that the continuous

phase will be the one in which the surfactant is most soluble. Another empirical method is the

Hydrophilic-Lipophilic Balance (HLB), which assigns a number to a surfactant based on its solu-

bility in water (Griffin, 1949). Oil-soluble (lipophilic) surfactants have an HLB number lower than

9, while water-soluble (hydrophilic) surfactants have an HLB number greater than 11. A general

rule-of-thumb is that a surfactant with an HLB ranging 4-6 will tend to form a water-in-oil (W/O)

emulsion and an HLB ranging 8-18 will tend to form an oil-in-water (O/W) emulsion; this method

correlates well with the Bancroft rule. These methods focus only on the nature of the surfactant

and do not consider the effects of other formulation variables (e.g., temperature, salinity of water,

oil phase composition).

37
A theoretical concept was introduced by Winsor (1954) that relates the surfactant with its

physico-chemical environment at the interface between water and oil phases, describing three ele-

mentary types of phase behavior. The Winsor ratio (R = ACO /ACW ) is defined as the ratio of

the net interaction energy of the surfactant with the oil phase (ACO ) to the net interaction energy

of the surfactant to the aqueous phase (ACW ). When the interaction of the surfactant with the

aqueous phase is greater than the interactions with the oil phase (R < 1), a type I phase behavior

results, consisting of a pure oil phase and an oil-in-water microemulsion phase with normal micelles

of surfactant solubilizing oil; this type of system will tend to form an O/W emulsion. If the inter-

action of the surfactant with the oil phase is greater than that of water (R > 1), a type II phase

behavior is exhibited, consisting of a pure aqueous phase and a water-in-oil microemulsion phase

with inverse micelles solubilizing water; this type of system will tend to form a W/O emulsion. The

type III phase behavior corresponds to equal interactions of the surfactant with oil and aqueous

phases (R = 1), consisting of three phases in equilibrium: a microemulsion containing the surfac-

tant which solubilizes large amounts of oil and water, and a pure oil and a pure water phases. A

schematic representation of the three Winsor phase behavior types is presented in Figure 2.10.

Figure 2.10: Schematic representation of phase behavior described by Winsor for different values
of the ratio of net interaction energy of the surfactant with the oil phase to the net interaction
energy of the surfactant to the aqueous phase (Salager, 2000a).

A change in the formulation variables can promote a change in the interaction energies of

the surfactant with the oil and aqueous phases, and therefore a change in the phase behavior. For

example, an increase in the salinity concentration of the water phase would decrease the interaction

energy of an anionic surfactant with the aqueous phase, thus promoting a change from type I to

type III, and a further increase in the salinity will eventually lead to a type II phase behavior

where all the surfactant molecules are in the oil phase. The Winsor R ratio can be used to make

38
qualitatively estimations of the changes in phase behavior, and ultimately in the emulsion type,

when the formulation variables are modified.

The surfactant affinity difference (SAD) can be considered as a generalized formulation parame-

ter that provides a quantitative approach of the relationship between the phase behavior described

by Winsor and the formulation variables (Salager, 1999). The concept of the effect of different

formulation variables in a single expression was presented by Salager et al. (1979) and Bourrel

et al. (1980) which was then called surfactant affinity balance (Salager et al., 1982). Using the

definition of surfactant affinity deviation or difference as the negative of the chemical potential,

SAD = µ◦w − µ◦o , where µ◦w and µ◦o are the standard chemical potentials of the surfactant in the

water and oil phases, respectively. When the surfactant affinity for the water phase equals its affin-

ity for the oil phase, SAD = 0, which is equivalent to a Winsor type III phase behavior (R = 1).

When SAD < 0, the affinity of the surfactant towards water dominates, forming O/W emulsions

(Winsor type I). For SAD > 0, the affinity of the surfactant is stronger for the oil phase, forming

W/O emulsions (Winsor type II). For ionic and nonionic (polyethoxylated) surfactants, the SAD

expressions are as follows,

SAD/RT = ln S − K [(E)ACN ] − f (A) + σs − aT (T − Tref ) (2.3)

SAD/RT = bS − K [(E)ACN ] − f (A) + αs − EON + cT (T − Tref ) (2.4)

where S is the salinity, K is a constant, ACN the alkane carbon number or its equivalent (EACN )

when the oil phase is not characterized by a pure hydrocarbon chain, f (A) is the contribution of the

type and concentration of co-surfactant like alcohol, σs and αs are the characteristic parameters of

the surfactants, EON is the number of ethylene oxide groups in nonionic surfactants, and (T −Tref )

is the temperature difference relative to some reference temperature (Salager, 1999). Each variable

results in a change of at least one of the interaction energies of the surfactant with either aqueous

or oil phase. For instance, an increase in salinity (S) or surfactant lipophilic group length (and thus

σs ) produces an increase in SAD, while an increase in the ACN of the oil phase or in temperature

produces an opposite effect decreasing SAD in eqn. 2.3 for ionic surfactants.

The effect of the formulation variables on the emulsion properties has been studied using the

SAD, considering similar amounts of water and oil and low quantities of surfactant (Salager, 1996).

39
Figure 2.11 summarizes the effect of the formulation variables on the type of emulsion, stability,

interfacial tension, viscosity, electrolytic conductivity and drop size, indicating the three phase

region (Winsor type III phase behavior) with a shaded band in the middle of each diagram. On

the abscissa of each portion Figure 2.11 is the formulation scan shown in the upper left subplot of

the figure. When the formulation variable is increased from SAD < 0, a transition in the emulsion

type from O/W to W/O and in the emulsion properties occur at SAD = 0, because the affinity of

the surfactant switches from hydrophilic to lipophilic. The emulsion stability and viscosity undergo

a minimum at SAD = 0, which is related to a low interfacial tension. The electrical conductivity

decreases drastically inside the three phase region, indicating the phase inversion.

Figure 2.11: Effect of formulation variables on type and properties of emulsions (Salager, 1999).

The previous representation of the effect of a single generalized formulation variable on the

emulsion properties was extended by Salager et al. (1983) to include the effect of the water-oil

ratio at constant surfactant concentration in the form of a bi-dimensional map. Figure 2.12 shows

a diagram of the bidimensional map, where the ordinate represents the generalized formulation

variable SAD and the abscissa represents the water-oil ratio (WOR). From the diagram, it is

possible to predict the emulsion type and the emulsion inversion based on the formulation variable

and the WOR. The formulation divides the map in two halves, with the upper half corresponding

40
to SAD > 0 suggesting a W/O emulsion, and a lower half corresponding to SAD < 0 suggesting an

O/W emulsion. The WOR divides the map in three vertical sections: the left section represents low

WOR (oil continuous emulsion); the right section represents high water content (water continuous

emulsion), and middle section will be controlled by the formulation. There are two sections where

extreme WOR value causes formation of an emulsion type different than the one suggested by

the formulation; these sections are associated with high instability and varying degrees of multiple

emulsions. For the lower left zone (B-) with oil in water in oil (o/W/O) emulsions, the formulation

suggests an O/W emulsion, but the excess oil becomes the continuous phase, hence resulting in a

multiple emulsion that satisfies both trends. Similarly the upper right zone (C+) with water in

oil in water (w/O/W) emulsions is where excess water is the continuous phase surrounding W/O

emulsion droplets.

Figure 2.12: Diagram of generalized formulation-composition bidimensional map (Salager et al.,


1983).

2.2.1 Hydrate formation in emulsions

In oil dominated flowline systems, water-in-oil emulsions play a critical role in the formation

and agglomeration of hydrate particles in pipelines. Previous studies have shown the formation of

hydrates at the interface of water droplets, forming hydrate shells around the water droplet. In

the study of gas hydrate formation with low water cuts (< 35vol%) in W/O emulsions using a

laboratory-scale high-pressure autoclave cell and an industrial-scale flow loop, Turner (2005) found

41
that dispersed water droplets convert directly to hydrate particles (formed from the gas dissolved

in the oil phase), maintaining the droplet size distribution during hydrate formation, and thus

suggesting the formation of a hydrate shell at the water droplet interface; a particle size analyzer,

based on the Focus Beam Reflectance Method (FBRM, Ruf et al., 2000) was used to measure the

water droplet and hydrate particle size distribution. To explain how the hydrates form in the water

interface, Taylor et al. (2007b) studied hydrate film growth at a flat hydrocarbon/water interface

using video microscopy combined with gas consumption measurements, using cyclopentane and

methane as hydrate formers. From the hydrate film growth experiments, they proposed a hydrate

film formation mechanism which consisted of three stages: (i) initial formation of a thin porous

hydrate film propagating across the interface, (ii) film thickening over time growing into the water

phase (shrinking core) and filling of the pores, and (iii) bulk conversion of hydrate film with filling

of the remaining pores. It was also proposed that the hydrate film formation mechanism could

be extrapolated to a water droplet interface following the same three stages and ending with full

conversion of water (inside the shell) into hydrate over a long period of time.

Taylor et al. (2007b) also developed a model to calculate hydrate growth in W/O emulsions

that considered the hydrate shell formation around the water droplets followed by the conversion

of the remaining water core to hydrate, in a similar mechanism to that described above. This

model considered the mass transfer limitations of methane across the boundary layer at the gas-oil

interface, through the oil phase and across the hydrate shell, to determine the hydrate growth rate

in the interior of the water droplet, with the assumption of constant droplet size and ignoring heat

transfer from the hydrate particles to the surroundings due to the exothermic heat of crystallization.

The methane diffusivity through the hydrate shell was obtained from experimental measurements.

The model gave a reasonably good fit to the experimental data, even though the initial hydrate

growth rate was under-estimated and the long term hydrate growth rate was over-estimated, these

errors were attributed to assumptions made in the development of the model.

2.2.2 Effects of hydrates on emulsion stability

The formation and dissociation of hydrates can affect the stability of emulsions, leading to the

generation of a free water phase from the agglomeration and coalescence of water dissociated from

hydrate particles. The presence of a free water phase increases the risk of forming a hydrate plug

42
in pipelines. Lachance et al. (2008) studied the effect of hydrate formation and dissociation in the

stability of W/O emulsions with different crude oils using differential scanning calorimetry (DSC).

From the comparison of results obtained with four crude oils, they found that hydrate formation

and subsequent dissociation destabilized W/O emulsions in oils with a low fraction of asphaltenes

or when a high concentration of resins prevented the deposition of asphaltenes onto the water

droplet interface, while oils with a higher fraction of asphaltenes resisted destabilization. They

proposed a hydrate-induced destabilization mechanism which consisted of a formation step, where

hydrate encrusted water droplets agglomerate, and a dissociation step, where the release of gas

molecules causes the interface to break resulting in coalescence of adjacent droplets. The resistance

to destabilization by oil with higher fractions of asphaltenes can be explained by the formation of

a rigid cross-linked network adsorbed at the droplet interface.

In high water content W/O emulsions (> 35vol%), the formation and dissociation of hydrates

also affect the emulsion stability, leading to an inversion of the emulsion from W/O to O/W or

the generation of multiple emulsions. Greaves et al. (2008) studied emulsions with water content

greater than 60 vol% in a high-pressure autoclave cell equipped with two particle size analyzer

probes (Focus Beam Reflectance Method, FBRM, and Particle Video Microscope, PVM). With

reference to Figure 2.12, these emulsions can be located in the upper half of the bi-dimensional

map (SAD > 0) and near the vertical inversion line. In a similar mechanism to that proposed

by Lachance et al. (2008), hydrate encrusted water droplets first agglomerated and then coalesced

during dissociation, causing the appearance of an aqueous continuous phase due to the large amount

of water in the system, destabilizing the W/O emulsion to form a w/O/W emulsion. Greaves et al.

(2008) also showed using in situ high pressure particle imaging analysis (from the PVM probe)

that hydrate dissociation can result in extensive hydrate agglomeration (see Figure 2.13 of PVM

images of hydrate agglomeration during hydrate dissociation), thereby confirming the calorimetric

results by Lachance et al. (2008) which indicated that hydrate dissociation can lead to emulsion

destabilization and hydrate particle agglomeration.

2.2.3 Agglomeration mechanism of hydrate particles

Different mechanisms have been suggested for the agglomeration of hydrate particles in oil dom-

inated systems. From the expected hydrophilic character of the hydrate surface, it was suggested

43
Figure 2.13: PVM images of 68 vol% of water dispersed in crude oil: a) initial W/O emulsion, b)
hydrate formation around water droplets, c) large hydrate agglomerates during hydrate dissociation,
c) w/O/W emulsion after hydrate dissociation (Greaves et al., 2008).

that capillary forces, formed by a liquid bridge between the particles, hold the particles together

(Austvik et al., 2000). The capillary bridge theory (Israelachvili, 1991) was used as an explanation

of hydrate particle agglomeration in investigations of micromechanical force measurements between

two hydrate particles. The inter-particle cohesion force increased when the temperature was raised

toward the melting temperature, and a quasi-liquid water layer formed at the particle surfaces

(Taylor et al., 2007a; Yang et al., 2004). The capillary bridge was also observed in the study of

micromechanical adhesion forces between hydrate particles and pipeline solid surfaces with a water

drop deposited on the solid surface (Aspenes et al., 2010a), increasing the adhesion force to more

than 10 times of that between hydrate particles, indicating possible deposition of hydrates on pipe

walls through this mechanism.

According to the capillary bridge theory, the capillary force between two hydrate particles is

proportional to the interfacial tension between the bridging liquid water and the continuous oil

phase. A decrease in the interfacial tension between water and oil decreases the capillary forces

and prevent the agglomeration of hydrate particles that may be separated by shear forces during

flow; this is one of the principles of hydrate anti-agglomeration agents (or anti-agglomerants, AA).

44
From studies of hydrate slurry viscosity (Fidel-Dufour et al., 2006), a hydrate agglomeration

mechanism was proposed which includes crystallization of a water droplet by the contact with a

hydrate particle, and growth of the resulting agglomerate with the subsequent contact of water

droplets. This mechanism was used for the development of a hydrate slurry rheology model by

Camargo and Palermo (2002), in which the viscosity of the hydrate slurry increases with agglom-

erate size. The adsorption of a surfactant on the hydrate particle surface, can prevent the contact

with water droplets and other hydrate particles by changing the wettability of the hydrate particle

surface to oil-wet and by creating steric repulsions; this is another principle of anti-agglomerants.

The complete elimination of interfacial agglomeration mechanisms of hydrate particles, like

liquid bridging, has been the principle of new technologies referred to as the stabilized cold flow

concept, in which flowable slurries of dry hydrate particles are generated (Gudmundsson, 2002;

Lund and Larsen, 2000; Talley et al., 2007; Wolden et al., 2005). The conversion of excess free

water into small and dry hydrate particles is the key to eliminate all chances of agglomeration in

the stabilized cold flow concept.

2.2.4 Formulation dependence of hydrate antiagglomeration agents

Anti-agglomerants are surfactants which have the purpose to disperse hydrate particles in the

oil phase and keep them apart, preventing agglomeration. Some mechanisms that contribute to

the anti-agglomeration of solid hydrate particles are common to stabilization of emulsions, like

steric repulsions. Other mechanisms involved in the anti-agglomeration of hydrates consider the

nature of the agglomeration phenomenon, which can be different from the emulsion rupture mech-

anism; studies on anti-agglomerants indicate that emulsification is not required for an effective

anti-agglomerant (York and Firoozabadi, 2008, 2009). From a theoretical analysis of the forces in

agglomeration and stabilization of hydrate particles (i.e., capillary and dispersion forces), Anklam

et al. (2008) concluded that an effective anti-agglomerant is one that: (a) reduces the size of hy-

drate particles through the reduction of water droplet size in the W/O emulsion, (b) decreases the

interfacial tension between water and oil phases to reduce capillary bridge forces, and (c) modifies

the wettability of the hydrate surface to oil-wet by increasing contact angle through the water

phase. These features for an effective anti-agglomerant agent can be attained considering the effect

of formulation variables on emulsion properties, tailoring the surfactant mixture to a particular

45
system, using as a reference a diagram like the one presented in Figure 2.11, where the formulation

conditions that provide minimum droplet size, low interfacial tension, and low emulsion viscosity

can be identified.

Effects of formulation variables on anti-agglomerant surfactants have been published in a limited

number of papers. York and Firoozabadi (2009) studied the effect of adding salt (NaCl and MgCl2 )

on the anti-agglomeration efficiency of an anionic quaternary ammonium salt and a non-ionic

rhamnolipid surfactant, in a model oil/water/tetrahydrofuran system (where tetrahydrofuran is a

model hydrate former). They observed that the addition of salt (NaCl or MgCl2 ) eventually results

in agglomeration of hydrates, with bivalent salts (MgCl2 ) being more detrimental, and non-ionic

biosurfactants being more effective at lower salt concentrations than the anionic surfactant. The

addition of both salts tested increased the SAD, making the surfactant more lipophilic with less

adsorption at the interface. With the anionic surfactant, agglomeration was observed at low salt

(NaCl and MgCl2 ) concentrations and high surfactant concentrations, restoring anti-agglomeration

behavior with the addition of salt until a point where agglomeration was observed again, indicating a

range of concentrations with good anti-agglomeration behavior. Others also studied the effect of salt

on anti-agglomerants (Gao, 2009), but reported an improvement in anti-agglomeration performance

when increasing salt concentration, which is attributed to thermodynamic inhibition of hydrates by

the increased concentration of salt in the unconverted water phase (York and Firoozabadi, 2008).

The addition of low concentrations of an alcohol, which acts as a co-surfactant, can change

the formulation improving the anti-agglomeration property of some surfactants. The addition of

methanol (MeOH, 0.5 – 5.0 wt.%) as a co-surfactant has been proven, at laboratory scale, to

decrease the required concentration of anti-agglomerant surfactant, maintaining the anti agglom-

eration properties of the system (Gao, 2009; York and Firoozabadi, 2008). In one study (York

and Firoozabadi, 2008), the improved anti-agglomeration characteristic was explained by a syn-

ergistic contribution of MeOH that increased the surfactant density at the interface, reducing

further the interfacial tension, and thereby decreasing water droplet size conducing to small hy-

drate particles. This change in the formulation modified the emulsion properties to coincide with

a minimum in droplet size as shown in Figure 2.11. In another study (Gao, 2009), it was shown

that anti-agglomeration was improved by adding 15 wt.% of MeOH and 0.5 wt.% of (un-identified)

anti-agglomerant to a system with 80% water cut, explaining the anti-agglomeration improvement

46
by the increase of MeOH concentration in the free water phase that thermodynamically inhibits

further formation of hydrates.

2.3 Summary

Typical hydrate prevention methods, based on shifting the hydrate equilibrium phase boundary

by adding thermodynamic inhibitors, focus on the treatment of the aqueous bulk phase requiring

large concentrations of methanol or monoethylene glycol. The gas and oil industry is currently

in a transition from hydrate avoidance to risk management with the use of inhibitors that act at

the interfaces of hydrate particles (kinetic hydrate inhibitors and anti-agglomerants), which has

been proven to reduce the required concentrations of chemical additives, i.e., a more economical

approach. This fact highlights the importance of considering surfaces and interfaces in the study

of gas hydrates.

Revisiting the conceptual model for hydrate formation in a multiphase flow system consisting

of water, oil, and gas presented in Figure 2.1, it is found that interfaces between water and hy-

drocarbon phases are ideal locations for gas hydrate formation, and that interaction mechanisms

between hydrate particles and solid surfaces may or may not result in the blockage of pipelines,

depending on several factors, such as oil composition (presence of natural surfactants), wettability

of solid surfaces, formulation variables and their effect on anti-agglomerants performance, presence

of free water, and particle size among others. Understanding the fundamental concepts of surface

chemistry, in particular surfactants (molecules with preference to interfaces) and emulsions (mix-

tures of two immiscible fluids with abundant surface area), is of great value for the research, design,

and development of novel techniques for the management of gas hydrates in oil and gas pipelines.

In this review the relationship between surface chemistry and gas hydrates was presented using

a number of studies available in the technical literature to provide a clear connection of the current

advances in the study of hydrates with interfacial phenomena.

47
CHAPTER 3

MULTIPHASE FLOW IN PIPELINES AND HYDRATE FORMATION

Observations in flow loop experiments of gas hydrate formation in high water cut systems showed

that pressure drop in pipelines depends on a complex flow regime involving multiple phases and solid

dispersion in liquid (Joshi et al., 2011a). These observations require merging two known phenomena

of multiphase flow: (i) multiphase (gas-liquid) flow regimes, which can be useful in estimating

pressure drop in the pipeline and surface area between water and hydrocarbon phases; and (ii)

slurry flow consisting of solid particles dispersed in the liquid phase during hydrate formation.

The purpose of this chapter is to present a review of fundamental concepts and models of steady-

state and transient multiphase flow in pipes, and identify state-of-the-art methods that could be

extended for the prediction of gas hydrate formation and transportability in flowlines.

This chapter starts with a review of single phase flow concepts, presented in Section 3.1. The

second section (Section 3.2) presents an extension of single phase flow concepts to multiple phases,

introducing the fundamental concepts used to describe multiphase flow in pipes. Section 3.3 presents

a discussion of multiphase flow modeling approaches. Finally, Section 3.4 presents the derivation of

a simple hydrodynamic slug flow model and its extension to predict hydrate formation, to illustrate

the effects of hydrate formation on multiphase flow.

3.1 A review of single-phase flow concepts

This section presents the derivation of the steady-state pressure-gradient equation for single-

phase liquid flow in pipes.

The one-dimensional mass conservation equation is,

∂ρ ∂ρv
=− (3.1)
∂t ∂x
For steady-state flow, the transient term disappears and the conservation of mass is given by,

∂ρv
=0 (3.2)
∂x

48
The conservation of momentum is,

∂ρv ∂ρv 2 ∂p πd
=− − − τw − ρg sin θ (3.3)
∂t ∂x ∂x Ap
The pressure-gradient equation is obtained by combining the mass and momentum balance

equations, and assuming steady-state flow,

∂p πd ∂v
= −τw − ρg sin θ − ρv (3.4)
∂x Ap ∂x
where p is pressure, x is the axial coordinate along the length of the pipe, τw is the wall shear

stress, d is the pipe internal diameter, Ap is the pipe cross sectional area, ρ is the fluid density, θ is

the pipe angle of inclination with respect to the horizontal, v is the fluid linear velocity. The first

term on the right side of equation 3.4 represents the friction or shear stress at the pipe wall (5 to

20% of total pressure drop), the second term is the pressure drop due to elevation change (80 to

95% of total pressure drop in vertical flow), the third term represents the change in velocity (i.e.,

kinetic energy), which could be negligible (Brill and Mukherjee, 1999).

The wall shear stress, τw , can be evaluated through the use of a dimensionless friction factor.

In this case, the Moody friction factor is used, defined as the ratio of wall shear to kinetic energy

of the fluid per unit volume,


τw
f= (3.5)
ρv 2 /8
Solving the wall shear stress from equation 3.5 and substituting into the frictional part of the

pressure-gradient equation,

f ρv 2 πd f ρv 2
 
∂p πd
= τw = = (3.6)
∂x f Ap 8 πd2 /4 2d
Next, the friction factor needs to be determined. For laminar flow, the friction factor is a

function of the Reynolds number (Re = ρvd/µ),


64
f= , for Re < 2000 (3.7)
Re
For turbulent flow, the friction factor is a function of Reynolds number and pipe wall character-

istics (i.e., relative roughness, ε/d). For smooth pipe (ε/d → 0) the friction factor can be evaluated

as a function of the Reynolds number only, using the following correlations,

f = 0.0056 + 0.5Re−0.32 , for 3000 < Re < 3 × 106 (Drew and McAdams, 1932) (3.8)

49
or,

f = 0.316Re−0.25 , for 3000 < Re < 105 (Blasius, 1908) (3.9)

For complete turbulence and fully rough wall pipe (i.e., region in Moody diagram where f is a

function of ε/d only), the friction factor can be evaluated using Nikuradse’s (1933) correlation,
 
1 2ε
√ = 1.74 − 2 log (3.10)
f d
In the transition region between laminar and fully turbulent flow, the friction factor is a function

of both the Reynolds number and relative roughness, and can be evaluated using the Colebrook

(1939) empirical equation,


 
1 2ε 18.7
√ = −2 log + √ (3.11)
f d Re f
Solving equation 3.11 for f requires an iterative procedure. A few iterations are needed if a

direct substitution method is used, where the calculated value is used as the next assumed value.

The initial guess value can be obtained from one of the explicit smooth-pipe equations or from the

following explicit approximation to the Colebrook equation (Zigrang and Sylvester, 1985),
  
1 2ε/d 5.02 2 ε/d 13
√ = −2 log − log + (3.12)
f 3.7 Re 3.7 Re
Initial values for roughness, ε, are needed for design calculations, some suggested values found

in Brill and Mukherjee’s (1999) monograph are:

• For new tubing, ε = 1.524 × 10−5 m

• Common value to generate pressure-gradient curves, ε = 4.572 × 10−5 m

• For tubing exposed to an environment that causes significant changes in roughness, very dirty

pipe, ε = 2.286 × 10−4 m

3.2 Multiphase flow concepts

The term multiphase flow refers to any fluid flow consisting of more than one phase. One of the

simpler cases of multiphase flow, which have been the subject of numerous studies, consists of two-

phase gas/liquid flow (Brill and Mukherjee, 1999). This case is used to introduce the fundamental

multiphase flow concepts and the different approaches used to model these systems. The presence

50
of dispersed phases within continuous phases presents a subsequent level of complexity. Three

phases exist during the transport of hydrocarbons in pipelines; gas, oil and water. Depending on

the amount and flow rates of each phase, different configurations can be observed; these will be

discussed with the modeling of transient multiphase flow in Section 3.3.3.

The holdup for each phase (gas or liquid) is defined as the fraction of pipe cross sectional area

(Ap ) or volume increment that is occupied by each phase.

Ak
Hk = (3.13)
Ap
where the subscript k denotes the phase, either gas (g) or liquid (L). In general, empirical correla-

tions for multiphase flow are based on the prediction of liquid holdup (HL ).

The sum of liquid and gas holdup equals one,

HL + Hg = 1 (3.14)

An important correlating parameter is the superficial velocity, which assumes that a given phase

occupies the entire pipe area, and it is defined as the volumetric flow rate of the phase (Qk ) divided

by the cross sectional area of the pipe,

Qk
vSk = (3.15)
Ap
The mixture velocity is defined as the sum of the superficial velocities.

QL + Qg
vm = = vSL + vSg (3.16)
Ap
The phase linear velocities are related to the superficial velocities and the liquid hold up by,
vSk
vk = (3.17)
Hk
The gas-liquid slip velocity is defined as the difference between the linear velocities,
vSg vSL
vs = vg − vL = − (3.18)
1 − HL HL
A condition of no-slip between the phases indicates that both the gas and liquid would flow at

the mixture velocity.

Flow pattern or flow regime is a classification of the different physical distributions of the

phases in the flow conduit. Figure 3.1 shows different flow regimes identified for gas/liquid flow in

horizontal pipes, the flow regimes for vertical pipes are shown in Figure 3.2.

51
Figure 3.1: Flow regimes for gas/liquid flow in a horizontal pipe (Brennen, 2005, p. 167).

Figure 3.2: Flow regimes for gas/liquid flow in a vertical pipe (Brennen, 2005, p. 170).

52
Another type of two-phase flow is solid/liquid mixtures or slurry flow, which consists of a

dispersion of solid particles in a continuous liquid phase. With increasing solid particle volume

fraction, different slurry flow regimes can occur, as illustrated in Figure 3.3 for sand and water.

From a homogeneous distribution of well-mixed solid particles in the liquid phase at low particle

concentration, to a heterogeneous distribution with faster sedimentation of larger particles gener-

ating a vertical gradient in the particle concentration, to a moving bed of particles, and finally to

a stationary bed where the packing of particles prevents its movement in the pipe.

Figure 3.3: Flow regimes for slurry (solid particles of sand dispersed in water) flow in a horizontal
pipe (Brennen, 2005, p. 169).

3.3 Multiphase flow modeling

The methods used to model multiphase flow vary widely according to the assumptions made

in their development. The homogeneous fluid model approach assumes that the two-phase flow

can be treated as single phase flow by using average properties for the mixture. In the separated-

flow approach, a steady-state momentum balance equation is developed for gas and liquid phases,

assuming that the two phases flow through separate conduits, and either equation can be used to

calculate pressure drop. The phenomenological or mechanistic approach used to develop steady-

state models represents a compromise between empirical correlations and conservation laws, where

basic physical laws are used to model flow phenomena, such as flow pattern prediction. The

transient two-phase flow modeling approach involves the simultaneous solution of conservation

equations of mass, momentum and energy for each phase (gas and liquid), and uses empirical

correlations and simplified constitutive relationships for flow parameters, such as interfacial friction,

fraction of liquid entrained in the free gas phase, and liquid holdup in the slug body.

53
3.3.1 Homogeneous fluid model approach

The single-phase pressure gradient equation can be modified for multiphase flow by considering

the fluids to be a homogeneous mixture, as follows,

dp f ρf vf2 dvf
= + ρf g sin θ + ρf vf (3.19)
dx 2d dx
where the subscript f denotes the homogeneous fluid. The homogeneous fluid properties are cal-

culated using different mixing rules depending on the author of the method (Brill and Mukherjee,

1999). The pressure-drop component caused by friction losses requires evaluation of a two phase

friction factor. The pressure-drop caused by elevation change is dependent on the two phase mixture

density. The pressure-drop caused by acceleration is normally negligible.

The assumption of homogeneous fluid ignore the fact that gas generally flows faster than liquid

(i.e., no-slip conditions), which results in under-prediction of the pressure drop because the liquid

holdup predicted in the pipe is smaller than the actual liquid holdup (Brill and Mukherjee, 1999).

Some methods based on empirical correlations were designed to account for the slip between the

gas and liquid phases, by calculating the liquid holdup as function of the multiphase flow regime.

Brill and Mukherjee (1999) presented a classification of different empirical correlations based on

the homogeneous fluid model approach, to predict pressure drop of multiphase flow in pipes. The

first class was called Category “a”, the correlations in this class assume the no-slip condition (gas

and liquid have equal velocity) and does not consider flow patterns. The properties of the mixture

are calculated based on the input gas/liquid ratio.

The second class was called Category “b”, the methods in this class consider slip between the

gas and liquid phases, but do not consider flow patterns. The methods in Category “b” offer

correlations for the two-phase friction factor and liquid holdup.

The third class was called Category “c”, the methods in this class consider gas-liquid slip and

flow patterns. A procedure to determine the flow pattern is required before the correlations for

liquid holdup and two-phase friction factor can be used. In these methods the acceleration pressure-

drop term is also dependent of flow pattern. Table 3.1 presents a list of the empirical correlations

for multiphase flow classified by Brill and Mukherjee (1999).

54
Table 3.1: Classification of empirical correlations for pressure drop prediction in multiphase flow
based on the homogeneous fluid model (Brill and Mukherjee, 1999).

Category Empirical correlation


Poettman and Carpenter (1952)
“a” Considers no-slip, not
Baxendell and Thomas (1961)
accounting for flow pattern
Fancher and Brown (1963)
“b” Considers slip, not Hagedorn and Brown (1965)
accounting for flow pattern Asheim (1986)
Duns and Ros (1963)
Orkiszewski (1967)
“c” Considers slip and flow Aziz et al. (1972)
pattern Chierici et al. (1974)
Beggs and Brill (1973)
Mukherjee and Brill (1983)

There is another approach for the homogeneous fluid model that, instead of using a friction factor

correlation uses a two-phase correction factor multiplier. This approach consists of estimating the

single-phase frictional pressure drop for the liquid phase only, and multiplying this by a two-phase

correction factor,
   
dp dp
= φ2LO (3.20)
dx f dx f,LO

where φ2LO is the two-phase correction factor, and the subscript LO denotes liquid only. The

superficial velocity of the liquid phase is used in the calculation of the single-phase frictional pressure

drop. An example of a two-phase correction factor is (Balasubramaniam et al., 2006),

µL − µg −1/4
  
ρL ρL − ρg
φ2LO = 1 + Hg 1 + Hg (3.21)
ρf ρg µg

3.3.2 Separated-flow model approach

The separated-flow model approach is specific for stratified flow. In the separated-flow model

approach the gas and liquid phases are assumed to flow in separate conduits with cross sectional

area equal to the area occupied by the phases. Steady-state momentum balance equations are

written for gas and liquid phases,

dp (τw,k Pk + τi Pi )
=− − ρk g sin θ (3.22)
dx Ak

55
where τw,k is the wall shear stress of phase k, evaluated assuming single-phase flow in a conduit

with cross sectional area equal to the area occupied by the phase (Ak ), Pk is the wetted perimeter

of phase k, τi is the interfacial shear stress, Pi is the width of the flat interface between the phases.

An empirical correlation based on the separate flow model approach is the Lockhart and Mar-

tinelli (1949) correlation. The Lockhart-Martinelli correlation was developed from horizontal flow

of air-liquid mixtures at near-atmospheric pressure. It assumes the acceleration term of the pres-

sure gradient equation (eqn. 3.19) negligible, the pressure drop component by elevation changes

vanishes, leaving only the friction component of the pressure drop equation. From the calculation

of “only gas” and “only liquid” reference flows they derived equations for the frictional pressure

gradient in the two-phase flow.


     
dp 2 dp 2 dp
= φGO = φLO (3.23)
dx f dx f,GO dx f,LO
The two-phase correction factors, φ2GO and φ2LO , are functions of the Lockhart-Martinelli pa-

rameter (X),

φ2GO = 1 + CX + X 2 ; φ2LO = 1 + CX −1 + X −2 (3.24)

The Lockhart-Martinelli parameter is defined as the ratio of the single-phase frictional pressure

gradients for the liquid and gas phases alone.

(dp/dx)f,LO
X= (3.25)
(dp/dx)f,GO

3.3.3 Transient multiphase flow in pipe modeling approach

The previous two-phase flow modeling approaches used separate empirical correlations for flow

regime, liquid holdup and friction factor, but it is important to note that these parameters are

physically inter-related. During the development of transient two-phase flow models an attempt

was made to unify the treatment of liquid holdup, pressure drop and flow regimes, as an integral

part of the model (Bendiksen et al., 1991).

The transient two-phase modeling approach involves the simultaneous solution of conservation

equations of mass, momentum and energy for each phase (gas and liquid), and uses empirical

correlations and simplified constitutive relationships for flow parameters, such as interfacial friction,

fraction of liquid entrained in the free gas phase, and liquid holdup in the slug body. Bendiksen

56
et al. (1991) presented a transient two-phase (gas and liquid) flow model that considered three

separate mass conservation equations for the gas, liquid bulk and liquid droplets dispersed in the

gas phase, which are coupled by interfacial mass transfer; a combined momentum equation for the

gas plus liquid droplets and a separate momentum equation for the liquid film; and a simplified

mixture energy equation considering all the phases as a homogeneous fluid. The mass conservation

equation for the gas phase is,

∂ 1 ∂
(Hg ρg ) = − (Ap Hg ρg vg ) + ψg + Sg (3.26)
∂t Ap ∂ x
For the liquid film the mass conservation equation is,

∂ 1 ∂ HL
(HL ρL ) = − (Ap HL ρL vL ) + ψg − ψe + ψd + SL (3.27)
∂t Ap ∂x HL + HD
The mass conservation equation for liquid droplets is,

∂ 1 ∂ HD
(HD ρL ) = − (Ap HD ρL vD ) + ψg + ψe − ψd + SD (3.28)
∂t Ap ∂ x HL + HD
where the phases are denoted by subscripts g, L and D for gas, liquid film and liquid droplets,

respectively; Sk is the external source/sink term of phase k; ψg is the mass transfer rate between

gas and liquid phases; ψe is the liquid droplet entrainment rate into the gas phase; and ψd is the

liquid droplet deposition rate into the liquid phase.

The linear momentum conservation equation for the gas phase is,

∂ ∂p 1 ∂  fg ρg vg2 Pg
(Hg ρg vg ) = −Hg − Ap Hg ρg vg2 −
∂t ∂x Ap ∂x 8 Ap
2
fi ρg vs Pi
− − Hg ρg g sin θ + ψg va − FD (3.29)
8 Ap

The linear momentum conservation equation for liquid droplets is,

∂ ∂p 1 ∂ 2

(HD ρL vD ) = −HD − Ap HD ρL vD − HD ρL g sin θ
∂t ∂x Ap ∂x
HD
−ψg va + ψe vi − ψd vD + FD (3.30)
HL + HD

The combined momentum equation for the gas plus liquid droplets result from adding equations

3.29 and 3.30,

57
 
∂ ∂p 1 ∂
Ap Hg ρg vg2 + Ap HD ρL vD
2

(Hg ρg vg + HD ρL vD ) = − (Hg + HD ) −
∂t ∂x Ap ∂x
2
fg ρg vg Pg 2
fi ρg vs Pi
− − − (Hg ρg + HD ρL ) g sin θ
8 Ap 8 Ap
HL
+ψg va + ψe vi − ψd vD (3.31)
HL + HD

Note that the gas/droplet drag terms (FD ) cancel in the combined gas/liquid droplet momentum

equation.

The linear momentum conservation equation for the liquid film is,

∂ ∂p 1 ∂  fL ρL vL2 PL fi ρg vs2 Pi
(HL ρL vL ) = −HL − Ap HL ρL vL2 − +
∂t ∂x Ap ∂x 8 Ap 8 Ap
HL
−HL ρL g sin θ − ψg va − ψe vi + ψd vD
HL + HD
∂HL
+HL d (ρL − ρg ) g cos θ (3.32)
∂x

Here vs is the slip velocity between the gas and liquid film; Pg , PL , and Pi are the wetted

perimeter of the gas, liquid and interface, respectively; fg , fL and fi are the friction factors of gas,

liquid and interface, respectively. The velocity, va , associated with the mass transfer rate term, ψg ,

depends on the mass transfer process:

• For evaporation from the liquid film, ψg > 0 and va = vL

• For evaporation from the liquid droplets, ψg > 0 and va = vD

• For condensation, ψg < 0 and va = vg

The droplet velocity is defined as,

vD = vg + v0D sin θ (3.33)

where v0D is the fall velocity of droplets. The interface velocity, vi , is approximated by the liquid

velocity, vL .

A pressure equation is obtained by expanding the mass conservation equations with respect to

pressure, temperature and composition, dividing the expansions by each phase density and adding

the equations.

58
     
Hg ∂ρg 1 − Hg ∂ρL ∂p 1 ∂
+ =− (Ap Hg ρg vg )
ρg ∂p T ρL ∂p T ∂t Ap ρg ∂x
 
1 ∂ 1 ∂ 1 1
− (Ap HL ρL vL ) − (Ap HD ρL vD ) + ψg −
Ap ρL ∂x Ap ρL ∂x ρg ρL
Sg SL SD
+ + + (3.34)
ρg ρL ρL

This pressure equation along with the momentum conservation equations can be solved simul-

taneously for pressure and phase velocities.

The total fluid mixture energy conservation equation is,

    
∂ 1 2 1 2
Hg ρg eg + vg + gh + HL ρL eL + vL + gh
∂t 2 2
    
1 2 ∂ 1 2
+HD ρL eD + vD + gh =− Hg ρg vg hg + vg + gh
2 ∂x 2
   
1 2 1 2
+HL ρL vL hL + vL + gh + HD ρL vD hD + vD + gh + hS + q (3.35)
2 2

where ek is internal energy per unit mass of phase k, hk is enthalpy of phase k, hS is enthalpy from

mass sources, q is heat transfer through the pipe wall, and h is elevation. The energy conservation

equation is used to determine the average fluid temperature along the length of a pipeline.

The mass transfer rate between gas and liquid phases is a function of the gas mass fraction

(Rs ),

   
∂Rs ∂p ∂Rs ∂p ∂x
ψg = +
∂p T ∂t ∂p T ∂x ∂t
    #
∂Rs ∂T ∂Rs ∂T ∂x
+ + (Hg ρg + HL ρL + HD ρL ) (3.36)
∂T p ∂t ∂T p ∂x ∂t

where Rs is defined as,

Hg ρg
Rs = (3.37)
(Hg ρg + HL ρL + HD ρL )
In the mass transfer rate equation (3.36), the first term represents the mass transfer caused

by a pressure change in the same section, the second term represents the mass transfer caused by

mass flowing out of the section driven by a pressure gradient, the third term represents the mass

transfer caused by a temperature change in the same section, and the fourth term represents the

59
mass transfer caused by mass flowing out of the section driven by a temperature gradient.

For this model, the fluid properties (density, compressibility, viscosity, surface tension, enthalpy,

heat capacity, and thermal conductivity) are generated a priori using a thermodynamic fluid prop-

erty package (e.g., Multiflash


R
, and PVTSim
R
), in the form of tables for a range of pressures and

temperatures. For a given pressure and temperature condition the properties are interpolated from

the tables.

Bendiksen et al. (1991) considered four multiphase flow regimes during the development of this

model, which are stratified, annular, bubble and slug. The transition between the flow regimes

is determined by assuming a continuous average gas holdup and selecting the flow regime with

the minimum gas velocity (i.e., a minimum slip constraint). Since friction factors and wetted

perimeters are dependent on flow regime, they implemented a series of proprietary semi-empirical

relations to calculate flow parameters such as: average wave height in stratified wavy flow, wall and

interfacial friction factors, droplet deposition and entrainment rates, slip velocity between gas and

liquid phases, slug bubble velocity, and gas fraction in liquid slugs.

The transient two-phase model presented by Bendiksen et al. (1991) have evolved into a com-

mercially available state-of-the-art transient multiphase flow model, called OLGA


R
, which consid-

ers more than two phases. This model accounts for three pre-defined phases, gas, oil and water,

which are distributed into different mass fields. Figure 3.4 presents a schematic representation of

the mass fields considered, which includes gas, oil and water as continuous phases, and dispersed

phases within continuous phases (i.e., oil and water droplets dispersed in continuous gas, water

droplets and gas bubbles dispersed in continuous oil, and oil droplets and gas bubbles dispersed in

continuous water). This allows for a closer representation of the production stream of hydrocarbon

pipelines.

Similarly to the two-phase model presented by Bendiksen et al. (1991), the OLGA
R
model

considers separate mass conservation equations for each mass field, a combined momentum con-

servation equation for each flowing layer (i.e., each continuous phase with its dispersed phases),

one pressure equation, and one mixture energy conservation equation. The model uses a sequential

splitting scheme to solve the conservation equations in subsequent steps for each time step, starting

with the simultaneous solution of the momentum and pressure equations to obtain pressure and

velocities, then solving the set of mass conservation equations to get phase masses, and finally the

60
Figure 3.4: Illustration of basic mass fields in a transient multiphase flow model, including gas, oil
and water as continuous phases, and dispersed phases within continuous phases (i.e., oil and water
droplets dispersed in continuous gas, water droplets and gas bubbles dispersed in continuous oil,
and oil droplets and gas bubbles dispersed in continuous water).

energy equation to solve for the mixture temperature.

A more flexible flow model framework was developed in the HORIZON JIP program for OLGA
R

that allows one to include the effects of a user-defined dispersed phase on the fluid flow behavior.

The new flow model framework adds two additional layers to the model, in the form of a bedding

layer and a deposited wall layer. A user-defined plug-in program can be coupled with OLGA
R

to define a new phase and calculate additional contributions to the conservation equations. The

plug-in consists of the following basic modules:

• The PVT module, used to calculate properties (such as density, viscosity, enthalpy, heat

capacity, and thermal conductivity) for the additional user-defined phases, and can modify

the properties of the pre-defined phases (gas, oil and water).

• The rheology module, used to calculate the apparent viscosity of a layer as a function of the

user-defined phase concentration in the layer, and the particle velocity.

• The flash module, used to calculate the mass transfer rate between the phases and mass fields.

• The entrainment/deposition module, used to calculate the mass transfer rate of the same

phase between different mass fields, for example, mass transfer from a continuous phase to a

dispersed phase into another continuous phase.

61
• The heat transfer module, used to calculate the heat transfer coefficient of a deposited wall

layer, and the corrected inner heat transfer coefficient; this only applies when the user-defined

phase can deposit on the pipe wall and the flash module accounts for this phenomena.

The user must set PVT data for user-defined phases, and can modify the properties of the pre-

defined phases (gas, oil and water) using the PVT module. It is possible to have balance equations

for statistical moments for each mass field, which are used for tracking particle size distribution.

The main purpose of the plug-in modules is to calculate contributions to the conservation

equations. A plug-in can contribute to the mass conservation equations through the inter-field mass

transfer terms, either by flashing (change of phase and mass-field) or entrainment/deposition (same

phase, change of mass-field). A plug-in contribution to the momentum conservation equations is

done through the calculation of modified viscosities for the flow layers, accounting for concentration

of particles. The modified viscosity will induce modifications in the friction factors. The particle

slip velocity with respect to the continuous phase may be calculated in the plug-in, which will

modify the resulting friction factor as well. The only contribution of a plug-in to the pressure

equation is through the mass transfer rate terms. OLGA


R
calculates one temperature for the fluid

and also calculates the temperature of the pipe wall, assuming radial heat conduction through

concentric wall layers. An internal heat transfer coefficient is calculated for the heat transfer from

the fluid to the inner pipe wall. The presence of an additional wall layer due to deposition of the

user-defined phase could add resistance to the overall heat transfer, and this is the contribution of

the plug-in to the energy conservation equation.

3.4 Effect of hydrate formation on multiphase flow

This section illustrates how multiphase flow characteristics for simple systems (gas and water)

may change when a gas hydrate solid phase is introduced. It has been observed experimentally

that, for a system already near the stratified/slug transition point, the formation of a small quantity

of hydrate may immediately lead to slug flow onset (Joshi, 2012). The ability to provide a first-

principles estimation of flow regime transitions and slugging dynamics represents a critical step

forward for overall multiphase flow modeling with hydrate suspensions.

A simple multiphase flow model, developed by Danielson (2011), is used as a starting point,

which is capable of predicting stable hydrodynamic slug flow based on fundamental multiphase flow

62
concepts. The hydrodynamic slug flow model is extended by including an additional mass balance

equation for the gas hydrate phase, allowing the introduction of the effects of gas hydrates as a

third phase. A transient hydrate kinetics model is used to calculate the gas hydrate growth rate,

and forms the basis for the mass transfer rate between the phases. The resulting model represents

an improvement on the commonly used steady-state assumption for both pressure drop and liquid

holdup in slugging systems, using instead the transient pressure drop and liquid holdup. Using fluid

properties and flow parameters, a mixture energy conservation equation is introduced to update the

temperature profile of the pipeline, accounting for the exothermic nature of the hydrate formation

reaction. The performance of the proposed model is compared against experimental data obtained

from experiments in an industrial scale flow loop.

3.4.1 Hydrodynamic slug flow model

A new approach was presented by Danielson (2011) to model two-phase flow, which is based on

the multiphase flow concepts presented in Section 3.2. The definition of slip velocity, vs = vg − vL ,

can be rearranged to give the holdup equation as follows,


vSg vSL
vs = − (3.38)
1 − HL HL
Substituting the gas superficial velocity by vSg = vm − vSL , multiplying by (1 − HL ), and then

multiplying by HL , the following holdup equation is obtained,

vs HL2 + (vm − vs ) HL − vSL = 0 (3.39)

The mass conservation equation for the liquid phase can be written as,

d (ρL Ap HL ) (ρL Ap vSL )IN − (ρL Ap vSL )OU T


= (3.40)
dt dx
Assuming the liquid density, pipe cross sectional area and section length constants, the mass

conservation equation simplifies to the volume conservation equation,

dHL (vSL )IN − (vSL )OU T


= (3.41)
dt dx
Using a moving reference frame that moves at the velocity of the gas bubble, vg , in the stratified

region of slug flow, the superficial liquid velocity is redefined as,

0
vSL = vSL + vg HL (3.42)

63
Assuming constant slug propagation is equal to the gas bubble velocity, vg , the volume conser-

vation equation can be written as,

dHL (v 0 + vg HL )IN − (vSL


0 +v H )
g L OU T
= SL (3.43)
dt dx
Considering a pipe segment divided into equally spaced sections and using the node numbering

convention shown in Figure 3.5, where node i is the solution node, the volume conservation equation

(3.43) can be expanded as follows using finite differences,


   (n−1)
(n) (n−1)
HL,i − HL,i 0
vSL,i 0
− vSL,i+1 + vg (HL,i−1 − HL,i )(n−1)
= (3.44)
∆t ∆x

Figure 3.5: Pipe segment divided into equally spaced sections, showing node numbering convention.

Note that n indicates the time step level of the temporal discretization and i indicates the node
0 ,
index for the spatial discretization. It is assumed that the superficial liquid velocity term, vSL

moves from right to left in the moving reference frame, and the vg HL terms move from left to right.

Using a first order upwind scheme to approximate the derivative of the superficial liquid velocity

terms in the moving reference frame, and a third order upwind scheme to approximate the derivate

of the vg HL terms, the volume conservation equation is,

 
(n) (n−1)
HL,i − HL,i (n−1)
0 0
∆x = vSL,i − vSL,i+1
∆t
(−HL,i−2 + 6HL,i−1 − 3HL,i − 2HL,i+1 )(n−1)
+vg (3.45)
6

Equation 3.45 can be solve explicitly for the liquid holdup in the new time step n,

(n) ∆t n 0
(n−1) 0
(n−1)
HL,i = HL,i + vSL,i − vSL,i+1
∆x )
(−HL,i−2 + 6HL,i−1 − 3HL,i − 2HL,i+1 )(n−1)
+vg (3.46)
6

64
The superficial liquid velocity at the moving reference frame could be expressed as a function

of the liquid holdup using the holdup equation (eqn. 3.39),

0 0
= vs HL2 + vm

vSL − vs HL − vg HL (3.47)

where the mixture velocity in the moving reference frame is,

0
vm = vm − vg (3.48)

From a drift-flux formulation, the mixture velocity in the moving reference frame can be ex-

pressed in terms of the drift-flux velocity, vo ,

0
vm = −vo (3.49)

Substituting,

0
vSL = vs HL2 − (vo + vs ) HL (3.50)

Note that the slip velocity is also a function of liquid holdup, and the following relationship

could be used,

vs = aHL2 + bHL + c (3.51)

where a, b and c are coefficients that can be determined from experimental data.

This hydrodynamic slug model was coded in Matlab, to reproduce the results reported by

Danielson (2011). Table 3.2 presents the input parameters used in the example. In this case,

a periodic boundary condition is used, where the fluids that exit at the right boundary cell is

reintroduced at the left boundary cell. The case is initialized with a liquid holdup of 0.4 and a

small perturbation (HL = 0.41) is introduced in the 20th node. Figure 3.6 shows the distribution of

liquid holdup along the pipeline length at four different times obtained using the hydrodynamic slug

model. The small perturbation in the interface between gas and fluid grows into a hydrodynamic

slug. The holdup in the stratified region is HL = 0.2692.

3.4.2 Extension of a simple hydrodynamic slug flow model for transient hydrate ki-
netics

This section presents the derivation of a three-phase (gas/liquid water/hydrate) hydrodynamic

slug flow model. This new model is an extension of the simple two-phase flow model presented in

the previous section (Section 3.4.1), to study the effect of hydrate formation on the flow regime in

65
Table 3.2: Input parameters for hydrodynamic slug flow model.

Input parameter Value


Gas superficial velocity (m/s) 3.0
Liquid superficial velocity (m/s) 1.0
Pipe diameter (m) 0.2
Pipe length (m) 500
Pipe section length, dx (m) 1.0
Slip velocity parameters a = 0; b = −4; c = 4
Drift flux velocity vo (m/s) 1.0
Time step (s) 0.04
Simulation time (s) 800

1 1

0.8 0.8
Liquid Holdup

Liquid Holdup

0.6 Small perturbation 0.6

0.4 0.4

0.2 0.2

0 0
0 100 200 300 400 500 0 100 200 300 400 500
Pipeline length (m) Pipeline length (m)
(a) Time = 0 s. (b) Time = 100 s.

1 1

0.8 0.8
Liquid Holdup

Liquid Holdup

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 100 200 300 400 500 0 100 200 300 400 500
Pipeline length (m) Pipeline length (m)
(c) Time = 500 s. (d) Time = 700 s.

Figure 3.6: Liquid holdup distribution along the pipeline length at four different times: (a) t = 0.04
s, (b) t = 100 s, (c) t = 500 s, (d) t = 700 s.

66
flowlines. The extended model includes a transient hydrate kinetics model (Turner et al., 2005),

allowing the prediction of hydrate formation, and is the base for mass transport between the phases.

For hydrate formation it is necessary to estimate the pressure and temperature conditions

along the flowline length. In this work, a simple pressure drop model based on the Lockhart and

Martinelli’s (1949) correlation is used to predict the pressure distribution along the pipeline and

an energy balance equation allows the prediction of the temperature.

In order to introduce gas hydrate as a third phase in the flow model, the multiphase flow

fundamental concepts presented in Section 3.2 are modified based on the conceptual picture shown

in Figure 3.7.

Figure 3.7: Conceptual picture used to define the phases and multiphase flow concepts.

First, the holdup for the gas hydrate phase (Hhyd ) is defined as the fraction of pipe cross

sectional area (Ap ) or volume increment that is occupied by the gas hydrate phase (Ahyd ),

Ahyd
Hhyd = (3.52)
Ap
The sum of the phases holdup equals one, Hg + Hw + Hhyd = 1. This phase summation allows

specifying the gas holdup in terms of the other two phases (Hg = 1 − Hw − Hhyd ).

The mixture velocity (vm ) is defined as the sum of the superficial velocities,
qg + qw + qhyd
vm = = vSg + vSw + vShyd (3.53)
Ap
The slip velocity is defined as the difference between the phase velocities, and we can define a

slip velocity between gas and liquid water (vs,(g−w) ) as,


vSg vsw
vs,(g−w) = vg − vw = − (3.54)
1 − Hw − Hhyd Hw

and a slip velocity between liquid water and hydrate (Us,(w−hyd) ) as,

67
vSw vShyd
vs,(w−hyd) = vw − vhyd = − (3.55)
Hw Hhyd
Equations 3.54 and 3.55 for the gas-water and water-hydrate slip velocities respectively, can be

rearranged to give a set of holdup equations (equations 3.56 and 3.57).

(1 − Hhyd ) vSw + Hw vShyd = vs,(g−w) Hw2 + vm − vs,(g−w) Hw + vs,(g−w) Hw Hhyd (3.56)




Hhyd vSw − Hw vShyd = vs,(w−hyd) Hw Hhyd (3.57)

The mass conservation equation for the water phase can be written as,

∂ (ρw Ap Hw ) ∂ (ρw Ap vSw )


= + ψw (3.58)
∂t ∂x
where ρw is the water density, and ψw is the mass transfer rate of water associated with the gas

hydrate reaction, and can be expressed as,

nH dmgas
ψw = (3.59)
∆x dt
where nH is the hydration number (kg of water/kg of gas), dmgas /dt is the gas mass consumption

rate during hydrate formation.

Assuming the water density, pipe cross sectional area, and section length are each separate

constants, the mass conservation equation simplifies to the volume conservation equation,

∂Hw ∂vSw 0
= + ψw (3.60)
∂t ∂x
where,

0 nH dmgas
ψw = (3.61)
ρw Ap ∆x dt
The water volume conservation equation (3.60) can be expanded using finite differences following

a similar procedure presented in the previous section (Section 3.4.1), yielding an explicit expression

for the liquid water holdup in the new time step n,

(n) (n−1) ∆t n 0 0
(n−1)
Hw,i = Hw,i + vSw,i − vSw,i+1
∆x )
(−Hw,i−2 + 6Hw,i−1 − 3Hw,i − 2Hw,i+1 )(n−1) 0
+vg + ∆tψw (3.62)
6

68
Following a procedure similar to the one used for the water phase, the mass conservation

equation for the gas hydrate phase can be expanded using finite differences resulting in the following

expression,

(n) (n−1) ∆t n 0 0
(n−1)
Hhyd,i = Hhyd,i + vShyd,i − vShyd,i+1
∆x )
(−Hhyd,i−2 + 6Hhyd,i−1 − 3Hhyd,i − 2Hhyd,i+1 )(n−1) 0
+vg + ∆tψhyd (3.63)
6

0
where ψhyd is the gas hydrate source term, which is related to the gas hydrate formation rate and

is defined as,

0 (nH + 1) dmgas
ψhyd = (3.64)
ρhyd Ap ∆x dt
The superficial velocities in equations 3.62 and 3.63 are a function of the phases holdup and

can be obtained from the simultaneous solution of equations 3.56 and 3.57.

Note that the gas-water slip velocity is considered to be function of liquid holdup, and the

following relationship is used,

vs,(g−w) = aHw2 + bHw + c (3.65)

where a, b, c are coefficients that can be determined from experimental data. No slip between

hydrates and water is considered in this model (vs,(w−hyd) = 0).

3.4.2.1 Transient hydrate kinetics model

The hydrate formation rate is calculated using the following intrinsic kinetics equation (Turner

et al., 2005),
 
dmgas k2
= −k1 exp As ∆Tsub (3.66)
dt Tsys
The intrinsic kinetics equation gives the rate of gas consumption during hydrate formation

(dmgas /dt) as a function of the intrinsic rate constants (k1 and k2 ) regressed from Vysniauskas

and Bishnoi (1983) data; the surface area (As ) between water and gas phases; and the subcooling

(∆Tsub ) as the thermal driving force. The subcooling is defined as the difference of the hydrate equi-

librium temperature and system temperature. Hydrate nucleation is assumed to occur immediately

after a specified subcooling.

69
The surface area between water and gas is calculated assuming stratified smooth flow in each

pipe section (Figure 3.8). The cross sectional area of the water layer (As ) as a function of pipe

radius (r) and wetted angle (α) is given by,

r2
As = (2α − sin 2α) (3.67)
2

Figure 3.8: Cross section of pipe with stratified smooth flow regime.

The relationship between water holdup (Hw ) and wetted angle in stratified flow is obtained by

the definition of water hold up,

2α − sin 2α
Hw = (3.68)

This is a non-linear function of the wetted angle. Given the liquid hold up, this equation can be

solved for the wetted angle using a numerical method. Alternatively, an explicit approximation for

the wetted angle as function of liquid hold up can be used to avoid iterative calculations (Biberg,

1999),
π 
α≈ 1 − (1 − Hw )1/3 + Hw1/3 (3.69)
2
Then, the surface area between water and gas in a pipe section is,

As = 2r sin(α)∆x (3.70)

3.4.2.2 Mixture energy conservation equation

The mixture energy conservation equation used in this work is based on a homogeneous fluid

model, where the gas, liquid and hydrate phases are treated as a pseudo-phase, and the properties

are the volume average of the independent phases (Cazarez-Candia and Vasquez-Cruz, 2005).

70
The energy balance equation includes terms for the transient behavior of the system enthalpy,

the convection of energy, the transient change in system pressure, fluid viscous dissipation, change

in potential energy, pipe wall heat transfer to the environment and a source term for the exothermic

nature of the hydrate formation reaction. The energy balance equation in terms of the pseudo-phase

enthalpy is,

∂h ∂h 1 ∂p ∂v ∂v q̇πd
+v = −v − vv − vg sin θ − + SE,hyd (3.71)
∂t ∂x ρ ∂t ∂t ∂x ρAp
where h is mixture enthalpy per unit mass, v is volume average velocity, ρ is volume average density,

p is system pressure, g is acceleration of gravity, θ is angle of inclination with respect to horizontal,

q̇ is the wall heat flux, d is pipe diameter, and SE,hyd is the energy source term and is related with

the exothermic nature of the hydrate formation reaction by,

nH + 1 dmgas
SE,hyd = ∆hhyd (3.72)
ρhyd Ap ∆x dt
where, ∆hhyd is the enthalpy change of hydrate formation (for methane hydrate, 4175.4×103 J/kg).

When hydrates are forming, the heat of formation (∆hhyd ) acts as an energy source inside a pipe

section.

Following the derivation presented in Appendix A, the energy balance equation (eqn. 3.71) can

be expressed explicitly in terms of system temperature (T ),

     
∂T ∂T ∂p ∂p 1 ∂p ∂v ∂v q̇πd
Cp +v −β +v − +v +v = −vg sin θ − + SE,hyd (3.73)
∂t ∂x ∂t ∂x ρ ∂t ∂t ∂x ρAp

where, Cp is the average heat capacity, β is the Joule-Thomson coefficient multiplied by Cp .

Equation 3.73 can also be discretized using finite differences obtaining an explicit solution for

the temperature in the solution node i at the new time step n (see Appendix A).

3.4.3 Example of application of the three-phase hydrodynamic slug flow model

The performance of the three-phase (gas/liquid water/hydrate) hydrodynamic slug flow model

presented in the previous section (Section 3.4.2) is tested against experimental data obtained from

an industrial scale flow loop (Joshi et al., 2011a). In this case, a periodic boundary condition is

used, where the fluids exiting the right boundary cell are reintroduced at the left boundary cell,

with the objective of modeling the flow behavior of the loop. The virtual connection between the

71
first and last cell represents, in terms of pressure, the positive displacement pump used in the flow

loop.

The flow loop is located inside a temperature controlled room, where the temperature can be

varied between 264 and 323 K. The average pressure in the flow loop is maintained constant at a

desired set point by injecting gas from a gas-filled piston accumulator connected to the flow loop.

Table 3.3 presents the input parameters used in the case study. The case is initialized with a water

holdup of 0.5. The temperature of the room housing the flow loop measured during the experiment

is an input to the model, and it is used to calculate the heat transfer between the fluids and the

environment outside the flow loop.

Table 3.3: Input parameters for three-phase (gas/liquid water/hydrate) hydrodynamic slug flow
model.

Input parameter Value


Initial loop temperature (K) 302.7
Initial loop pressure (MPa) 6.9
Initial gas superficial velocity (m/s) 1.7
Initial water superficial velocity (m/s) 0.8
Pipe diameter (m) 0.0972
Pipe length (m) 94
Pipe section length, dx (m) 1.0
Gas-water slip velocity parameters a = 3.5; b = −7.0; c = 4.2
Water-hydrate slip velocity (m/s) 0.0
Time step (s) 0.04
Simulation time (s) 6000

We assume that there is no slip between water and hydrate, i.e., the slip velocity is equal to zero.

The hydrate equilibrium conditions (pressure and temperature), assuming pure methane as the gas

phase (Figure 3.9), were calculated using CSMGem (Ballard and Sloan, 2004a). The nucleation

of hydrates was observed at a subcooling of 1.1 K in the flow loop experiment. In the simulation,

hydrates are assumed to nucleate when this subcooling is reached.

Figure 3.10 shows the distribution of water and hydrate holdup along the flowline length at four

different times obtained using the three-phase hydrodynamic slug model. At the beginning of the

simulation, the system was forced to remain in a stratified flow regime and no hydrates (Hhyd = 0)

are present in the system, as shown in Figure 3.10(a). Upon hydrate nucleation, the solid hydrate

72
12

10

Pressure (MPa)
L−H

6 L−H−V

4 L−V

2
270 275 280 285 290
Temperature (K)

Figure 3.9: Hydrate equilibrium curve for pure methane (V) and liquid water (L) calculated with
CSMGem.

particles act as a perturbation that promotes a flow regime transition from stratified to slug flow;

the beginning of this transition is shown Figure 3.10(b). The amount of hydrates continues to

increase with time, as shown in Figure 3.10(c) and Figure 3.10(d).

The temperature predicted by the model is compared against measured data from the flow

loop experiment (Figure 3.11). Note that the input to the model was an approximation to the

ambient temperature measured during the experiment (curve labeled as “Ambient”), and the fluid

temperature was matched by adjusting the overall heat transfer coefficient. With the three-phase

hydrodynamic slug flow model, it was possible to match the cooling rate of fluids inside the loop

and to capture the exothermic nature of the hydrate formation reaction shown as an increase in

temperature at the point of hydrate formation onset.

3.5 Summary

This chapter presented a review of the fundamental concepts of steady-state and transient

multiphase flow modeling. In order to understand the origin of multiphase flow modeling techniques

it is necessary to review the single-phase flow pressure-gradient equation, which resulted from the

analytical solution of the mass and momentum conservation equations. There is one empirical

component in the pressure gradient equation for single-phase flow, which is the dimensionless

friction factor correlation.

73
1 1

0.8 0.8
Hold−up

Hold−up
0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 20 40 60 80 0 20 40 60 80
Pipeline length (m) Pipeline length (m)
(a) Time = 40 s. (b) Time = 3230 s.

1 1

0.8 0.8
Hold−up

Hold−up

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 20 40 60 80 0 20 40 60 80
Pipeline length (m) Pipeline length (m)
(c) Time = 4000 s. (d) Time = 6000 s.

Figure 3.10: Distribution of water (continuous blue line) and hydrate (segmented green line) holdup
along the pipeline length at four different times, (a) 40 s, (b) 3230 s, (c) 4000 s, (d) 6000 s, showing
the transition from stratified to slug flow upon hydrate formation.

74
Fluid
300 Ambient
Experiment fluid

Temperature (K)
Experiment ambient

290

280 Hydrate
formation onset

270
0 0.5 1 1.5
Time (hr)
Figure 3.11: Comparison of predicted results and experimental data of fluid and ambient temper-
ature as function of time.

The ultimate objective in flow modeling is to estimate the pressure drop in the pipeline, an

important element for the design of pipeline systems to guarantee the effective transport of fluids

from one point (e.g., wellhead) to another (e.g., offshore platform).

In general, the multiphase flow modeling methods are extensions to the well known single-phase

flow models. The complexity of the multiphase flow requires empirical correlations to estimate the

effect of unknown mechanisms affecting the pressure gradient. The homogeneous fluid modeling

method considers the multiphase mixture as a homogeneous phase and uses modeling approaches

similar to that of single phase flow, but this method typically underestimate the pressure drop.

Other methods consider separate flow and try to estimate the flow regimes based on predefined

flow regime maps. Most of the flow regime maps were developed for air/water flow at near atmo-

spheric pressure conditions, which could not be representative of gas/oil systems encountered in

offshore production. The transient multiphase modeling method involves the simultaneous solution

of conservation equations of mass, momentum and energy for each phase, and uses empirical cor-

relations and simplified constitutive relationships for flow parameters, such as interfacial friction,

entrainment of dispersed phases in continuous phases, and liquid holdup in the slug body. The

transient multiphase flow modeling method form the basis of the commercially available OLGA
R

transient multiphase flow simulator, which is considered the standard for multiphase flow model-

75
ing in the oil and gas industry. It was shown how the flexible flow model framework of OLGA
R

allows to define new phases and introduce additional contributions to the conservation equations,

representing the effect of the new phase.

A new simple modeling approach based on fundamental concepts of multiphase flow was re-

viewed in the final section of this chapter, and is known as the hydrodynamic slug flow model. This

hydrodynamic slug flow model predicts the formation of liquid slugs. The hydrodynamic slug flow

model was extended including an additional mass balanced equation for the gas hydrate phase,

and an energy conservation equation. The hydrate formation rate was calculated using a transient

hydrate kinetics model. The mass transfer between the phases (gas, water and hydrate) is a func-

tion of the hydrate formation rate, and was included as a source/sink term in the mass balance

equations.

The extended hydrodynamic slug flow model was applied to simulate a flow loop experiment,

where the temperature of the fluids inside the flow loop is decreased until hydrates start to form.

The model captured the flow regime transition, from stratified to slug flow, observed in flow loop

experiments upon hydrate formation. The temperature behavior of the system was also captured

with the extended hydrodynamic slug flow model, matching the cooling rate of fluids inside the

loop and the exothermic nature of the hydrate formation reaction that increases the temperature

at the point of hydrate formation onset.

76
CHAPTER 4

FLOW LOOP EXPERIMENTS

Two large scale flow loop facilities were available during the development of this study, the

ExxonMobil Upstream Research Company flow loop located at Friendswood, Texas, and the Uni-

versity of Tulsa hydrate flow loop located at Tulsa, Oklahoma (the loop was originally designed

and constructed at the Marathon Oil Technology Center in Littleton, Colorado). This chapter

presents a description of the two large scale flow loops, the experimental procedure and materials

used, and the experimental matrix that represents the available data. These data are critical to

the development of the conceptual and mathematical models described in Chapter 5.

4.1 ExxonMobil flow loop

Figure 4.1 shows a schematic of the ExxonMobil flow loop used to study hydrate formation in

multiphase flow systems. The ExxonMobil flow loop has an internal diameter of 3.826 inches (4

in. nominal size) with an axial length of 314 ft (95 m). The fluids are displaced by a multiphase

sliding-vane pump, specially designed to minimize shear on hydrate particles, which can operate

at rotation speeds between 300 to 1500 rpm (equivalent to 0.8 to 3.8 m/s fluid mixture velocities).

The flow loop is located inside a temperature controlled room, which can operate in the range

from 90 to 25 o F (273 to 269 K). The flow loop can be operated at pressures up to 2000 psi (12.41

MPa). A piston-driven gas accumulator with maximum volume of 120 gallons is connected to the

flow loop near the suction side of the pump, and allows gas injection into the loop to maintain

constant pressure across the loop during hydrate formation. The position of the piston inside the

gas accumulator is controlled by a hydraulic power unit (HPU). The flow loop incorporates a port

where a particle size analyzer probe (FBRM or PVM) can be installed for visualization of entrained

gas bubbles, droplets and hydrate particles in the continuous phase. A mass flow and density meter

is also installed in the flow loop.

The parameters that can be measured in this flow loop are average pressure of the loop, pressure

drop across the loop, temperature inside the loop, and gas injection into the loop. The loop

pressure is measured with an accuracy of 0.06 psig using a pressure transducer located on the line

77
Figure 4.1: Schematic of ExxonMobil flow loop.

connecting the gas accumulator to the loop. The pressure drop across the loop is measured using

a differential pressure transducer with an accuracy of 0.12 psi, which is located across the pump.

The temperature is measured using two thermocouples with an accuracy of 0.02 o F, located at the

intake and discharge of the pump toward the center of the pipe. The gas injected into the loop is

indirectly measured by the position of the piston, which can be related to the accumulator volume.

The flow loop incorporates two viewports, each equipped with a digital camera and a light source

on the opposite side for video recording. The viewports are cubic expansion boxes with vertical

circular viewing area of 4 inches in diameter.

The experimental procedure followed for the tests performed at the ExxonMobil flow loop

consisted of two stages. The first stage consists of charging fluids into the loop by adding the

desired amount of oil and water, and pressurizing with methane gas to the desired operation

pressure. The fluids are circulated in the loop using a pump speed of 600 rpm for at least two

hours, to mix and saturate the phases at the ambient temperature. The second stage consists of

setting the experimental conditions, by setting the constant operation pressure in the HPU, the

pump speed, and the cold room temperature. When a test is complete, the temperature is increased

above 70 o F (294 K) and held at this temperature for the next test.

Two sets of experiments were performed in the ExxonMobil flow loop for this study. The first

set of experiments was performed in the Fall of 2010, and consisted of pure water (100% water

cut) and methane gas experiments to study the hydrate formation mechanisms in water-dominated

78
systems. Table 4.1 presents the parameters of the experiments performed in the Fall of 2010. The

liquid loading values used in this set of experiments were 50, 75 and 90 % of the flow loop volume;

the liquid loading is the fraction of flow loop volume occupied by liquid phases. The mixture

velocities used in the experiments were 1.0, 1.75 and 2.0 m/s; the mixture velocity of fluids inside

the flow loop is set by the pump speed. In all the experiments the operation pressure was set

to 1000 psi (6.89 MPa) and maintained constant by injecting gas from the gas accumulator. The

temperature of the flow loop room is decreased from an initial ambient temperature to a set point

of 34 o F (274 K).

Table 4.1: Experiments performed in the ExxonMobil flow loop in Fall 2010.

Test # Liquid loading Water cut Mixture velocity


(%) (%) (m/s)
XoM-2010-1 50 100 1.0
XoM-2010-2 50 100 1.75
XoM-2010-3 50 100 2.5
XoM-2010-4 75 100 1.0
XoM-2010-5 75 100 1.75
XoM-2010-6 75 100 2.5
XoM-2010-7 90 100 2.5

The second set of experiments was performed in the Fall of 2011, and consisted of a variety

of water- and oil-dominated systems. For the oil-dominated systems Conroe oil or King Ranch

Condensate were used as the oil phase. Table 4.2 presents the parameters of the experiments

performed in the Fall of 2011. The liquid loading values used in this set of experiments were 50,

75, 80, 90 and 95 % of the flow loop volume. The mixture velocities used in the experiments were

1.0, 1.75, 2.0, 2.5 and 3.0 m/s. In all the experiments the operation pressure was set to 1000 psi

(6.89 MPa) and maintained constant by injecting gas from the gas accumulator. The temperature

of the flow loop room is decreased from an initial ambient temperature to a set point of 34 o F (274

K).

Additional data from oil-dominated experiments (Table 4.3) performed in the ExxonMobil flow

loop is available from previous studies (Turner, 2005).

79
Table 4.2: Experiments performed in the ExxonMobil flow loop in Fall 2011.

Test # Liquid loading Water cut Mixture velocity Oil


(%) (%) (m/s)
XoM-2011-1 50 100 1.0 N/A
XoM-2011-2 50 100 1.75 N/A
XoM-2011-3 75 100 0.75 N/A
XoM-2011-4 75 100 1.0 N/A
XoM-2011-5 75 100 2.0 N/A
XoM-2011-6 95 100 1.0 N/A
XoM-2011-7 95 100 2.0 N/A
XoM-2011-8 95 100 2.5 N/A
XoM-2011-9 50 15 1.0 Conroe
XoM-2011-10 50 15 1.75 Conroe
XoM-2011-11 50 15 2.0 Conroe
XoM-2011-12 50 15 2.5 Conroe
XoM-2011-13 50 15 3.0 Conroe
XoM-2011-14 75 15 1.0 Conroe
XoM-2011-15 75 15 1.75 Conroe
XoM-2011-16 75 15 2.5 Conroe
XoM-2011-17 90 15 1.0 Conroe
XoM-2011-18 90 15 1.75 Conroe
XoM-2011-19 90 15 2.5 Conroe
XoM-2011-20 50 90 1.0 Conroe
XoM-2011-21 50 90 1.75 Conroe
XoM-2011-22 50 90 2.5 Conroe
XoM-2011-23 50 75 1.0 Conroe
XoM-2011-24 50 75 2.5 Conroe
XoM-2011-25 50 25 1.0 King Ranch condensate
XoM-2011-26 50 25 1.75 King Ranch condensate
XoM-2011-27 50 25 2.5 King Ranch condensate
XoM-2011-28 80 25 1.0 King Ranch condensate
XoM-2011-29 50 25 1.75 King Ranch condensate
XoM-2011-30 50 25 2.5 King Ranch condensate
XoM-2011-31 50 50 1.0 King Ranch condensate
XoM-2011-32 50 50 1.75 King Ranch condensate
XoM-2011-33 50 50 2.5 King Ranch condensate

Table 4.3: Additional experimental data from the ExxonMobil flow loop (Turner, 2005).

Test # Liquid loading Water cut Mixture velocity Oil


(%) (%) (m/s)
XoM-2004-1 60 35 0.7 Conroe
XoM-2004-2 70 37 2.3 Conroe
XoM-2004-3 70 37 3.5 Conroe

80
4.2 Tulsa University flow loop

The hydrate flow loop available at the University of Tulsa in Tulsa, Oklahoma, was originally

designed and constructed at the Marathon Oil Technology Center in Littleton, Colorado. Figure 4.2

shows a schematic of this hydrate flow loop used to study hydrate formation in multiphase flow

systems. The University of Tulsa hydrate flow loop has an internal diameter of 2.9 inches (3 in.

nominal size) with an axial length of 164 ft (50 m). The fluids are displaced by a Leistritz twin-

screw multiphase pump, which can operate at rotation speeds up to 2000 rpm (equivalent to 3.7

m/s fluid mixture velocities). The flow loop has a cooling jacket over most of its length. Glycol

is used as the coolant fluid, and the glycol temperature can be controlled between the range from

32 o F to 80 o F (273 to 300 K) by measuring the glycol inlet and outlet temperatures. The flow

loop can be operated at pressures up to 2000 psi (12.41 MPa). A high-pressure piston pump is

connected to the loop at the discharge side of the pump, and allows gas injection into the loop to

maintain constant pressure across the loop during hydrate formation. A gamma densitometer is

installed to measure density data of the fluids.

The parameters that can be measured in this flow loop are pump suction and discharge pressures,

pressure drop, temperature, fluid density and gas injection into the loop. The pump suction and

discharge pressures are measured with an accuracy of 0.06 psig using pressure transducers. The

pressure drop is measured across each section of the loop and across the pump, using differential

pressure transducers with an accuracy of 0.025 psi. Several thermocouples are installed on the

outside pipe wall and measure the temperature of the pipe surface with an accuracy of 0.01 o F.

The gas injected into the loop is indirectly measured by the volume of oil pumped into the other

side of the piston of the high-pressure piston pump. The mass of gas injected is calculated using

equations of state and the gas composition. The equations of state available to calculate the mass

of gas injected are Peng-Robinson, Redlich-Kwong and Benedict-Webb-Rubin. The flow loop has

four viewports at the end of each section, each equipped with a digital camera for video recording.

Each view port has three sapphire windows at a 120o angle of each other around the pipe.

The experimental procedure used for the tests performed at the University of Tulsa flow loop

is similar to the one used for the ExxonMobil flow loop. In the first stage, fluids are charged into

the flow loop, adding oil, water and pressurizing with gas at ambient temperature. In the second

81
Figure 4.2: Schematic of University of Tulsa hydrate flow loop.

stage the experimental conditions (pressure, pump speed and final temperature) are set.

Table 4.4 presents the parameters of experiments performed in the University of Tulsa flow loop.

The experimental data from these experiments were obtained from previous studies (Boxall, 2009;

Joshi, 2012). The first five experiments consisted in oil-dominated systems using one model oil

(Citgo) and three different crude oils (Buttermilk, Caratinga and Troika). The liquid loading used

in the oil-dominated experiments was 50%. The mixture velocity used was 1.2 m/s. The first three

experiments in Table 4.4 were performed at a constant pressure of 2000 psi (13.79 MPa), while the

experiments using Troika oil were at constant volume with an initial pressure of 1500 psi (10.34

MPa). In the constant volume experiments no gas is injected into the loop, and the system pressure

is allowed to decrease during cooling (due to contraction of gas phase) and hydrate formation (due

to consumption of gas phase). The temperature of the system is decreased to a final temperature

between 41 and 43 o F (278 and 279 K).

4.3 Summary

This chapter presented a description of the two large scale flow loop facilities used in this

study; the ExxonMobil Upstream Research Company flow loop located at Houston, Texas, and

the University of Tulsa hydrate flow loop located at Tulsa, Oklahoma. The experiments performed

at the flow loop facilities allow one to study the effects of hydrate formation in oil and water-

dominated multiphase flow systems. The multiphase flow observed in the flow loops is a physical

82
Table 4.4: Available experimental data from the University of Tulsa hydrate flow loop (Joshi,
2012).

Test # Liquid Water Mixture Pressure Temperature Oil


loading cut velocity
(%) (%) (m/s) (psig) (o F)
TU-2004-1 50 25 1.2 2000 42 Citgo
TU-2004-2 50 25 1.2 2000 42 Buttermilk
TU-2004-3 50 25 1.2 2000 43 Caratinga
TU-2004-4 50 25 1.2 1500 41 Troika
TU-2004-5 50 50 1.2 1500 41 Troika
TU-2004-6 50 100 1.2 2000 45 N/A
TU-2004-7 50 100 0.2 2000 45 N/A
TU-2004-8 50 100 0.2 2000 45 N/A
TU-2009-1 50 100 1.2 1500 45 N/A
TU-2009-2 75 100 1.2 1500 45 N/A
TU-2009-3 75 100 0.7 1500 45 N/A
TU-2009-4 90 100 1.2 1500 45 N/A

simulation of the expected multiphase flow in the pipelines of subsea transport facilities. The flow

loops provide a controlled environment, where the pressure, temperature and mixture fluid velocity

can be specified. The system conditions can be placed inside the hydrate formation region by

decreasing the temperature. The amount of hydrates formed can be estimated by measuring the

gas consumption (or gas injected into the flow loop to maintain constant pressure). Measurements

of pressure drop across the flow loop can be used to estimate the transportability of the fluids inside

the pipeline as function of the hydrate volume fraction, until a hydrate plug is formed. Next chapter

(Chapter 5) presents conceptual and mathematical models developed from the data obtained from

the experiments performed in these flow loops. These data is also used for the validation of the

mathematical models.

83
CHAPTER 5

MATHEMATICAL MODEL FOR HYDRATE FORMATION

Three different systems of oil and gas production facilities, such as, oil-dominated systems (oil

as hydrate carrier phase), water-dominated systems (water as hydrate carrier phase), and gas-

dominated systems (hydrate film growth on pipe walls), have been considered for the development

of conceptual and mathematical models of gas hydrate formation in pipelines. A set of conceptual

pictures was developed to explain the physical phenomena of gas hydrate formation in these sys-

tems. Different hydrate formation mechanisms were identified for the different systems encountered

in oil/gas flowlines. For oil-dominated systems, dispersed water droplets are converted into hydrate

particles, and the agglomeration of these hydrate particles increase the slurry viscosity leading to

a plug. Experimental observations in flow loop experiments of water-dominated systems, at the

ExxonMobil flow loop, showed that pressure drop in the pipe depends on a flow regime transi-

tion involving multiple phases and solid dispersion in the liquid phase. For gas-dominated systems,

hydrates deposit and grow on pipe walls, narrowing the flow diameter and increasing surface rough-

ness.

This chapter presents a description of the three sub-models that constitute the comprehensive

hydrate formation model, including the mathematical models’ contributions to the multiphase flow

transient conservation equations presented in Section 3.3.3. This work presents improvements to

the hydrate particle aggregation module used for oil-dominated systems, based on experimental

data, which account for temperature, particle-particle contact time, excess water, and the presence

of surface active compounds. The model for water-dominated systems consists of a mass transfer-

based hydrate growth model and hydrate plugging criterion, based on fluid velocity and liquid

holdup. The model for gas-dominated systems consists of a combined heat and mass transfer

model for hydrate film growth on pipe walls.

5.1 Model for hydrate formation in oil-dominated systems

A previous version of the model for gas hydrate formation in oil-dominated systems was pre-

sented in Section 1.2 of the introductory chapter. The gas hydrate formation mechanism in oil-

84
dominated systems was described using the conceptual model shown in Figure 5.1 (reproduced

from Figure 1.8), where water is assumed to be entrained in the continuous oil phase as water

droplets, and the surface area between water and hydrocarbon phases is calculated using the Hinze

(1955) correlation. An alternative approach to calculate the surface area of water droplets is based

on an empirical correlation developed by Boxall (2009) for mean droplet size as function of fluid

properties and flow regime. Hydrate nucleation is assumed to occur immediately after a specified

subcooling (∆Tsub = Thyd eq − Tsystem ), and the hydrate growth is calculated using an intrinsic first

order kinetics equation with an adjustable rate constant (Boxall et al., 2009), or using the transport

model developed by Davies (2009) based on mass and heat transfer resistances. Once the hydrate

particle has formed, the particle agglomerate size is calculated using a steady-state balance between

the inter-particle cohesion force and fluid flow shear forces (Camargo and Palermo, 2002), assuming

a constant cohesion force. The viscosity of the slurry is calculated using a modified Mills equation

(Camargo and Palermo, 2002; Mills, 1985), using the effective volume fraction of the agglomerated

particles. A hydrate plug is identified by a large viscosity increase, which causes an unacceptably

large pressure drop in the line, prohibiting flow.

Figure 5.1: Conceptual model for hydrate formation in oil-dominated multiphase flow systems
consisting of gas, oil, and water; adapted from Turner (2005), with input from J. Abrahamson (U.
Canterbury, Christchurch, NZ).

Improvements have been made to the agglomeration section of the conceptual model by remov-

ing the assumption of a constant particle cohesion force and introducing a dynamic cohesion force

model. The dynamic cohesion force model is based on particle capillary attraction due to a liquid

bridge between particles and sintering of the liquid bridge.

85
5.1.1 Description of dynamic particle cohesion force model

The discussion of agglomeration mechanisms of hydrate particles presented in Section 2.2.3

showed that a liquid bridge between hydrate particles can be responsible for hydrate particle ag-

glomeration in oil-dominated systems. The liquid bridge forms from a liquid layer present on the

hydrate particle surface when the particle is surrounded by a secondary immiscible phase (i.e.,

the oil phase). Similar liquid layers have been observed on the surface of ice (Bienfait, 1992;

Döppenschmidt and Butt, 2000) and are considered thermodynamically stable, since the presence

of the liquid layer reduces the free energy of the system.

Based on the schematic representation of the capillary liquid bridge between two hydrate par-

ticles, shown in Figure 5.2, Aman et al. (2011) derived an equation for the particle cohesion force

(Fa ) from the specific free energy of the liquid bridge,

Fa 2πσ cos θp 2πfslv



= H
+ 2πσ sin (α) sin (θp + α) − (5.1)
R 1 + 2d 2αR∗ + 2Hα

where R∗ is the harmonic mean radius of particle radii R1 and R2 , σ is the water-oil interfacial

tension, θp is the wetting angle of water on the particle surface, H is the total height of the liquid

bridge, d is the immersion depth (i.e., how far the bridge extends into the particle), α is the

embracing angle, and fslv is a generic three phase free energy term. The first term represents

the capillary pressure across the liquid bridge, the second term represents the contribution of the

surface tension of the liquid bridge interfacial area, and the third term represents the specific free

energy of the three phase contact line. The last term is typically small, in comparison to the first

two terms, and can be neglected.

For the implementation of the model, it is assumed that hydrate particles are spherical and that

all unconverted water is evenly distributed on the surface of the hydrate particles as a uniform water

layer. As additional water is converted to hydrate (estimated by the kinetic or transport-limited

hydrate formation models mentioned above), less water is available to participate in the capillary

bridge between hydrate particles, resulting in a decrease of cohesive force. The liquid bridge height

(H) is defined as a thermodynamic stable liquid layer plus the excess water. The thermodynamic

stable liquid layer is assumed to be equal to that measured for ice, approximately 50 nm thick

(Döppenschmidt and Butt, 2000). The wetting angle of water on the hydrate particle surface (θp )

is assumed to be equal to 29o (a value measured by Asserson et al., 2009 for Freon hydrates in

86
Figure 5.2: Illustration of the capillary liquid bridge between two hydrate particles (Aman et al.,
2011).

brine). The embracing angle (α) is assumed to be equal to 0.1o . This value of embracing angle was

calculated by Aman et al. (2011) considering the baseline cohesive force of 4.3 mN/m measured for

cyclopentane hydrate, with a water/cyclopentane interfacial tension of 51 mN/m and the assumed

values for liquid layer thickness and wetting angle of water on hydrate surface. Rabinovich et al.

(2005) presented relations to estimate the volume (V ) and immersion depth (d) of the capillary

liquid bridge between particles,

V = πR2 α2 H + 0.5πR3 α4 (5.2)

r !
H 2V
d= −1 + 1+ (5.3)
2 πRH 2
From cyclopentane hydrate cohesion experiments using a micromechanical force apparatus,

Aman et al. (2011) observed that inter-particle cohesion forces increased with contact time, at-

tributing this behavior to the liquid bridge converting into hydrate, a phenomenon called sintering.

They observed that the cohesion force started to increase when the contact time was greater than

30 seconds. The required force to fracture the solid hydrate bridge is defined by the tensile strength

of hydrate and the area of the solid bridge,

FS = τt Ai (5.4)

87
where FS is the inter-particle force attributed to sintering, τt is the hydrate tensile strength, and

Ai is the cross sectional area of the solid bridge at the weakest point (the middle).

Due to the increase in measured cohesion force with contact time, Aman et al. (2011) deter-

mined that the hydrate bridge area (Ai ) must be increasing over the course of contact time, while

the hydrate tensile strength is a constant mechanical property. The cyclopentane hydrate tensile

strength (τt ) was estimated to be 0.91 MPa, for the baseline cohesive force of 4.3 mN/m. Using

the estimated hydrate tensile strength, they calculated the hydrate bridge area corresponding to

experimental cohesive force measurements for different contact times, in terms of the radius of con-

tact (χ). Then, the time-dependence of the hydrate bridge area was approximated by the following

expression, regressed from the experimental radius of contact (χ),

χ = 1.2147t0.1249 (5.5)

The dynamic hydrate particle cohesion force model results from combining the capillary cohesion

force equation (5.1) and the sintering cohesion force equation (5.4), considering that the capillary

force dominates below about 30 s, and the sintering force dominates above 30 s,
 2πσ cos θ
p

H
+ 2πσ sin (α) sin (θp + α) , tcontact < 30 s
FIP = 1 + 2d (5.6)
2
τt π 1.2147t0.1249 , tcontact ≥ 30 s

where FIP is the inter-particle cohesion force, and tcontact is the contact time between hydrate

particles. Figure 5.3 shows the results of this model compared against experimental data for

cohesion forces as function of contact time of cyclopentane hydrate particles. The model offers a

good prediction of the experimental data.

The dynamic cohesion force model, represented by equation (5.6), is based on five key quanti-

ties: (i) water-oil interfacial tension; (ii) hydrate surface wettability; (iii) hydrate particle radius;

(iv) volume of liquid water available, arising from either a thermodynamic quasi-liquid layer or

unconverted water; and (v) contact time between the hydrate particles. The water-oil interfacial

tension is a characteristic property of the system, and allows one to account for the presence of

surface active compounds (or surfactants). Surfactants that adhere to the liquid bridge interface

may reduce the interfacial tension, hence reducing the cohesion force between hydrate particles

(Aman et al., 2010). Surfactants can also affect the hydrate surface wettability, as was observed

by Aman et al. (2012) during micromechanical force measurements between cyclopentane hydrate

88
Figure 5.3: Comparison of experimental measurements (points) and model prediction (solid line)
of cyclopentane hydrate cohesive forces as function of contact time (Aman et al., 2011).

particles in the presence of different carboxylic acids. However, there are no means available to

predict the effect of surfactants on hydrate surface wettability, so this parameter is assumed to be

constant in the model. The volume of liquid water available to form a liquid bridge has a direct

effect on the magnitude of the cohesion force. As the liquid water is converted into hydrate, the

liquid bridge volume decreases, resulting in a decrease of cohesive force. The sintering of the liquid

bridge starts when the particles have been in contact for a period greater than 30 seconds.

5.1.2 Implementation of the model into a transient multiphase flow simulator

The model for hydrate formation in oil-dominated systems was implemented in this thesis work

in the OLGA
R
transient multiphase flow simulator, through the development of a user-defined

plug-in program. Using the available OLGA


R
flow model framework it is possible to define a

hydrate phase that will be dispersed into the continuous oil phase. Figure 5.4 shows a schematic

representation of the mass fields considered for oil-dominated systems, which includes gas and oil as

continuous phases, water and hydrate as dispersed phases in the oil continuous phase, and oil and

water may be dispersed as droplets in the continuous gas phase. The user-defined plug-in program

calculates additional contributions to the conservation equations using the PVT, flash and rheology

modules.

89
Figure 5.4: Illustration of mass fields considered in the oil-dominated hydrate formation model
coupled with the OLGA R
transient multiphase flow simulator, including gas and oil as continuous
phases, water and hydrate is assumed to be dispersed in the oil phase, and oil and water may be
dispersed as droplets in the continuous gas phase.

The PVT module calculates the properties for the hydrate phase, making additional contri-

butions to the mass, momentum and energy conservation equations. Typical values of hydrate

structure II (Sloan and Koh, 2008a) are assigned for density (914 kg/m3 ), heat capacity (498.83

J/(kg K)), thermal conductivity (0.7 W/(m K)). The enthalpy of the hydrate phase (hhyd ) is a

function of the enthalpy of the gas (hg ) and water (hw ) phases at the system temperature and

pressure, and the heat of reaction (∆hhyd ), to account for the exothermic nature of the hydrate

formation reaction,

hg + nH hw + ∆hhyd
hhyd = (5.7)
1 + nH
where nH is the hydration number (water to gas ratio in the hydrate phase).

The flash module calculates the gas consumption rate during hydrate formation using the in-

trinsic first order kinetics equation with an adjustable rate constant (Boxall et al., 2009) or using

the transport model developed by Davies (2009) based on mass and heat transfer resistances. The

flash module makes direct contributions to the mass conservation equations by calculating the mass

transfer rates from gas and water phases to the hydrate phase during hydrate formation, and during

hydrate dissociation where gas and water are released from the hydrate phase. The mass transfer

rates are used to update the phase masses of gas, water and hydrate.

90
The rheology module calculates the apparent viscosity of the oil layer as a function of the hydrate

particles agglomerate diameter, making contributions to the momentum conservation equation. The

hydrate particles agglomerate diameter is calculated using a force balance between the hydrate

particles cohesion force and the flow shear forces, based on the Camargo and Palermo (2002)

model. The particles cohesion force is calculated using the dynamic cohesion force model described

in Section 5.6. The flow shear forces are calculated in the OLGA
R
model. The effective viscosity

of the oil phase, considering the hydrate particle dispersion, is calculated using a modified Mills

equation (Camargo and Palermo, 2002; Mills, 1985), based on the effective volume fraction of the

agglomerated particles. The modified viscosity will induce modifications in the friction factors.

5.1.3 Validation of the model against flow loop data

This section presents the comparison of prediction results obtained with the model for hydrate

formation in oil-dominated systems against data from flow loop experiments.

Figure 5.5 to Figure 5.7 show results corresponding to three different experiments, one performed

at the ExxonMobil flow loop and two at the University of Tulsa flow loop, using three different

oils. These three cases show good agreement between model prediction and the experimental data.

The intrinsic kinetics equation was used to estimate the hydrate formation rate, using a correction

factor of 0.002. Figure 5.5 shows the hydrate volume fraction as function of time of an ExxonMobil

flow loop constant pressure experiment using Conroe oil, 70% liquid loading, 37% water cut, and

2.3 m/s fluid mixture velocity. The hydrate formation starts around 0.7 hours with a high hydrate

formation rate, which decreases slightly in a few minutes. The hydrate formation rate remains

constant for a long portion of the experiment, and slows down at the end, when almost all the

water is converted into hydrate.

Figure 5.6 shows the cumulative gas injection into the loop (a measure of the amount of hy-

drates formed) as function of time of a University of Tulsa flow loop constant pressure experiment,

using Caratinga oil, 50% liquid loading, 25% water cut and 1.2 m/s fluid mixture velocity. The

gas injection in the first five hours of the experiment is required to maintain constant pressure

inside the loop, due to contraction of the gas phase during cooling of the flow loop; there is no

hydrate formation in this period. After five hours in the experiment, the hydrate formation onset

corresponds to a discontinuity in the slope of the cumulative gas injection curve, where the rate of

91
0.1

Hydrate volume fraction


0.08 CSMHyK
oil−dominated
model
0.06

0.04 Experiment

0.02

0
0 0.5 1 1.5 2
Time (hr)

Figure 5.5: Plot of hydrate volume fraction as function of time, comparing experimental data
and oil-dominated model prediction for an ExxonMobil flow loop constant pressure experiment
with Conroe oil, 70% liquid loading, 37% water cut, and 2.3 m/s fluid mixture velocity (Test #
XoM-2004-2).

gas injection increases due to gas consumption during hydrate formation. The gas injection stops

when all the water is converted into hydrate. The intrinsic kinetics equation was used to estimate

the hydrate formation rate, using a kinetic rate correction factor of 0.002, which also offers a good

prediction of the hydrate formation rate in this University of Tulsa flow loop experiment.

5 CSMHyK
Cumulative gas injection

oil−dom.
4 model
into loop (kg)

3
Experiment
2

0
0 5 10 15
Time (hr)

Figure 5.6: Plot of cumulative gas injection into the loop as function of time, comparing ex-
perimental data and oil-dominated model prediction for a University of Tulsa flow loop constant
pressure experiment with Caratinga oil, 50% liquid loading, 25% water cut, and 1.2 m/s fluid
mixture velocity (Test # TU-2004-3).

Figure 5.7 shows the flow loop pressure as function of time of a University of Tulsa flow loop

constant volume experiment using Troika oil, 50% liquid loading, 25% water cut, and 1.2 m/s

92
fluid mixture velocity. Since this is a constant volume experiment, the pressure decreases due to

contraction of the gas phase during cooling of the flow loop. Upon the hydrate formation onset, a

discontinuity in the rate of change of the pressure is observed, with a greater decrease in the pressure

due to consumption of the gas phase during hydrate formation. Considering the different conditions

of these three experiments, the model predicts with good agreement the behavior observed.

14
Flow Loop Pressure (MPa)

13 CSMHyK
oil−dominated
model
12

Experiment
11

10
0 5 10 15 20
Time (hr)

Figure 5.7: Plot of flow loop pressure as function of time, comparing experimental data and
oil-dominated model prediction for an University of Tulsa flow loop constant volume experiment
with Troika oil, 50% liquid loading, 25% water cut, and 1.2 m/s fluid mixture velocity (Test #
TU-2004-4).

Additionally, the prediction performance of hydrate formation in oil-dominated systems was

compared against selected experimental data obtained at the ExxonMobil flow loop in the Fall of

2011 (Table 4.2). This set of experiments includes a range of low and high water cuts that can be

used to test the limits of model applicability. Three test cases with low water cut are presented;

Test XoM-2011-9 with Conroe oil, liquid loading of 50 vol.%, and 15% water cut; Test XoM-2011-14

with Conroe oil, liquid loading of 75 vol.%, and 15% water cut; and Test XoM-2011-25, with King

Ranch condensate, 50 % liquid loading, and 25% water cut. One test case with a higher water

cut of 75 vol.% (Test XoM-2011-23) is presented that seems to indicate the transition from oil to

water-dominated systems. The thermal response observed in the flow loop experiments during the

cooling stage was captured by adjusting the overall heat transfer coefficient (an input to the model).

An overall heat transfer coefficient of 80 W/(m2 K) (14 BTU/(ft2 hr K)) was used to match the

cooling stage of the ExxonMobil flow loop experiments. The intrinsic kinetics equation was used

to estimate the hydrate formation rate, using a correction factor of 0.002.

93
Figure 5.8 shows a plot of temperature as function of time of Test XoM-2011-9, with Conroe

oil, 50% liquid loading, 15% water cut, and 1 m/s fluid mixture velocity, where the green curve

is the ambient temperature outside the flow loop, the blue dashed curve is the measured fluid

temperature, and the red curve is the fluid temperature predicted by the oil-dominated hydrate

formation model. The experiment starts with a cooling stage, decreasing the ambient temperature.

There is heat transfer from the fluids inside the pipe to the cooled air outside the loop. In the

model, an overall heat transfer coefficient of 80 W/(m2 K) (14 BTU/(ft2 hr K)) was used to match

the cooling stage of the experiment. Upon the hydrate formation onset, around 2 hours, an increase

in the fluid temperature is observed due to the exothermic nature of the hydrate formation reaction,

which is also captured by the model.

310
CSMHyK
oil−dominated
Temperature (K)

300 model

290
Experiment

280 Ambient
temp.

0 1 2 3 4
Time (hr)

Figure 5.8: Plot of fluid and ambient temperature as function of time, comparing experimental
data and oil-dominated model prediction for an ExxonMobil flow loop experiment with Conroe oil,
50% liquid loading, 15% water cut, and 1 m/s fluid mixture velocity (Test # XoM-2011-9).

Figure 5.9 shows the predicted fluid distribution, in terms of volume fraction, inside the flow loop

as function of time for Test XoM-2011-9, the experimental hydrate volume fraction is also shown

(gray dashed curve). The volume fraction of total liquid and gas remains constant throughout

the experiment, while the water volume fraction decreases after hydrates start to form. Water is

almost completely consumed after 4 hours. The model underpredicts the hydrate formation rate

in this low water cut experiment, with a 20% relative error in the amount of hydrates with respect

to the experimental value at the end of the experiment. The intrinsic kinetics equation was used

to estimate the hydrate formation rate, using a correction factor of 0.002.

94
Total liquid

Volume fraction
0.4 Gas

Experiment hyd.
0.2 Water

Hydrate
0
0 1 2 3 4
Time (hr)

Figure 5.9: Plot of phases volume fraction as function of time, for an ExxonMobil flow loop
experiment with Conroe oil, 50% liquid loading, 15% water cut, and 1 m/s fluid mixture velocity.
The estimated experimental hydrate volume fraction is also shown for comparison with the model
prediction (Test # XoM-2011-9).

For Test XoM-2011-14 with a higher liquid loading of 75 vol.% and same low water cut of 15% as

in the previous case, the model also capture the exothermic response at the hydrate formation onset

after the experimental cooling period, as observed in Figure 5.10. The predicted fluid distribution

as function of time is shown in Figure 5.11; the experimental hydrate volume fraction is also

shown (gray dashed curve). The volume fraction of gas remains almost constant throughout the

experiment, with a slight decrease during the hydrate formation. Note that this is a constant

pressure experiment and gas is being injected into the loop to maintain a constant pressure. The

water volume fraction decreases after hydrates start to form. In this experiment, water is not

completely consumed because of the greater amount of liquids. The model overpredicts the hydrate

volume fraction with a relative error of 10% with respect to the experimental value, for this low

water cut experiment. However, the predicted hydrate formation rate seems to be in agreement

with the experiment, since the hydrate volume fraction curves are nearly parallel.

With a different oil phase (King Ranch condensate) used in Test XoM-2011-25, with 50 % liq-

uid loading, 25% water cut, and 1 m/s fluid mixture velocity, the model for hydrate formation in

oil-dominated systems approximates the hydrate formation rate at earlier formation times (Fig-

ure 5.12). For the hydrate volume fraction estimates, there is an increase in the hydrate formation

rate observed in the experiment (at around 2 hours) that is not capture by the model (see Fig-

ure 5.12). However, in general the model does capture the trend of the experimental data and can

95
310
CSMHyK
oil−dominated

Temperature (K)
300 model

290 Experiment

280
Ambient
temp.

0 1 2 3
Time (hr)

Figure 5.10: Plot of fluid and ambient temperature as function of time, for an ExxonMobil flow
loop experiment with Conroe oil, 75% liquid loading, 15% water cut, and 1 m/s fluid mixture
velocity (Test # XoM-2011-14).

0.4

Gas
0.3
Volume fraction

0.2
Water
Hydrate

0.1
Experimental
hyd.

0
0 0.5 1 1.5 2 2.5 3 3.5
Time (hr)

Figure 5.11: Plot of phases volume fraction as function of time, for an ExxonMobil flow loop
experiment with Conroe oil, 75% liquid loading, 15% water cut, and 1 m/s fluid mixture velocity.
The estimated experimental hydrate volume fraction is also shown for comparison with the model
prediction (solid lines) (Test # XoM-2011-14).

96
be used for estimation purposes, though underpredicts hydrate volume fraction at later times.

0.2

Hydrate volume fraction


0.15 Experiment

0.1

CSMHyK
0.05 oil−dominated
model

0
0 1 2 3 4
Time (hr)

Figure 5.12: Plot of hydrate volume fraction as function of time, comparing experimental data
with predictions using the oil-dominated hydrate formation model for an ExxonMobil flow loop
experiment with King Ranch condensate, 50% liquid loading, 25% water cut, and 1 m/s fluid
mixture velocity (Test # XoM-2011-25).

For Test XoM-2011-23, with Conroe oil, 50% liquid loading, 75% water cut and 1 m/s fluid

mixture velocity, the plot of hydrate volume fraction as function of time (Figure 5.13) shows good

agreement between the predicted and measured hydrate formation rate, up to about 3 hours, where

the model experiences some instabilities in the calculated hydrate volume fraction. The assumption

of all water dispersed as droplets in a continuous oil phase, results in all hydrate particles being

carried by the oil phase. At a certain hydrate concentration the model predicts a huge increase

in the oil effective viscosity, predicting a hydrate plug, which results in the instabilities observed.

This is related with the assumption of maximum packing fraction of spherical particles in the

oil phase, which is reached considering the small amount of oil in this case. The 75% water cut

represents a limiting condition for the oil-dominated model, where the main assumptions made in

the development of the model are no longer valid. This high water cut value could represent the

transition to a water-dominated system, which is studied in the next section.

5.2 Model for hydrate formation in water-dominated systems

Water-dominated systems are characterized by a free and continuous water layer. As the oil/gas

fields mature, the produced water content increases, excess water cannot be further dispersed in

the oil, and a free water layer may form. Some oil and gas will also be dispersed in the free water

97
0.5
CSMHyK
oil−dominated

Hydrate volume fraction


0.4 model

0.3

0.2
Experiment
0.1

0
0 1 2 3 4 5
Time (hr)

Figure 5.13: Plot of hydrate volume fraction as function of time, comparing experimental data
and oil-dominated model prediction for an ExxonMobil flow loop experiment with Conroe oil, 50%
liquid loading, 75% water cut, and 1 m/s fluid mixture velocity (Test # XoM-2011-23).

layer due to shear forces.

A simplified scenario that considers a liquid aqueous phase and a hydrocarbon gaseous phase is

used for the development of a conceptual model for hydrate formation in high water cut systems.

For modeling purposes, the problem is divided into three stages (see Figure 5.14). The first stage

consists of estimating surface area between water and gas phases, which can be obtained from

reported multiphase flow correlations that estimate the entrainment of gas bubbles in the continuous

water phase. Second, assuming immediate hydrate nucleation to occur after a specified subcooling,

the hydrate growth rate can be calculated using a mass transfer-based growth model. Third, the

plugging criterion can be estimated as function of hydrate volume fraction, fluid velocity and liquid

holdup.

Figure 5.14: Conceptual picture for hydrate formation in water-dominated systems consisting of
gas and water phases.

98
To complement the stages identified in Figure 5.14 for the mechanism of hydrate formation in

water-dominated systems, it is required to merge two known phenomena of multiphase flow. First,

the determination of multiphase gas–liquid flow regimes, as illustrated in Figure 5.15, can be useful

in estimating pressure drop in the pipeline and surface area between water and gas phases. Second,

solid particles that are dispersed in the liquid phase upon hydrate formation resemble slurry flow.

Figure 5.15: Schematic illustration of gas-liquid flow regimes in horizontal pipe; adapted from
Brennen (2005).

With increasing solid particle volume fraction, four different slurry flow regimes can occur in

water–hydrate systems, as illustrated in Figure 5.16 (Brennen, 2005). The first solid–liquid flow

regime consists of a homogeneous distribution of well-mixed solid particles in the liquid phase

at low particle concentration. The second solid–liquid flow regime consists of a heterogeneous

distribution with faster sedimentation of larger particles generating a vertical gradient in the particle

concentration. The third solid–liquid flow regime consists of a moving bed of particles. The fourth

solid–liquid flow regime consists of a stationary bed of packed particles.

The hydrate formation mechanism in water-dominated systems differs from the one proposed for

oil-dominated systems. Joshi et al. (2011b) studied hydrate-hydrate cohesion forces in water using a

micromechanical force apparatus, concluding that the magnitude of hydrate cohesion forces in water

are relatively low and cannot explain particle agglomeration as part of the plugging mechanism.

Particle agglomeration is a major part of the hydrate plugging mechanism in oil-dominated systems.

Further studies of water-dominated systems by Joshi et al. (2011a) using the ExxonMobil flow loop

99
Figure 5.16: Schematic illustration of slurry flow (solids dispersion in liquid phase) regimes; adapted
from Brennen (2005) and Kitanovski and Poredos (2002).

showed that pressure drop in the pipe depends on a complex flow regime transition, involving

multiple phases and solid dispersion in the liquid phase. Three different regions were identified

based on pressure drop measurements across the flow loop after hydrate nucleation, as shown in

Figure 5.17. Region I consists of constant pressure drop upon hydrate formation, which is associated

with a homogeneous distribution of hydrate particles in the liquid phase. Region II consists of a

sharp increase in pressure drop with increasing hydrate concentration. The hydrate concentration

at which the pressure drop starts increasing is identified as Φtransition , and denotes the transition

to a heterogeneous distribution of hydrate particles in the liquid phase. Region III consists of large

fluctuations in pressure drop, representing a plugged condition in the flow loop. It is hypothesized

that a particle bed and a wall deposit forms in Region III, increasing pressure drop.

An improved representation of the hydrate formation mechanism in water-dominated systems

results when the multiphase gas–liquid flow regimes are combined with slurry flow regimes, as il-

lustrated in Figure 5.18. The proposed conceptual model for water-dominated systems starts with

a homogeneous distribution of hydrate particles in a stratified wavy gas–liquid flow regime. In-

creasing the hydrate volume fraction, the system evolves to a heterogeneous distribution of hydrate

particles in the liquid phase, due to buoyant forces. An experimentally observed increase in hydrate

volume fraction marks the transition to a more complicated flow regime, which exhibits combined

characteristics of slug flow with a stationary bed at the tail of the slug and a heterogeneous distri-

bution of hydrate particles into the liquid slug; this seems to be responsible for large fluctuations

100
Figure 5.17: Typical behavior of measured pressure drop observed after hydrate formation onset
in high water cut flow loop experiments; adapted from Joshi et al. (2011a).

in the pressure drop leading to a hydrate plug, preventing fluid flow.

From high water cut experiments performed in the ExxonMobil flow loop varying the pump

speed, Joshi et al. (2011a) observed that the transition from Region I to Region II is irreversible

with respect to mixture velocity, suggesting that the transition hydrate concentration (Φtransition )

would be an indication of eventual plugging with continued hydrate growth. This transition may

be used as a plugging criterion. A correlation to predict Φtransition as a function of the mixture

Reynolds number and liquid loading was developed by Grasso (presented in the August 2012 CSM

hydrate consortium meeting) from experimental data obtained at the ExxonMobil and University

of Tulsa flow loops at constant mixture velocities (Figure 5.19). The mixture Reynolds number is

defined as,

ρw vm d
Re = (5.8)
µw
This Φtransition correlation can be applied as a prediction tool to estimate the maximum amount

of hydrates that a system is able to flow without a plugging risk. This correlation is applicable

between two boundaries, the lower boundary is the deposition velocity and the upper boundary is

the packing volume fraction of sphere particles.

101
Figure 5.18: Conceptual model for hydrate formation mechanism in water-dominated systems
consisting of an aqueous phase, a gaseous phase and solid hydrate particles, showing transition of
flow regimes with increasing hydrate volume fraction.

Figure 5.19: Correlation of transition hydrate concentration (Φtransition ) as function of Reynolds


number (Re) and liquid loading (LL) in vol.%, developed from experimental data obtained at the
ExxonMobil and University of Tulsa flow loops at constant mixture velocities; presented in the
August 2012 CSM hydrate consortium meeting by Giovanny Grasso.

102
5.2.1 Description of the model

A mass transfer model proposed by Skovborg and Rasmussen (1994) is used to predict the gas

consumption rate during hydrate formation in water-dominated systems,

dmgas
= −kmass As (Csol − Csol−hyd ) (5.9)
dt
The mass transfer model uses as driving force for hydrate formation, the difference between gas

solubility in water (Csol ) and gas concentration in water in equilibrium with hydrates (Csol−hyd ).

The mass transfer model relies on two key quantities: the gas-water surface area (As ) and the mass

transfer coefficient (kmass ).

The gas-water surface area is determined using empirical correlations that depend on the mul-

tiphase flow regime (stratified flow and slug flow). For stratified flow, the gas-water surface area is

calculated assuming a stratified smooth interface, according to the illustration shown in Figure 5.20.

The cross sectional area of the water layer (Aw ) as a function of pipe radius (r) and wetted angle

(α) is given by,

r2
Aw = (2α − sin 2α) (5.10)
2

Figure 5.20: Cross section of pipe with stratified smooth flow regime.

The cross sectional area of the pipe (Ap ) is,

Ap = πr2 (5.11)

103
The relationship between liquid holdup (HL ) and wetted angle (α) in stratified flow is obtained

by the definition of liquid hold up (HL ),

Aw 2α − sin 2α
HL = = (5.12)
Ap 2π
This is a non-linear function of the wetted angle. Given the liquid hold up, this equation can be

solved for the wetted angle using a numerical method. Alternatively, an explicit approximation for

the wetted angle as function of liquid hold up can be used to avoid iterative calculations (Biberg,

1999),
π 1/3

α≈ 1 − (1 − HL )1/3 + HL (5.13)
2
Then, the gas-water surface area per unit volume for stratified flow is,

2r sin (α) ∆x 2 sin (α)


As = 2
= (5.14)
πr ∆x πr
For slug flow, the major contribution of gas-water surface area is at the liquid slug, where the

dissipation of energy at the front of the liquid slug generates surface area by the entrainment of

gas bubbles in the liquid slug, as illustrated in Figure 5.21 (Andreussi and Bendiksen, 1989). The

gas bubble dispersion in the water layer can be calculated using a correlation for liquid holdup in

the slug (HLs ) as a function of the mixture velocity (Gregory et al., 1978),
1
HLs = (5.15)
vm 1.39

1+ 8.66
where vm is the mixture velocity.

Figure 5.21: Illustration of gas bubbles entrainment in the liquid slug, due to energy dissipation
in the front of the liquid slug; adapted from Andreussi and Bendiksen (1989).

There is an alternative correlation to estimate the liquid holdup in the slug by Gomez et al.

(2000), which can be used in horizontal and upward vertical flow and is a function of the slug

Reynolds number (ReLs ) and pipe inclination angle in radians (θ),

HLs = 1.0 exp − 0.45θ + 2.48 × 10−6 ReLs



(5.16)

104
The liquid slug Reynolds number is defined as,

ρL vm d
ReLs = (5.17)
µL
The gas bubble size in the slug body can be estimated using a correlation for horizontal pipelines

under turbulent liquid flow conditions based on the theory proposed by Levich (1962), as presented

by Hesketh et al. (1987). The maximum stable gas bubble diameter is given by,

σ 0.6
   0.5 
d
db,max = 1.38 (W ecrit )0.6 (5.18)
ρc0.3 ρd0.2 µc0.1 1.1
vSc
where W ecrit is the critical Weber number, which is equal to 1.0 if the forces tending to break up

and the forces tending to restore the gas bubble are equal (i.e., equilibrium bubble size); σ is the

interfacial tension; ρc is the density of the continuous phase, ρd is the density of disperse phase, µc

is the viscosity of continuous phase; d is the internal pipe diameter; and vSc is superficial velocity

of the continuous phase.

The Sauter mean diameter (db,32 ) is related to the interfacial area of the dispersed phase and

should be used for the purposes of mass and heat transfer calculations,

db,32 = 0.62db,max (5.19)

The number of bubbles per unit volume (Nbubbles ) is obtained from the liquid slug void fraction

(1 − HLs ) divided by the volume of one gas bubble, considering the Sauter mean diameter,

6(1 − HLs )
Nbubbles = (5.20)
πd3b,32
Then the gas-water surface area per unit volume for slug flow is,

As = πd2b,32 Nbubbles (5.21)

The mass transfer coefficient is calculated using a correlation presented by Hanratty (1991),

vL∗
kmass = 0.5 (5.22)
NSC
where vL∗ is the characteristic liquid phase velocity given by,
r
τi
vL∗ = (5.23)
ρL
and NSC is the Schmidt number defined as,
µL
NSC = (5.24)
ρL Dgas−water

105
where Dgas−water is diffusivity coefficient of gas in water. Here the gas is assumed to be mostly

methane and the diffusivity coefficient of methane in water can be calculated using the Wilke-Chang

correlation (Kudsen et al., 1999).

The interfacial stress is calculated using,

ρg (vg − vL )2
τi = fi (5.25)
2
The interfacial friction factor (fi ) is assumed constant and equal to 0.022.

The concentrations of methane in water (Csol and Csol−hyd ), used in equation 5.9, are calculated

a priori using a commercial thermodynamic program (e.g., Multiflash


R
or PVTSim
R
) for a range of

pressure and temperature values. A look-up table is used to interpolate the methane concentrations

given a system pressure and temperature.

5.2.2 Implementation of the model into a transient multiphase flow simulator

The model for hydrate formation in water-dominated systems was implemented in this thesis

work in the OLGA


R
transient multiphase flow simulator, through the development of a user-defined

plug-in program, that allows to define a gas hydrate phase in the OLGA
R
’s flow model framework.

Figure 5.22 shows a schematic representation of the mass fields considered for water-dominated

systems, which includes gas and water as continuous phases, and hydrate and gas as dispersed

phases in the water continuous phase. The user-defined plug-in program calculates additional

contributions to the multiphase flow conservation equations using the PVT and flash modules.

Figure 5.22: Illustration of mass fields considered in the water-dominated hydrate formation
model coupled with the OLGA R
transient multiphase flow simulator, including gas and water as
continuous phases, and hydrate and gas as dispersed phases in the water continuous phase.

106
The PVT module calculates the properties for the hydrate phase, making additional contri-

butions to the mass, momentum and energy conservation equations. Typical values of hydrate

structure I (Sloan and Koh, 2008a) are assigned for density (910 kg/m3 ), heat capacity (2080 J/(kg

K)), thermal conductivity (0.49 W/(m K)). The enthalpy of the hydrate phase (hhyd ) is a function

of the enthalpy of the gas (hg ) and water (hw ) phases at the system temperature and pressure,

and the heat of reaction (∆hhyd ), to account for the exothermic nature of the hydrate formation

reaction,

hg + nH hw + ∆hhyd
hhyd = (5.26)
1 + nH
where nH is the hydration number (water to gas ratio in the hydrate phase).

The flash module calculates the gas consumption rate during hydrate formation using the mass

transfer model (equation 5.9). The flash module makes direct contributions to the mass conservation

equations by calculating the mass transfer rates from gas and water phases to the hydrate phase

during hydrate formation, and the opposite during hydrate dissociation where gas and water are

released from the hydrate phase. The mass transfer rates are used to update the phase masses of

gas, water and hydrate.

5.2.3 Validation of the model against flow loop data

The prediction performance of the CSMHyK water-dominated model was compared against

experimental data obtained using the ExxonMobil flow loop for high water cut systems. Ten high

water cut experiments were selected for predictions using the water-dominated hydrate formation

model (Table 5.1). These experiments cover the range of high water cut values from 100 to 75

vol.%, and different mixture velocities from 0.75 to 2.5 m/s. The cases with 100 vol.% water cut,

only included hydrocarbon gas (methane) and liquid water, no liquid hydrocarbon phase was loaded

into the loop.

The water-dominated hydrate formation model showed excellent agreement with experimental

data for two cases with 50 vol.% liquid loading and 100 vol.% water cut; Test XoM-2011-1 with 1

m/s fluid mixture velocity, and Test XoM-2010-2 with 2.5 m/s fluid mixture velocity. Figure 5.23

to Figure 5.25 show the comparison between the experimental data and simulation results from

Test XoM-2011-1 with 50% liquid loading, 100% water cut, and 1 m/s fluid mixture velocity.

107
Table 5.1: Experiments from the ExxonMobil flow loop selected for predictions using the water-
dominated hydrate formation model.

Test # Liquid loading Water cut Mixture velocity


(%) (%) (m/s)
XoM-2010-2 50 100 2.5
XoM-2011-1 50 100 1.0
XoM-2011-3 75 100 0.75
XoM-2011-4 75 100 1.0
XoM-2011-5 75 100 2.0
XoM-2011-20 50 90 1.0
XoM-2011-21 50 90 1.75
XoM-2011-22 50 90 2.5
XoM-2011-23 50 75 1.0
XoM-2011-24 50 75 2.5

Figure 5.23 shows the temperature of the system as function of time; in the model, an overall heat

transfer coefficient of 80 W/(m2 K) (14 BTU/(ft2 hr K)) was used to match the cooling stage of

the experiment. This case did not show a noticeable temperature increase at the time of hydrate

nucleation. Figure 5.24 shows the gas injected into the loop, required to maintain a constant

pressure inside the flow loop, as function of time. The initial increase in gas injection corresponds

to the cooling of fluids, with a contraction of the gas phase. There is a change in the slope of

the gas injection curve that indicates the hydrate formation onset. The slope of the gas injection

curve during hydrate formation can be related to the hydrate formation rate. There is a good

match between the experimental data and the simulation results. Figure 5.25 shows the hydrate

volume fraction as function of time with the calculated Φtransition . When the hydrate volume

fraction is greater than Φtransition the system enters in Region II and the pressure drop increases,

eventually leading to a hydrate plug. A deviation of the estimated hydrate volume fraction from

the experimental values above the Φtransition is observed, since the model does not considers the

phenomena affecting the pressure drop, which tend to decrease the hydrate formation rate.

For Test XoM-2010-2, with a higher fluid mixture velocity (2.5 m/s), the experiment exhibited

a slight increase in temperature at the point of hydrate formation onset, which is captured by the

model, as shown in Figure 5.26. Figure 5.27 shows the gas injected into the loop, with an excellent

agreement between the model prediction and the experimental values during the cooling stage of the

108
310
CSMHyK

Temperature (K)
300 water−dominated
model
290
Experiment

280
Ambient (input)
270
0 0.5 1 1.5 2 2.5
Time (hr)
Figure 5.23: Plot of fluid and ambient temperature as function of time, comparing experimental
data with predictions using the model for hydrate formation in water-dominated systems for an
ExxonMobil flow loop experiment with 50% liquid loading, 100% water cut, and 1 m/s fluid mixture
velocity (Test # XoM-2011-1).

10 CSMHyK
Gas injected (kg)

water−dominated
8
model
6

4
Experiment
2

0
0 0.5 1 1.5 2 2.5
Time (hr)
Figure 5.24: Plot of gas injected into flow loop as function of time, comparing experimental
data with predictions using the model for hydrate formation in water-dominated systems for an
ExxonMobil flow loop experiment with 50% liquid loading, 100% water cut, and 1 m/s fluid mixture
velocity (Test # XoM-2011-1).

109
Hydrate volume fraction
0.25
Experiment
0.2
CSMHyK

in liquid
0.15 water−dominated
model
0.1
Φtransition
0.05

0
0 0.5 1 1.5 2 2.5
Time (hr)
Figure 5.25: Plot of hydrate volume fraction as function of time, comparing experimental data with
predictions using the model for hydrate formation in water-dominated systems for an ExxonMobil
flow loop experiment with 50% liquid loading, 100% water cut, and 1 m/s fluid mixture velocity
(Test # XoM-2011-1), also showing the calculated Φtransition .

experiment (i.e., gas injection required to maintain constant pressure due to gas phase contraction

during cooling). Upon hydrate formation the gas injection rate increases, and the model matches

the measured gas injection rate. Figure 5.28 shows the hydrate volume fraction as function of time

with the calculated Φtransition . Similar to the previous case, the model predicts a hydrate formation

rate in agreement with the experimental values up to the transition hydrate concentration, where

predicted and measured values diverge. In this case, the hydrate formation rate increases slightly

above Φtransition .

CSMHyK
Temperature (K)

300
water−dominated
model
290
Experiment

280 Ambient (input)

270
0 0.5 1 1.5 2
Time (hr)
Figure 5.26: Plot of fluid and ambient temperature as function of time, comparing experimental
data with predictions using the model for hydrate formation in water-dominated systems for an
ExxonMobil flow loop experiment with 50% liquid loading, 100% water cut, and 2.5 m/s fluid
mixture velocity (Test # XoM-2010-2).

110
20

Gas injected (kg)


15
CSMHyK
water−dominated
10
model

Experiment
0
0 0.5 1 1.5 2
Time (hr)
Figure 5.27: Plot of gas injected into flow loop as function of time, comparing experimental
data with predictions using the model for hydrate formation in water-dominated systems for an
ExxonMobil flow loop experiment with 50% liquid loading, 100% water cut, and 2.5 m/s fluid
mixture velocity (Test # XoM-2010-2).
Hydrate volume fraction

Experiment
in liquid

0.4

0.2 CSMHyK
water−dom. Φtransition
model
0
0 0.5 1 1.5 2
Time (hr)
Figure 5.28: Plot of hydrate volume fraction as function of time, comparing experimental data with
predictions using the model for hydrate formation in water-dominated systems for an ExxonMobil
flow loop experiment with 50% liquid loading, 100% water cut, and 2.5 m/s fluid mixture velocity
(Test # XoM-2010-2), also showing the calculated Φtransition .

111
For experiments with liquid hydrocarbon (Conroe oil), the model also offers a good approxima-

tion of the experimental data. Results for tests XoM-2011-21 (90 vol.% water cut) and XoM-2011-24

(75 vol.% water cut) are shown in Figure 5.29 and Figure 5.30. In both cases the model predicts

with acceptable agreement the thermal behavior of the system, and underpredicts the amount of

hydrates formed by 10% relative error.

310

Hydrate volume fraction


Temperature (K)

Experiment
300 CSMHyK
CSMHyK water−dominated
water−dominated

in liquid
0.4 model
290 model

0.2
280
Φtransition
Ambient Experiment
270 0
0 1 2 3 0 1 2 3
Time (hr) Time (hr)
(a) Temperature (b) Hydrate volume fraction

Figure 5.29: Comparison of experimental data with predictions using the model for hydrate forma-
tion in water-dominated systems for Test XoM-2011-21 with Conroe oil, 50% liquid loading, 90%
water cut, and 1.75 m/s fluid mixture velocity: (a) ambient and fluid temperature as function of
time, (b) hydrate volume fraction as function of time, including the calculated Φtransition .

To summarize the prediction errors of the high water cut cases studied, Table 5.2 presents the

measured and predicted hydrate transition concentrations (Φtransition ), and the relative percentage

error. The hydrate transition concentration is considered a criterion to determine the hydrate plug

formation in the pipeline. The maximum prediction error for the hydrate transition concentration

obtained from the ten cases studied is about 12%, with an average error of 4.5%.

5.3 Model for hydrate formation in gas-dominated systems

Gas-dominated systems may be characterized by a predominant gaseous phase with small

amounts of liquid hydrocarbon (i.e., condensate) and water. Hammerschmidt (1934) was the first to

determine that gas hydrates were plugging natural gas transmission lines well above the ice forma-

tion temperatures. Dorstewitz and Mewes (1994) published the first flow loop study investigating

gas systems. Hydrates were observed to first form on the pipe wall at the water-gas interface. The

112
Hydrate volume fraction
CSMHyK
Temperature (K)

300
water−dominated Experiment
model

in liquid
290 0.4
Experiment

280 0.2 Φtransition


CSMHyK
water−dominated
Ambient
model
270 0
0 1 2 3 0 1 2 3
Time (hr) Time (hr)
(a) Temperature (b) Hydrate volume fraction

Figure 5.30: Comparison of experimental data with predictions using the model for hydrate forma-
tion in water-dominated systems for Test XoM-2011-24 with Conroe oil, 50% liquid loading, 75%
water cut, and 2.5 m/s fluid mixture velocity: (a) ambient and fluid temperature as function of
time, (b) hydrate volume fraction as function of time, including the calculated Φtransition .

Table 5.2: Summary of prediction error of hydrate volume fraction at the hydrate transition
concentration (Φtransition ), for the ten high water cut cases studied.

Test # Experimental Predicted Relative Error


(Φtransition ) (Φtransition )
(vol.%) (vol.%) (%)
XoM-2010-2 22.59 26.69 3.26
XoM-2011-1 8.16 8.87 8.70
XoM-2011-3 6.05 5.99 0.98
XoM-2011-4 8.20 9.17 11.83
XoM-2011-5 18.41 20.35 10.54
XoM-2011-20 8.97 8.91 0.66
XoM-2011-21 17.81 17.36 2.53
XoM-2011-22 25.38 26.37 3.90
XoM-2011-23 9.17 8.97 2.19
XoM-2011-24 26.38 26.30 0.30

113
hydrate layer then grew along the pipe wall, until the entire perimeter of the pipe was covered with

hydrates. Hatton and Kruka (2002) conducted a flow loop study, observing hydrate buildup at the

pipe wall, as well as hydrate formation in the bulk liquid in a horizontal 3-inch pipe. The observed

buildup on the pipe wall suggests a possible plugging mechanism in gas pipelines.

Rao et al. (2011) presented a laboratory scale study of methane hydrate deposition and growth.

Using a high pressure Jerguson visual cell and maintaining a continuous flow of water-saturated

methane gas, they were able to quantify hydrate growth on the outside of a small cold pipe placed

inside the visual cell. From the measurements of thickness and volume of the hydrate deposits as

function of time, Rao et al. (2011) developed a model based on heat and mass balances.

Figure 5.31 shows a proposed conceptual model for gas hydrate formation in gas-dominated

systems, consisting of four main stages. In the first stage, based on the assumption of immediate

hydrate nucleation after a specified subcooling, gas hydrate forms on the pipe wall at the gas-water

interface. Free or condensed water may be present from the saturated gas phase due to a decrease

in temperature. A layer of gas hydrates grow to cover the perimeter of the pipe. The second

stage consists of the inward growth of annular hydrate deposits (Dorstewitz and Mewes, 1994;

Hatton and Kruka, 2002). An increase in shear forces, related to increased flow velocities because

of narrowed conduit diameters, could promote sloughing of hydrate deposits from the pipe wall,

releasing hydrate particles that can accumulate in a jamming process plugging the pipe (Hatton

and Kruka, 2002). However, more fundamental work is required to understand the conditions for

sloughing of hydrates deposits from the pipe wall and subsequent jamming to form a plug. These

last two phenomena are currently under investigation in the Center for Hydrate Research at the

Colorado School of Mines.

5.3.1 Description of the model

The model for hydrate formation in gas-dominated systems was specially designed for cases

with a water-saturated gas phase flowing through a pipe without free liquid phases. There is no

free water initially in the system, all the water is dissolved in the gas phase. If the pipe surface

temperature falls below the dew point temperature, water condenses on the cold pipe surface.

The amount of liquid water condensing out of the water-saturated gas phase is determined by the

decrease in water concentration in the gas phase, through a phase equilibrium calculation.

114
Figure 5.31: Conceptual picture for hydrate formation in gas-dominated systems consisting of gas,
some condensate and water; adapted from Lingelem et al. (1994).

The condensed liquid water is immediately converted to gas hydrate. The hydrate growth

rate is calculated using an intrinsic kinetic equation, of first order in temperature departure from

equilibrium,
 
dmgas k2
= −k1 exp As ∆Tsub (5.27)
dt Tsystem
The intrinsic kinetics equation gives the rate of gas consumption during hydrate formation as

a function of the kinetics rate constants (k1 and k2 ), the surface area (As ) between water and

the gas phase, and the subcooling (∆Tsub ) as the thermal driving force. Vysniauskas and Bishnoi

(1983) measured the gas consumption rate during the formation of methane hydrate at different

temperatures and subcooling, in the absence of either mass or heat transfer resistances. Their

experimental data, were used to regressed the intrinsic rate constants, k1 = 7.3548 × 1017 and

k 2 = −13600 K.

The amount of hydrate phase formed is evenly distributed in a deposited layer on the pipe wall.

The hydrate deposit is assumed to have a constant porosity. The porosity of the hydrate deposit

(φhyd deposit ) is used to determine the deposit thickness, through the definition of the hydrate deposit

density (ρhyd deposit ) as,

ρhyd deposit = (1 − φhyd deposit )ρhyd + φhyd deposit ρgas (5.28)

The presence of a hydrate deposited layer on the pipe wall will add thermal resistance to the

overall heat transfer between the fluid and the inner pipe wall. The heat transfer coefficient of

115
the hydrate deposit (HT Chyd deposit ) is calculated, assuming a steady-state heat balance across

the hydrate wall layer, as the hydrate thermal conductivity (khyd ) divided by the hydrate deposit

thickness (δhyd deposit ),

khyd
HT Chyd deposit = (5.29)
δhyd deposit

5.3.2 Implementation of the model into a transient multiphase flow simulator

The model for hydrate formation in gas-dominated systems was implemented in this thesis work

in the OLGA
R
transient multiphase flow simulator, through the development of a user-defined plug-

in program, that allows one to define a gas hydrate phase in the OLGA
R
’s flow model framework.

(Figure 5.32) shows a schematic representation of the mass fields considered for gas-dominated

systems, which includes a water-saturated gas phase and a hydrate film deposited on the pipe

wall. The user-defined plug-in program calculates additional contributions to the multiphase flow

conservation equations using the PVT, flash and heat transfer modules.

Figure 5.32: Illustration of the mass fields considered in the gas-dominated hydrate formation
model coupled with the OLGA R
transient multiphase flow simulator, including a gas phase satu-
rated with water and hydrate film deposited on the pipe wall.

The PVT module calculates the properties for the hydrate deposited layer based on properties

of the hydrate and gas phases. The PVT module makes contributions to the mass, momentum

and energy conservation equations. Typical values of hydrate structure II (Sloan and Koh, 2008a)

are assigned for the hydrate phase properties, such as, density (914 kg/m3 ), heat capacity (498.83

J/(kg K)), thermal conductivity (0.5 W/(m K)). The density of the hydrate deposit is calculated

using equation 5.28. The enthalpy of the hydrate phase (hhyd ) is a function of the enthalpy of the

gas (hg ) and water (hw ) phases at the system temperature and pressure, and the heat of reaction

116
(∆hhyd ), to account for the exothermic nature of the hydrate formation reaction,

hg + nH hw + ∆hhyd
hhyd = (5.30)
1 + nH
where nH is the hydration number (water to gas ratio in the hydrate phase).

The flash module calculates the gas consumption rate during hydrate formation using the in-

trinsic kinetics equation 5.27. The amount of hydrates formed is limited by the available amount

of condensed water. The flash module makes direct contributions to the mass conservation equa-

tions by calculating the mass transfer rates from gas and water phases to the hydrate phase during

hydrate formation, and the opposite during hydrate dissociation where gas and water are released

from the hydrate phase. The mass transfer rates are used to update the phase masses of gas, water

and hydrate.

The heat transfer module calculates the heat transfer coefficient of a deposited hydrate layer

using equation 5.29, and the corrected inner heat transfer coefficient as,

HT Cinner,OLGA · HT Chyd deposit


HT Cinner = (5.31)
HT Cinner,OLGA + HT Chyd deposit
OLGA
R
calculates one temperature for the fluid and also calculates the temperature of the pipe

wall, assuming radial heat conduction through concentric wall layers. An internal heat transfer

coefficient (HT Cinner,OLGA ) is calculated for the heat transfer from the fluid to the inner pipe wall.

The presence of a hydrate wall layer adds thermal resistance to the overall heat transfer, and this

is the contribution of the plug-in to the energy conservation equation.

This model has not been compared against experimental data, because of the scarcity of ex-

perimental data for gas-dominated systems. This represents the first attempt to implement the

capability of predicting hydrate film growth on pipe wall in a transient multiphase flow simulator,

and should be considered as a first pass model. In the next chapter, an example application of this

model is presented, in the context of water-saturated gas flowing through a horizontal pipeline,

showing the potential effect of hydrate film growth on pressure drop.

5.4 Summary

This chapter shows how the CSMHyK model have been improved and extended to water and

gas-dominated systems, through the development of user-defined plug-ins for OLGA


R
. The new

flexible flow model framework available in OLGA


R
allows one to include the effects of an additional

117
gas hydrate phase on the fluid flow behavior. The plug-in program can be coupled with OLGA
R
,

defining a hydrate phase and calculating additional contributions to the conservation equations.

A PVT module is used to calculate properties for the hydrate phase, a flash module is used to

calculate mass transfer rates for the mass conservation equations, and a heat transfer module is

used to calculate the heat transfer coefficient of a deposited hydrate layer, affecting the energy

equation.

The hydrate formation model for oil-dominated systems includes a new dynamic cohesion force

model. The model predicts the hydrate volume fraction measured in flow loop experiments with

good agreement. The water-dominated model uses a mass transfer-based hydrate growth model

to calculate the amount of hydrate formed in the system. A correlation to estimate the hydrate

transition concentration as function of Reynolds number and liquid loading is used as a plugging

criterion. The model predictions matched experimental data obtained in the ExxonMobil flow

loop. The gas-dominated model was also implemented as a plug-in for OLGA
R
. The model for

gas-dominated systems was specially designed for cases with a water-saturated gas phase flowing

through a pipe without free liquid phases. Water condenses on the cold pipe surface and imme-

diately is converted to gas hydrate. The gas-dominated hydrate formation model has not been

compared against experimental data due to the lack of such data, and should be considered as a

first pass model for hydrate film growth in gas-dominated systems.

118
CHAPTER 6

APPLICATION OF THE HYDRATE MATHEMATICAL MODEL

Offshore explorations in deeper and colder waters impose more challenging scenarios to the flow

assurance of the produced streams. High pressures and low temperatures of operation of subsea

production facilities with longer tiebacks will promote the formation of natural gas hydrates. Gas

hydrates form in the presence of appropriate quantities of gas and water, and are considered one of

the most challenging problems in deep subsea facilities because of their rapid formation compared

to other solid deposits (Sloan, 2005).

A transient gas hydrate model that predicts when and where gas hydrates form in flowlines will

have significant utility for the flow assurance engineer in the oil and gas industry. By predicting

gas hydrate formation and transportability, the gas hydrate model can be applied to design and

optimize oil and gas transport facilities, focusing on prevention, management, or remediation of

gas hydrates in flowlines.

This chapter presents examples of applications of the hydrate formation models developed in

this work and presented in Chapter 5 for oil, water and gas-dominated systems. The hydrate

plugging risk is studied, using the typical geometry and fluid properties of an offshore well from

the Caratinga field located in the Campos basin, Brazil (provided by Petrobras).

6.1 Estimating hydrate plugging risk in oil-dominated systems

This section focuses on the application of the hydrate formation model for oil-dominated sys-

tems, to study the hydrate plugging risk of an offshore well. Three different periods of the well

life are considered: an early stage with low produced gas/oil ratio (GOR) and low water cut, a

middle stage with an increased GOR, and a late stage with higher GOR and higher water cut.

The hydrate plugging risk is estimated from the calculation of three performance measures (pres-

sure drop along the flowline, hydrate volume fraction, and hydrate slurry relative viscosity) in an

attempt to quantify the plugging risk. Once the hydrate plugging risk is estimated for the three

different periods of the well life, the gas hydrate model is used to optimize the concentration of

thermodynamic hydrate inhibitor (THI) to be injected in order to reach a safe operation condition.

119
In this study, ethanol is considered as the THI because it is the typical hydrate inhibitor used in

South America (because of the large industrial production of ethanol in the region).

Additionally, the hydrate model for oil-dominated systems is used to study the effect of hydrates

in transient shut-in/restart operations of the offshore well and the performance of hydrate inhibition

in terms of ethanol concentration. In general, ethanol injection decreased the hydrate-plugging risk

in steady-state and transient operations.

6.1.1 Offshore well hydrate risk assessment

This section presents the approach followed to estimate the hydrate plugging risk in an offshore

well using the gas hydrate prediction model for oil-dominated systems. The simulation case consists

of typical well/flowline/riser geometry and fluid properties from the Caratinga field located at the

Campos Basin in Brazil. Figure 6.1 shows the geometry in terms of horizontal length and depth

from sea level, which consists of a slanted well with a diameter of 5 in., a departure of 2000 m, and

a depth of 1800 m. The well is connected to a straight flowline with a slight uphill slope, a diameter

of 6 in., and a horizontal length of approximately 13 km, followed by the bend into the riser that

leads up to the platform. Table 6.1 presents a summary of the Caratinga crude-oil properties and

composition measured by Sjöblom et al. (2010).

0
Platform
500 Wellhead

1000
Depth (m)

Flowline
1500

2000 Well
2500

3000
0 5 10 15
Horizontal distance (km)

Figure 6.1: Geometry of well, flowline, and riser, based on typical geometries from the Caratinga
field located at the Campos Basin in Brazil.

As a first approach to estimate hydrate plugging risk, a pressure versus temperature diagram,

shown in Figure 6.2, was constructed overlaying the flowline operation conditions under steady-state

120
Table 6.1: Properties and composition of Caratinga crude oil (Sjöblom et al., 2010).

Property Value
Density (kg/m3 ) 914.0
Viscosity dead oil (Pa.s) 0.262
Interfacial tension (mN/m) 23
Asphaltene content (%) 6.2
Saturates content (%) 39.8
Aromatic content (%) 39.8
Resin content (%) 14.3

flow with hydrate equilibrium curves at different ethanol concentrations. Whenever the flowline

operating conditions curve lies to the left of the hydrate equilibrium curves is considered a region

for hydrate formation. It can be observed that a portion of the flowline operating conditions curve

is located inside the hydrate formation region, confirming the feasibility of forming hydrates in this

system, and that more than 30 wt% ethanol is required to totally inhibit the system. However,

this diagram does not provide information about the amount of hydrates formed and hydrate

transportability; hence nothing can be said about the actual plugging risk level of this system.

14
Ethanol concentration (wt%)
40 30 20 10 0
12 Wellhead
Pressure (MPa)

630 m 210 m
10 3.2 km
Flowline operation
8 conditions
6 14.1 km

4
2
Topsides
0
260 280 300 320
Temperature (K)
Figure 6.2: Diagram of pressure versus temperature showing flowline operation conditions during
steady-state flow and hydrate equilibrium curves at different ethanol concentrations. Points on
flowline operation conditions curve indicate distance from wellhead.

121
The hydrate formation model for oil-dominated systems coupled with the OLGA
R
transient

multiphase flow simulator can be used to determine the flowline operating conditions and estimate

the amount and transportability of gas hydrates at different periods of the well life. During the

well life, an increase in the GOR and in the water cut is expected, which represents an increase

in the amount of fundamental components for gas hydrate formation. This means that different

periods in the well life are characterized by different magnitudes of GOR and water cut, resulting in

different amounts and transportability of hydrates in the system. On the basis of this information,

hydrate remediation techniques can also differ for different periods in the well life. In this study,

we consider three different scenarios representing periods in the well life: (i) an early stage with

low GOR and low water cut (Case 1), (ii) a middle stage with an increased GOR (Case 2), and

(iii) a late stage with higher GOR and higher water cut (Case 3).

The hydrate plugging risk is determined from the calculation of three performance measures

(∆Pf lowline is the pressure drop along the flowline, Φhyd is the hydrate volume fraction in pipe, and

µr is the hydrate slurry relative viscosity) and is classified in three qualitative levels as follows:

• Low-risk: ∆Pf lowline < 2 MPa (300 psi); Φhyd < 0.10; µr < 10. At this level, formed hydrates

are considered to be easily transported through the pipeline and up to the riser.

• Intermediate-risk: 2 MPa (300 psi) < ∆Pf lowline < 3.45 MPa (500 psi); 0.10 < Φhyd < 0.40;

10 < µr < 100. At this level, hydrates could still flow through the pipeline but represent a

risk where there are restrictions or changes in flow direction, where hydrates can accumulate,

jam, and plug the line.

• High-risk: ∆Pf lowline > 3.45 MPa (500 psi); Φhyd > 0.40; µr > 100. At this level, a very

viscous hydrate slurry forms, fluid flow stops, and the pipeline plugs.

Note that the performance measure values used to delimit the three levels of hydrate plugging

risk in this study were determined from the current version of the gas hydrate model and for the

considered well/flowline/riser geometry, and may or may not underestimate the effect of hydrates

on fluid flow in more general scenarios. These values could be adjusted by improving the accuracy

of the model with further knowledge of the mechanisms of hydrate formation and transportability

from future studies.

122
Case 1: Low Production GOR and Low Water Cut

Case 1 represents an early stage in the well life characterized by a liquid loading of 90 vol%, a

water cut of 30%, and a production GOR of 570 scf/STB. The calculated temperature distribution

along the flowline length is presented in Figure 6.3 with the hydrate equilibrium temperature.

The warm fluids flowing from the well cool down at the beginning of the flowline because of heat

transfer to the ocean (ambient temperature of 277 K). Hydrates start to form when the fluids

temperature decreases 3.6 K below the hydrate equilibrium temperature (i.e., a subcooling of 3.6

K). The hydrate heat of formation raises the temperature of the system. Then the heat transfer to

the ocean balance with the heat of formation, making the system temperature equal to the hydrate

equilibrium temperature, in a condition known as heat transfer limited. When all the free gas phase

is converted into hydrates, the hydrate formation reaction stops. The system cools down until the

seafloor temperature is reached, with no further hydrate formation. Then, the system temperature

increases and hydrates dissociate flowing up the riser.

340
Temperature (K)

320

300 Hydrate equilibrium temperature

280

Fluids temperature
260
0 5 10 15
Pipeline Length (km)

Figure 6.3: Temperature distribution along the pipeline length, showing system temperature
and hydrate equilibrium temperature corresponding to the steady-state flow predicted by the oil-
dominated gas hydrate formation model for Case 1.

The plot of the fluids distribution shown in Figure 6.4 supports the explanation previously

presented for the system temperature behavior. It can be observed at the beginning of the flowline

that the liquid volume fraction started at 0.9 (i.e., 90 vol% liquid loading) and increases up to a

100% liquid loading in the flowline, because of free gas being consumed during hydrate formation.

Some hydrates continue to form, due to the pressure drop along the flowline resulting in gas evolving

123
from the oil phase. Gas hydrates are transported and melted up the riser. Figure 6.5 shows the

ratio of hydrate slurry viscosity to the viscosity of the continuous oil phase; a maximum hydrate

slurry viscosity of 3.4 times the oil viscosity can be observed, with a location that coincides with

an accumulation of hydrates at the base of the riser in Figure 6.4.

Volume Fraction 0.8 Total liquids

0.6

0.4
Water
0.2
Gas Hydrate
0
0 5 10 15
Pipeline Length (km)

Figure 6.4: Fluid distribution in terms of volume fraction along the pipeline length, corresponding
to the steady-state flow predicted by the oil-dominated gas hydrate formation model for Case 1.

3.5
Slurry Relative Viscosity

2.5

1.5

1
0 5 10 15
Pipeline Length (km)

Figure 6.5: Hydrate slurry relative viscosity along the pipeline length, corresponding to the steady-
state flow predicted by the oil-dominated gas hydrate formation model for Case 1.

In general, Case 1 is considered to have a low risk of hydrate plugging because the calculated

flowline pressure drop of 1.6 MPa (230 psi) is less than the specified limit of 2 MPa (300 psi),

the maximum hydrate fraction in pipe of 0.09 is less than 0.10, and the maximum hydrate-slurry

124
relative viscosity of 3.4 is less than 10.

Case 2: Increased GOR

Case 2 represents a middle stage in the well life characterized by a liquid loading of 79 vol%,

a water cut of 32%, and a production GOR of 894 scf/STB. Figure 6.6 shows the calculated tem-

perature distribution along the flowline length and the hydrate equilibrium temperature. Similar

to that of Case 1, the warm fluids flowing from the well cool down at the beginning of the flowline

until a subcooling temperature of 3.6 K is reached when hydrates nucleate. The hydrate heat of

formation raises the temperature, and the system becomes heat transfer limited until all the ad-

ditional gas is converted into hydrates, consuming the water phase. When the hydrate formation

reaction stops, the system cools down, reaching the seafloor temperature with no further hydrate

formation. Then, the system temperature increases and hydrates dissociate, flowing up the riser.

340
Temperature (K)

320

300 Hydrate equilibrium temperature

280

Fluids temperature
260
0 5 10 15
Pipeline Length (km)

Figure 6.6: Temperature distribution along the pipeline length, showing system temperature
and hydrate equilibrium temperature corresponding to the steady-state flow predicted by the oil-
dominated gas hydrate formation model for Case 2.

Figure 6.7 shows a plot of fluid distribution along the pipeline length. It can be observed at

the beginning of the flowline that the liquid volume fraction started at nearly 0.8 and increases

until the flowline is filled by liquid; simultaneously, the gas volume fraction is reduced because

of a decrease in temperature and gas consumption during hydrate formation. In this case, water

is consumed almost completely during hydrate formation because of the increased GOR from the

well. Hydrates are melted on their way up the riser. But now a greater amount of hydrates, when

compared to Case 1, is present in the flowline, and could represent a higher risk of accumulating,

125
jamming, and plugging during flow through a restriction or change in flow direction. Figure 6.8

shows the ratio of hydrate slurry viscosity to the viscosity of the continuous oil phase along the

flowline length; a maximum hydrate slurry viscosity of 8.9 times the oil viscosity can be observed.

0.8 Total liquids

Volume Fraction 0.6

0.4 Hydrate
Gas
0.2
Water
0
0 5 10 15
Pipeline Length (km)

Figure 6.7: Fluid distribution in terms of volume fraction along the pipeline length, corresponding
to the steady-state flow predicted by the oil-dominated gas hydrate formation model for Case 2.

10
Slurry Relative Viscosity

0
0 5 10 15
Pipeline Length (km)

Figure 6.8: Hydrate slurry relative viscosity along the pipeline length, corresponding to the steady-
state flow predicted by the oil-dominated gas hydrate formation model for Case 2.

Case 2 is considered to have an intermediate risk of plugging with hydrates because the calcu-

lated flowline pressure drop of 2.7 MPa (398 psi) is between the specified lower limit of 2 MPa and

the upper limit of 3.45 MPa, and the maximum hydrate fraction in pipe of 0.30 is greater than

0.10.

126
Considering the previous results and the estimated hydrate plugging risk, the gas hydrate model

for oil-dominated systems was used to study the performance of THI injection, with the objective

of determining the optimum ethanol concentration needed to lower the hydrate plugging risk of

Case 2 to a low risk level. Figure 6.9 shows three plots of the performance measures (∆Pf lowline ,

Φhyd , and µr ) used to establish the hydrate plugging risk as function of ethanol concentration. It

can be observed that the injection of ethanol reduces the performance measures up to an ethanol

concentration of 30 wt%. The pressure drop along the flowline has a minimum value at 30 wt%,

while the hydrate volume fraction in pipe and the hydrate slurry relative viscosity become constant,

indicating a complete dissociation of hydrates. The ethanol injection reduces the hydrate plugging

risk to a safer operation point if the injected concentration is greater than or equal to 30 wt% of

ethanol.

2.75 0.4 10
Pressure drop along flowline (MPa)

Hydrate slurry relative viscosity


Hydrate volume fraction

2.7 8
0.3

2.65 6
0.2
2.6 4

0.1
2.55 2

2.5 0 0
0 10 20 30 40 0 10 20 30 40 0 10 20 30 40
Ethanol concentration (wt%) Ethanol concentration (wt%) Ethanol concentration (wt%)

Figure 6.9: Performance measures (∆Pf lowline , Φhyd , and µr ) used to establish the hydrate plugging
risk as a function of ethanol concentration, predicted by the oil-dominated gas hydrate formation
model for Case 2.

Case 3: Higher GOR and Higher Water Cut

Case 3 represents a late stage in the well life characterized by a liquid loading of 67.4 vol%,

a water cut of 52.5%, and a production GOR of 1,677 scf/STB. Figure 6.10 shows the calculated

temperature distribution along the flowline length and the hydrate equilibrium temperature at the

plugged condition. A great portion of the flowline length is at very low temperatures. The system

temperature increases and follows the hydrate equilibrium temperature in the direction of the riser.

Figure 6.11 shows the plot of fluids distribution along the pipeline length. It can be observed

that the liquid loading remains fairly high and the gas volume fraction decreases slightly, but the

127
310

300

Temperature (K)
Hydrate equilibrium temperature
290

280

270 Fluids temperature

260
0 5 10 15
Pipeline Length (km)

Figure 6.10: Temperature distribution along the pipeline length, showing system temperature
and hydrate equilibrium temperature corresponding to the steady-state flow predicted by the oil-
dominated gas hydrate formation model for Case 3.

water volume fraction decreases from the beginning of the flowline until it is completely consumed

in the hydrate formation reaction, resulting in a massive accumulation of hydrates in the middle of

the flowline. Figure 6.12 shows the ratio of hydrate slurry viscosity to the viscosity of the continuous

oil phase (in a logarithmic scale) along the pipeline length; it can be observed that the maximum

hydrate slurry viscosity reaches values greater than 1,000 times the oil viscosity, confirming the

formation of a hydrate plug.

0.8
Volume Fraction

Total liquids

0.6 Hydrate

0.4
Gas
0.2 Water

0
0 5 10 15
Pipeline Length (km)

Figure 6.11: Fluid distribution in terms of volume fraction along the pipeline length, corresponding
to the steady-state flow predicted by the oil-dominated gas hydrate formation model for Case 3.

In Case 3, the viscous hydrate slurry formed prevents the fluid from flowing. This case has

a high risk of plugging with hydrates because the calculated flowline pressure drop of 9 MPa is

128
4
10

Slurry Relative Viscosity


3
10

2
10

1
10

0
10
0 5 10 15
Pipeline Length (km)

Figure 6.12: Hydrate slurry relative viscosity along the pipeline length, corresponding to the
steady-state flow predicted by the oil-dominated gas hydrate formation model for Case 3.

greater than 3.45 MPa, the maximum hydrate fraction in pipe of 0.55 is greater than 0.40, and the

maximum hydrate-slurry viscosity is too high.

The oil-dominated hydrate formation model was used to study the performance of THI injection

at the conditions of Case 3, with the objective of determining the optimum ethanol concentration

needed to lower the hydrate plugging risk to a low risk level. Figure 6.13 shows three plots of the

performance measures (∆Pf lowline , Φhyd , and µr ) used to establish the hydrate plugging risk as

a function of ethanol concentration. It can be observed that the injection of ethanol reduces the

performance measures, preventing the formation of a hydrate plug. For 10 wt% of ethanol injected,

the models estimates a minimum value of pressure drop along the flowline for this systems, but

the amount of hydrates (about 40 vol.%) and slurry relative viscosity (> 10) indicates that the

system is at an intermediate hydrate plugging risk. The pressure drop along the flowline increases

with further increase in ethanol injection, which is related to the increase in liquid content and

the released gas from hydrate dissociation that changes the multiphase flow regime. The hydrate

volume fraction and the hydrate slurry relative viscosity decrease with increasing injected ethanol

concentration up to an ethanol concentration of 30 wt%, where hydrates are completely dissociated.

Ethanol injection prevents formation of a hydrate plug and remediates the hydrates completely if

the injected concentration is greater than or equal to 30 wt% of ethanol, reducing the hydrate

plugging risk.

129
4
0.8
Pressure drop along flowline (MPa)
10 10

Hydrate slurry relative viscosity


Hydrate volume fraction
3
8 0.6 10

2
6 0.4 10

1
4 0.2 10

0
2 0 10
0 10 20 30 0 10 20 30 0 10 20 30
Ethanol concentration (wt%) Ethanol concentration (wt%) Ethanol concentration (wt%)

Figure 6.13: Performance measures (∆Pf lowline , Φhyd , and µr ) used to establish the hydrate
plugging risk as a function of ethanol concentration, predicted by the oil-dominated gas hydrate
formation model for Case 3.

6.1.2 Well restart procedure optimization

Considering Case 2, which showed an intermediate risk of plugging with hydrates, the gas

hydrate prediction model for oil-dominated systems is used to study the effect of hydrates in

transient shut-in/restart operations and the system performance as a function of THI concentration

in the shut-in/restart procedure. Two valves are included in the model to control the shut-in and

restart operations, one valve located at the wellhead and another located at the platform end.

During the shutdown of production, the valves are closed for 50 hours, allowing the gravitational

segregation of fluids in the system toward a hydrostatic and thermal equilibrium. Figure 6.14

shows the system temperature distribution after 50 hours of shut-in and the corresponding hydrate

equilibrium temperature for Case 2 without ethanol injection. A maximum subcooling of 12 K

is observed over a great portion of the flowline. However, the amount of hydrates formed during

the rearrangement of fluids in the flowline following the shutdown of production is not significant,

as shown by the fluid distribution profile in Figure 6.15, which is similar to the one presented in

Figure 6.7 for the steady-state operation of the well. An accumulation of water toward the low spot

near the wellhead and the migration of the gas toward the topside in the riser can be observed;

some accumulation of unconverted water also takes place at the base of the riser, which is separated

from the gas by an oil layer.

130
30

Hydrate equilibrium temperature

Temperature (ºC)
20

10

0
Fluids temperature

−10
0 5 10 15
Pipeline Length (km)

Figure 6.14: Temperature distribution along the pipeline length, showing system temperature and
hydrate equilibrium temperature, after 50 hours of shut-in, predicted by the oil-dominated gas
hydrate formation model for Case 2 without ethanol.

1
Total liquids
0.8
Volume Fraction

0.6

0.4 Hydrate

0.2 Gas
Water

0
0 5 10 15
Pipeline Length (km)

Figure 6.15: Fluids distribution in terms of volume fraction along the pipeline length, after 50
hours of shut-in, predicted by the oil-dominated gas hydrate formation model for Case 2 without
ethanol.

131
After 50 hours of shut-in, the production is restarted following the valve opening schedule shown

in Figure 6.16. This valve opening schedule is similar to an optimized valve opening schedule

designed to minimize hydrate formation upon restart by Zerpa et al. (2011) using the same hydrate

prediction model, where production is restarted using a valve opening of 1% for 60 minutes, then

increasing progressively until the valves are fully open.

1
Valve opening (fraction)
0.8

0.6

0.4

0.2

0
0 50 100 150
Time (min)

Figure 6.16: Valve opening schedule used during the restart of production.

During the restart of production at different THI concentrations (from 0 to 40 wt% of ethanol),

it was found that an accumulation of hydrates was formed by the initial movement of fluids, and

this accumulation is transported through the flowline by a wave front. Figure 6.17 shows the

evolution of the restart of production in terms of pressure drop along the flowline as a function of

time (left plot) and the hydrate volume fraction at a fixed point in the middle of the flowline as a

function of time (right plot), for Case 2 with 20 wt% ethanol injected before the shut-in and during

the restart. It can be observed that pressure drop increases with time, exhibiting a large peak at

approximately 200 minutes, which coincides with the large peak of hydrate volume fraction. This

peak represents the hydrate accumulation that travels along the flowline transported by the wave

front and is followed by a bank of fluids with no hydrates; then, the system equilibrates, reaching

a stable value. This hydrate accumulation that moves through the flowline upon restart represents

a critical condition and is used to quantify the hydrate plugging risk.

The system performance during the restart procedure is evaluated for different ethanol con-

centrations using, as performance measures, the maximum values reached during the initial stage

of restart of the following variables: pressure drop along the pipeline, hydrate-volume fraction,

132
5 0.25

Hydrate volume fraction


Pressure drop (MPa) 4 0.2

3 0.15

2 0.1

1 0.05

0 0
0 500 1000 0 500 1000
Time (min) Time (min)

Figure 6.17: Evolution of the production restart in terms of pressure drop along the pipeline (left
plot) and hydrate volume fraction (right plot) at a point in the middle of the pipeline with respect
to time, predicted by the oil-dominated hydrate formation model for Case 2 with 20 wt% ethanol
injected before the shut-in and during the restart.

and hydrate-slurry relative viscosity. Figure 6.18 shows three plots of the performance measures

(max {∆Pf lowline (t)}, max {Φhyd (t)}, and max {µr (t)}) used to establish the hydrate plugging risk

as a function of ethanol concentration. It can be observed that the injection of ethanol reduces the

performance measures. The maximum pressure drop along the pipeline exhibits a minimum value

at 20 wt% ethanol injected, and increases with further increase in ethanol injection, which is related

to the balance between the decrease in hydrate slurry viscosity and the increase in liquid loading by

the inhibitor injection. The maximum hydrate volume fraction in the pipe decreases and remains

nearly constant between 10 and 20 wt% ethanol, then increases, showing a local maximum at 25

wt% ethanol, and decreases with further increase in ethanol concentration. The hydrate slurry

relative viscosity shows a behavior similar to that of the hydrate volume fraction, exhibiting a local

maximum at 25 wt% ethanol. Ethanol injection improves the system performance upon restart,

reducing the hydrate plugging risk. A concentration of 20 wt% ethanol seems to be the optimum

concentration to use in transient operations.

133
Max. Hydrate slurry relative viscosity
3
5.5 0.4 10

Max. Hydrate volume fraction


Max. Pressure drop (MPa)

0.35
2
5 10
0.3

0.25 1
4.5 10
0.2

0
4 10
0 10 20 30 40 0 10 20 30 40 0 10 20 30 40
Ethanol concentration (wt%) Ethanol concentration (wt%) Ethanol concentration (wt%)

Figure 6.18: Performance measures — max {∆Pf lowline (t)}; max {Φhyd (t)}; and max {µr (t)} —
used to evaluate the hydrate plugging risk as a function of ethanol concentration in transient restart
operations, predicted by the oil-dominated hydrate formation model for Case 2.

6.2 Estimating hydrate plugging risk in water-dominated systems

The hydrate formation model for water-dominated systems, presented in Section 5.2, was applied

to estimate the hydrate plugging risk in an offshore well. The geometry of the well/flowline/riser

used was presented in Figure 6.1. The fluid properties used are from the Caratinga field (Table 6.1),

with a high produced water cut of 85 vol.%. Two simulation cases are studied, the first case

considers a bare pipe without thermal insulation, and the second case accounts for a thermal

insulation layer on the flowline.

6.2.1 Uninsulated case

The flowline operation conditions for the uninsulated case are plotted in a pressure versus

temperature diagram, including the hydrate equilibrium curve for this system, to determine if this

subsea flowline enters the hydrate formation region (Figure 6.19). Produced reservoir fluids transfer

heat to the cold environment resulting in the fluids being cooled down, and the system enters in

the hydrate formation region (left of hydrate equilibrium curve), with a significant portion of the

pipeline being within hydrate forming conditions. Then, fluids flow up the riser with a significant

decrease in pressure, until the system is out of the hydrate formation region near the platform.

Figure 6.20 shows the temperature distribution along the length of the pipeline, predicted by

the water-dominated hydrate formation model. The temperature of the produced fluids decreases

below the hydrate equilibrium temperature, and hydrates start to form when the system reaches

134
20
Hydrate
equilibrium Wellhead

Pressure (MPa)
15 curve 590 m from wellhead

Flowline operation
10 conditions

5
Topsides

0
280 300 320
Temperature (K)
Figure 6.19: Diagram of pressure versus temperature showing flowline operation conditions dur-
ing steady-state flow, and hydrate equilibrium curve, predicted by the water-dominated hydrate
formation model for the uninsulated case.

a subcooling of 3.6 K. The temperature remains constant during hydrate formation, and decreases

around 5 km of pipeline length when all free gas is converted into hydrate (see Figure 6.21).

Then, the fluids temperature decreases further and equals the ambient temperature. The hydrate

equilibrium temperature decreases because of the decrease in system pressure in the riser, and the

system leaves the hydrate stable region.

Figure 6.21 shows the predicted fluid distribution in terms of volume fraction along the pipeline

length. The volume fraction of gas and water decreases while hydrate increases, and all the free

gas is consumed. The maximum amount of hydrates in the pipeline is 35 vol.%, approximately.

Hydrates start to dissociate in the riser near the platform.

The hydrate plugging risk is established by comparing the hydrate transition concentration

(Φtransition , the hydrate volume fraction transitioning to plugging) with the predicted hydrate

volume fraction (see Figure 6.22). For steady-state operations, if the maximum predicted hydrate

volume fraction is lower than Φtransition , the fluids can be produced with the presence of hydrates

in the line without risk of plugging. Figure 6.22 shows the distribution of hydrate volume fraction

along the pipeline with the corresponding Φtransition , calculated using the pipeline conditions. In

this case, the estimated hydrate volume fraction is greater than Φtransition , and the system is

135
340

320

Temperature (K)
Hydrate equilibrium temperature
300

280

260 Fluids temperature

240
0 5 10 15
Pipeline Length (km)

Figure 6.20: Temperature distribution along the pipeline length, showing system temperature and
hydrate equilibrium temperature corresponding to the steady-state flow, predicted by the water-
dominated hydrate formation model for the uninsulated case.

0.8
Volume Fraction

0.6
Water

0.4 Hydrate
Gas

0.2 Oil

0
0 5 10 15
Pipeline Length (km)

Figure 6.21: Fluid distribution in terms of volume fraction along the pipeline length, correspond-
ing to the steady-state flow, predicted by the water-dominated hydrate formation model for the
uninsulated case.

136
considered to have a high hydrate plugging risk.

The Φtransition correlation is a function of mixture Reynolds number and liquid loading. Since

the mixture Reynolds number is a function of fluid properties, it is also a function of temperature.

The decrease of Φtransition at the beginning of the flowline is related to the sharp decrease in

temperature observed in Figure 6.20. The continuing decrease in Φtransition is related with the

consumption of the gas phase, which reduces the mixture velocity. Φtransition reaches a constant

value when all the free gas phase is converted into hydrate.

0.8
Volume Fraction

0.6

0.4 Hydrate

0.2 Φtransition

0
0 5 10 15
Pipeline Length (km)

Figure 6.22: Hydrate volume fraction along the pipeline length, showing the calculated Φtransition
as plugging criterion, corresponding to the steady-state flow predicted by the water-dominated
hydrate formation model for the uninsulated case.

6.2.2 Thermal insulated case

In this case, a thermal insulation layer is included on the quasi-horizontal portion of the pipeline,

reducing the overall heat transfer coefficient of the system by 46%. The thermal insulation reduces

the rate of heat transfer from the fluids to the external environment, as observed in Figure 6.23,

where the system temperature reaches the hydrate equilibrium temperature in a longer distance

from the wellhead than in the uninsulated case. Hydrates start to form when the system reaches

a subcooling of 3.6 K, and the temperature remains constant during hydrate formation. Hydrates

start to dissociate in the riser, with the increase in ambient temperature, and the system temper-

ature follows the hydrate equilibrium temperature.

Figure 6.24 shows the predicted fluid distribution in terms of volume fraction along the pipeline

length. The volume fraction of gas and water decreases while the hydrate volume fraction increases.

137
340

320

Temperature (K)
Hydrate equilibrium temperature
300

280
Fluids temperature
260

240
0 5 10 15
Pipeline Length (km)

Figure 6.23: Temperature distribution along the pipeline length, showing system temperature and
hydrate equilibrium temperature corresponding to the steady-state flow, predicted by the water-
dominated hydrate formation model for the thermal insulated case.

Not all the free gas is consumed in this case, because of the lower heat transfer rate to the envi-

ronment, reducing the hydrate formation rate. The maximum amount of hydrates in the pipeline

is 24 vol.%, approximately.

0.8

Water
Volume Fraction

0.6

0.4 Gas
Hydrate
0.2
Oil

0
0 5 10 15
Pipeline Length (km)

Figure 6.24: Fluid distribution in terms of volume fraction along the pipeline length, corresponding
to the steady-state flow, predicted by the water-dominated hydrate formation model for the thermal
insulated case.

For the thermal insulated case, the hydrate volume fraction is lower than the calculated hydrate

transition concentration (see Figure 6.25). This case can be considered to have a lower hydrate

plugging risk than the uninsulated case.

138
1

0.8

Volume Fraction
0.6

Φtransition
0.4

0.2
Hydrate

0
0 5 10 15
Pipeline Length (km)

Figure 6.25: Hydrate volume fraction along the pipeline length, showing the calculated Φtransition
as plugging criterion, corresponding to the steady-state flow predicted by the water-dominated
hydrate formation model for the thermal insulated case.

6.3 Hydrate film growth in a gas-dominated horizontal pipeline

The hydrate formation model for gas-dominated systems was applied to a case with water-

saturated natural gas flowing through a horizontal pipeline of length equal to 5 km and inner

diameter of 5 inches. Figure 6.26 shows the estimated thickness of the hydrate deposit along the

pipeline length for three different times during hydrate formation, with a maximum hydrate deposit

thickness of 12 mm formed over 100 days. The increase in deposit thickness is associated with an

increase in pressure drop across the pipeline, as shown in Figure 6.27. In this case, the simulation

was set with a constant mass flow rate at the inlet and a constant outlet pressure. In order to

maintain the constant mass flow rate, it is required to increase the inlet pressure (see Figure 6.27).

Figure 6.28 shows the deposit thickness along the pipeline length at 100 days for different levels

of water saturation of the gas phase. The hydrate deposit thickness increases with increasing water

saturation of the gas phase.

6.4 Summary

This chapter presented examples of application of the three versions of CSMHyK developed in

this work for oil, water, and gas-dominated systems. The hydrate formation model for oil-dominated

systems was used to estimate the hydrate plugging risk of an offshore well with geometry and fluid

properties typical of those in the Caratinga field located in the Campos basin, Brazil, allowing

139
14

Thickness of gas hydrates


deposited at wall (mm)
12 100 days

10

8
60 days
6

4 30 days

0
0 1000 2000 3000 4000 5000
Pipeline length (m)

Figure 6.26: Plot of gas hydrate deposit thickness along the pipeline length at 30, 60 and 100 days,
showing an increase in deposit thickness with time, predicted by the hydrate formation model for
gas-dominated systems.

8.4
100 days
8.2
Pressure (MPa)

60 days
8

30 days
7.8

7.6

7.4
0 1000 2000 3000 4000 5000
Pipeline length (m)

Figure 6.27: Plot of system pressure along pipeline length, showing an increase in inlet pressure
with time required to maintain a constant mass flow rate at the inlet of the pipeline, predicted by
the hydrate formation model for gas-dominated systems.

140
25

Thickness of gas hydrates


deposited at wall (mm)
20 300 ppm

15
200 ppm
10

5 150 ppm

0
0 1000 2000 3000 4000 5000
Pipeline length (m)

Figure 6.28: Plot of hydrate deposit thickness along the pipeline length at 100 days, for different
levels of water saturation of the gas phase; 150, 200 and 300 ppm of water in the gas phase, predicted
by the hydrate formation model for gas-dominated systems.

the estimation of the hydrate plugging risk at three different periods of the well life. The hydrate

plugging risk was estimated from the calculation of three performance measures (pressure drop

along the flowline, hydrate volume fraction in pipe, and hydrate slurry relative viscosity) in an

attempt to quantify the plugging risk. The hydrate plugging risk depends on the amount of water

and gas flowing from the well; therefore, the hydrate plugging risk could increase with well life

because different periods in the well life are characterized by different magnitudes of GOR and

water cut.

Additionally, the hydrate formation model for oil-dominated systems was applied to study the

performance of ethanol injection, as a thermodynamic hydrate inhibitor, in steady-state flow and

transient shut-in/restart operations. In general, ethanol injection decreased the hydrate plugging

risk in steady-state and transient operations. The application of the oil-dominated hydrate forma-

tion model proved to be useful in determining the optimal ethanol concentration that minimized

the hydrate plugging risk.

The hydrate formation model for water-dominated systems was applied to the same well/flow-

line/riser geometry from the Caratinga field with and without thermal insulation, and with a high

water cut of 85 vol.%. The hydrate plugging risk of the water-dominated cases was estimated using

a correlation to estimate the hydrate transition concentration as function of Reynolds number and

liquid loading. The hydrate transition concentration represents the hydrate concentration at which

141
an irreversible flow pattern transition occurs that eventually leads to the hydrate plug formation.

The hydrate formation model for water-dominated systems allows one to estimate the thermal

insulation needed to reduce the hydrate plugging risk in offshore production facilities.

The model for gas-dominated systems was specially designed for cases with a water-saturated

gas phase flowing through a pipe without free liquid phases. A case with water-saturated natural

gas flowing through a horizontal pipeline was studied. The hydrate deposit thickness increases

with time, which causes an increase in the pressure drop across the pipeline. An increase in water

saturation of the gas phase increases the thickness of the hydrate deposit. These results provide

the basis and motivation for future experimental studies of gas-dominated systems for validation

of this first pass model, which represents the first model to address gas-dominated systems.

142
CHAPTER 7

SUMMARY AND CONCLUSIONS

This dissertation presents a compilation of the latest knowledge in the areas of flow assurance

and gas hydrates. Flow assurance is an applied science that focuses on the design and optimization

of facilities and procedures to maintain non-interrupted flow in transport facilities of the oil and

gas industry, in wells and pipelines to the sales point. Gas hydrates are crystalline compounds (i.e.,

solid phase), that form by combining small gas molecules with liquid water at high pressures and

low temperatures. The rapid formation of gas hydrates is a challenging problem in flow assurance,

particularly in subsea transport facilities of the offshore oil and gas industry, where high pressure

and low temperature are normal operating conditions.

The prediction of hydrate formation conditions using thermodynamic calculations (with ex-

cellent accuracy) is not sufficient to estimate hydrate plugging risk in transport facilities. The

objective of this research was to provide a solution to estimate hydrate plugging risk, by extending

and improving a first-pass hydrate formation model developed at the Center for Hydrate Research

of the Colorado School of Mines, to predict gas hydrate formation, dissociation and transportability

in oil and gas flowlines. The resulting prediction tool may enhance the design process of production

facilities and flow assurance procedures, through the estimation of hydrate plugging risk in specific

scenarios. Through this thesis work, the initial hydrate model has been transformed from a kinetic

model for oil-dominated systems only, to a comprehensive modeling approach which can be applied

to oil-, water- and gas-dominated systems.

The main conclusions that can be drawn from the research conducted for this thesis are:

1. Improved flow assurance predictive models for gas hydrates can be constructed by including

interphase mass transfer, interphase momentum transfer, reaction kinetic heat generation,

and modified phase properties (e.g., effective viscosity, corrected heat transfer coefficients),

to the governing equations of one-dimensional transient multiphase flow in pipelines.

2. Fluid-fluid and fluid-solid interfaces and the bulk flow of multiple phases in pipelines are key

parameters that must be considered in developing the gas hydrate models.

143
3. Different hydrate formation mechanisms between oil and water-dominated systems were iden-

tified from experiments in large scale flow loop and other laboratory scale apparatuses (e.g.,

micromechanical force apparatus).

4. The hydrate formation model for oil-dominated systems was improved by the implementation

of a new dynamic cohesion force model. The dynamic cohesion force model is based on particle

capillary attraction due to a liquid bridge between particles and sintering of the liquid bridge

(i.e., liquid bridge converting into hydrate). The model predicts well the hydrate volume

fraction measured in flow loop experiments.

5. The CSMHyK model was extended to include water-dominated systems by implementation

of a mass transfer-based hydrate growth model and hydrate plugging criterion, based on fluid

velocity and liquid holdup. The model predicts the amount of hydrate formed, while the

onset of hydrate plug formation can be predicted using the correlation of hydrate transition

concentration (hydrate concentration at which an irreversible flow pattern transition occurs

that eventually leads to the hydrate plug formation). The model predictions matched hydrate

formation rates from experimental data from gas-water flow loop experiments, and the model

predicted fairly well the amount of hydrates observed in high water cut (75% and 90% water

cut) experiments with crude oil. The model predicted the experimentally observed hydrate

transition concentration with a maximum error of 12% (average error of 4.5%).

6. The model for hydrate formation in gas-dominated systems was specially designed for cases

with a water-saturated gas phase flowing through a pipe without free liquid phases. The gas-

dominated model consists of a combined heat and mass transfer model to predict hydrate film

growth on pipe walls. The gas-dominated hydrate formation model has not been compared

against experimental data, and should be considered as a first pass model for hydrate film

growth in gas-dominated systems.

7. In oil-dominated systems, the hydrate-plugging risk increases with the amount of water and

gas produced from the well; therefore, the hydrate-plugging risk could increase with well

life, which is characterized by increasing GOR and water cut. Prediction of the amount of

hydrates formed and their transportability through pipelines results in an estimate of hydrate

144
plugging risk.

8. In gas-dominated systems, the hydrate deposit thickness increases with time, which causes

an increase in the pressure drop across the pipeline. An increase in water saturation of the

gas phase increases the thickness of the hydrate deposit.

9. The gas hydrate models developed in this thesis can be used in the design of flow assurance

strategies.

145
CHAPTER 8

SUGGESTIONS FOR FUTURE RESEARCH

In this work, a gas hydrate model originally designed to predict hydrate formation in oil-

dominated systems was improved and extended to water and gas-dominated systems. The models

developed in this work represent a significant advancement for the prediction of hydrate plugging

risk in the pipelines of oil and gas transport facilities, and can be used as a tool to design flow

assurance strategies. However, the accuracy of these models in representing hydrate formation

mechanisms in multiphase flow can be improved with further research.

The hydrate formation model for oil-dominated systems is based on the kinetics for hydrate

structure I, where the hydrate kinetic rate constants were regressed from experimental data of

methane hydrates. The kinetics for structure I may not be applicable in pipeline modeling, where

hydrate structure II is typically expected. Further investigations are needed to determine how the

kinetics for hydrate structure I affect the accuracy of the predictions, and to include the kinetics

for the hydrate structure II formed with heavier components (e.g., propane, iso-butane) which are

more applicable for typical pipeline plugging predictions.

It is suggested to continue the development of CSMHyK models as user-defined plug-in programs

that can be coupled to the OLGA


R
transient multiphase flow simulator. The flexible flow model

framework of OLGA
R
allows defining an additional gas hydrate phase and calculating additional

contributions to the conservation equations. Improved pressure drop predictions could be obtained

through modifications of contributions to momentum and pressure conservation equations. Models

for phenomena like bedding or deposition of hydrate particles, which were not considered in this

work, could improve pressure drop predictions. Alternatively, the development of improved hydrate

bedding/agglomeration/deposition models can be tested in simplified multiphase flow models, such

as the simple hydrodynamic slug flow model presented in Section 3.4.2.

An attempt to quantify the hydrate plugging risk of a subsea pipeline with an oil-dominated

system was presented in Chapter 6. Three performance measures (pressure drop, hydrate volume

fraction and hydrate slurry relative viscosity) were calculated to determine the hydrate plugging

risk. Depending on the values of these parameters the hydrate plugging risk was classified in low,

146
intermediate, and high plugging risk. The limitation of this approach is that the different risk levels

are specific to a particular system geometry and fluid properties. A more general quantification

parameter for hydrate plugging risk is desirable, which should be independent of the characteristics

of a given system. The quantification parameter could be developed using a dimensionless number

that combines parameters such as pressure drop, hydrate volume fraction, slurry relative viscosity,

and flow velocity, which can be then applied to different scenarios to quantify hydrate plugging

risk.

The gas and oil industry is currently in a transition from using hydrate avoidance methods to risk

management methods. The use of inhibitors that act at the interfaces of hydrate particles, such as

kinetic hydrate inhibitors (KHI) and anti-agglomerant (AA) surfactants, have been proven to reduce

the required concentrations of chemical additives in a more economical and efficient flow assurance

strategy. The effect of AA surfactants may be considered through the dynamic cohesion force

model by a decrease in water-oil interfacial tension. However, this has not been compared against

experimental data. On the other hand, the purpose of KHIs is to delay the hydrate nucleation.

Currently, the stochastic nature of hydrate nucleation, even with the injection of kinetic inhibitors,

has made its implementation into the hydrate formation model difficult. Introducing the capabilities

of modeling the effect of AAs and KHIs on the hydrate formation models could be a subject of

future research.

The current work has put more emphasis on the hydrate formation models for oil, water and

gas-dominated systems, and has overlooked the importance of dissociation models. Hydrate dis-

sociation represents an important phenomenon that is present in every subsea transport facility,

since hydrates will dissociate when flowing in the vertical riser toward the platform. Also, after

the formation of a hydrate plug, the hydrate dissociation mechanism forms an important part of

the remediation strategy. Most of the flow loop experiments should have dissociation data after

the hydrate formation stage. Analysis of the dissociation data could result in the suggestion of

mechanisms that can be translated into mathematical models. The models for hydrate dissociation

can be easily coupled with the hydrate formation models in the same computational program.

The transition from oil to water-dominated systems needs to be studied in more detail. Based

on the flow loop experiments performed in this work, the case with 75% water cut represents the

transition between oil and water-dominated systems for the values of water cut tested. Additional

147
flow loop experiments are needed that include several water cut values between 60 and 90%, which

could help in defining hydrate formation mechanisms during the transition from oil to water-

dominated systems. The transition from oil to water-dominated systems could be correlated with

laboratory scale experiments (e.g., bottle tests, high pressure autoclave tests). The dependence of

this transition to specific crude oils could be studied as well.

Hydrate deposition on the pipe wall was considered in the hydrate formation model for gas-

dominated systems, which considered a gas phase saturated with water flowing in the pipeline (i.e.,

single phase flow). However hydrate deposition on the pipe wall in multiphase flow has not been

considered in this study. In the models developed in this work, it is assumed that hydrate particles

form in the liquid phases and can either agglomerate (oil-dominated systems) or promote a flow

regime transition (water-dominated systems). But deposition of hydrate particles on a pipe wall

surface was not accounted for in the oil and water-dominated models. Hydrate particles deposition

on a pipe wall is possible through adhesion forces between the hydrate particle and pipe surface,

as showed by Aspenes et al. (2010a) from measurement of micromechanical adhesion forces.

148
REFERENCES CITED

Alboudwarej, H., Huo, Z., Kempton, E.C., 2006. Flow-assurance aspects of subsea systems design
for production of waxy crude oils, in: SPE Annual Technical Conference and Exhibition, Society
of Petroleum Engineers, San Antonio, Texas, USA, 24–27 Sep. Doi: 10.2118/103242-MS.

Aman, Z.M., Brown, E.P., Sloan, E.D., Sum, A.K., Koh, C.A., 2011. Interfacial mechanisms gov-
erning cyclopentane clathrate hydrate adhesion/cohesion. Physical Chemistry Chemical Physics,
13(44):19796–19806. Doi: 10.1039/C1CP21907C.

Aman, Z.M., Dieker, L.E., Aspenes, G., Sum, A.K., Sloan, E.D., Koh, C.A., 2010. Influence of
model oil with surfactants and amphiphilic polymers on cyclopentane hydrate adhesion forces.
Energy & Fuels, 24(10):5441–5445. Doi: 10.1021/ef100762r.

Aman, Z.M., Sloan, E.D., Sum, A.K., Koh, C.A., 2012. Lowering of clathrate hydrate cohesive forces
by surface active carboxylic acids. Energy & Fuels, 26(8):5102–5108. Doi: 10.1021/ef300707u.
http://pubs.acs.org/doi/pdf/10.1021/ef300707u.

Andreussi, P., Bendiksen, K., 1989. An investigation of void fraction in liquid slugs for horizontal
and inclined gas–liquid pipe flow. International Journal of Multiphase Flow, 15(6):937 – 946.
Doi: 10.1016/0301-9322(89)90022-0.

Anklam, M.R., York, J.D., Helmerich, L., Firoozabadi, A., 2008. Effects of antiagglomerants on the
interactions between hydrate particles. AIChE Journal, 54(2):565–574. Doi: 10.1002/aic.11378.

Asheim, H., 1986. Mona, an accurate two-phase well flow model based on phase slippage. SPE
Production Engineering, 1(3):221–230. Doi: 10.2118/12989-PA.

Aspenes, G., Dieker, L.E., Aman, Z.M., Høiland, S., Sum, A.K., Koh, C.A., Sloan, E.D., 2010a.
Adhesion force between cyclopentane hydrates and solid surface materials. Journal of Colloid
and Interface Science, 343(2):529–536. Doi: 10.1016/j.jcis.2009.11.071.

Aspenes, G., Høiland, S., Borgund, A.E., Barth, T., 2010b. Wettability of petroleum pipelines:
Influence of crude oil and pipeline material in relation to hydrate deposition. Energy & Fuels, 24
(1):483–491. Doi: 10.1021/ef900809r.

Asserson, R.B., Hoffmann, A.C., Høiland, S., Asvik, K.M., 2009. Interfacial tension measurement
of freon hydrates by droplet deposition and contact angle measurements. Journal of Petroleum
Science and Engineering, 68(3-4):209–217. Doi: 10.1016/j.petrol.2009.06.012.

Austvik, T., 1992. Hydrate formation and behavior in pipes. Ph.D. thesis. Norwegian University of
Science and Technology. Trondheim, Norway.

149
Austvik, T., Li, X., Gjertsen, L.H., 2000. Hydrate plug properties: Formation and removal of
plugs. Annals of the New York Academy of Sciences, 912(1):294–303. Doi: 10.1111/j.1749-
6632.2000.tb06783.x.

Ayers, R.R., 2003. Thermally managed flowline engineering. Technical Report - DeepStar CTR
5307. DeepStar.

Aziz, K., Govier, G.W., Fogarasi, M., 1972. Pressure drop in wells producing oil and gas. Journal
of Canadian Petroleum Technology, 11(3). Doi: 10.2118/72-03-04.

Bai, Y., Bai, Q., 2005. Subsea systems engineering, in: Subsea Pipelines and Risers - Part III:
Flow Assurance. Elsevier. chapter 17, pp. 263 – 276.

Balasubramaniam, R., Rame, E., Kizito, J., Kassemi, M., 2006. Two Phase Flow Modeling: Sum-
mary of FlowRegimes and Pressure Drop Correlations inReduced and Partial Gravity. Tech-
nical Report NASA/CR2006-214085. NASA - Glenn Research Center. Cleveland, Ohio, USA.
(http://hdl.handle.net/2060/20060008906, accessed: 10/19/2012).

Ballard, A.L., Sloan, E.D., 2004a. The next generation of hydrate prediction: Part III. gibbs energy
minimization formalism. Fluid Phase Equilibria, 218(1):15–31. Doi: 10.1016/j.fluid.2003.08.005.

Ballard, L., Sloan, E.D., 2004b. The next generation of hydrate prediction IV: A compari-
son of available hydrate prediction programs. Fluid Phase Equilibria, 216(2):257–270. Doi:
10.1016/j.fluid.2003.11.004.

Bancroft, W.D., 1912. The theory of emulsification, i. The Journal of Physical Chemistry, 16(3):
177–233. Doi: 10.1021/j150129a001.

Baxendell, P.B., Thomas, R., 1961. The calculation of pressure gradients in high-rate flowing wells.
Journal of Petroleum Technology, 13(10):1023–1028. Doi: 10.2118/2-PA.

Becher, P., 2001. Emulsion stability: simple theories, in: Emulsions: Theory and Practice. Oxford
University Press, New York. chapter 4, Third edition. pp. 119–148.

Beggs, D.H., Brill, J.P., 1973. A study of two-phase flow in inclined pipes. Journal of Petroleum
Technology, 25(5):607–617. Doi: 10.2118/4007-PA.

Bendiksen, K.H., Maines, D., Moe, R., Nuland, S., 1991. The dynamic two-fluid model OLGA:
Theory and application. SPE Production Engineering, 6(2):171–180. Doi: 10.2118/19451-PA.

Biberg, D., 1999. An explicit approximation for the wetted angle in two-phase stratified pipe flow.
Canadian Journal of Chemical Engineering, 77(6):1221–1224. Doi:10.1002/cjce.5450770619.

Bienfait, M., 1992. Roughening and surface melting transitions: consequences on crystal growth.
Surface Science, 272(1-3):1–9. Doi: 10.1016/0039-6028(92)91415-8.

150
Blasius, H., 1908. Grenzschichten in flssigkeiten mit kleiner reibung (boundary layersin fluidswith-
littlefriction). Z. Math. Phys., 56.

Borgund, A.E., Høiland, S., Barth, T., Fotland, P., Askvik, K.M., 2009. Molecular analysis of
petroleum derived compounds that adsorb onto gas hydrate surfaces. Applied Geochemistry, 24
(5):777–786. Doi: 10.1016/j.apgeochem.2009.01.004.

Bourrel, M., Salager, J.L., Schechter, R.S., Wade, W.H., 1980. A correlation for phase behavior of
nonionic surfactants. Journal of Colloid and Interface Science, 75(2):451–461. Doi: 10.1016/0021-
9797(80)90470-1.

Boxall, J., 2009. Hydrate plug formation from less than 50% water content water-in-oil emulsions.
Ph.D. thesis. Colorado School of Mines. Golden, Colorado.

Boxall, J., Davies, S., Koh, C.A., Sloan, E.D., 2009. Predicting when and where hydrate plugs
form in oil-dominated flowlines. SPE Projects, Facilities & Construction, 4(3):80–86. Doi:
10.2118/129538-PA.

Brennen, C.E., 2005. Fundamentals of multiphase flow. Cambridge University Press, New York,
NY.

Brill, J.P., Mukherjee, H., 1999. Multiphase Flow in Wells. SPE Monograph Series, Society of
Petroleum Engineers, Richardson, Texas.

Camargo, R., Palermo, T., 2002. Rheological properties of hydrate suspensions in an asphaltenic
crude oil, in: Mori, Y.H. (Ed.), 4th International Conference on Gas Hydrates, ICGH, Yokohama,
Japan, 19-23 May. pp. 880–885.

Cazarez-Candia, O., Vasquez-Cruz, M.A., 2005. Prediction of pressure, temperature, and velocity
distribution of two-phase flow in oil wells. Journal of Petroleum Science and Engineering, 46(3):
195–208. Doi: 10.1016/j.petrol.2004.11.003.

Chierici, G.L., Ciucci, G.M., Sclocchi, G., 1974. Two-phase vertical flow in oil wells - prediction of
pressure drop. Journal of Petroleum Technology, 26(8):927–938. Doi: 10.2118/4316-PA.

Clausse, M., Peyrelasse, J., Heil, J., Boned, C., Lagourette, B., 1981. Bicontinuous structure zones
in microemulsions. Nature, 293(5834):636–638. Doi: 10.1038/293636a0.

Coelho, R.R., Hovell, I., Rajagopal, K., 2012. Elucidation of the functional sulphur chemical struc-
ture in asphaltenes using first principles and deconvolution of mid-infrared vibrational spectra.
Fuel Processing Technology, 97:85–92. Doi: 10.1016/j.fuproc.2011.12.041.

Colebrook, C.F., 1939. Turbulent flow in pipes, with particular reference to the transition region
between the smooth and rough pipe laws. J. Inst. Civil Engrs, 11(4):133 – 156. Doi: 10.1680/i-
joti.1939.13150.

151
Creek, J.L., Subramanian, S., Estanga, D., 2011. New method for managing hydrates in deepwater
tiebacks, in: 2011 Offshore Technology Conference, Offshore Technology Conference, 2-5 May.
Doi: 10.4043/22017-MS.

Danielson, T.J., 2011. A simple model for hydrodynamic slug flow, in: 2011 Offshore Tech-
nology Conference, Offshore Technology Conference, Houston, Texas, USA, 2-5 May. Doi:
10.4043/21255-MS.

Davies, S., 2009. The role of transport resistances in the formation and remediation of hydrate
plugs. Ph.D. thesis. Colorado School of Mines. Golden, Colorado, USA.

Davis, L., 2011. Offshore tide on the rise, in: E&P Magazine. Hart Energy, Houston, Texas, USA.
(http://www.epmag.com/item/Offshore-Tide-The-Rise 92764, accessed: 09/17/2012).

de Boer, R.B., Leerlooyer, K., Eigner, M.R.P., van Bergen, A.R.D., 1995. Screening of crude oils
for asphalt precipitation: Theory, practice, and the selection of inhibitors. SPE Production &
Operations, 10(1):55–61. Doi: 10.2118/24987-PA.

Dickakian, G., Seay, S., 1988. Asphaltene precipitation primary crude exchanger fouling mechanism.
Oil & Gas Journal, 86(10):47–50.

Dieker, L.E., Aman, Z.M., George, N.C., Sum, A.K., Sloan, E.D., Koh, C.A., 2009. Microme-
chanical adhesion force measurements between hydrate particles in hydrocarbon oils and their
modifications. Energy & Fuels, 23(12):5966–5971. Doi: 10.1021/ef9006615.

Dieker, L.E., Taylor, C.J., Koh, C.A., Sloan, E.D., 2008. Micromechanical adhesion force measure-
ments between cyclopentane hydrate particles, in: 6th International Conference of Gas Hydrates,
ICGH, Vancouver, British Columbia, Canada, 6-10 Jul.

Döppenschmidt, A., Butt, H.J., 2000. Measuring the thickness of the liquid-like layer on ice surfaces
with atomic force microscopy. Langmuir, 16(16):6709–6714. Doi: 10.1021/la990799w.

Dorstewitz, F., Mewes, D., 1994. The influence of heat transfer on the formation of hydrate layers in
pipes. Int. J. Heat and Mass Transfer, 37(14):2131–2137. Doi: 10.1016/0017-9310(94)90314-X.

Drew, T. B., K.E.C., McAdams, W.H., 1932. The friction factor for clean round pipes. Trans.
AIChE, 28(1932):56–72.

Duns, H., Ros, N.C.J., 1963. Vertical flow of gas and liquid mixtures in wells, in: 6th World
Petroleum Congress, World Petroleum Congress, Frankfurt, Germany, 19-26 Jun.

Durbut, P., 1999. Surface activity. Marcel Dekker, New York. chapter 3. Surfactant Science, pp.
47–97. Doi: 10.1201/b10985-4.

Englezos, P., Kalogerakis, N., Dholabhai, P.D., Bishnoi, P.R., 1987. Kinetics of gas hydrate for-
mation from mixtures of methane and ethane. Chemical Engineering Science, 42(11):2659–2666.
Doi: 10.1016/0009-2509(87)87016-1.

152
Esumi, K., 2006. Adsolubilization of organic pollutants, in: Somasundaran, P. (Ed.), Encyclopedia
of Surface and Colloid Science: Absorption - Adsirotuib. Taylor & Francis Group. volume 1, 2nd
edition. pp. 124–127.

Fancher, G.H., Brown, K.E., 1963. Prediction of pressure gradients for multiphase flow in tubing.
SPE Journal, 3(1):59–69. Doi: 10.2118/440-PA.

Fidel-Dufour, A., Gruy, F., Herri, J.M., 2006. Rheology of methane hydrate slurries during their
crystallization in a water in dodecane emulsion under flowing. Chemical Engineering Science, 61
(2):505–515. Doi: 10.1016/j.ces.2005.07.001.

Gao, S., 2009. Hydrate risk management at high watercuts with anti-agglomerant hydrate in-
hibitors. Energy & Fuels, 23(4):2118–2121. Doi: 10.1021/ef8009876.

Gomez, L.E., Shoham, O., Taitel, Y., 2000. Prediction of slug liquid holdup: horizontal to upward
vertical flow. International Journal of Multiphase Flow, 26(3):517–521. Doi: 10.1016/S0301-
9322(99)00025-7.

Greaves, D., Boxall, J., Mulligan, J., Sloan, E.D., Koh, C.A., 2008. Hydrate formation from
high water content-crude oil emulsions. Chemical Engineering Science, 63(18):4570–4579. Doi:
10.1016/j.ces.2008.06.025.

Gregory, G.A., Nicholson, M.K., Aziz, K., 1978. Correlation of the liquid volume fraction in the
slug for horizontal gas-liquid slug flow. International Journal of Multiphase Flow, 4(1):33–39.
Doi: 10.1016/0301-9322(78)90023-X.

Griffin, W.C., 1949. Classification of surface active agents by hlb. J. Soc. Cosmet. Chem., 1(5):
311–326.

Gudmundsson, J.S., 2002. Cold flow hydrate technology, in: Mori, Y.H. (Ed.), 4th International
Conference on Gas Hydrates, ICGH, Yokohama, Japan, 19-23 May. pp. 912–916.

Guo, B., Song, S., Chacko, J., Ghalambor, A., 2005. Chapter 15 - flow assurance, in: Offshore
pipelines. Gulf Professional Publishing, Burlington. chapter 15, pp. 169–214. Doi: 10.1016/B978-
075067847-6/50072-X.

Hagedorn, A.R., Brown, K.E., 1965. Experimental study of pressure gradients occurring during
continuous two-phase flow in small-diameter vertical conduits. Journal of Petroleum Technology,
17(4):475–484. Doi: 10.2118/940-PA.

Hagesaether, L., Lunde, K., Nygaard, F., Eidsmoen, H., 2007. Flow-assurance modeling: Reality
check and aspects of transient operations of gas/condensate pipelines. SPE Projects, Facilities
& Construction, 2(1):1–17. Doi: 10.2118/108834-PA.

Hammami, A., Ratulowski, J., 2007. Precipitation and deposition of asphaltenes in production
systems: A flow assurance overview, in: Mullins, O.C., Sheu, E.Y., Hammami, A., Marshall,
A.G. (Eds.), Asphaltenes, Heavy Oils, and Petroleomics. Springer, New York. chapter 23, pp.
617–660. Doi: 978-0-387-68903-6.

153
Hammerschmidt, E.G., 1934. Formation of gas hydrates in natural gas transmission lines. Industrial
& Engineering Chemistry, 26(8):851–855. Doi: 10.1021/ie50296a010.

Hanratty, T.J., 1991. Separated flow modelling and interfacial transport phenomena. Applied
Scientific Research, 48(3):353–390. Doi: 10.1007/BF02008206.

Hatton, G.J., Kruka, V.R., 2002. Hydrate Blockage Formation - Analysis of Werner Bolley Field
Test Data. Technical Report. DeepStar.

Hesketh, R.P., Fraser Russell, T.W., Etchells, A.W., 1987. Bubble size in horizontal pipelines.
AIChE Journal, 33(4):663–667. Doi: 10.1002/aic.690330414.

Hinze, J.O., 1955. Fundamentals of the hydrodynamic mechanism of splitting in dispersion pro-
cesses. AIChE Journal, 1(3):289–295. Doi: 10.1002/aic.690010303.

Hoekstra, L.L., Vreeker, R., Agterof, W.G.M., 1992. Aggregation of colloidal nickel hydroxycar-
bonate studied by light-scattering. Journal of Colloid and Interface Science, 151(1):17–25. Doi:
10.1016/0021-9797(92)90234-d.

Høiland, S., Askvik, K.M., Fotland, P., Alagic, E., Barth, T., Fadnes, F., 2005. Wettability of freon
hydrates in crude oil/brine emulsions. Journal of Colloid and Interface Science, 287(1):217–225.
Doi: 10.1016/j.jcis.2005.01.080.

Israelachvili, J., 1991. Intermolecular and surface forces. Academic Press Inc., San Diego, CA,
USA. 2nd edition.

Joshi, S., 2012. Experimental investigation and modeling of gas hydrate formation in high water
cut producing oil pipelines. Ph.D. thesis. Colorado School of Mines. Golden, Colorado, USA.

Joshi, S., Rao, I., Zerpa, L.E., Webb, E., Lafond, P., Sloan, E.D., Sum, A.K., Koh, C.A., 2011a.
Understanding hydrate plug formation from high water cut systems with a four inch flowloop,
in: 7th International Conference of Gas Hydrates, ICGH, Edinburgh, Scotland, 17-21 Jula.

Joshi, S., Sloan, E.D., Koh, C.A., Sum, A.K., 2011b. Micromechanical adhesion force measure-
ments between cyclopentane hydrate particles in water, in: 7th International Conference of Gas
Hydrates, ICGH, Edinburgh, Scotland, 17-21 Julb.

Kaczmarski, A.A., Lorimer, S.E., 2001. Emergence of flow assurance as a technical discipline
specific to deepwater: Technical challenges and integration into subsea systems engineering, in:
2001 Offshore Technology Conference, Offshore Technology Conference, Houston, Texas, USA,
30 Apr-3 May. Doi: 10.4043/13123-MS.

Kalogerakis, N., Jamaluddin, A.K.M., Dholabhai, P.D., Bishnoi, P.R., 1993. Effect of surfactants
on hydrate formation kinetics, in: SPE International Symposium on Oilfield Chemistry, Society
of Petroleum Engineers, New Orleans, Louisiana, 2-5 Mar. Doi:10.2118/25188-ms.

154
Kanicky, J.R., Lopez-Montilla, J.C., Pandey, S., Shah, O.D., 2001. Surface chemistry in the
petroleum industry, in: Holmberg, K. (Ed.), Handbook of Applied Surface and Colloid Chem-
istry. John Wiley & Sons. chapter 11, pp. 251–267.

Kitanovski, A., Poredos, A., 2002. Concentration distribution and viscosity of ice-slurry in het-
erogeneous flow. International Journal of Refrigeration, 25(6):827–835. Doi: 10.1016/S0140-
7007(01)00091-3.

Kudsen, J.G., Hottel, H.C., Sarofim, A.F., Wankat, P.C., Knaebel, K.S., 1999. Heat and mass trans-
fer, in: Perry, R.H., Green, D.W., Maloney, J.O. (Eds.), Perry’s Chemical Engineers’ Handbook.
McGraw-Hill, 7th edition. pp. 5–51.

Lachance, J.W., Sloan, E.D., Koh, C.A., 2008. Effect of hydrate formation/dissociation on emulsion
stability using dsc and visual techniques. Chemical Engineering Science, 63(15):3942–3947. Doi:
10.1016/j.ces.2008.04.049.

Lee, J., Shin, C., Lee, Y., 2010. Experimental investigation to improve the storage poten-
tials of gas hydrate under the unstirring condition. Energy & Fuels, 24(2):1129–1134. Doi:
10.1021/ef901020g.

Lee, Y.S., 2008. Reverse micelles, in: Self-assembly and nanotechnology - A force balance approach.
John Wiley & Sons. chapter 4, pp. 95–96.

Leporcher, E.M., Peytavy, J.L., Mollier, Y., Sjöblom, J., Labes-Carrier, C., 1998. Multi-
phase transportation: Hydrate plugging prevention through crude oil natural surfactants, in:
SPE Annual Technical Conference and Exhibition, Society of Petroleum Engineers, 27-30 Sep.
Doi:10.2118/49172-ms.

Levich, V.G., 1962. Physicochemical hydrodynamics. Prentice-Hall, Englewood Cliffs, N.J., USA.

Lingelem, M.N., Majeed, A.I., Stange, E., 1994. Industrial experience in evaluation of hydrate
formation, inhibition, and dissociation in pipeline design and operation. Annals of the New York
Academy of Sciences, 715(1):75–93. Doi: 10.1111/j.1749-6632.1994.tb38825.x.

Lockhart, R.W., Martinelli, R.C., 1949. Proposed correlation of data for isothermal two-phase,
two-component flow in pipes. Chemical Engineering Progress, 45(1):3943.

Lund, A., Larsen, R., 2000. Conversion of water to hydrate particles theory and application, in:
14th Symposium on Thermophysical Properties, Elsevier, Boulder, Colorado, .

Luo, Y.T., Zhu, J.H., Fan, S.S., Chen, G.J., 2007. Study on the kinetics of hydrate formation in a
bubble column. Chemical Engineering Science, 62(4):1000–1009. Doi: 10.1016/j.ces.2006.11.004.

McCain, W.D., 1990. Components of naturally occurring petroleum fluids, in: The properties of
petroleum fluids. PennWell. chapter 1, 2nd edition edition. p. 40.

155
McLean, J.D., Kilpatrick, P.K., 1997. Effects of asphaltene aggregation in model heptane-toluene
mixtures on stability of water-in-oil emulsions. Journal of Colloid and Interface Science, 196(1):
23–34. Doi:10.1006/jcis.1997.5177.

McLean, J.D., Spiecker, P.M., 1998. The role of petroleum asphaltenes in the stabilization of water-
in-oil emulsions, in: Mullins, O.C., Sheu, E.Y. (Eds.), Structure and Dynamics of Asphaltenes.
Plenum, New York. chapter 12, pp. 377–422.

Mills, P., 1985. Non-newtonian behavior of flocculated suspensions. Journal de Physique Lettres,
46(7):301–309. Doi: 10.1051/jphyslet:01985004607030100.

Mitchell, D.L., Speight, J.G., 1973. The solubility of asphaltenes in hydrocarbon solvents. Fuel, 52
(2):149–152. Doi: 10.1016/0016-2361(73)90040-9.

Mokhatab, S., Towler, B., 2009. Wax prevention and remediation in subsea pipelines and flowlines.
World Oil, 230(11).

Mukherjee, H., Brill, J., 1983. Liquid holdup correlations for inclined two-phase flow. Journal of
Petroleum Technology, 35(5):1003–1008. Doi: 10.2118/10923-PA.

Nikuradse, J., 1933. Laws of flow in rough pipes. VDI Forschungsheft.

Orkiszewski, J., 1967. Predicting two-phase pressure drops in vertical pipe. Journal of Petroleum
Technology, 19(6):829–838. Doi: 10.2118/1546-PA.

Poettman, F.H., Carpenter, P.G., 1952. The multiphase flow of gas, oil, and water through vertical
flow strings with application to the design of gas-lift installations, in: Drilling and Production
Practice, American Petroleum Institute, Wichita, Kansas, USA, Mar.

Preston, W.C., 1948. Some correlating principles of detergent action. Journal of Physical Chemistry,
52(1):84–97. Doi: 10.1021/j150457a010.

Rabinovich, Y.I., Esayanur, M.S., Moudgil, B.M., 2005. Capillary forces between two spheres
with a fixed volume liquid bridge: theory and experiment. Langmuir, 21(24):10992–10997. Doi:
10.1021/la0517639.

Rao, I., Sloan, E.D., Koh, C.A., Sum, A.K., 2011. Laboratory experiments and modeling for hydrate
formation and deposition from water saturated gas systems, in: 7th International Conference of
Gas Hydrates, ICGH, Edinburgh, Scotland, 17-21 Jul.

Ronalds, B.F., 2005. Applicability ranges for offshore oil and gas production facilities. Marine
Structures, 18(3):251–263. Doi: 10.1016/j.marstruc.2005.06.001.

Ruf, A., Worlitschek, J., Mazzotti, M., 2000. Modeling and experimental analysis of PSD mea-
surements through FBRM. Particle & Particle Systems Characterization, 17(4):167–179. Doi:
10.1002/1521-4117(200012)17:4<167::AID-PPSC167>3.0.CO;2-T.

156
Salager, J.L., 1996. Guidelines for the formulation, composition and stirring to attain desiredemul-
sion properties, in: Chattopadhyay, A.K., Mittal, K.L. (Eds.), Surfactants in solution. Marcel
Dekker, New York. Surfactant Science. chapter 16, pp. 261–295.

Salager, J.L., 1999. Microemulsions, in: Broze, G. (Ed.), Handbook of detergents - Part A: Prop-
erties. Marcel Dekker, New York. Surfactant Science. chapter 8, pp. 253–302.

Salager, J.L., 2000a. Emulsion properties and related know-how to attain them, in: Nielloud, F.,
Marti-Mestres, G. (Eds.), Pharmaceutical emulsions and suspensions. Marcel Dekker, New York.
Drugs and Pharmaceutical Science. chapter 3, pp. 73–125.

Salager, J.L., 2000b. Formulation concepts for the emulsion maker, in: Nielloud, F., Marti-Mestres,
G. (Eds.), Pharmaceutical emulsions and suspensions. Marcel Dekker, New York. Drugs and
Pharmaceutical Science. chapter 2, pp. 19–72.

Salager, J.L., Anton, R., Sabatini, D., Harwell, J., Acosta, E., Tolosa, L., 2005. Enhancing sol-
ubilization in microemulsionsstate of the art and current trends. Journal of Surfactants and
Detergents, 8(1):3–21. Doi: 10.1007/s11743-005-0328-4.

Salager, J.L., Loaiza-Maldonado, I., Minana-Perez, M., Silva, F., 1982. Surfactant-oil-water sys-
tems near the affinity inversion part i: Relationship between equilibrium phase behavior and
emulsion type and stability. Journal of Dispersion Science and Technology, 3(3):279 – 292. Doi:
10.1080/01932698208943642.

Salager, J.L., Minana-Perez, M., Perez-Sanchez, M., Ramfrez-Gouveia, M., Rojas, C.I., 1983.
Surfactant-oil-water systems near the affinity inversion part III: The two kinds of emul-
sion inversion. Journal of Dispersion Science and Technology, 4(3):313 – 329. Doi:
10.1080/01932698308943373.

Salager, J.L., Morgan, J.C., Schechter, R.S., Wade, W.H., Vasquez, E., 1979. Optimum formulation
of surfactant/water/oil systems for minimum interfacial tension or phase behavior. Society of
Petroleum Engineers Journal, 19(2):107–115. Doi: 10.2118/7054-PA.

Sandrea, I., Sandrea, R., 2007. Global offshore oil: Exploration trends show continued promise in
world’s offshore basins. Oil & Gas Journal, 105(9).

Sarica, C., Panacharoensawad, E., 2012. Review of paraffin deposition research under multiphase
flow conditions. Energy & Fuels, 26(7):3968–3978. Doi: 10.1021/ef300164q.

Schramm, L.L., 2005. Interfacial energetics, in: Emulsions, Foams, and Suspensions: Fundamentals
and Applications. Wiley-VCH, Germany. chapter 3, pp. 53–97.

Scriven, L.E., 1976. Equilibrium bicontinuous structure. Nature, 263(5573):123–125. Doi:


10.1038/263123a0.

157
Serbutoviez, S., 2012. Offshore hydrocarbons, in: Panorama. IFP Energies nouvelles,
Rueil-Malmaison Cedex, France. (http://www.ifpenergiesnouvelles.com/publications/notes-de-
synthese-panorama/panorama-2012, accessed: 09/17/2012).

Shah, D.O., 1977. The world of surface science. Chemical Engineering Education, American Sociey
of Engineering Education.

Sjöblom, J., Aske, N., Auflem, I.H., Brandal, Ø., Havre, T.E., Sæther, Ø., Westvik, A., Johnsen,
E.E., Kallevik, H., 2003. Our current understanding of water-in-crude oil emulsions.: Recent
characterization techniques and high pressure performance. Advances in Colloid and Interface
Science, 100-102:399–473. Doi: 10.1016/S0001-8686(02)00066-0.

Sjöblom, J., Hemmingsen, P.V., Kallevik, H., 2007. The role of asphaltenes in stabilizing water-
in-crude oil emulsions, in: Mullins, O.C., Sheu, E.Y., Hammami, A., Marshall, A.G. (Eds.),
Asphaltenes, Heavy Oils, and Petroleomics. Springer, New York. chapter 21, pp. 549–587. Doi:
10.1007/0-387-68903-6 21.

Sjöblom, J., Øvrevoll, B., Jentoft, G., Lesaint, C., Palermo, T., Sinquin, A., Gateau, P., Barré,
L., Subramanian, S., Boxall, J., Davies, S., Dieker, L., Greaves, D., Lachance, J., Rensing,
P., Miller, K., Sloan, E.D., Koh, C.A., 2010. Investigation of the hydrate plugging and non-
plugging properties of oils. Journal of Dispersion Science and Technology, 31(8):1100 – 1119.
Doi: 10.1080/01932690903224698.

Skovborg, P., Rasmussen, P., 1994. A mass transport limited model for the growth of methane
and ethane gas hydrates. Chemical Engineering Science, 49(8):1131–1143. Doi: 10.1016/0009-
2509(94)85085-2.

Sloan, E.D., 2005. A changing hydrate paradigm - from apprehension to avoidance to risk man-
agement. Fluid Phase Equilibria, 228-229:67–74. Doi: 10.1016/j.fluid.2004.08.009.

Sloan, E.D., 2011. Chapter one - introduction, in: Sloan, E.D., Koh, C.A., Sum, A.K. (Eds.),
Natural Gas Hydrates in Flow Assurance. Gulf Professional Publishing, Boston. chapter 1, pp.
1–11.

Sloan, E.D., Koh, C.A., 2008a. Clathrate hydrates of natural gases. CRC Press, Boca Raton, FL.
3rd edition.

Sloan, E.D., Koh, C.A., 2008b. Estimation techniques of phase equilibria of natural gas hydrates,
in: Clathrate hydrates of natural gases. CRC Press, Boca Raton, FL. chapter 4, 3rd edition. pp.
189–256.

Sloan, E.D., Koh, C.A., 2008c. Hydrates in production, processing, and transportation, in:
Clathrate hydrates of natural gases. CRC Press, Boca Raton, FL. chapter 8, 3rd edition. pp.
643–683.

158
Solans, C., Infante, M., Garcia-Celma, M., Salager, J., 1996. Quantifying the concept of physico-
chemical formulation in surfactant-oil-water systems state of the art, in: Trends in Colloid and
Interface Science X. Springer Berlin / Heidelberg. volume 100 of Progress in Colloid and Polymer
Science, pp. 137–142.

Spiecker, P.M., Gawrys, K.L., Trail, C.B., Kilpatrick, P.K., 2003. Effects of petroleum resins on
asphaltene aggregation and water-in-oil emulsion formation. Colloids and Surfaces A: Physico-
chemical and Engineering Aspects, 220(13):9–27. Doi: 10.1016/S0927-7757(03)00079-7.

Sullivan, A.P., Kilpatrick, P.K., 2002. The effects of inorganic solid particles on water and crude
oil emulsion stability. Industrial & Engineering Chemistry Research, 41(14):3389–3404. Doi:
10.1021/ie010927n.

Tajima, H., Kiyono, F., Yamasaki, A., 2009. Direct observation of the effect of sodium dodecyl
sulfate (SDS) on the gas hydrate formation process in a static mixer. Energy & Fuels, 24(1):
432–438. Doi: 10.1021/ef900863y.

Talley, L.D., Turner, D.J., Priedeman, D.K., 2007. Method of generating a non-plugging hydrate
slurry. Patent WO/2007/095399. ExxonMobil Upstream Research Company. USA.

Taylor, C.J., Dieker, L.E., Miller, K.T., Koh, C.A., Sloan, E.D., 2007a. Micromechanical adhesion
force measurements between tetrahydrofuran hydrate particles. Journal of Colloid and Interface
Science, 306(2):255–261. Doi: 10.1016/j.jcis.2006.10.078.

Taylor, C.J., Miller, K.T., Koh, C.A., Sloan, E.D., 2007b. Macroscopic investigation of hydrate film
growth at the hydrocarbon/water interface. Chemical Engineering Science, 62(23):6524–6533.
Doi: 10.1016/j.ces.2007.07.038.

Turner, D., 2005. Clathrate hydrate formation in water-in-oil dispersions. Ph.D. thesis. Colorado
School of Mines. Golden, Colorado, USA.

Turner, D.J., Boxall, J., Yang, S., Kleehammer, D.M., Koh, C.A., Miller, K.T., Sloan, E.D., 2005.
Development of a hydrate kinetic model and its incorporation into the OLGA2000 R
transient
multiphase flow simulator, in: 5th International Conference on Gas Hydrates, ICGH, Trondheim,
Norway, 13-16 Jun. pp. 1231–1240.

van der Waals, J.H., Platteeuw, J.C., 1959. Clathrate solutions, in: Prigogine, I. (Ed.), Advances
in Chemical Physics I. John Wiley & Sons, Inc., pp. 1–57.

Verruto, V.J., Kilpatrick, P.K., 2008. Water-in-model oil emulsions studied by small-angle neutron
scattering: Interfacial film thickness and composition. Langmuir, 24(22):12807–12822. Doi:
10.1021/la802095m. http://pubs.acs.org/doi/pdf/10.1021/la802095m.

Vysniauskas, A., Bishnoi, P.R., 1983. A kinetic study of methane hydrate formation. Chemical
Engineering Science, 38(7):1061–1072. Doi: 10.1016/0009-2509(83)80027-X.

159
Wang, J., Creek, J.L., Buckley, J.S., 2006. Screening for potential asphaltene problems, in: SPE An-
nual Technical Conference and Exhibition, Society of Petroleum Engineers, San Antonio, Texas,
USA, 24-27 Sep. Doi: 10.2118/103137-MS.

Winsor, P.A., 1954. Solvent properties of amphiphilic compounds. Butterworths Scientific Publica-
tions, London.

Wolden, M., Lund, A., Oza, N., Makogon, T., Argo, C., Larsen, T., 2005. Cold flow black oil
slurry transport of suspended hydrate and wax solids, in: 5th International Conference on Gas
Hydrates, ICGH, Trondheim, Norway, 13-16 Jun. pp. 1101–1106.

Yang, S.O., Kleehammer, D.M., Huo, Z., Sloan, E.D., Miller, K.T., 2004. Temperature dependence
of particle-particle adherence forces in ice and clathrate hydrates. Journal of Colloid and Interface
Science, 277(2):335–341. Doi: 10.1016/j.jcis.2004.04.049.

York, J.D., Firoozabadi, A., 2008. Alcohol cosurfactants in hydrate antiagglomeration. Journal of
Physical Chemistry B, 112(34):10455–10465. Doi: 10.1021/jp8017265.

York, J.D., Firoozabadi, A., 2009. Effect of brine on hydrate antiagglomeration. Energy & Fuels,
23(6):2937–2946. Doi: 10.1021/ef800937p.

Zerpa, L.E., Sloan, E.D., Sum, A., Koh, C., 2011. Generation of best practices in flow assurance
using a transient hydrate kinetics model, in: 2011 Offshore Technology Conference, Offshore
Technology Conference, Houston, TX, USA, 2-5 May. Doi: 10.4043/21644-MS.

Zhang, C.S., Fan, S.S., Liang, D.Q., Guo, K.H., 2004. Effect of additives on formation of natural
gas hydrate. Fuel, 83(16):2115–2121. Doi: 10.1016/j.fuel.2004.06.002.

Zhong, Y., Rogers, R.E., 2000. Surfactant effects on gas hydrate formation. Chemical Engineering
Science, 55(19):4175–4187. Doi: 10.1016/S0009-2509(00)00072-5.

Zigrang, D.J., Sylvester, N.D., 1985. A review of explicit friction factor equations. Journal of
Energy Resources Technology, 107(2):280–283. Doi: 10.1115/1.3231190.

160
APPENDIX - DERIVATION OF THE MIXTURE ENERGY CONSERVATION EQUATION

The mixture energy conservation equation presented in Section 3.4.2.2 is based on a homoge-

neous fluid model, where the gas, liquid and hydrate phases are treated as a pseudo-phase. The

average density of the mixture is calculated as,

ρ = ρg Hg + ρw Hw + ρhyd Hhyd (A.1)

and the average velocity is calculated as,

ρg Hg vg + ρw Hw vw + ρhyd Hhyd vhyd ρg vSg + ρw vSw + ρhyd vShyd


v= = (A.2)
ρ ρ
The energy balance equation in terms of the pseudo-phase enthalpy is,

∂h ∂h 1 ∂p ∂v ∂v q̇πd
+v = −v − vv − vg sin θ − + SE,hyd (A.3)
∂t ∂x ρ ∂t ∂t ∂x ρAp
where h is mixture enthalpy per unit mass, v is volume average velocity, ρ is volume average density,

p is system pressure, g is acceleration of gravity, θ is angle of inclination with respect to horizontal,

q̇ is the wall heat flux, d is pipe diameter, and SE,hyd is the energy source term and is related with

the exothermic nature of the hydrate formation reaction by,

nH + 1 dmgas
SE,hyd = ∆hhyd (A.4)
ρhyd Ap ∆x dt
where, ∆hhyd is the enthalpy change of hydrate formation (for methane hydrate, 4175.4×103 J/kg).

When hydrates are forming, the heat of formation (∆hhyd ) acts as an energy source inside a pipe

section.

The following is a procedure to express the mixture energy conservation equation (eqn. A.3)

explicitly in terms of temperature. From the thermodynamic theory, entropy (s) is a function of

pressure (p) and temperature (T ).


   
∂sk ∂sk
dsk = dTk + dpk (A.5)
∂Tk pk ∂pk Tk

The specific enthalpy is also a function of temperature and pressure,


 
∂hk
dhk = Cp,k dTk + dpk (A.6)
∂pk Tk

161
For simple compressible systems of fixed chemical composition subject to an internally reversible

process, the first law of thermodynamics can be written as,

dhk υk dpk
dsk = − (A.7)
Tk Tk

where, υk is specific volume and k denotes either gas or liquid.

Substituting A.6 into A.7,


"  #
Cp,k 1 ∂hk
dsk = dTk + − υk dpk (A.8)
Tk Tk ∂pk Tk

From equations A.5 and A.8, the following equation can be obtained,
 
∂sk Cp,k
= (A.9)
∂Tk pk Tk
From Maxwell’s relationships,
   
∂υk ∂sk
=− (A.10)
∂Tk pk ∂pk Tk

Substituting A.9 and A.10 into A.5,


 
Cp,k ∂υk
dsk = dTk − dpk (A.11)
Tk ∂Tk pk

Substituting A.11 into A.7,


 
Cp,k ∂υk dhk υdpk
dTk − dpk = − (A.12)
Tk ∂Tk pk Tk Tk
Solving for dhk ,
"   #
∂υk
dhk = Cp,k dTk + υk − Tk dpk (A.13)
∂Tk pk

Deriving equation A.13 with respect to pressure at constant enthalpy gives,


  "   #
dTk 1 ∂υk
ηk = = Tk − υk (A.14)
dpk hk Cp,k ∂Tk pk
where, η is the Joule-Thomson coefficient.

Substituting A.14 into A.13

dhk = Cp,k dTk − ηk Cp,k dpk (A.15)

162
Combining A.15 with A.3, the energy balance equation in terms of temperature is obtained,

     
∂T ∂T ∂p ∂p 1 ∂p ∂v ∂v q̇πd
Cp +v −β +v − +v +v = −vg sin θ − + SE,hyd (A.16)
∂t ∂x ∂t ∂x ρ ∂t ∂t ∂x ρAp

where,

ρg Hg ηg Cp,g + ρw Hw ηw Cp,w
β= (A.17)
ρ
ρg Hg Cp,g + ρw Hw Cp,w
Cp = (A.18)
ρ
The Joule-Thomson coefficient for the gas phase can be estimated using the following correlation,
 
18
0.0048823Tpc 2
−1
Tpr
ηg = (A.19)
ppc Cp,g γg
The Joule-Thomson coefficient for the liquid water phase can be estimated from,
1
ηw = − (A.20)
ρw Cp,w
The energy balance equation in terms of temperature can be re-written as,

   
∂T ∂T 1 ∂p β ∂p ∂p v ∂v ∂v vg sin θ q̇πd SE,hyd
+v − − +v + +v =− − + (A.21)
∂t ∂x ρCp ∂t Cp ∂t ∂x Cp ∂t ∂x Cp Cp ρAp Cp

The energy balance equation in terms of temperature can be discretized applying a first-order

downstream scheme for spatial derivatives and a first-order upstream scheme for time derivatives,

considering a fully-implicit numerical scheme, as follows,

(n+1) (n)
" (n+1) (n+1)
# " (n+1) (n)
#
Ti − Ti Ti − Ti−1 1 pi − pi
+v −
∆t ∆x ρCp ∆t
(n+1) (n+1) (n+1) (n+1)
" (n+1) (n)
# " (n+1) (n)
#
β pi − pi pi − pi−1 v vi − vi vi − vi−1
− +v + +v
Cp ∆t ∆x Cp ∆t ∆x
vg sin θ q̇πd SE,hyd
=− − + (A.22)
Cp Cp ρAp Cp

Re-ordering,

163
      
1 v (n+1) 1 β 1 v (n+1) v 1 v (n+1)
+ T − + + pi + + v
∆t ∆x i ρCp ∆t Cp ∆t ∆x Cp ∆t ∆x i
(n)  
Ti v (n+1) 1 β (n) βv (n+1) v (n) vv (n+1)
= + Ti−1 − + pi − pi−1 + vi + v
∆t ∆x ρCp ∆t Cp ∆t Cp ∆x Cp ∆t Cp ∆x i−1
vg sin θ q̇πd SE,hyd
− − + (A.23)
Cp Cp ρAp Cp

For this model we assume that pressure and velocity profile are known and the only unknown
(n+1)
is Ti , which is solved explicitly and sequentially from node 1 to n.

 −1     
(n+1) 1 v 1 β 1 v (n+1) v 1 v (n+1)
Ti = + + + pi − − v
∆t ∆x ρCp ∆t Cp ∆t ∆x Cp ∆t ∆x i
(n)  
Ti v (n+1) 1 β (n) βv (n+1)
+ + Ti−1 − + pi − p
∆t ∆x ρCp ∆t Cp ∆t Cp ∆x i−1

v (n) vv (n+1) vg sin θ q̇πd SE,hyd
+ v + v − − + (A.24)
Cp ∆t i Cp ∆x i−1 Cp Cp ρAp Cp

164

You might also like