You are on page 1of 188

Copyright

by

Venkateswaran Sriram Pudugramam

2013
The Thesis Committee for Venkateswaran Sriram Pudugramam
certifies that this is the approved version of the following thesis:

A numerical study of CO2 -EOR with emphasis on mobility


control processes: Water Alternating Gas (WAG) and foam

APPROVED BY

SUPERVISING COMMITTEE:

Supervisor:
Mojdeh Delshad

Kamy Sepehrnoori
A numerical study of CO2 -EOR with emphasis on mobility
control processes: Water Alternating Gas (WAG) and foam

by

Venkateswaran Sriram Pudugramam, B.Tech.

Thesis
Presented to the Faculty of the Graduate School of
The University of Texas at Austin
in Partial Fulfillment
of the Requirements
for the Degree of

Master of Science in Engineering

The University of Texas at Austin


August 2013
Dedication

To my parents,

who have been the reason for whatever I have achieved till date.
Acknowledgments

I would like to express my sincere thanks to my research advisor, Dr. Mojdeh


Delshad, for showing the confidence and giving me an opportunity to work on CO2 -
EOR. Her steer, advice and guidance to move me past the innumerous occasions when I
felt stuck was invaluable. Very few people have a calming influence under pressure and
I feel grateful to have worked with one of the best research advisors. I would also like
to acknowledge Dr. Abdoljalil Varavei and Mohammad Reza Beygi for their help during
my simulation work and for the stimulating technical discussions we had. In addition, I
would like to show my gratitude to Dr. Kamy Sepehrnoori for serving as a reader of my
thesis and for providing valuable advice and feedback.
I would like to thank the entire faculty at the Department of Petroleum and
Geosystems Engineering at UT Austin for being such wonderful teachers and inspirational
researchers. I would also like to thank the staff members at the Center for Petroleum
and Geosystems Engineering, including Joanna Castillo, Esther Barrientes, Frankie Hart.
Funding for this research has been provided by DOE-NETL under contract number DE-
FE0005952. Thanks to CMG, Eclipse, PVTSim for providing the resources needed to
complete this research. I have had a wonderful two years with my labmates Zhitao,
Haomin, Neil, Hariharan, Mohammad, Peter both working on research as well as following
sports. I will also miss the company of my friends at UT who have made my graduate
study enjoyable. Special thanks to Hemanth who helped me compile my thesis in LATEX.
Finally, my sincere thanks to my family for their unconditional love and unwavering
support.

v
Abstract

A numerical study of CO2 -EOR with emphasis on mobility


control processes: Water Alternating Gas (WAG) and foam

Venkateswaran Sriram Pudugramam, M.S.E


The University of Texas at Austin, 2013

Supervisor: Mojdeh Delshad

CO2 enhanced oil recovery (CO2 -EOR) in residual oil zones has emerged as a vi-
able technique to maximize both the oil production and carbon storage. Most CO2 field
projects suffer from inadequate sweep because of high mobility of CO2 compared to the
oil. Gas conformance techniques have the potential to further improve the effectiveness
of CO2 -EOR projects.
The choice of mobility control to improve the sweep efficiency is critical and sim-
ulation studies with hysteretic relative permeability and mechanistic foam model can
assist in the choice of technique and optimization of the process for each reservoir. Two
promising mobility control practices of Water Alternating Gas (WAG) and foam are eval-
uated using the in-house compositional gas reservoir simulator (DOE-CO2 ).
The effect of hysteresis and cycle dependent relative permeability on WAG and
foam injections incorporating a new three-phase hysteresis model has been investigated.
Simulations are performed with and without hysteresis to assess the impact of the satu-
ration history and saturation path on gas entrapment, fluid injectivity and oil recovery.
The foam assisted technique in CO2 -EOR processes has also been investigated. Here
foam is generated in-situ by injecting surfactant solution with CO2 rather than directly
injecting foam. A simplified yet mechanistic population-balance model implemented in
vi
the in-house simulator has been applied to test the impact of foam. The results have
been compared with an empirical foam model which is the standard model in commercial
simulators.
Simulations have been performed on actual field models for selection and opti-
mization of the CO2 injection scheme, quantifying the impact of hysteresis, depicting the
effectiveness of CO2 -EOR process as against a surfactant flood, the effectiveness of foam
assisted floods and insights into low tension gas flooding process. All the above analyses
have also been performed on layer cake models with properties replicating the Permian
Basin carbonate reservoirs and Gulf Coast sandstone reservoirs.
Hysteresis shows an improvement in oil recovery of gas injection schemes where
flow reversal takes place. Foam has been found to be effective and the models show lower
CO2 utilizations factors compared to the case without foam.

vii
Table of Contents

Acknowledgments v

Abstract vi

List of Tables xii

List of Figures xiv

Abbreviations and Nomenclature xix

Chapter 1. Introduction 1
1.1 Research Objectives and Tasks . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Description of Chapters . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

Chapter 2. Background and Literature review 5


2.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.1 CO2 flooding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.2 Mobility control processes: WAG and foam . . . . . . . . . . . . . 6
2.2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.1 Relative permeability models . . . . . . . . . . . . . . . . . . . . 6
2.2.2 Hysteresis models . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.3 Impact of hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2.4 Foam models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.5 Field studies: CO2 flooding . . . . . . . . . . . . . . . . . . . . . 12

Chapter 3. Description of the Reservoir Simulator - DOE-CO2 17


3.1 Description of UTCOMP . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Advanced Features in DOE-CO2 . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.1 Foam module . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.2 Relative permeability module . . . . . . . . . . . . . . . . . . . . 19
viii
3.2.3 Fluid property and phase behavior module . . . . . . . . . . . . . 20
3.2.4 Geomechanics module . . . . . . . . . . . . . . . . . . . . . . . . 20

Chapter 4. Application of a New Cycle Dependent Three Phase Rela-


tive Permeability Model 22
4.1 Existing Three Phase Relative Permeability Models . . . . . . . . . . . . 22
4.2 Existing Hysteresis Models . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.3 Description of the New Cycle Dependent Three Phase Relative Permeabil-
ity Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.3.1 Relative permeability model (UTKR3P) . . . . . . . . . . . . . . 24
4.3.2 Three Phase Hysteresis model (UTHYST) . . . . . . . . . . . . . 26
4.4 Validation with Experimental Data . . . . . . . . . . . . . . . . . . . . . 27
4.5 Sensitivity Analysis of Various Parameters . . . . . . . . . . . . . . . . . 28
4.5.1 Sensitivity to parameters kr0 , C1 and C2 . . . . . . . . . . . . . . 29
4.5.2 Sensitivity to parameters ‘b’ . . . . . . . . . . . . . . . . . . . . . 30
4.5.3 Sensitivity to trapped gas phase . . . . . . . . . . . . . . . . . . . 30
4.6 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 30

Chapter 5. Application of a Mechanistic Foam Model to Foam Assisted


CO2 floods 36
5.1 Existing Foam Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.2 Description of the Mechanistic Foam Model . . . . . . . . . . . . . . . . 38
5.3 Validation with Experimental Data . . . . . . . . . . . . . . . . . . . . . 39
5.4 Metrics for Foam Performance Quantification . . . . . . . . . . . . . . . 40
5.5 Application to Numerical Simulation Models . . . . . . . . . . . . . . . . 40
5.5.1 Foam Assisted WAG flood . . . . . . . . . . . . . . . . . . . . . . 41
5.5.2 Simultaneous injection of gas and water with surfactant . . . . . . 42
5.5.3 Sensitivities to key foam parameters . . . . . . . . . . . . . . . . 43
5.6 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 44

Chapter 6. Field-Scale Case Study - Field C 57


6.1 Field Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2 Model Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
ix
6.3 Waterflood Simulation using Eclipse, UTCHEM and DOE-CO2 . . . . . 59
6.4 Selection of a Gas Injection Scheme: WAG vs Continuous Flood . . . . . 60
6.5 Sensitivity: Production Constraints in WAG . . . . . . . . . . . . . . . . 61
6.6 Sensitivity: Cycle Lengths in WAG . . . . . . . . . . . . . . . . . . . . . 61
6.7 Sensitivity: Hysteresis Model . . . . . . . . . . . . . . . . . . . . . . . . 62
6.8 Benchmark Cases and Comparison with Eclipse . . . . . . . . . . . . . . 63
6.9 Surfactant Flood Simulation . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.10 Comparison of Surfactant vs. CO2 Flood . . . . . . . . . . . . . . . . . . 64
6.11 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 64

Chapter 7. Field-Scale Case Study - Field L 90


7.1 Reservoir Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.2 Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7.3 WAG Injection Scheme (Base Case) . . . . . . . . . . . . . . . . . . . . . 91
7.4 Study 1: Mobility Control Processes - FAWAG vs. WAG . . . . . . . . . 92
7.4.1 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 92
7.5 Study 2: Foam Models - Empirical vs. Mechanistic Foam Models . . . . 92
7.5.1 Empirical foam model parameters . . . . . . . . . . . . . . . . . . 93
7.5.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 94
7.6 Study 3: FAWAG vs. Low Tension Gas Flooding . . . . . . . . . . . . . 94
7.6.1 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 95
7.7 Study 4: Benchmarking between DOE-CO2 and STARS . . . . . . . . . . 95
7.7.1 STARS empirical foam model . . . . . . . . . . . . . . . . . . . . 96
7.7.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 96
7.8 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 97

Chapter 8. Synthetic Field Case Studies 113


8.1 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
8.2 Carbonates (Permian Basin) . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.2.1 Model description and properties . . . . . . . . . . . . . . . . . . 115
8.2.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 116
8.3 Sandstones (Gulf Coast) . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
x
8.3.1 Model description and properties . . . . . . . . . . . . . . . . . . 117
8.3.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . 118
8.4 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 119

Chapter 9. Summary, Conclusions and Recommendations 128


9.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 128
9.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

Appendix: Input Files 132

References 158

xi
List of Tables

Table 3.1: Summary of the available relative permeability options in DOE-


CO2 (DOE-CO2 user guide (2013); Beygi, 2014) . . . . . . . . . . 21
Table 3.2: Summary of the available hysteresis options in DOE-CO2 (DOE-
CO2 user guide (2013); Beygi, 2014) . . . . . . . . . . . . . . . . . 21
Table 4.1: Summary of reservoir properties used in 2D simulation cases . . . 32
Table 4.2: Relative permeability properties for the base case (Case 1A) . . . 32
Table 4.3: Relative permeability properties that are varied for various cases . 32
Table 5.1: Summary of reservoir properties used in the synthetic case . . . . 45
Table 5.2: Summary of UTKR3P relative permeability parameters used in the
synthetic case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Table 5.3: Oil composition for fluid used in the synthetic model . . . . . . . 45
Table 5.4: Foam parameter values in the local equilibrium foam model . . . 46
Table 5.5: Comparison of FAWAG and WAG simulation results . . . . . . . 46
Table 5.6: Comparison of simultaneous injection - foam and no foam results 46
Table 5.7: Parameter values for sensitivity cases . . . . . . . . . . . . . . . . 46
Table 5.8: Results of sensitivity simulations . . . . . . . . . . . . . . . . . . 47
Table 6.1: Summary of reservoir properties for the Field C model . . . . . . 66
Table 6.2: Composition for the fluid in Field C . . . . . . . . . . . . . . . . . 66
Table 6.3: Comparison between Continuous CO2 and WAG cases . . . . . . 66
Table 6.4: Production and injection BHP constraints - WAG models . . . . . 67
Table 6.5: Sensitivity in production constraints - WAG models . . . . . . . . 67
Table 6.6: Sensitivity in cycle lengths - WAG models . . . . . . . . . . . . . 68
Table 6.7: Sensitivity in hysteresis - WAG models . . . . . . . . . . . . . . . 68
Table 7.1: Summary of reservoir properties for the Field L model . . . . . . 98
Table 7.2: Oil composition for the fluid in Field L . . . . . . . . . . . . . . . 98
Table 7.3: Foam parameters for the mechanistic model - Field L . . . . . . . 98
Table 7.4: Comparison between FAWAG (mechanistic) and WAG models . . 99
Table 7.5: Empirical foam model parameters . . . . . . . . . . . . . . . . . . 99
xii
Table 7.6: Comparison between foam models - mechanistic (IFOAM3) and
empirical (IFOAM2) . . . . . . . . . . . . . . . . . . . . . . . . . 99
Table 7.7: Comparison between models - FAWAG (IFOAM3) and low tension
gas flooding (IFOAM3+ICAP) . . . . . . . . . . . . . . . . . . . 100
Table 8.1: Summary of reservoir properties - carbonate reservoir case . . . . 120
Table 8.2: Summary of relative permeability end points - carbonate reservoir
case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Table 8.3: Comparison of results - carbonate reservoir case . . . . . . . . . . 120
Table 8.4: Summary of reservoir properties - sandstone reservoir case . . . . 121
Table 8.5: Comparison of results - sandstone reservoir case . . . . . . . . . . 121

xiii
List of Figures

Figure 1.1: Incremental oil production from gas EOR processes in US. . . . . 4
Figure 2.1: Schematic of the CO2 flooding process (from NETL website). . . 16
Figure 2.2: Fluid movement in WAG [from Jarrell et al. (2002)] . . . . . . . 16
Figure 4.1: Effect of parameter ‘b2 ’ on the curvature of phase 2 iso-perms
where phase 2 refers to oil phase. . . . . . . . . . . . . . . . . . . 33
Figure 4.2: Calculated vs. experimental relative permeability values (Sam-
ple15 Oak 1991) where plots 4.2a, 4.2b, 4.2c show the match for
phases 1(water), 2(oil) and 3(gas). . . . . . . . . . . . . . . . . . 33
Figure 4.3: UTKR3P model with path dependency (Kr031 = 6 Kr032). Here
phase 3 can refer to gas phase. . . . . . . . . . . . . . . . . . . . 34
Figure 4.4: UTKR3P model path dependency (bj 6= 0) . . . . . . . . . . . . 34
Figure 4.5: Phase-3 relative permeability using UTKR3P with UTHYST for
the effect of trapped phase-3 . . . . . . . . . . . . . . . . . . . . 35
Figure 4.6: Saturation and trapped saturation for Phase 3 in two WAG cycles 35
Figure 5.1: Cumulative oil recovery for N2 foam - experimental and simulation
results (Naderi et al., 2013) . . . . . . . . . . . . . . . . . . . . . 47
Figure 5.2: Pressure drop for N2 foam vs. pore volumes of fluid injected ex-
perimental and simulation results (Naderi et al., 2013) . . . . . . 48
Figure 5.3: Permeability (mD) profile for the layer cake synthetic model . . . 48
Figure 5.4: Effluent volume fraction for FAWAG case . . . . . . . . . . . . . 49
Figure 5.5: Effluent volume fraction for WAG case . . . . . . . . . . . . . . 49
Figure 5.6: Recycling fraction - FAWAG vs. WAG . . . . . . . . . . . . . . 50
Figure 5.7: Oil recovery (fraction of OOIP) as a function of total injected fluid
pore volumes - FAWAG vs. WAG . . . . . . . . . . . . . . . . . 50
Figure 5.8: Cumulative gas injected - FAWAG vs. WAG . . . . . . . . . . . 51
Figure 5.9: Cumulative water injected - FAWAG vs. WAG . . . . . . . . . . 51
Figure 5.10: Effluent volume fraction for simultaneous injection with foam . . 52
Figure 5.11: Effluent volume fraction for simultaneous injection without foam 52
Figure 5.12: Recycling fraction for simultaneous injection - foam vs. no foam 53

xiv
Figure 5.13: Oil recovery (fraction of OOIP) for simultaneous injection - foam
vs. no foam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 5.14: BHP of the injector for simultaneous injection - foam vs. without
foam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Figure 5.15: Recycling fraction for smaller generation rate coefficient compared
to the base case . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Figure 5.16: Recycling fraction for lower coalescence rate coefficient compared
to the base case . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Figure 5.17: Recycling fraction for smaller viscosity coefficient compared to the
base case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Figure 5.18: Recycling fraction for smaller limiting capillary pressure compared
to the base case . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Figure 6.1: Permeability in the X direction (mD) - Field C . . . . . . . . . . 69
Figure 6.2: Porosity map - Field C . . . . . . . . . . . . . . . . . . . . . . . 69
Figure 6.3: Initial oil saturation - Field C . . . . . . . . . . . . . . . . . . . . 70
Figure 6.4: Location of wells - Field C . . . . . . . . . . . . . . . . . . . . . 71
Figure 6.5: Schematic of wells - Field C . . . . . . . . . . . . . . . . . . . . . 71
Figure 6.6: Oil-water relative permeability curves for waterflood match - Field C 72
Figure 6.7: Gas-oil relative permeability curves for waterflood match - Field C 72
Figure 6.8: Oil recovery (fraction of remaining OIP) comparison for 5 year -
waterflood . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Figure 6.9: Average reservoir pressure comparison for 5 year - waterflood . . 73
Figure 6.10: Water injection rates comparison for 5 year - waterflood . . . . . 74
Figure 6.11: Water production rates comparison for 5 year - waterflood . . . . 74
Figure 6.12: Effluent volume fraction for the continuous CO2 flood . . . . . . 75
Figure 6.13: Effluent volume fraction for the WAG CO2 flood . . . . . . . . . 75
Figure 6.14: Recycling fraction for Field C - continuous CO2 and WAG floods 76
Figure 6.15: Oil recovery for Field C - continuous CO2 and WAG . . . . . . . 76
Figure 6.16: Oil recovery for Field C - sensitivity to production constraints . 77
Figure 6.17: Recycling for Field C - sensitivity to production constraints . . . 77
Figure 6.18: Gas injection rates for Field C - sensitivity to production constraints 78
Figure 6.19: Recycling fraction for Field C - sensitivity to cycle lengths . . . . 78
Figure 6.20: Gas injection rates for Field C - sensitivity to cycle lengths . . . 79
xv
Figure 6.21: Oil recovery fraction for Field C - sensitivity to cycle lengths . . 79
Figure 6.22: Carlson’s hysteresis model with Corey relative permeability model 80
Figure 6.23: Recycling fraction for Field C - sensitivity to hysteresis . . . . . 81
Figure 6.24: Gas injection rates for Field C - sensitivity to hysteresis . . . . . 81
Figure 6.25: Oil recovery (fraction of remaining OIP) for Field C - sensitivity
to hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Figure 6.26: Oil-water relative permeability curves for CO2 flood match . . . 82
Figure 6.27: Gas-oil relative permeability curves for CO2 flood match - Field C 83
Figure 6.28: Oil recovery (fraction of remaining OIP) comparison: continuous
CO2 flood - Field C . . . . . . . . . . . . . . . . . . . . . . . . . 83
Figure 6.29: Pressure comparison: continuous CO2 flood - Field C . . . . . . 84
Figure 6.30: Gas injection volume comparison: continuous CO2 flood - Field C 84
Figure 6.31: Water production volume comparison: continuous CO2 flood -
Field C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Figure 6.32: Oil and water rates: surfactant flood - Field C . . . . . . . . . . 85
Figure 6.33: Oil recovery fraction and pressure: surfactant flood - Field C . . 86
Figure 6.34: Surfactant injection and production: surfactant flood - Field C . 86
Figure 6.35: Water injection and production: surfactant flood - Field C . . . . 87
Figure 6.36: Oil rates for waterflood, WAG-CO2 flood, and surfactant flood -
Field C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Figure 6.37: Oil recovery fraction versus time for waterflood, WAG-CO2 flood,
and surfactant flood - Field C . . . . . . . . . . . . . . . . . . . 88
Figure 6.38: Oil recovery fraction versus pore volume injection for waterflood,
WAG-CO2 flood and surfactant flood - Field C . . . . . . . . . . 88
Figure 6.39: Average pressure for waterflood, WAG-CO2 flood and surfactant
flood - Field C . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Figure 7.1: Permeability in the X direction (mD) - Field L . . . . . . . . . . 100
Figure 7.2: Initial oil saturation distribution for Field L . . . . . . . . . . . . 101
Figure 7.3: Oil-water relative permeability curves matched with the tabular
values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Figure 7.4: Gas-oil relative permeability curves matched with the tabular values102
Figure 7.5: Cumulative oil produced and pressure vs. time - WAG . . . . . . 102
Figure 7.6: Water and gas production rates vs. time - WAG . . . . . . . . . 103
xvi
Figure 7.7: Water and gas injection rates vs. time - WAG . . . . . . . . . . 103
Figure 7.8: CO2 production and injection rates vs. time - WAG . . . . . . . 104
Figure 7.9: Effluent volume fraction vs. time - WAG . . . . . . . . . . . . . 104
Figure 7.10: Recycling fraction for Field L - FAWAG (IFOAM3 - mechanistic)
vs. WAG (No foam) . . . . . . . . . . . . . . . . . . . . . . . . . 105
Figure 7.11: Oil recovery (fraction of OOIP) for Field L - FAWAG (IFOAM3 -
mechanistic) vs. WAG (No foam) . . . . . . . . . . . . . . . . . 105
Figure 7.12: BHP of the injector for Field L - FAWAG (IFOAM3 - mechanistic)
vs. WAG (No foam) . . . . . . . . . . . . . . . . . . . . . . . . . 106
Figure 7.13: Permeability reduction factor for Field L - Mechanistic vs. Em-
pirical foam models . . . . . . . . . . . . . . . . . . . . . . . . . 106
Figure 7.14: BHP of the injector for Field L - (IFOAM3 - mechanistic) versus
(IFOAM2 - empirical) . . . . . . . . . . . . . . . . . . . . . . . . 107
Figure 7.15: Recycling fraction for Field L - (IFOAM3 - mechanistic) versus
(IFOAM2 - empirical) . . . . . . . . . . . . . . . . . . . . . . . . 107
Figure 7.16: Oil recovery (fraction of OOIP) for Field L - (IFOAM3 - mecha-
nistic) vs. (IFOAM2 - empirical) . . . . . . . . . . . . . . . . . . 108
Figure 7.17: Recycling fraction for Field L - (IFOAM3 - FAWAG) vs. (IFOAM3+
ICAP - low tension gas flooding) . . . . . . . . . . . . . . . . . . 108
Figure 7.18: Oil recovery (fraction of OOIP) for Field L - (IFOAM3 - FAWAG)
vs. (IFOAM3+ICAP - low tension gas flooding) . . . . . . . . . 109
f
Figure 7.19: STARS vs. DOE-CO2 model - gas mobility reduction factor (krg /krg )109
Figure 7.20: Oil recovery - STARS vs. DOE-CO2 foam models . . . . . . . . 110
Figure 7.21: Gas injection rates - STARS vs. DOE-CO2 foam models . . . . . 110
Figure 7.22: Gas production rates - STARS vs. DOE-CO2 foam models . . . 111
Figure 7.23: Inverse of foam mobility reduction factor - STARS foam model . 111
Figure 7.24: Foam mobility reduction factor - DOE-CO2 foam model . . . . . 112
Figure 8.1: Permeability in the X-direction (mD) - carbonate case . . . . . . 122
Figure 8.2: Porosity distribution in the carbonate case . . . . . . . . . . . . 122
Figure 8.3: Recycling fraction comparison for the carbonate reservoir cases . 123
Figure 8.4: Comparison of the injector bottom-hole pressure for the carbonate
reservoir cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Figure 8.5: Oil recovery (fraction of OOIP) comparison for the carbonate
reservoir cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
xvii
Figure 8.6: Permeability in the X-direction (mD) - sandstone reservoir case . 124
Figure 8.7: Porosity distribution in the sandstone reservoir case . . . . . . . 125
Figure 8.8: Recycling fraction comparison for the sandstone reservoir cases . 125
Figure 8.9: Comparison of the injector bottom-hole pressure for the sandstone
reservoir cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Figure 8.10: Oil recovery (fraction of OOIP) comparison for the sandstone
reservoir cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Figure 8.11: Effluent volume fraction for the carbonate reservoir case with WAG127
Figure 8.12: Effluent volume fraction for the sandstone reservoir with WAG . 127

xviii
Abbreviations and Nomenclature

Abbreviations:
EOR : Enhanced Oil Recovery
EOS : Equation Of State
FAWAG : Foam Assisted Water Alternating Gas
IMPEC : Implicit Pressure Explicit Concentration
IMPES : Implicit Pressure Explicit Saturation
MMP : Minimum Miscibility Pressure
OOIP : Original Oil In Place
PVT : Pressure Volume Temperature
ROZ : Residual Oil Zones
STARS : Commercial foam simulator of CMG Ltd.
UTCHEM : Chemical flooding simulator of UT Austin
UTCOMP : Gas flooding simulator of UT Austin
UTHYST : Hysteresis model developed by Beygi et al. (2013)
UTKR3P : Relative permeability model developed by Beygi et al. (2013)
WAG : Water Alternating Gas injection
Nomenclature:
α : Proportionality constant for gas viscosity
β : Constant to change the curvature of residual/trapped saturation
φ : Porosity
bj : Parameter relating the two- and three-phase residual saturation
C1 and C2 : Relative permeability curvature parameters
Cs : Surfactant concentration
k1 and k−1 : Foam generation and coalescence rate coefficients
xix
krj : Relative permeability of phase ‘j’
0
krj : End point relative permeability of phase ‘j’
Lj : Land coefficient for phase ‘j’
nf : Bubble density of free foam
nt : Bubble density of trapped foam
Qb : Source/sink for foam bubble
Sj : Saturation of phase ‘j’
Sjt : Trapped saturation of phase ‘j’
Sjc : Connate/critical saturation of phase ‘j’
Sirj : Residual saturation of phase ‘i’ w.r.to phase ‘j’ in two-phase flow
2P
Sjr : Two-phase residual saturation of phase ‘j’
3P
Sjr : Three-phase residual saturation of phase ‘j’
Sjmax : Maximum saturation of phase ‘j’
u : Darcy velocity
v : Interstitial velocity

xx
Chapter 1

Introduction

Primary production from a hydrocarbon reservoir can only recover a small fraction
(10%) of the initial hydrocarbon in place using the reservoir’s natural energy drive. In
the next phases, fluids such as water or gas are injected to extract the hydrocarbon. The
injection of water for sweeping the hydrocarbon as well as maintaining or raising pressure
comes under secondary production. This phase can recover as high as 20-30% additional
hydrocarbon. The majority of the remaining oil is trapped by capillary forces, bypassed
due to reservoir heterogeneities or unfavorable mobility of the injected fluid to displace
reservoir oil. Therefore, a significant fraction of the remnant oil is available as a target
for Enhanced Oil Recovery (EOR) processes.

One of the types of EOR processes is Gas EOR which includes CO2 , N2 , air and
hydrocarbon flooding in a miscible or immiscible mode. This is a commonly used EOR
for light and medium API oils. The main advantage of miscible gas flooding is its better
microscopic sweep leading to less residual oil in the pores. CO2 flooding is one of the
most widely used gas flooding processes. Over 75% of the incremental oil production
in 2010 in the US from gas flooding processes comes from CO2 flooding (see Figure 1.1
from Koottungal, 2010). As per the Oil and Gas Journal (2010), CO2 flooding (miscible)
produced 272,109 bbl/d from 109 projects and CO2 flooding (immiscible) 9,160 bbl/d
from 5 projects. When combined, they accounted for 5.1% of the total US oil production.
The other advantage of CO2 flooding is in its ability to store CO2 in the subsurface
formations reducing the global carbon footprint.

The major drawback associated with gas injection is its poor volumetric sweep

1
efficiency, as a result of which the gas does not contact a lot of oil, and incremental
recovery is low. This can happen due to channeling (flow through high permeability
layers), viscous fingering (viscosity difference between oil and gas) and gas overriding
(density difference of the injected gas and the reservoir oil). So, mobility and conformance
issues are the most serious concerns which diminish the effectiveness of a CO2 flood.

1.1 Research Objectives and Tasks

Mobility control of CO2 involves overcoming the difference in densities between


injected CO2 and reservoir oil. While sweep and conformance control could be addressed
by infill drilling, pattern drilling, horizontal wells and packers for isolation, the main
focus of this research is mobility control of the injected CO2 using the techniques of
Water Alternating Gas (WAG) injection and foam injection. Predictive simulation of
these processes is therefore critical to optimally develop the field using CO2 flooding. It
is critical to develop robust, scalable, and mechanistic numerical modeling tools because
of the multiple scales of the various interacting processes in the reservoir, the basin scale
of the reservoir, the need for long time predictions, and the potential for significant
coupling between geomechanics and flow. The current research will therefore look at two
important aspects for the simulation of mobility control processes in CO2 flooding.

The first one is relative permeability of the different phases. The performance of
CO2 injection in producing the oil is a strong function of relative permeability and capil-
lary pressure functions. The fact that relative permeabilities and capillary pressures are
functions of both saturation and its history, known as hysteresis, is well established. In
the case of WAG simulations, where flow reversal of the different phases occurs frequently,
hysteresis is a major factor. This study will therefore look at the existing relative perme-
ability and hysteresis models, utilize the best model available and quantify the impact
of various parameters on oil recovery.

2
The second aspect is the modeling for foam process. Foam can be described
as gas bubbles separated by liquid films and stabilized in the presence of surfactant
molecules. Modeling foam can be done by an empirical approach where the gas relative
permeability is modified based on an estimation of the foam strength. This study will
look at mechanistic foam models where the foam bubble density is computed based on
foam generation and coalescence mechanisms and foam rheology is taken into account.
Benchmarking will also be done for the mechanistic foam models with empirical foam
models.

1.2 Description of Chapters

In the second chapter, a brief description of the CO2 flooding and the mobility
control methods such as WAG and foam are provided. Also, an extensive literature
review on relative permeability models, foam models and field-scale case studies are
provided. The third chapter briefly describes the features of UTCOMP and later the
advanced features in DOE-CO2 . The fourth chapter describes UTKR3P, a new three
phase hysteretic relative permeability model implemented in DOE-CO2 and sensitivities
of various parameters in the model. Also described is UTHYST, the hysteresis model
which could be used with any three phase relative permeability model modeling trapped
phase saturations.

The fifth chapter describes the local equilibrium approximation of the population
balance foam model by Chen et al. (2010) which is implemented in DOE-CO2 . Also,
simulations to show the impact of foam and the sensitivities of various parameters in the
foam model are carried out.

In the sixth chapter, the EOR processes of surfactant flooding and CO2 flooding
for a field-scale model are compared and the best development scheme for the field is
suggested. The study also focusses on various sensitivities of CO2 flooding including

3
production constraints, half cycle length in WAG and hysteresis behavior. The results
are also benchmarked with the ones from EclipseT M , a widely used commercial simulator.

In the seventh chapter, simulations of foam flooding for a field-scale model are
presented. The study also compares the empirical foam model with the mechanistic
foam model. Insights into low tension gas flooding, a novel gas flooding process, are also
presented. The results of the foam flooding DOE-CO2 are also benchmarked with the
ones from STARST M , a widely used commercial simulator for foam simulations.

In the eighth chapter, the mobility control processes of WAG and foam are com-
pared in two different and popular development settings: Permian Basin carbonates and
Gulf Coast sandstones. The ninth chapter summarizes the findings of this research and
provides recommendations on the way forward.

Figure 1.1: Incremental oil production from gas EOR processes in US.

4
Chapter 2

Background and Literature review

2.1 Background
2.1.1 CO2 flooding

One of the popular miscible gas flooding processes is CO2 flooding. This process
is commonly used to recover oil from reservoirs in which the initial pressure has been
depleted through primary production and possibly water-flooding. Water is injected into
the reservoir until pressure is restored to a desired level, and then CO2 is introduced into
the reservoir through these same injection wells. As the CO2 is forced into the reservoir
a zone of miscible CO2 and light hydrocarbons forms a front that is soluble with the oil,
making it easier to move toward production wells. Production is from an oil bank that
forms ahead of the miscible front. As reservoir fluids are produced through production
wells, the CO2 reverts to a gaseous state and provides a ‘gas lift’ similar to that of original
reservoir natural gas pressure. See Figure 2.1 for a pictorial representation of the CO2
flooding process.

Depending on the reservoir fluids at reservoir conditions, the displacement of oil


by CO2 can be either miscible or immiscible. Miscibility is achieved when two fluids
combine and form a single phase; one fluid can completely displace the other phase
leaving no residual behind. The advantages of immiscible injection are that the pressure
requirement is not large, solvents are not expensive and oil can still be recovered. The
recovery mechanisms associated with miscible injection are swelling of oil, reduction of
oil viscosity, developing miscibility (IFT reduction) and gravity drainage. In the miscible
process, the pressure has to be above the Minimum Miscibility Pressure (MMP) which

5
depends on temperature, pressure, solvent purity and the molecular weight of the heavy
components of the crude oil. CO2 has a density like liquid but a viscosity like gas in the
reservoir at supercritical conditions and so it is a preferred solvent.

2.1.2 Mobility control processes: WAG and foam

Large solvent (CO2 ) mobility promotes instabilities and amplifies the effect of
heterogeneities. Hence, one of the flood designs in CO2 flooding is WAG where the
initial CO2 slug is typically followed by alternate water and CO2 injection - the water
serving to improve sweep efficiency and minimize the amount of CO2 required for the
flood. This alternate injection counters the effects of water under-riding and gas over-
riding. This scheme tries to minimize the gas recycling in high permeability layers and
improve sweep efficiency. The injection can also be done in a tapered fashion where the
CO2 slug size decreases and the water slug size increases as cycles progress. A schematic
of the fluid displacement in WAG can be seen in Figure 2.2.

Foam is a dispersion of gas in a continuous liquid phase, which can constitute a


minor fraction of foam. The gas phase is discontinuously organized in gas bubbles which
contact each other by thin liquid films, called lamellae. Foam can affect the oil recovery
in three ways compared to gas or WAG (Farajzadeh et al., 2012): (1) by stabilizing
the displacement process as the displacing fluid (gas or foam) viscosity increases; (2) by
blocking the high-permeable swept zones and diverting the fluid into the unswept zones;
and (3) by reducing the capillary forces via reducing the interfacial tensions due to the
presence of a surfactant.

2.2 Literature Review


2.2.1 Relative permeability models

One of the earliest published relative permeability models is by Corey et al. (1956).
Corey and co-workers presented the results of three-phase relative permeability tests on
6
nine water-wet Berea sandstone samples and also showed a method to calculate the oil
and water relative permeabilities from the easily measured gas relative permeabilities.
Land (1968) presented the relative permeability functions for the two- and three-phase
systems for the imbibition cycle for water wet rocks. Also an empirical relation between
residual non-wetting phase saturation after water imbibition and initial non-wetting phase
saturation is presented.

Stone (1970) presented a model (Stone I) for interpolating between two phase
oil-water and oil-gas curves to obtain three phase oil relative permeabilities. In this
model, the relative permeabilities of water and gas depend only on the concerned phase
saturation. Stone (1973) also improved (Stone II) on his earlier model by incorporating
residual oil data in the three phase region. Aziz and Settari (1979) normalized Stone I
as the original model was valid only if the end point was unity.

Delshad et al. (1987) presented the two- and three-phase relative permeabilities
of micellar fluids. They showed that the relative permeability of each phase is a func-
tion of its own saturation and adding polymer only had a slight effect on the relative
permeabilities of brine and microemulsion. Baker (1988) presented a true-linear inter-
polation between equal relative permeabilities on a ternary plot by straight lines. Both
linear and saturation-weighted interpolation was shown to obtain better fits to the avail-
able three-phase data. Delshad and Pope (1989) made an extensive comparison of seven
three-phase models and also proposed a parameter based model in which the two-phase
relative permeability of oil does not explicitly appear in the three-phase equations.

Dria (1993) measured two and three phase relative permeabilities for CO2 flooding
conditions and concluded that significant differences exist between three-phase gas, oil,
and brine relative permeabilities when the gas is CO2 as against N2 . Jerauld (1997)
developed a three-phase relative permeability correlation that incorporates hysteresis in
gas, oil and water as well as the dependence of relative permeability on composition and

7
interfacial tension. This model was developed for Prudhoe Bay but was concluded that
it could be applied to other mixed-wet reservoirs with changes in the input parameters.

Blunt (2000) presented a three-phase relative permeability model by writing the


relative permeabilities as unique functions of the flowing saturation. This model accounts
for changes in composition, saturation paths, trapping of all phases and any sequence of
saturation changes for a water-wet media. The concept of layer drainage allows the oil
relative permeability to be extrapolated to low saturation. Fayers et al (2000) presented
a model based on the Baker saturation weighted interpolation but modified for the oil
phase through saturation rescaling. This model has inherent hysteresis behavior and is
also extended to allow for compositional changes in the oil and gas phases.

Hustad (2000) presented a formulation incorporating hysteresis and miscibility in


both capillary pressure and relative permeability simultaneously. Yuan and Pope (2012)
presented a model to avoid the phase identification problem by formulating a relative
permeability model based on the continuous Gibbs free energy. Beygi et al. (2013)
presented a cycle dependent three phase hysteretic relative permeability model which is
discussed in detail in the later chapters.

2.2.2 Hysteresis models

Land (1968) presented a method to calculate the imbibition relative permeability


based on free and trapped saturation. This was compared with experimental data for
alundum and Berea sandstone samples. Killough (1976) presented a hysteresis model per-
mitting smooth transitions in either direction between drainage and imbibition based on
bounding curves and a Killough parameter to calculate imbibition relative permeability.

Carlson (1981) presented a method to calculate the imbibition relative permeabil-


ity based on Lands parameter. The data required to calculate the imbibition curve are
the drainage curve, the historical maximum non-wetting saturation and an additional
point on an experimental imbibition curve. Fayers and Mathews (1984) considered hys-
8
teresis in oil phase by computing the residual oil saturation based on the trapped gas
saturation. Larsen and Skauge (1998) developed hysteresis models for both the wetting
and the non-wetting phases. This is an irreversible model in which the relative perme-
ability continuously changes as cycles of imbibition and drainage progress. This model
also utilizes the Land parameter.

The hysteresis model developed by Egermann et al. (2000) is based on fractal


theory and involves defining trapping and un-trapping coefficients. This model computed
different Land coefficients for different cycles. Spiteri et al. (2005) have presented a new
model of trapping and waterflood relative permeability base on pore-network simulations.
This model captures non-monotonicity of the intitial-residual curves and convexity of the
waterflood relative permeability curves for oil-wet media. The hysteresis model developed
by Beygi et al. (2013) has monotonically increasing trapped saturations and involves
defining a dynamic Land coefficient dependent on the cycles. This model is dealt in
detail in the later chapters.

2.2.3 Impact of hysteresis

A lot of literature has been published showing the impact of hysteresis on compo-
sitional simulation. Christensen et al. (1998) showed by simulation results that recovery
by WAG injection may be underestimated in the compositional model due to lack of pos-
sibility for cycle dependent relative permeability. The results showed lower gas-oil ratio
and higher oil recovery for the case with cycle dependent relative permeabilities com-
pared to the case with two-phase hysteresis models. Spiteri et al. (2004) have showed
that the Land trapping coefficient decreases in successive cycles during flow reversals.
Also, the three-phase hysteresis models yielded higher recovery predictions due to re-
duced mobility of the gas phase during water injection. They concluded that depending
on the interpolation models used, the difference in recovery could be as high as 15 %.

Ghomian et al. (2008) showed the impact of WAG parameters (WAG ratio and
9
slug size), hysteresis and reservoir heterogeneity on the recovery fraction. Gas trapping
due to hysteresis affected the relative permeabilities and resulted in an improvement in
sweep efficiency. Masalmeh and Wei (2010) modeled three phase flow, hysteresis and
compositional effects on relative permeabilities for a reservoir with different scales of
heterogeneity. They concluded that the combination of hysteresis and IFT dependent
relative permeabilities had the most significant impact on WAG injection schemes in
heterogenous reservoirs. Shahverdi et al. (2011) showed how the three phase relative
permeability models failed to predict the continued production of oil after breakthrough
of gas seen in WAG experiments. The experiments showed strong cyclic hysteresis for
oil, water and gas relative permeabilities in both water-wet and mixed-wet systems.

2.2.4 Foam models


2.2.4.1 Population balance models

One of earliest work on population balance method of foam flow description was
by Patzek (1985). In that work, the effect of bubble size was incorporated accounting for
changes in foam texture caused by mechanisms which create, destroy or transfer bubbles.
The effects of capillary snap-off, bubble division, coalescence, diffusion evaporation as
well as bubble trapping and mobilization were taken into account. Chang et al. (1990)
conducted foam coreflood experiments with CO2 foam and proposed correlations for the
relationship between gas mobility and gas fraction, interstitial gas velocity, surfactant
concentration and oil saturation. They showed that the experimental results agree well
with the correlations.

Friedmann et al. (1990) proposed a mathematical model that included the prin-
cipal mechanisms that govern foam displacement in porous media. The effect of foam
on gas-phase relative permeability and apparent viscosity for both static or continuous
gas foams and strong or discontinuous gas foams are modeled. Kovscek et al. (1995)
presented a mechanistic model for foam displacement in porous media incorporating

10
the pore-level mechanisms of foam generation, coalescence and transport. This model
employed capillary pressure dependent kinetic expressions for lamellae generation and
coalescence and trapping of lamellae. Chen et al. (2010) proposed a simplified approx-
imation to the population-balance method for computing number density, or number
concentration, of foam bubbles per unit volume of flowing gas. This model provides an
appropriate prediction of field-scale foam applications.

2.2.4.2 Empirical and semi-empirical models

Marfoe et al. (1987) modified the gas viscosity to incorporate the foam effect in
the reservoir simulations. Their model implicitly applies the foam bubble size correlated
with foamed gas velocity and available flowing water content. Patzek and Myhill (1989)
presented a simple model of steam foam transport where the only adjustable parameter
is an effective surfactant partition coefficient that accounts for surfactant losses and
inefficiencies of foam generation. This model was then applied to the Kern River Bishop
pilot.

Islam and Farouq Ali (1990) expanded the Marfoe et al. model by adding the
detrimental effect of oil in the foam texture along with the effect of absolute permeability
and pressure gradient on foam. Fisher et al. (1990) proposed an empirical foam model for
the coalescence regime where the destruction of foam lamellae depends on gas/surfactant
solution flowrates, fluid mobilities and rock capillary pressure. Mohammadi et al. (1993)
showed that a reservoir simulator with empirical foam modeling capabilities provides a
useful tool for interpretation of the obtained field information in conjunction with known
geological characteristics. This approach could adequately capture most of the foam
behavior without resorting to detailed mechanisms of foam bubble creation/destruction.

11
2.2.4.3 Other models

Zhou and Rossen (1995) presented a fractional flow analysis extrapolated from
laboratory data for foams at the limiting capillary pressure. Chou (1990) presented a
statistical network model to describe the behavior of foam in porous media. The theory
is based on pore throat and pore body size distribution functions which characterize the
converging/diverging nature of the pore geometry and the assumption that foam lamellae
are generated by snap-off in pores with an aspect ratio greater than a critical value.

Cheng et al. (2000) compared the behavior of foam based on two models in
two different simulators. They also presented a fitting scheme for benchmarking the
parameters of both the models with one another. Farajzadeh et al. (2012) reviewed the
mechanisms and theories such as disjoining pressure, coalescence and drainage, entering
and spreading of oil, oil emulsification, pinch-off to explain the impact of oil on foam
stability in the bulk and porous media. Also, ideas on the improvement of foam stability
and longevity in the presence of oil are presented. Naderi et al. (2013) compared the
performance of N2 foams with CO2 foams and also benchmarked the simulation results
with the experimental values.

2.2.5 Field studies: CO2 flooding

A lot of literature has been published over the years regarding field studies in CO2
flooding. The following listed shows a breadth of the types of studies published.

Hoiland et al. (1986) published the study on West Sussex Shannon reservoir in
the Rocky Mountain for a CO2 flood pilot. The study demonstrated the ability of CO2
to mobilize and displace additional volumes through a miscible injection scheme after a
pre-flush of water inspite of adverse permeability profile and complex reservoir faulting.
Beliveau (1987) presented the planning and design of a CO2 pilot in the Midale field in
Canada. The study also highlights the best pilot design as an integration of monitoring
and producing type pilots. Dawson et al. (1989) present the simulations results of
12
hydrocarbon miscible flood in the Sadlerocht and Sag River sandstones in the Prudhoe
Bay Unit. The simulation study was determine EOR sensitivity to reservoir description,
solvent injection rate and solvent bank size and finalize the injection strategy.

Hervey and Iakovakis (1991) summarize a miscible CO2 flood in the Rangely
Weber Sand Unit in Colorado. The project was designed for a 30 % slug of CO2 in a
1:1 WAG ratio with incremental recovery of 34.5 % of production. Stein et al. (1992)
described the performance of a pilot and a unit-wide CO2 flood in the Slaughter Estate
Unit, Slaughter field in the Permian Basin. The performance and design of both the
projects are compared to get insight into the process, the impact of the flood variables
and the effect of project scale.

Surguchev and Korbol (1992) offer an insight into screening for WAG injection
strategies. Evaluation of miscibility, effect of injection parameters on injectivity and
different injection strategies such as non-stationary and cyclic injection is also presented.
Masoner and Wackowski (1995) summarize the miscible CO2 project in the Rangely
Weber Sand Unit in the Rocky Mountains. They conclude that WAG tapering and
reduced half cycles have proven to be effective and operational demands such as gas
recompression limits, high breakthrough rates can significantly alter expectations on a
tertiary project.

Harpole and Hallenbeck (1996) review the CO2 flood design and field performance
in the East Vacuum Grayburg San Andres Unit. The field has variable CO2 strategy
with the WAG ratio varying due to variability in geology requiring pattern balancing,
injection conformance and managing large swings in gas production rates. Friedmann
et al. (1997) describe the development and testing of a foam based gel technology for
conformance control in the Rangely Weber Unit in Colorado. The results showed that
injection well and reservoir responses were encouraging and CO2 breakthrough times
increased in some offset producers.

13
Jeschke et al. (2000) present the results of a study of the CO2 flood potential
of California oil reservoirs and the possible sources of CO2 for California EOR projects.
Goodyear et al. (2003) present the impact of compositional effects and operating tem-
perature on offshore CO2 injection performance for WAG and crestal injection schemes.
Temperature plays a significant role in offshore projects where parts of a reservoir may
have been cooled by sea water injection. Genetti et al. (2003) summarize the field
development history of Salt Creek field in the Permian Basin where successive imple-
mentation of infill drilling and improved recovery processes such as CO2 flooding have
increased to recovery to over 50 %. Wassmuth and Hodgins (2005) describe the applica-
tion of CO2 flooding in a vuggy/fractured carbonate formation in the Weyburn reservoir
in Saskatchewan. Additional improvement in oil recovery was obtained by implementing
conformance control methods such as foams, gels and gel-foams.

Holtz (2008) summarizes the application of CO2 flooding in different settings


of Gulf Coast sandstone reservoirs such as barrier/strandplain, submarine fan and flu-
vial/deltaic sandstones. Several different flood types have been applied, with the miscible
and the gravity stable types dominating. Sahin et al. (2008) document the best practices
in the design and operation of a full-field immiscible CO2 injection project in Bati Raman
oil field in Turkey. Different techniques to improve sweep such as gel treatment, chemi-
cally augmented water injection, injector-producer pattern modification and downspacing
have been tried in the field. Ghomian et al. (2008) investigate the economic incentives
needed to encourage oil companies to store CO2 in oil reservoirs. The results indicated
that sandstone reservoirs were in greater need for CO2 incentives compared to carbonate
reservoirs.

Sweatman et al. (2009) summarize how the oil and gas industry has achieved
great success in engineering the process to capture, transport and inject CO2 in EOR
projects. Merchant (2010) reviewed many CO2 projects all over the US and presents the
results when HCPV injected is much more than 80 % in a field. The study concluded that
14
incremental recovery could be as high as 26 % when slug sizes exceeded 190 %. Basry
et al. (2011) present the challenges faced, lessons learnt during the implementation,
operation and monitoring of the first miscible CO2 -EOR pilot project in a heterogenous
carbonate oil reservoir in Abu Dhabi, UAE. Rosman et al. (2011) present the results of
an immiscible WAG study for the Dulang, an offshore oilfield in Malaysia. The study
resulted in significant cost savings from the reduction in number of infill wells required,
increasing the overall value of the project.

Mollaei and Delshad (2011) developed a forecasting tool which provided time evo-
lutions of key WAG flood parameters such as average oil saturation, cumulative recovery,
oil cut, oil rate and volumetric sweep for miscible/immiscible WAG EOR processes. Vic-
tor (2011) presents the impact of geological structure on the sweep efficiency and reservoir
oil recovery in a small scale heterogenous model using CO2 injection.

15
Figure 2.1: Schematic of the CO2 flooding process (from NETL website).

Figure 2.2: Fluid movement in WAG [from Jarrell et al. (2002)]


16
Chapter 3

Description of the Reservoir Simulator - DOE-CO2

DOE-CO2 is an isothermal, three-dimensional, four-phase, compositional, Equa-


tion of State (EOS) simulator. The simulator (DOE-CO2 ) will be capable of simulating
various recovery processes (i.e., primary, secondary water-flooding, and miscible and im-
miscible gas flooding). This simulator is an advanced version of UTCOMP, an in-house
compositional gas flooding simulator developed at The University of Texas at Austin
(Chang, 1990). The following section briefly describes the features of UTCOMP and
later the advanced features in DOE-CO2 .

3.1 Description of UTCOMP

UTCOMP is an isothermal, three-dimensional, EOS compositional reservoir sim-


ulator. The formulation of UTCOMP is based on the volume-balanced approach (Acs et
al., 1985) with some modifications, which was detailed in the work of Chang et al. (1990).
The solution scheme is analogous to IMPES. Four-phase flow behavior can be modeled
using UTCOMP. The non-aqueous fluid properties are modeled using the Peng-Robinson
(PR) EOS (Peng and Robinson, 1978) or a modified version of the Redlich-Kwong (RK)
EOS (Turek et al., 1984).

For the discretization of the component mass-balance equation, a higher-order


finite difference method, as well as the conventional one-point upstream weighting scheme,
is used for numerical dispersion and grid orientation control. Physical dispersion is
modeled using the full dispersion tensor, and the elements of the dispersion tensor contain
contributions from two sources: molecular diffusion and mechanical dispersion. Either

17
constant bottom-hole pressure or constant flow rate well conditions can be specified for
both vertical and horizontal wells. Well rates and transmissibilities are treated explicitly.
A constant or variable time-stepping can be chosen.

A variable-width cross-section option is also available, which accommodates the


simulation of two-dimensional reservoir cross sections with radial flow near injection and
production wells and any arbitrary two-dimensional geometry between wells using either
pressure- or rate-specified boundary conditions.

Tracer capabilities can also be modeled using UTCOMP. Polymer flood capability
is added as an extension of the tracer option. Surfactant-enhanced remediation processes
for aquifers contaminated with dense non-aqueous phase liquids can as well as be modeled
using the dilute-surfactant option, as an extension of the tracer option.

To design and simulate processes using foam, Rossen and co-workers (Cheng et
al., 2000) have developed a foam model for both flow regimes where the gas relative
permeability is modified for the effect of foam. The foam model incorporated in UT-
COMP is based on the model of Cheng et al. (2000). Although foam alters both gas
relative permeability and viscosity in complex ways, for simplicity of computations, the
foam model used here assigns all of the reductions of gas mobility due to foam to the gas
relative permeability. The models of Vassenden and Holt (1998) and in the CMG-STARS
commercial reservoir simulator (Cheng et al. 2000) use a similar approach.

By utilizing the same compositional equations originally present in UTCOMP,


a black-oil model is developed. The black-oil model reads the PVT properties such as
formation volume factors for oil, gas and water as look-up tables. Other properties such
as solution gas in water and oil, relative permeability, capillary pressures, viscosity, and
total compressibility are also provided in the form of look-up tables.

The effect of asphaltene precipitation has been implemented using Nghiem’s model
(Nghiem et al., 1997). Also implemented in the simulator are models to compute the

18
change of porosity and permeability due to the asphaltene deposition.

The capability of evaluating CO2 storage in subsurface reservoirs, for designing


and operating geologic CO2 disposal systems has been added to UTCOMP (Vikas, 2002).
For the specific problem of CO2 sequestration in deep saline reservoirs, the interaction
between brine and CO2 has been added. For viscosity the critical volume of H2 O and for
density the computed Peng Robinson EOS density are modified for salinity. A salinity
correction for the solubility of CO2 in brine is also added to the existing implementation
of Henry’s law to simulate CO2 EOR.

3.2 Advanced Features in DOE-CO2

This is a special version developed under the DOE grant DE-FE0005952 entitled
Development of an Advanced Simulator to Model Mobility Control and Geomechanics
during CO2 Floods. This includes foam model for mobility control, three phase com-
positional relative permeability models with the hysteresis option, advanced three phase
hydrocarbon flash phase behavior, geomechanics module, corner point grid, visualization,
and efficient solvers. Some of these enhancements are discussed below.

3.2.1 Foam module

In addition to an empirical foam model (Cheng et al., 2000) a local-equilibrium


approximation method to the population balance modeling is also implemented as an op-
tion where the foam texture is approximated by placing foam-generation and coalescence
rates in equilibrium (Chen et al., 2010). This foam-texture approximation increases the
applicability of population balance models in the full-field three-dimensional studies.

3.2.2 Relative permeability module

Along with all the widely used relative permeability models, hysteresis models are
included for accurate modeling of the processes in which the saturation direction and/or
19
saturation path changes. The existing hysteresis models such as: Carlson’s two-phase
hysteresis model, Fayers and Matthews’ oil hysteresis model, and Larsen and Skauge’s
three-phase hysteresis model are included. Further, a new three-phase hysteresis model
(UTHYST) is developed and implemented (see Chapter 4 for complete description). A
list of the available relative permeability options and the hysteresis model options are
summarized in Tables 3.1 and 3.2 (Beygi et al., 2013; Beygi, 2014).

3.2.3 Fluid property and phase behavior module

In the simulator, the phases that may flow simultaneously are assumed to be (i)
aqueous phase, (ii) oleic, (iii) gas phase, and (iv) a secondary non-aqueous phase. For
the purpose of phase equilibrium calculations (flash calculation), only water is allowed in
the aqueous phase and does not involve in the flash calculations. A Gibbs stability test
in the model is included to perform a phase identification test to consistently label each
phase for subsequent property calculations such as relative permeability, viscosity, den-
sity, interfacial tension, and capillary pressure. The time step strategy is an IMPEC-type
method (implicit pressure and explicit concentration). The gridblock pressure is solved
first using the explicit dating of saturation-dependent terms. Subsequently, the material
balance equations are solved explicitly for the total concentration of each component.
The physical dispersion term is also included in the governing equations.

3.2.4 Geomechanics module

The compositional flow model is a multi-component, four-phase Equation of State


(EOS) compositional model (Chang, 1990). As part of this module, the fluid flow equa-
tion is coupled with porous media deformation. An isothermal condition for the coupled
fluid-solid system is assumed and no chemical reactions between fluid components and
solid medium are considered (Wang, 2014). The compositional flow model in the coupled
compositional flow and linear poro-elasticity model is similar to the procedure in Chen

20
et al. (1995) and Gai (2004).

Table 3.1: Summary of the available relative permeability options in DOE-CO2 (DOE-
CO2 user guide (2013); Beygi, 2014)

Table 3.2: Summary of the available hysteresis options in DOE-CO2 (DOE-CO2 user
guide (2013); Beygi, 2014)

21
Chapter 4

Application of a New Cycle Dependent Three Phase


Relative Permeability Model

Mobilizing stranded and trapped oil from the Residual Oil Zones (ROZs) is the
most significant means of producing oil from the transition zones. The performance
of CO2 injection in producing the oil is a strong function of relative permeability and
capillary pressure functions. The fact that relative permeabilities and capillary pressures
are functions of both saturation and its history, known as hysteresis, is well established. It
is well known that water injectivities often decrease substantially following injection of a
gas slug. The reason for the decrease is that gas is trapped by the water and hence reduces
the relative permeability of water. Hence estimating the trapped phase saturation and
thereby modeling the relative permeability of the phases including saturation direction
dependency because hysteresis behavior is critical for CO2 -EOR processes where flow
reversal takes place i.e. WAG.

4.1 Existing Three Phase Relative Permeability Models

There are numerous parameters that impact relative permeability: saturations,


saturation history (direction and path), wettability, pore structure, interfacial tension or
more generally the trapping number and oil/gas compositions. In three-phase flow, for a
specified phase with a fixed saturation, there are an infinite number of combinations of
the saturations for three-phase flow. So, measuring all these combinations is practically
infeasible. A conventional approach is to estimate three-phase flow relative permeability
based on measured two-phase relative permeabilities. Beygi et al. (2013) provides a

22
chronological list of some of the three-phase relative permeability models. Most of the
existing three phase relative permeability models have saturation-averaged interpolation
method in lieu of other mathematical models, viz. capillary models, statistical models,
and network models. Several investigators [Fayers and Matthews (1984); Baker (1988);
Delshad and Pope (1989); Oak (1990 and 1991); Skjveland and Kleppe (1992); Hustad
and Holt (1992); Balbinski et al. (1999); Pejic and Maini (2003); Spiteri et al. (2004);
Ahmadloo et al. (2009)] have identified various limitations of classical three-phase relative
permeability models such as Stone I and Stone II, especially for non-water wet conditions
or at low oil saturation.

4.2 Existing Hysteresis Models

There are numerous empirical models to capture saturation direction and path de-
pendency in relative permeability and in capillary pressure. A common feature proposed
by early two-phase hysteresis models [for example Carlson (1981)] is the assumption of
reversibility of the relative permeability curves where the primary imbibition curve is
representative of any subsequent drainage processes. However, this assumption is not
validated with experimental results for multi-cycle processes like WAG.

The three-phase hysteresis models, however, predict cycle-dependent relative per-


meability curves where each cycle has the scanning curves for the drainage and imbibition
cycles. These models predict an irreversible hysteresis behavior imposed by saturation
direction dependency. It is observed that both gas and water relative permeabilities
decrease in each WAG cycles at the same saturation. Larsen and Skauge (1998) pub-
lished an empirical model based on their immiscible- and miscible- WAG experiments.
Their model relaxes the assumption of the scanning curves reversibility resulting in a
cycle-dependent relative permeability model for the gas phase. They also proposed a
three-phase water relative permeability model for water-wet media based on saturation-
averaged interpolation between primary and secondary waterflood relative permeabilities.
23
For oil wet media, the water entrapment by gas is only considered in the Egermann three-
phase hysteresis model (Egermann et al., 2000). Oil relative permeability in subsequent
WAG cycles increases because trapped oil saturation decreases with a higher likelihood
of simultaneous water and gas trapping.

A major deficiency of the existing models is in the limitation in accounting for


hysteresis when one phase disappears/appears because of mass transfer between the
phases. Moreover, the currently available three-phase hysteresis models, by their nature,
add complexity to compositional simulations.

4.3 Description of the New Cycle Dependent Three Phase Rel-


ative Permeability Model
4.3.1 Relative permeability model (UTKR3P)

A three-phase relative permeability model should capture the essential physical


behavior and be as simple as possible to be implemented in compositional reservoir
simulators. The proposed model by Beygi et al. (2013) has generalized the approach
that Jerauld (1997) used for the gas phase to all of the phases. For any phase j, the
relative permeability is given by
C1j
0 (1 + C2j ) S̄j
krj = krj   j = 1, 2 or 3 (4.1)
C1j 1+ C1
1 + C2j S̄j 2j

where the normalised equation is defined as

Sj − min (Sj , Sjt )


S¯j = (4.2)
1 − 3i=1 Sir3P
P

In the following equations, the phase index can take the generic values of 1, 2 and
3 or ‘j’, ‘m’ and ‘l’. Equation 4.1 results in an S-shape curve which can be transformed
to an exponential Corey-type model based on the exponents C1 and C2 . At small phase
saturations, the model approaches Corey-type behavior leading to good agreement with

24
experimental results. The existence of a continuous oil layer, for example, in intermediate-
wet rocks can be modeled effectively using the proposed approach. At high saturations,
however, the second term in the denominator of Eq. 4.1 dominates and dampens the
sharp increase in relative permeability. Multiple relative permeability curves can be
generated for each phase based on the saturation path. The power exponent of the
second term in the denominator of Equation 4.1 is chosen to be the maximum acceptable
value resulting in a non-negative slope of the relative permeability curve.

The trapped saturation Sjt in the normalized saturation (Eq. 4.2) is computed
3P
based on the hysteresis model and Sjt ∈ [Sjc , Sjr ]. Residual saturations are modeled
using a modified correlation proposed by Yuan and Pope (2012). The definition of residual
saturation has been modified from the original model to an effective saturation:

3P 2P
  
Sjr = min Sj , max Sjc , Sjr (1 − bj (Sm − Smc ) (Sl − Slc )) m 6= l 6= j (4.3)

The fitting parameter ‘b’ for each phase depends on rock wettability and phase
composition. Figure 4.1 depicts the impact of b2 on the curvature of the phase 2 (oil)
isoperms. Large positive values of b2 result in convex isoperms while negative values of
‘b’ result in concave isoperms. Many other models do not have this essential flexibility.
3P ∗
The two-phase residual saturation Sjr assumed either as the minimum of Sjrm
∗ ∗ ∗
and Sjrl or estimated by a saturation-weighted interpolation between Sjrm and Sjrl (use
Equation 4.4 for linear interpolation). If the optional compositional consistency method
is not used, the latter method can mitigate discontinuities associated with the appearance
or disappearance of a phase during compositional simulations, as noted by Fayers (1989)
and Balbinski et al. (1999).

Three-phase parameters, i.e. endpoint relative permeability kr0 and curvatures C1


and C2 , are estimated by a linear saturation-weighted interpolation between the two-
phase parameters.
(Sm − Smc ) Fjm + (Sl − Slc ) Fjl
Fj = (4.4)
1 − Sj − Smc − Slc
25
where F represents any of the two phase parameters (kr0 , C1 , or C2 ).

4.3.2 Three Phase Hysteresis model (UTHYST)

The UTHYST model (Beygi et al., 2013) also offers a simple approach to calculate
irreversible hysteresis behavior for any phase. The hysteretic relative permeability uses
the following equation in multi-cycle processes where the impact of saturation direction
and saturation path is considered.

krj = krj [Sj , (Sjt )]n (4.5)

This hysteresis model can be used with any relative permeability model that
includes the trapped phase saturation. The monotonically increasing trapped saturation
for different phases is preserved even in miscible multi-cycle processes as follows:
 α
Smc j
(Sjt )n = (Sjt )E,n−1 + ∆Sjt I
(4.6)
Sm

where (Sjt )E,n−1 is the trapped saturation corresponding to the end of previous cycle,
∆Sjt is the trapped saturation for the current cycle, and set of parentheses introduces the
effect of conjugate phase saturation at the start of the current cycle on phase entrapment.
Trapped saturation depends on the maximum phase saturation and the dynamic Land’s
coefficient and is calculated using a modification to Jerauld’s model (Jerauld, 1997).

 max  max Sj , Sjmax − Sjc
Sjt Sj , Lj = Sjc + (4.7)
 max
 1+ Lβj
1 + Lj max Sj , Sj − Sjc j

The trapped saturation in different cycles is calculated as follows.

(Sjt )E,n−1 = Sjt Sjmax , Lj E,n−1


 
(4.8)
h i
∆Sjt = Sjt Sjmax − (Sj )E,n−1 , (Lj )n (4.9)

The trapping behavior during multi-cycle WAG processes may not be reproduced
using a constant Lands coefficient. Element et al. (2003) used an in-situ saturation
26
monitoring technique to mitigate laboratory artifacts within their experiments on multi-
cycle water-alternate-CO2 floods in the aged Berea sandstone. They showed that Land’s
coefficient varies during different hysteresis cycles and that none of available hysteresis
models can accurately predict the observed three-phase hysteresis. Spiteri and Juanes
(2006) analyzed the two-phase data of Oak (1991) with a wide range of permeability and
found that Land’s coefficient depends on rock permeability and the pair of fluids used.
They attributed this behavior to the difference in advancing contact angle for different
fluids. Shahverdi et al. (2011) carried out miscible WAG floods using water- and mixed-
wet cores and stated that using two-phase Land’s coefficient to calculate trapped gas
saturation results in poor agreement with the measured data. The UTHYST model,
thereby, addresses a cycle-dependent Land’s coefficient in a multi-cycle process.

1 1
Lj = ∗
− P3 (4.10)
Sjr − Sjc 1 − i=1 Sic

where Sjr is the multiphase residual saturation corrected for compositional changes at
high trapping number where the trapped saturation decreases due to mass transfer be-
tween the trapped and the displacing fluids (Beygi et al. 2013). The denominator of the
second part of Eq. 4.10 refers to the maximum phase saturation based on the minimum
attainable saturation of the other two phases flowing with phase ‘j’ (Smc and Slc ).

4.4 Validation with Experimental Data

In general, the UTKR3P model captures the trend in relative permeabilities for
water- and oil-wet rocks even in the saturation ranges where phase relative permeability
is a function of its own saturation. To quantify the effectiveness of the UTKR3P model
compared to measured three-phase data, we used the standard error of estimate (SEE)
as s
calc expr
2
krj − krj
SEEj = (4.11)
Nj

27
where the superscripts calc and expr denote calculated and measured relative perme-
ability and ‘Nj ’ is number of experimental data points for each phase ‘j’. Figure 4.2
shows the trend in measured and calculated relative permeability for different phases.
There is a good agreement between the measured and calculated relative permeabilities
considering only one parameter (bj ) is used and a very narrow range of saturations was
seen in the experiment. Matching ‘b’ parameters and the SEEs are given for each phase
in Figure 4.2.

4.5 Sensitivity Analysis of Various Parameters

In this section we discuss the simulation results for a synthetic, two-dimensional


reservoir. The effect of the various parameters in UTKR3P and UTHYST on the flow be-
havior is depicted. The simulations have been performed using DOE-CO2 compositional
gas flooding simulator. A 2D homogenous reservoir was constructed replicating the high-
permeability, intermediate-wet sample (Sample 15) of Oak (1991). Eight-components
were used to represent light oil with Peng-Robinson equation-of-state. The simulation
model has two wells as a quarter of five spot pattern with injector and producer located
at the opposite sides of the reservoir model. A WAG injection scheme is employed for
3 cycles with each cycle consisting of 15 days of CO2 injection followed by 15 days of
water injection. The cumulative injection after 90 days is around 1.0 HCPV. Both the
injection and the production wells are pressure constrained and the reservoir pressure is
maintained at the initial pressure throughout the simulation. A summary of the reservoir
properties is given in Table 4.1. The injection and production bottomhole pressures are
maintained at 2000 psi and 1000 psi respectively.

The values of the UTKR3P model parameters for the base case are given in Table
4.2. For the sensitivity cases, only the varied parameters are summarized in Table 4.3.
In all the analyses below, since the UTKR3P model is phase independent, the phases
have been numbered ‘1’, ‘2’, and ‘3’. For these simulations, however, they represent
28
water, oil and gas, respectively. The phase relative permeabilities for the grid cell closest
to the injector have been plotted in all the figures. The impact of the saturation path
dependency on the relative permeability has been incorporated by a couple of parameter
sets in the UTKR3P model. One is the parameter set of kr0 , C1 and C2 for each of the
two-phase combinations denoted as 12, 13, 21, 23, 31 and 32. The second is the ‘b’
parameter for each phase (b1 , b2 , b3 ), which relates the two-phase residual saturations
to the three-phase values. The impact of each set of parameters is studied independently.

4.5.1 Sensitivity to parameters kr0 , C1 and C2

To simplify the analysis, only the two phase end point relative permeability has
been varied to show the impact of saturation path dependency on three phase relative
permeabilities. The model Case 2A, as described in Tables 4.2 and 4.3 are constructed
for this purpose. Figure 4.3 from Case 2A shows the saturation-averaged three phase
relative permeability values for the hysteretic phase and discretized by time for the first
WAG cycle (30 days). Also plotted in Figure 4.3 are the corresponding two phase curves.
Tracking the gas saturation close to the wellbore shows that from the start of gas injection,
three phase flow occurs and the gas saturation in the gridlock increases and reaches a
maximum within the first 2.5 days. The water saturation decreases and reaches the
connate water saturation as shown by the gas relative permeability following the two-
phase gas oil curve and producing most of the oil during 2.5 - 15 days. Thereafter, as
water is injected, the gas saturation starts to decrease and follows the two phase gas-water
curve until the gas saturation goes to zero at 30 days. This analysis shows how a simple
difference in end point permeability of a phase with respect to the two different phases
can manifest the saturation path dependency. The capability of the model to switch
seamlessly from two phases to three-phase and back to two-phase prediction is depicted
by the simulation run. This analysis is for showing the saturation path dependency with
different end-point values. Once the other parameters (C1 and C2 ) are included in the

29
analysis, the impact becomes more pronounced.

4.5.2 Sensitivity to parameters ‘b’

Figure 4.4 shows the results from the base case (Case 1A) and Case 2B. The only
difference in the models is the ‘b’ parameter, which is non-zero in Case 2B for one of the
phases. The ‘b’ parameter impacts the three phase residual saturations which in-turn
modifies the three phase relative permeability. The first cycle overlaps for both the cases
and the relative permeability shifts slightly in the subsequent cycles as the saturation
direction is reversed.

4.5.3 Sensitivity to trapped gas phase

Trapped gas is one of the major contributors of hysteretic behavior in WAG


injection. In this model (Case 3), the hysteretic model (UTHYST) has been applied.
The difference between this model (Case 3) and the base case (Case 1A) is the trapped
gas which is taken into account in Case 3. By following the red vertical line in Figure 4.5,
it is clear that by going through the increasing and decreasing phase saturation cycles,
the relative permeability decreases because of the phase trapping. Figure 4.6 shows
the phase saturation and the trapped saturation as a function of time. The trapped
saturation increases during the decreasing saturation cycle due to capillary trapping at
the pore level. The trapped saturation remains constant during the increasing saturation
cycle. This hysteretic behavior is also one of the major causes for the reduced water
injectivity seen in field cases of WAG injection.

4.6 Summary and Conclusions

A new three-phase relative permeability model (UTKR3P) incorporating hystere-


sis and compositional dependency has been developed, tested against experimental data
for different wettability conditions, and implemented into a compositional reservoir sim-
30
ulator. This model is significantly more flexible than previously published three phase
models.

The hysteresis model (UTHYST) is an extension of the Land’s trapping model but
with a dynamic coefficient. The model overcomes some of the limitations of published
three-phase hysteresis models for non-water wet rocks and mitigates some of the com-
plexity and difficulties associated with commonly applied models in numerical reservoir
simulators. The model is validated using multi-cyclic three-phase water-alternating-gas
experimental data for non-water wet rocks.

31
Table 4.1: Summary of reservoir properties used in 2D simulation cases

Table 4.2: Relative permeability properties for the base case (Case 1A)

Table 4.3: Relative permeability properties that are varied for various cases

32
Figure 4.1: Effect of parameter ‘b2 ’ on the curvature of phase 2 iso-perms where phase 2
refers to oil phase.

Figure 4.2: Calculated vs. experimental relative permeability values (Sample15 Oak
1991) where plots 4.2a, 4.2b, 4.2c show the match for phases 1(water), 2(oil) and 3(gas).

33
Figure 4.3: UTKR3P model with path dependency (Kr031 6= Kr032). Here phase 3 can
refer to gas phase.

Figure 4.4: UTKR3P model path dependency (bj 6= 0)

34
Figure 4.5: Phase-3 relative permeability using UTKR3P with UTHYST for the effect
of trapped phase-3

Figure 4.6: Saturation and trapped saturation for Phase 3 in two WAG cycles

35
Chapter 5

Application of a Mechanistic Foam Model to Foam


Assisted CO2 floods

CO2 flooding processes have experienced poor sweep efficiency despite the favor-
able characteristics of CO2 . Foam is one of the ways to mitigate this problem. The basic
idea is that the mobility of a gas flowing through a porous medium is reduced when it is
dispersed within a liquid as foam. The benefit of foam will be much higher if foam can be
formed in the reservoir in situ rather than injecting foam which only has a near wellbore
impact. Foam can be generated in-situ by injected slugs of water mixed with surfactant
and slugs of gas either in sequence or simultaneously. The surfactant injected along with
water stabilizes the lamellae thus trapping gas and reducing the gas permeability. Since
gas mobility is reduced in these zones, gas is diverted to other and unswept zones. The
main effect of foam is to suppress frontal instability and decrease gas mobility farther
from the wellbore because of in-situ generation.

5.1 Existing Foam Models

From literature, the approaches for modeling foam can be classified into two ma-
jor types. The first is the semi-empirical approach where gas mobility in the presence
of foam is a function of gas rate, oil saturation, surfactant concentration and other fac-
tors. These models modify gas mobility by a mobility reduction factor computed as a
function of the above listed parameters. Some of the models under this category include
Patzek and Myhill (1989), Fisher et al. (1990), Zhou and Rossen (1995), Rossen (1996).
The second approach is more mechanistic and thus complex and uses the population

36
balance model to incorporate foam texture and foam generation and coalescence into the
expression for gas mobility. These models track the bubble density changes caused by
coalescence, generation, and transport. Several investigators have applied this approach
(Patzek (1988), Chang et al. (1990), Friedmann et al. (1991), Fergui et al. (1995),
Kovscek et al. (1995), Chen et al. (2010)). The model parameters for these population
balance methods need to be obtained by fitting coreflood foam tests. Bertin et al. (1998)
proposed a simplified approach referred to as local equilibrium where the bubble pop-
ulation correlation is expressed as a function of porosity, permeability, gas saturation,
limiting capillary pressure, and flowing foam fraction. The local equilibrium approach is
a good approach of mechanistic foam simulation at a reduced computational cost.

The foam model based on the work of Cheng et al. (2000) is also incorporated
into DOE-CO2 . In this model (also known as Pc∗ model), foam forms if (i) surfactant is
present and its concentration is above some threshold value (Cs∗ ) and (ii) water saturation
exceeds a threshold value of Sw∗ . The high quality or coalescence regime corresponds to
Sw = Sw∗ . If Sw > Sw∗ , foam reduces gas mobility by a large constant factor and this
corresponds to the low quality regime. The foam model is described as follows

If Sw < Sw∗ − ε or Cs < Cs∗ then f


krg = krg (5.1)
If Sw∗ − ε < Sw < Sw∗ + ε and Cs < Cs∗ then
f krg (5.2)
krg = ∗ +ε)
(R−1)(Sw −Sw
1+ 2ε
krg
If Sw > Sw∗ + ε or Cs ≥ Cs∗ then f
krg = (5.3)
R
where Cs is the surfactant concentration, Cs∗ is the threshold surfactant concentration
f
for foam formulation, krg is the effective gas relative permeability modified for foam, R is
the maximum foam mobility reduction factor and Sw∗ is a threshold water concentration.
The foam parameter R can be modified according to gas flow rate to allow for shear
thinning behavior of foam in low quality regime as follows
 σ−1
ug
R = Rref (5.4)
ug,ref
37
Here ug is gas volumetric flux, Rref is the values of R at a reference gas volumetric flux,
and σ is the conventional power-law exponent. For Newtonian foam behavior, σ = 1,
and for shear thinning behavior, σ < 1.

5.2 Description of the Mechanistic Foam Model

The local equilibrium model by Chen et al. (2010) has been implemented in
DOE-CO2 compositional gas flooding simulator. The model can be described as follows
 net generation term

∂ ∂
z }| {
[Φ (Sf nf + St nt )] + (uf nf ) = ΦSg k1 |v~w | |v~f |1/3 − k−1 |v~f | nf  + Qb (5.5)
 
|∂t {z } |∂x {z } | {z
generation
} | {z }
coalescence
|{z}
source/sink
accumulation term advection

where φ is the porosity, S is the phase saturation, n is the bubble density, u is the Darcy
velocity, v is the interstitial velocity, k is the rate coefficient, Qb is a source/sink term for
bubble and the subscripts f and t represent flowing and trapped foam. The subscripts
w and g represent water and gas, 1 for foam generation and -1 for foam coalescence.

The generation rate constant k1 is defined as


h  n ω i
f
k1 = k10 1 − (5.6)
n∗

where ω is a constant determining the shape of inverse proportionality of foam generating


sites to pre-existing gas bubbles and n∗ is an upper limit for the concentration of foam
bubbles related to pore size.

The coalescence rate constant k−1 varies significantly with the local capillary pres-
sure and surfactant formulation as
 2
0 Pc
k−1 = k−1 (5.7)
Pc∗ − Pc
0
where scaling factor k−1 is assumed to be a constant and Pc∗ is the limiting capillary
pressure for foam coalescence.

38
Experimental investigation of various surfactants suggests the following functional
form for Pc∗ versus surfactant concentration (Khatib et al., 1988):
 
∗ ∗ Cs
Pc = Pc,max tanh (5.8)
Cs0

where Pc,max is a maximum limiting value for Pc∗ and Cs0 is a reference surfactant con-
centration for strong net foam generation.

For flowing foam, the gas viscosity is modified as (Khatib et al., 1988; Falls et al.,
1988):
αnf
µf = µg + (5.9)
|vf |1/3
where µg is the viscosity of gas without foam, vf is the foam velocity and α is the
proportionality constant dependent primarily on the surfactant solution. If the net foam
generation is zero, the foam generation and coalescence are in equilibrium. Thus Equation
5.5 can be written as
net generation term
z
 }| {

k1 1 − nf
 0h  ω i
1/3


|v~w | |v~f | − k −1 | v
~f | nf
=0 (5.10)

| n {z | {z
} coalescence term} 
generation term

And finally Equation 5.10 is rearranged as:

n∗ k−1 |v~f |2/3


ω
ω
nωf + 0
nf − n∗ = 0 (5.11)
k1 |v~w |

Solving for this equation yields bubble density and with the above equations, the
gas viscosity and hence its mobility can be computed.

5.3 Validation with Experimental Data

Experimental validation of the foam local equilibrium model with simulation re-
sults has been done by Naderi et al. (2013). They compared experimentally measured
values of cumulative oil production Qoil (t) and pressure drop (δP/L) with simulation for
39
foam experiments with CO2 and N2 . The corresponding results for N2 foam are shown in
Figures 5.1 and 5.2. The results show a fair match between experimental and simulated
values validating the local equilibrium approximation model.

5.4 Metrics for Foam Performance Quantification

Two metrics which are used to quantify the performance of any mobility control
technique for gas flooding processes are as follows:
Gas utilization factor: It is defined as the amount of gas injected for every barrel of
incremental oil produced due to gas injected. It is measured in units of Mscf/bbl. It
quantifies the efficiency of the gas displacement process. The lower the utilization factor,
the better and more cost effective is the displacement process. This is not computed for
a simultaneous injection scheme with water and gas injected at the same time.

Cumulative gas injected (M scf )


Gas utilization f actor = (5.12)
Incremental oil produced (bbl)

Recycling fraction: It is defined as the ratio of the cumulative injectant (CO2 ) produced
to the cumulative injected gas (CO2 ). This has no units and signifies the amount of gas
injected that is recycled. The lower the recycling fraction, less gas is recycled and the
displacement is more efficient.

Cumulative injectant (CO2 ) produced


Recycling f raction = (5.13)
Cumulative gas (CO2 ) injected

The above two metrics will be used subsequently in all the simulation results.

5.5 Application to Numerical Simulation Models

A layer cake synthetic model has been constructed for the purpose of studying
the impact of foam on simulations. The model is a 2D reservoir discretized into 60 x 1 x
10 gridblocks with a constant porosity of 0.2. A heterogeneous permeability profile has
been created for the layer cake model specifically for making gas flooding unfavorable.
40
Gas has a tendency to override due to its lower density and sweep the shallower layers
better, so high permeability layers are included at the top to make the natural sweep
unfavorable. The permeability profile is shown in Figure 5.3. A summary of the reservoir
properties is given in Table 5.1. The relative permeabilities have been defined with the
UTKR3P model as described in Section 4.3 and the parameter values are shown in Table
5.2.

An eight component PVT equation of state (EOS) model has been assigned to the
layer cake model. This PVT model is from one of the field scale model used for simulations
discussed in Chapter 6. The reservoir oil is a relatively light with ◦ AP I=30; bubble
point pressure (Pbp )=1091 psia; solution gas ratio (Rs =270 scf/stb) and Temperature
(T) =220◦ F . An Equation-of-state model has been obtained for the reservoir oil based
on the detailed PVT study and Peng-Robinson equation of state with Peneloux correction
was employed. The minimum miscibility pressure (MMP) is around 2850-2900 psi with
CO2 . The composition of the oil is shown in Table 5.3. The endpoint mobility ratio is
about 2 for the water-oil and about 37 for the CO2 -oil indicating the displacement will
be unfavorable. The viscosities of CO2 , oil and water viscosities are around 0.06, 1.9 and
0.3 cp at the initial reservoir conditions.

The mechanistic foam model as described in Section 5.2 is used for modeling foam
and the values of all the parameters are shown in Table 5.4. An injector and a producer
well are placed at two sides of the model along X axis (injector at gridblock of X=1 and
producer at gridblock of X=60). The wells are completed over the entire model thickness.

5.5.1 Foam Assisted WAG flood

To evaluate the impact of a foam assisted WAG scheme on the flood performance,
two cases are considered. The first case is a WAG scheme in which a slug of CO2 is
injected for 2 months followed by a slug of water for 2 months. The second case is a
foam assisted WAG (FAWAG) scheme in which a slug of CO2 is injected for 2 months
41
followed by a slug of surfactant and water for 2 months. The cycle is repeated 3 times in
both cases. The injector and the producer are bottomhole pressure constrained at 6500
psia and 600 psia respectively.

5.5.1.1 Results and discussion

Comparison of FAWAG and WAG results shows the effluent CO2 concentration
is almost the same in the first cycle (see Figures 5.4 and 5.5). The effluent CO2 concen-
tration in the FAWAG model is lower in the second and the third cycles (120-360 days).
This is the effect of foam which is generated after the slug of water with surfactant is
injected in the first cycle. The FAWAG case also has much lower cumulative gas injected,
lower cumulative gas produced, lower cumulative water injected and lower CO2 utiliza-
tion factors compared to the WAG model with almost the same oil recovery as shown in
Table 5.5. Figure 5.6 shows that the FAWAG model has lower recycling fraction. Figure
5.7 shows that the FAWAG model has almost the same oil recovery with much less pore
volumes of fluid injected . The mobility reduction factor continuously varies with time
for very grid block but the range is between 1 and 500. The cumulative gas injected and
the cumulative water injected for the FAWAG and WAG model are plotted in Figures
5.8 and 5.9. The FAWAG model has lower gas injected due to the effect of foam and
lower water injected as the trapped gas reduces the water injectivity too.

5.5.2 Simultaneous injection of gas and water with surfactant

In a simultaneous injection scheme, the gas and the water are injected simultane-
ously displacing the oil. Two cases are considered, the first one in which gas and water
are injected for a period of 12 months. The second case in which water is mixed with
surfactant and simultaneously injected with gas for 12 months. The injector is rate con-
strained with a gas injection rate of 250 Mscf/d and water injection rate of 200 bbl/d.
The producer is pressure constrained at 600 psia.

42
5.5.2.1 Results and discussion

Comparison of the results between the simulations with and without foam shows
that foam has a lower effluent CO2 concentration and later breakthrough at 210 days
compared to 145 days for the case without foam (see Figure 5.10 and 5.11). The foam
case has slightly higher oil recovery but much lower recycling fraction as shown in Table
5.6. Figure 5.12 shows the trend of recycling fraction for both cases. Figure 5.13 shows
that the foam has higher oil recovery at the same total gas/water pore volumes injected.
Since the wells are constrained by the rate, the impact of foam will be shown in the
bottom hole pressure (BHP) of the injectors. Figure 5.14 shows that the foam case has
higher BHP as there is more resistance to gas injection due to foam generation and an
increase in gas viscosity (lower gas mobility).

5.5.3 Sensitivities to key foam parameters

In this section, we repeat the simulations for different values of generation rate
coefficient, coalescence rate coefficient, limiting capillary pressure, and effective viscosity
coefficient. The aim is to see which of these parameters have a stronger impact on the
foam performance (recycling fraction). The base case is the simultaneous injection of
gas and water with surfactant. Sensitivity simulations are performed by running cases
individually with smaller values for each of the above four parameters as shown in Table
5.7.

5.5.3.1 Results and discussion

For three of the parameters namely generation rate coefficient, limiting capillary
pressure, and effective viscosity coefficient, a smaller value leads to a weaker foam as
evident from the higher recycling ratios as seen in Table 5.8 and Figures 5.15 through
5.18. For the coalescence coefficient, the recycling ratio is almost the same as the base
case. It can also be inferred that generation rate coefficient, limiting capillary pressure

43
and effective viscosity coefficient have more impact on foam strength compared to the
coalescence rate coefficient for the range of variables used in this study.

5.6 Summary and Conclusions

The population balance foam model by Chen et al. (2010) has been implemented
into DOE-CO2 , the in-house compositional gas flooding simulator. Experimental val-
idation for N2 and CO2 foams with simulation data shows that the local equilibrium
approximation works well for modeling foam texture (Naderi et al., 2013). Simulations
with layer cake models conclude that foam has a significant benefit over WAG and si-
multaneous injection schemes for the cases with unfavorable permeability heterogeneity.

44
Table 5.1: Summary of reservoir properties used in the synthetic case

Table 5.2: Summary of UTKR3P relative permeability parameters used in the synthetic
case

Table 5.3: Oil composition for fluid used in the synthetic model

45
Table 5.4: Foam parameter values in the local equilibrium foam model

Table 5.5: Comparison of FAWAG and WAG simulation results

Table 5.6: Comparison of simultaneous injection - foam and no foam results

Table 5.7: Parameter values for sensitivity cases

46
Table 5.8: Results of sensitivity simulations

Figure 5.1: Cumulative oil recovery for N2 foam - experimental and simulation results
(Naderi et al., 2013)

47
Figure 5.2: Pressure drop for N2 foam vs. pore volumes of fluid injected experimental
and simulation results (Naderi et al., 2013)

Figure 5.3: Permeability (mD) profile for the layer cake synthetic model

48
Figure 5.4: Effluent volume fraction for FAWAG case

Figure 5.5: Effluent volume fraction for WAG case

49
Figure 5.6: Recycling fraction - FAWAG vs. WAG

Figure 5.7: Oil recovery (fraction of OOIP) as a function of total injected fluid pore
volumes - FAWAG vs. WAG

50
Figure 5.8: Cumulative gas injected - FAWAG vs. WAG

Figure 5.9: Cumulative water injected - FAWAG vs. WAG

51
Figure 5.10: Effluent volume fraction for simultaneous injection with foam

Figure 5.11: Effluent volume fraction for simultaneous injection without foam

52
Figure 5.12: Recycling fraction for simultaneous injection - foam vs. no foam

Figure 5.13: Oil recovery (fraction of OOIP) for simultaneous injection - foam vs. no
foam

53
Figure 5.14: BHP of the injector for simultaneous injection - foam vs. without foam

Figure 5.15: Recycling fraction for smaller generation rate coefficient compared to the
base case

54
Figure 5.16: Recycling fraction for lower coalescence rate coefficient compared to the
base case

Figure 5.17: Recycling fraction for smaller viscosity coefficient compared to the base case

55
Figure 5.18: Recycling fraction for smaller limiting capillary pressure compared to the
base case

56
Chapter 6

Field-Scale Case Study - Field C

In this chapter, the results of compositional simulation for a field-scale model


are presented. The aim of this study is to compare the EOR processes of surfactant
flooding and CO2 flooding and suggest the best development scheme for the field. The
study also focusses on CO2 flooding in detail comparing a continuous gas flooding with
a WAG injection. Sensitivities are also performed on well production constraints, half
cycle length in WAG, and hysteresis behavior. The results from DOE-CO2 simulator are
also compared with those from EclipseT M . Surfactant flooding simulation for the same
field is performed using UTCHEM, an in-house chemical flooding simulator. Finally, the
surfactant flooding is compared to CO2 flooding and recommendations are given as to
which EOR process is better for the field.

6.1 Field Introduction

The field under consideration, Reservoir C, is a sector model based on an actual


field with a 15 year history of primary and waterflooding production. The reservoir is
undergoing primary depletion through horizontal producers and a secondary recovery
project area in the field is being developed as a line-drive waterflood, utilizing horizontal
injectors drilled parallel and between the existing producers. CO2 flooding and surfactant
flooding are being considered and reservoir simulation studies are conducted to test the
feasibility of these two EOR processes.

The reservoir lithology is carbonate with a thickness of only 12 ft at a depth of


around 6000 ft. The average permeability of the reservoir is 7 mD with an average poros-

57
ity of 15%. The ratio Kv/Kh is 0.25 and the reservoir temperature is 220◦ F . The initial
pressure is 3900 psia at the end of 15 years of primary production and waterflooding.
The field currently has 6 producers and 12 injectors; four of which were producers at
the start of the field production and recompleted as injectors. The deviated wells are in
a line drive pattern of 3 injector lines and 2 producer lines intersecting both geological
layers. The average initial water saturation in the reservoir is 0.55. A summary of the
reservoir properties is shown in Table 6.1.

6.2 Model Introduction

A 3D Cartesian model is constructed discretized into 198x83x4 gridblocks and


incorporating the gradual dipping structure of the reservoir. The size of the gridblock
is 200 ft and 50 ft in the X and Y direction respectively with variable thickness for the
4 numerical layers. The permeability, porosity, and oil saturation maps are exported
directly from the simulation in Eclipse provided by the oil company. These maps are
also shown in Figures 6.1 through 6.3. The average permeability is low (7 mD) and the
pay thickness is small (12 ft). So the effect of gravity override for the injected gas will
be minimal.

The reservoir fluid is a relatively light oil with ◦ AP I=30; Pbp =1091psia; Rs =270
scf/stb and T=220◦ F . An 8 component Equation-of-state (EOS) model has been ob-
tained for the reservoir oil based on the detailed PVT study and Peng-Robinson equa-
tion of state with Peneloux correction was employed. The MMP is around 2850-2900
psi with CO2 . The composition of the oil is shown in Table 6.2. The endpoint mobility
ratio is about 2 for the water-oil fluid pair and about 37 for the CO2 -oil pair indicat-
ing the displacement will be unfavorable especially for CO2 displacing oil. Thus, some
type of conformance control to have a stable displacement of oil such as WAG might be
beneficial.

58
There are a total of 18 active wells in the sector model with 6 producers and 12
injectors. All the wells are completed as long deviated wells. The well skin has been used
in the form of a PI multiplier in each wellblock based on the primary and waterflood field
history match. An overview of the wells and their locations are shown in Figures 6.4 and
6.5.

For the base case, Stone-II relative permeability model is used with the parameters
as per the 4 relative permeability curved measured and provided in the core analysis
report. The parameters of Stone-II model are varied to obtain a good match with the
measured data. The original oil-water and gas-liquid curves are shown in Figures 6.6 and
6.7.

6.3 Waterflood Simulation using Eclipse, UTCHEM and DOE-


CO2

This study is to compare the waterflood recovery between the various simulators
used for this field study. These simulators are different in their functionality and applica-
tion and hence there needs to be a common basis to compare the results. The first step is
to compare the results of waterflooding using these three reservoir simulators before the
more complex surfactant and CO2 floods. The waterflood simulations are compared for
a period of 5 years. One major difference in these simulators is the handling of relative
permeabilities. UTCHEM is used for chemical flooding where there is no capability for
gas PVT (i.e Bo , Rs etc.). Therefore, oil recovery and water rates are based on the
oil-water relative permeability curves. DOE-CO2 and Eclipse simulators are composi-
tional simulators and hence use both oil-water and oil-gas relative permeability curves.
Eclipse scales the relative permeability tables to arrive at identical relative permeabilities
at maximum saturations of the given phase (i.e. connate/critical saturations of the other
phases). The relative permeabilities were varied until similar recoveries were obtained
with different simulators using two-phase relative permeability tables provided in Eclipse
59
as the base. Stone-II three phase relative permeability model was used in Eclipse and
DOE-CO2 simulators. UTCHEM does not model the gas and Corey relative permeabil-
ity model is the only option available. The relative permeabilities used in the different
simulators are shown in Figures 6.6 and 6.7. The oil recovery, pressure, water injection
rate, and water production rate from the benchmarked models are shown in Figures 6.8
through 6.11. The results show a reasonably good match.

6.4 Selection of a Gas Injection Scheme: WAG vs Continuous


Flood

This study is to compare the two types of gas flooding namely continuous CO2 gas
flood and WAG. In the continuous CO2 flood, water was injected for 1 year followed by a
continuous CO2 flood for 4 years. In the WAG scheme, a slug of CO2 is injected for 180
days followed by a slug of water for 180 days. This cycle is repeated 5 times for a total
of 5 years. The CO2 injection is rate constrained in both the models and the maximum
rate constraint is 3800 Mscf/day per well. This is corresponding to a 4% HCPV/month
of injected fluid. Water injection wells and the production wells are BHP constrained.

The results from both the models are summarized in Table 6.3. The effluent CO2
concentration is higher in the continuous CO2 model compared to the WAG model (see
Figures 6.12 and 6.13). The WAG model has a higher recycling ratio than the continuous
CO2 model initially but reduces at later times (see Figure 6.14). The oil recovery is clearly
higher for the WAG injection compared to the continuous CO2 injection with the same
pore volumes of fluid injected (see Figure 6.15). Also, from an economic point of view the
WAG injection is more efficient since less volume of gas is injected with a higher volume
of water injected for a higher oil recovery. Ultimately, the selection of the gas injection
scheme depends on the economics, the production philosophy of the oil company and
the operational difficulties amongst other factors. But as an initial estimate, the WAG
process is bound to be more economic where a cheaper injectant (water) replaces a costlier
60
injectant (CO2 ).

6.5 Sensitivity: Production Constraints in WAG

After establishing WAG as an improved CO2 flooding process, this study is to


test the sensitivities of production constraints mainly the injection rate of CO2 . Three
simulation cases are set up and compared, one with a maximum gas injection rate of
2280 Mscf/day per well, another with a BHP constrained gas injection in addition to the
WAG case in the earlier section with a maximum injection rate of 3800 Mscf/day per
well. The production wells are pressure constrained at different pressures ranging from
600-2300 psia. The injection BHP for the wells range from 5000-6500 psia. A summary
of the wells and their BHP’s are provided in Table 6.4.

The results are summarized in Table 6.5. The oil recoveries are very similar for
three cases (Figure 6.16). The recycling ratio also follows a similar profile in all the three
cases (see Figure 6.17). The gas injection rates are plotted in Figure 6.18. The instability
seen in this plot for the WAG-2280 and WAG-3800 models are numerical issues related to
the solver . In the current version of DOE-CO2 , for a rate constrained well, the solution
is explicitly solved and when the rate exceeds the maximum value specified, the well
is switched to a pressure constrained well in the following timestep. This issue will be
addressed in future version by solving the well constraint implicitly rather than lagging
one time step.

6.6 Sensitivity: Cycle Lengths in WAG

This study is to test the sensitivities of cycle lengths of the WAG process. Three
models are constructed, one being a straight WAG scheme with a half cycle length of
180 days, the second being a straight WAG with a half cycle length of 90 days and the
third being a tapered WAG with decreasing CO2 cycle length and increasing water cycle

61
length. The injection of CO2 in all the models is rate constrained at 2280 Mscf/day per
well. The production wells are pressure constrained at different pressures ranging from
600-2300 psia (see Table 6.4).

The results are summarized in Table 6.6. The recycling ratio is almost the same
with small differences due to injection cycle lengths (see Figure 6.19). The gas injection
rates are plotted in Figure 6.20 are almost the same with the only difference being in the
cycle lengths of injection. As explained in an earlier section, the numerical instability is
due to the explicit calculation. The oil recovery in Figure 6.21 shows that the 90-day half
cycle length case has a slightly better recovery than the other two cases. But the 90-day
half cycle scheme may not be preferred from an operational perspective as it would lead
to more frequent switching of injection wells from CO2 to water and more surveillance.

6.7 Sensitivity: Hysteresis Model

This study is to quantify the impact of hysteresis in a WAG injection scheme for
this field. There are many hysteresis models available, as discussed in Chapter 4, and in
this study, Carlson’s model has been considered. The Carlson’s model can be visualized
as shown in Figure 6.22. The Carlson hysteresis model is a reversible model where
the secondary cycles (drainage and imbibition) follow the same curve as the primary
drainage. Three numerical cases are set up with the same WAG injection scheme, one
without hysteresis, another with hysteresis using a Land’s coefficient of 2.0 (i.e. weak
hysteresis effect) and one with a Land’s coefficient of 1.25 (i.e. strong hysteresis effect).
The hysteresis is considered only for the gas phase.

The results from these models are summarized in Table 6.7. The recycling ratio
and the gas injection rate is almost the same as shown in Figures 6.23 and 6.24. The
oil recovery in Figure 6.25 shows that the simulation including hysteresis has a slightly
better recovery in comparison with the simulation without hysteresis. Also, the runs with

62
hysteresis have lower cumulative water injected. This phenomenon of water injectivity
loss due to gas trapping is seen in many WAG field projects.

6.8 Benchmark Cases and Comparison with Eclipse

To benchmark the results of the CO2 WAG injection from DOE-CO2 with that
from the commercial simulator EclipseT M , first, a continuous CO2 injection is simulated.
As mentioned before, the two phase relative permeability was modified in order to match
Eclipse waterflood results. The relative permeabilities used are shown in Figures 6.26
and 6.27. The results are shown in Figures 6.28 through 6.31. The main matching
parameter is oil recovery as a function of pore volumes of total fluid injected. The
relative permeabilities have been varied to obtain a similar oil recovery vs. pore volume
plot. The same relative permeabilities are then used for the CO2 WAG simulations
when compared with surfactant flooding. The comparison shows that the Eclipse model
produces the same oil as the DOE-CO2 model with lesser volumes of injected gas. Despite
the effort to match the effect of relative permeabilities, the results conclude that the non-
directional end-point scaling that Eclipse applies on the Stone II relative permeability
tables makes the results different to those obtained with DOE-CO2 . Nevertheless, a
reasonably good match has been obtained with oil recovery vs. pore volumes injected
and hence oil recovery comparison can be carried out in further simulation runs.

6.9 Surfactant Flood Simulation

Field C is also under consideration for a surfactant flood as an EOR process.


Such a flood will involve a slug of surfactant to mobilize the residual oil and then a bank
of water to displace the oil towards the producers. In this specific case, a slug of 0.17
PV of 1% surfactant is injected followed by a slug of water till the end of 5 years. The
water injected after the surfactant slug is around 0.22 PV with a cumulative injection
of 0.39 PV. These optimal slug sizes have been chosen based on sensitivity simulations
63
have been performed. The injected salinity in this simulation is 0.7875 meq/ml. This
surfactant flood is run without polymer for mobility control since the permeability is too
low and polymer injection cause significant injectivity loss. Therefore, the recovery would
reflect an unstable surfactant displacement process. The impact is that the macroscopic
sweep efficiency would be lower than a SP (surfactant-polymer) process and possibly the
waterflood because of the water relative permeability enhancement as a result of residual
oil mobilization. The oil recovery from the 5-year surfactant flood is 11.3% with an
incremental recovery of 2.5% over a 5-year waterflood. The surfactant breaks through at
around 1300 days. The results of the surfactant flood are shown in Figures 6.32 through
6.35.

6.10 Comparison of Surfactant vs. CO2 Flood

Comparing the surfactant flood with WAG-CO2 , we find that the CO2 flood yields
a higher recovery at a faster rate compared to the surfactant flood. One of the main
reasons could the higher injectivity of the gas compared to the surfactant leading to a
higher PV injection during the period of 5 years. Due to the lower injectivity, the bank of
incremental oil mobilized by the surfactant compared to the waterflood breaks through
at around 1000 days (see Figure 6.36). Though the oil recovery and oil rates are higher
for the CO2 flood, the decision of which EOR process to apply could be made only after
an economic analysis of the investment and returns for both the processes. This will also
depend on the existing surface facilities and possible sourcing for CO2 . The oil recovery
and pressure of the chemical and CO2 injection against the waterflood are shown in
Figures 6.37 through 6.39.

6.11 Summary and Conclusions

Field scale simulations were performed for Field C to simulate CO2 flooding pro-
cess. A comparison of continuous CO2 flood injection with a WAG flood shows that
64
the WAG scheme has lower recycling and higher oil recovery with much less quantity
of gas injected. Sensitivities were performed for the WAG with respect to production
constraints, cycle length, and hysteresis. The simulation results seem to be less sensitive
to rate constraints of injection. Simulations show that a shorter half cycle length for
the WAG injection yields a slightly higher oil recovery but may not be preferred due
to added operational difficulties. A stronger hysteresis effect (higher gas trapping) on
the gas phase leads to higher trapping and better oil recovery though the difference in
recovery was small in this case. An effort was also made to benchmark the results of gas
flooding between DOE-CO2 and Eclipse. Then a case of surfactant flooding was run after
benchmarking waterflood for UTCHEM, DOE-CO2 , and Eclipse. The surfactant flood
has a slower response due to the lower injectivities compared with a gas flooding. The
incremental recovery with a WAG-CO2 flood was higher and faster than with a surfactant
flood.

65
Table 6.1: Summary of reservoir properties for the Field C model

Table 6.2: Composition for the fluid in Field C

Table 6.3: Comparison between Continuous CO2 and WAG cases

66
Table 6.4: Production and injection BHP constraints - WAG models

Table 6.5: Sensitivity in production constraints - WAG models

67
Table 6.6: Sensitivity in cycle lengths - WAG models

Table 6.7: Sensitivity in hysteresis - WAG models

68
Figure 6.1: Permeability in the X direction (mD) - Field C

Figure 6.2: Porosity map - Field C

69
Figure 6.3: Initial oil saturation - Field C

70
Figure 6.4: Location of wells - Field C

Figure 6.5: Schematic of wells - Field C

71
Figure 6.6: Oil-water relative permeability curves for waterflood match - Field C

Figure 6.7: Gas-oil relative permeability curves for waterflood match - Field C

72
Figure 6.8: Oil recovery (fraction of remaining OIP) comparison for 5 year - waterflood

Figure 6.9: Average reservoir pressure comparison for 5 year - waterflood

73
Figure 6.10: Water injection rates comparison for 5 year - waterflood

Figure 6.11: Water production rates comparison for 5 year - waterflood

74
Figure 6.12: Effluent volume fraction for the continuous CO2 flood

Figure 6.13: Effluent volume fraction for the WAG CO2 flood

75
Figure 6.14: Recycling fraction for Field C - continuous CO2 and WAG floods

Figure 6.15: Oil recovery for Field C - continuous CO2 and WAG

76
Figure 6.16: Oil recovery for Field C - sensitivity to production constraints

Figure 6.17: Recycling for Field C - sensitivity to production constraints

77
Figure 6.18: Gas injection rates for Field C - sensitivity to production constraints

Figure 6.19: Recycling fraction for Field C - sensitivity to cycle lengths

78
Figure 6.20: Gas injection rates for Field C - sensitivity to cycle lengths

Figure 6.21: Oil recovery fraction for Field C - sensitivity to cycle lengths

79
Figure 6.22: Carlson’s hysteresis model with Corey relative permeability model

80
Figure 6.23: Recycling fraction for Field C - sensitivity to hysteresis

Figure 6.24: Gas injection rates for Field C - sensitivity to hysteresis

81
Figure 6.25: Oil recovery (fraction of remaining OIP) for Field C - sensitivity to hysteresis

Figure 6.26: Oil-water relative permeability curves for CO2 flood match

82
Figure 6.27: Gas-oil relative permeability curves for CO2 flood match - Field C

Figure 6.28: Oil recovery (fraction of remaining OIP) comparison: continuous CO2 flood
- Field C

83
Figure 6.29: Pressure comparison: continuous CO2 flood - Field C

Figure 6.30: Gas injection volume comparison: continuous CO2 flood - Field C

84
Figure 6.31: Water production volume comparison: continuous CO2 flood - Field C

Figure 6.32: Oil and water rates: surfactant flood - Field C

85
Figure 6.33: Oil recovery fraction and pressure: surfactant flood - Field C

Figure 6.34: Surfactant injection and production: surfactant flood - Field C

86
Figure 6.35: Water injection and production: surfactant flood - Field C

Figure 6.36: Oil rates for waterflood, WAG-CO2 flood, and surfactant flood - Field C

87
Figure 6.37: Oil recovery fraction versus time for waterflood, WAG-CO2 flood, and
surfactant flood - Field C

Figure 6.38: Oil recovery fraction versus pore volume injection for waterflood, WAG-CO2
flood and surfactant flood - Field C

88
Figure 6.39: Average pressure for waterflood, WAG-CO2 flood and surfactant flood -
Field C

89
Chapter 7

Field-Scale Case Study - Field L

In this chapter, the results of compositional simulation for a field-scale model are
presented. The main emphasis in this chapter is to study the mobility control process
of foam. The mechanistic foam model by Chen et al. (2010) has been implemented into
DOE-CO2 and one of the aims of this study is to test the robustness of the foam model in
its application to field-scale simulations. The results have been compared with existing
empirical foam model by Cheng et al. (2000) based on the work by Rossen et al. (1994) in
which the foam strength is a function of water saturation and gas flow rate depending on
the foam regime. An attempt has also been made to compare this with the foam model
in ST ARS T M simulator which is similar in several aspects to the Pc* model by Rossen
et al. (1994). The model takes into account for the effect of surfactant concentration
and oil saturation on foam stability. Finally, simulations have been performed on a gas
flooding scheme alternating with the surfactant injected in the water cycle to generate
foam in situ leading to higher oil recovery.

7.1 Reservoir Description

The field under consideration, Field L (for purposes of this thesis) is an actual field
with a 50 year history of primary production and waterflooding. It is currently under a
feasibility study for a gas storage project with foam as the mobility control mechanism
to increase the storage capacity and produce the remaining oil. The reservoir lithology
is sandstone with a thickness of 40 ft. The area under consideration has a channel and a
non-channel area. The permeability of the channel is 400 mD with an effective porosity

90
of 0.208 and the non-channel area has a permeability of 10 mD with an effective porosity
of 0.128. The ratio of Kv/Kh is 0.005 and the temperature is 194◦ F . The initial pressure
is assumed to be close to the saturation pressure of 1050 psia. The field has two wells
producing at very low rates of less than 50 bbls/day at the current date. A summary of
the reservoir properties is shown in Table 7.1.

7.2 Model Description

A 3D Cartesian model is constructed and discretized into 46x54x6 grid blocks and
incorporating the anticline structure of the reservoir. The size of the grid block is 264
ft in the X and Y directions with variable thickness for the 6 layers. A 13 component
EOS model is used. The Equation-of-state model is based on a detailed PVT study and
Peng-Robinson equation of state with Peneloux correction employed. The composition
of the oil is shown in Table 7.2. The saturation pressure is estimated to be 1075 psia.
The permeability and the current oil saturation distribution are shown in Figures 7.1
and 7.2. The relative permeabilities have been matched by Corey function as shown in
Figures 7.3 and 7.4.

7.3 WAG Injection Scheme (Base Case)

A section of the field with grid block ranges (X: 20 to 26, Y: 11 to 17) has
been selected as a pilot area for testing the feasibility of a foam assisted WAG injection
scheme. Four production wells and one injection well are used in an inverted 5-spot
pattern pilot. In the base case simulation, a simple WAG scheme is simulated for a
period of 3 years. A slug of CO2 is injected for 2 months followed by a slug of water
for 2 months constituting a cycle. This cycle is repeated for 3 years. The injection wells
are rate constrained with limits on BHP. CO2 injection rate has been designed to be
1.5% of the hydrocarbon pore volume (HCPV) in the pilot per month resulting in an
injected rate of 507 Mscf/day. WAG ratio has been assumed to be 2.0 resulting in a
91
water injection rate of 3%HCPV/month equivalent to 474 stb/d. The production wells
are BHP constrained at 95 psia.

The results for the WAG injection scheme are shown in Figures 7.5 through 7.9.
The plotted results are cumulative oil; water, oil and gas production rates; pressure;
water and gas injection rates and effluent volume fraction.

7.4 Study 1: Mobility Control Processes - FAWAG vs. WAG

In this case study, foam is evaluated as the mobility control process. The only
difference with the base case is that 5000ppm of surfactant is injected along with the
water in the water half-cycle. The foam model used in this case is the mechanistic foam
model by Chen et al. (2010). The model parameters are summarized in Table 7.3.

7.4.1 Results and discussion

The results from the Foam Assisted WAG (FAWAG) model are compared to the
ones from the base case. The comparison is summarized in Table 7.4. The effluent CO2
volume fraction is consistently lower in the FAWAG compared to the WAG injection.
The FAWAG case has a slightly higher recycling ratio but shows a greatly declining
trend versus pore volumes injected (see Figure 7.10). The oil recovery is almost similar for
much less pore volumes of fluid injected (see Figure 7.11). Since this is a rate constrained
model, the impact of foam will be shown in the BHP of the injector. Figure 7.12 shows
that the FAWAG has higher BHP because of the higher resistance or effective viscosity
of foam.

7.5 Study 2: Foam Models - Empirical vs. Mechanistic Foam


Models

Along with the mechanistic foam model as applied in the above section, DOE-CO2
also has an option of an empirical P c∗ foam model dependent mainly on water saturation
92
with thresholds for water saturation and surfactant concentration. In this case study an
attempt has been made to replicate the foam behavior as seen in the above section with
the empirical foam model.

7.5.1 Empirical foam model parameters

The empirical foam model by Cheng et al. (2000) based on the work by Rossen
et al. (1994) can be described as follows

If Sw < Sw∗ − ε or Cs < Cs∗ then f


krg = krg (7.1)

If Sw∗ − ε < Sw < Sw∗ + ε and Cs < Cs∗ then


f krg (7.2)
krg = ∗ +ε)
(R−1)(Sw −Sw
1+ 2ε
krg
If Sw > Sw∗ + ε or Cs ≥ Cs∗ then f
krg = (7.3)
R
where Cs is the surfactant concentration, Cs∗ is the threshold surfactant concentration
f
for foam formulation, krg is the effective gas relative permeability modified for foam, R is
the maximum foam mobility reduction factor and Sw∗ is a threshold water concentration.
The above formulation shows that the gas mobility reduction increases sharply from 1 to
a value of R for the water saturation ranging from Sw∗ − ε to Sw∗ + ε. This dependence
of R on water saturation is beneficial in an attempt to replicate the foam behavior of
mechanistic foam model.

From the mechanistic foam model simulation results, the trend of R versus Sw is
plotted and empirical model parameters are fitted to the same plot as shown in Figure
7.13. This provides an initial estimate of the foam parameters to run the simulation.
All other constraints in the mechanistic foam FAWAG model remain the same. The
assumptions are simplistic but the best that can be done to arrive at an equivalent
model. With the above methodology, the parameter values for the empirical foam model
are shown in Table 7.5.

93
7.5.2 Results and discussion

In the figures hereafter, the empirical foam model as IFOAM2 and mechanistic
foam model case is named as IFOAM3. The results from the IFOAM2 model are com-
pared to the ones from the IFOAM3. The comparison is summarized in Table 7.6. The
effluent CO2 volume fraction is mostly similar in both the models. The BHP of the injec-
tor in both the models is also the same (see Figure 7.14). The IFOAM2 model has a lower
recycling ratio implying that the empirical foam model gives a stronger compared to the
mechanistic model. At the same time, the IFOAM2 model has a higher cumulative gas
injected implying that the empirical model foam gives a weaker foam compared to the
mechanistic one. This gives mixed conclusions about the foam strength in the empirical
model (see Figures 7.15 and 7.16). This implies that the assumptions made to arrive at
an equivalent empirical model from the mechanistic foam model are not justified.

7.6 Study 3: FAWAG vs. Low Tension Gas Flooding

In this case study, the process of foam assisted WAG is compared against a newer
recovery technique called low tension gas flooding. In low tension gas flooding, two dif-
ferent surfactants are injected in the water slug of the WAG scheme. The function of
one surfactant is to form foam in-situ in the presence of gas. The function of the other
surfactant is to lower the interfacial tension between oil/water to reduce the residual oil
saturation leading to higher oil recoveries. Technically, these two surfactants are com-
pletely different in their formulation and have to be defined separately in the simulation.
In an ideal scenario for maximizing recovery, one should inject the foaming surfactant
in the initial stages of gas injection. Once foam is formed and gas mobility is reduced,
one should switch to the low IFT surfactant which then mobilizes the residual oil and
increases oil recovery.

In the current capability of DOE-CO2 , the surfactant has been treated as a tracer

94
for the purposes of foam simulation and it is not possible to switch over from one sur-
factant functionality to the other during the simulation. As a preliminary analysis, for
quantifying the benefit of a low tension gas flooding scheme, the same surfactant has
been assumed to act in both ways and low IFT relative permeability curves are supplied
in the model to estimate the incremental benefit. The simulation is carried out for 3
cycles of CO2 and water injection for a total period of 1 year.

7.6.1 Results and discussion

In the figures hereafter, the low tension gas flooding model will be referred as
IFOAM3+ICAP and the mechanistic foam model will be referred as IFOAM3. The
results from the IFOAM3+ICAP model are compared to the ones from IFOAM3. The
comparison is summarized in Table 7.7. The effluent CO2 volume fraction is mostly
similar in both the models. The IFOAM3+ICAP model has a lower recycling ratio with
higher oil recovery at smaller pore volumes of fluid injected implying that low tension
gas flooding has significant benefits over a simple FAWAG process (see Figures 7.17 and
7.18).

7.7 Study 4: Benchmarking between DOE-CO2 and STARS

This study is to benchmark the foam model in DOE-CO2 with that in ST ARS T M ,
one of the most widely used commercial simulators for foam flooding. ST ARS T M has an
option for population balance foam simulation but the more widely applied model is an
empirical one similar in many aspects to the empirical model in DOE-CO2 . This study
makes an attempt to create an equivalent model in ST ARS T M and compare the results
with the DOE-CO2 model.

95
7.7.1 STARS empirical foam model

In the ST ARS T M foam model, the gas relative permeability is multiplied by a


factor FM which is inversely related to a product of many factors as explained below.

f 1
krg = krg ∗F M ; F M = (7.4)
1 + F M M OB ∗ F 1 ∗ F 2 ∗ F 3 ∗ F 4 ∗ F 4 ∗ F 5 ∗ F 6 ∗ F 7

where F M M OB is the maximum gas mobility reduction factor, F 1 through F 6 cor-


respond to the effect of surfactant concentration, oil saturation, capillary number, gas
capillary number, oil mole fraction, and salt concentration respectively on foam strength.
F 7 is the factor which corresponds to the effect of water saturation on foam. F 7 can be
described as follows:

arctan(EP DRY ∗ (Sw − F M DRY ))


F 7 = 0.5 + (7.5)
π

where F M DRY is a dryout factor varying between 0-1 and EP DRY is a scaling factor
dependent on the porous media structure.

For getting an equivalent model to the empirical model in DOE-CO2 , the fol-
lowing are the values of the parameters used: F 2 to F 6 = 1 (since the effects are not
considered in DOE-CO2 foam model) EP SU RF =75; F M SU RF =0.9E-4; EP DRY =50;
F M DRY =0.36; F M M OB=90. The foam mobility reduction factor plotted as a func-
tion of water saturation using the above parameters is shown in Figure 7.19.

7.7.2 Results and discussion

Firstly, the simulators are benchmarked against waterflood scheme. Then the
results of foam simulation from the ST ARS T M model are compared with those of the
empirical foam model in DOE-CO2 . The ST ARS T M model has a 10% higher oil recovery
compared to the empirical foam model in DOE-CO2 as shown in Figure 7.20. The gas
injection and production rates are shown in Figures 7.21 and 7.22. The gas injection rates
in STARS model are higher leading to the conclusion that the foam is weaker compared
96
to the DOE-CO2 model. The foam models do not yield identical foam strength evident
from the oil recovery despite making the models equivalent with a selection of parameter
values. The FM parameter (inverse of the reduction factor) from the STARS model and
foam mobility reduction factor from the DOE-CO2 model are shown in Figures 7.23 and
7.24. Though they look similar in profile, the models themselves do not yield the same
recovery.

7.8 Summary and Conclusions

Field scale simulations were performed on Field-L to test the impact of foam as
a mobility control process. The foam assisted WAG (FAWAG) scheme has lower CO2
utilization factor and lower recycling ratio compared to the WAG scheme. The mechanis-
tic foam model is compared with the empirical foam model in DOE-CO2 . Results show
that assumptions made to arrive at an equivalent empirical model from the mechanistic
foam model are not adequate. Simulations also show that low tension gas flooding has
significant benefits over a simple FAWAG process. Simulations to benchmark the foam
modeling in DOE-CO2 with a commercial simulator (STARS) show that the results give
very similar trend with some differences.

97
Table 7.1: Summary of reservoir properties for the Field L model

Table 7.2: Oil composition for the fluid in Field L

Table 7.3: Foam parameters for the mechanistic model - Field L

98
Table 7.4: Comparison between FAWAG (mechanistic) and WAG models

Table 7.5: Empirical foam model parameters

Table 7.6: Comparison between foam models - mechanistic (IFOAM3) and empirical
(IFOAM2)

99
Table 7.7: Comparison between models - FAWAG (IFOAM3) and low tension gas flooding
(IFOAM3+ICAP)

Figure 7.1: Permeability in the X direction (mD) - Field L

100
Figure 7.2: Initial oil saturation distribution for Field L

Figure 7.3: Oil-water relative permeability curves matched with the tabular values

101
Figure 7.4: Gas-oil relative permeability curves matched with the tabular values

Figure 7.5: Cumulative oil produced and pressure vs. time - WAG

102
Figure 7.6: Water and gas production rates vs. time - WAG

Figure 7.7: Water and gas injection rates vs. time - WAG

103
Figure 7.8: CO2 production and injection rates vs. time - WAG

Figure 7.9: Effluent volume fraction vs. time - WAG

104
Figure 7.10: Recycling fraction for Field L - FAWAG (IFOAM3 - mechanistic) vs. WAG
(No foam)

Figure 7.11: Oil recovery (fraction of OOIP) for Field L - FAWAG (IFOAM3 - mecha-
nistic) vs. WAG (No foam)

105
Figure 7.12: BHP of the injector for Field L - FAWAG (IFOAM3 - mechanistic) vs. WAG
(No foam)

Figure 7.13: Permeability reduction factor for Field L - Mechanistic vs. Empirical foam
models

106
Figure 7.14: BHP of the injector for Field L - (IFOAM3 - mechanistic) versus (IFOAM2
- empirical)

Figure 7.15: Recycling fraction for Field L - (IFOAM3 - mechanistic) versus (IFOAM2 -
empirical)

107
Figure 7.16: Oil recovery (fraction of OOIP) for Field L - (IFOAM3 - mechanistic) vs.
(IFOAM2 - empirical)

Figure 7.17: Recycling fraction for Field L - (IFOAM3 - FAWAG) vs. (IFOAM3+ ICAP
- low tension gas flooding)

108
Figure 7.18: Oil recovery (fraction of OOIP) for Field L - (IFOAM3 - FAWAG) vs.
(IFOAM3+ICAP - low tension gas flooding)

f
Figure 7.19: STARS vs. DOE-CO2 model - gas mobility reduction factor (krg /krg )

109
Figure 7.20: Oil recovery - STARS vs. DOE-CO2 foam models

Figure 7.21: Gas injection rates - STARS vs. DOE-CO2 foam models

110
Figure 7.22: Gas production rates - STARS vs. DOE-CO2 foam models

Figure 7.23: Inverse of foam mobility reduction factor - STARS foam model

111
Figure 7.24: Foam mobility reduction factor - DOE-CO2 foam model

112
Chapter 8

Synthetic Field Case Studies

In this chapter, the mobility control processes of WAG and foam are compared in
two different reservoir settings: Permian Basin Carbonates and Gulf Coast Sandstones.
Nearly 66% of all CO2 flooding projects in the US are in the Permian Basin and the rest
is mostly in the Gulf Coast sandstones. WAG is one of the most widely used processes
in the field for mobility control. But there could be scenarios in which gas breakthrough
is quick even after application of WAG. This could possibly be due to the water under-
riding and gas over-riding. The foam process is beneficial as it decreases the mobility
of gas achieving a better conformance. Here we use foam in the form of foam assisted
WAG (FAWAG) in both of the above settings. The goal is to create a model with similar
conditions to quantify the possible incremental benefit of using foam over WAG.

Also, a novel technique is to use the surfactant as an agent for reducing residual
oil saturation as well as generating foam. This process is also known by low tension gas
flooding. The surfactants used in foam are usually dilute and the ones used for residual
saturation are low IFT surfactants. In an actual scenario, both these surfactants are
different and are injected at different times in the production period; a slug of low IFT
surfactant is first injected followed by a FAWAG process with dilute surfactant for foam.
This capability of modeling two different types of surfactant processes simultaneously is
not implemented in DOE-CO2 but we can quantify the preliminary benefit assuming one
surfactant can function as both of them. The incremental benefit of using low tension
gas flooding over FAWAG is quantified in both of the settings.

113
8.1 Literature Review

This section presents a brief review of the literature published regarding Permian
Basin carbonates and Gulf Coast sandstone reservoirs.

Perry (1982) presents an overview of the miscible CO2 process in the Weeks Island
field on a Gulf Coast pier-cement type salt dome. The study shows that the displacement
follows a gravity stable displacement and that the oil column thickness has increased to
57 ft from an initial 23 ft. Johnston (1988) presents a review of the early pilot history
and a simulation history match of pilot performance in the Weeks Island field. The study
concluded that gravity stable flooding is very effective in these high permeable, highly
step oil reservoirs. Senocak et al. (2008) focus on the heterogeneity measures in terms
of recovery efficiency and utilization rate in a mature CO2 flood in the Little Creek field.
Mapping of well-by-well heterogeneity and Lorenz coefficient seemed to provide better
insights than mapping permeability, porosity or thickness. Davis et al. (2011) present
the results of CO2 flooding in Hastings field, Mississippi. The study concluded that the
success of the CO2 flood was based on the ability to maintain miscibility and moving the
CO2 injectors away from the oil-water contact.

Hill et al. (1994) present an overview of the CO2 flood design in the San Andres
carbonate formation, South Welch Unit of the Welch field. A compositional model study
of the pilot performance was used to design a CO2 processing facility and sensitivities
such as WAG ratios, cycle lengths and total slug sizes were investigated. Hsu et al. (1997)
present a field-scale fully compositional CO2 flood simulation study in the Wasson Denver
Unit in the Permian Basin capable of capturing areal variations on an individual well
basis. Transition zones were also included and the study pinpointed the best shut-in wells
to return to production, identified infill and horizontal drilling locations and quantified
the injectant losses. Brinkman et al. (1998) summarized the miscible CO2 flood project
for the Sharon Ridge Canyon Unit and identified the best patterns for CO2 flooding,

114
and selected a strategy for sizing the final flood area and timing of future expansions.
Oudinot et al. (2009) presented the CO2 injection performance in San Juan basin, New
Mexico. Honarpour et al. (2010) looked at CO2 flooding in the residual oil zones in
Seminole San Andres unit in the Permian Basin and better history match was obtained
through improved rock and fluid characterization.

8.2 Carbonates (Permian Basin)

CO2 flooding is one of the most important EOR recovery processes in the US and
the carbonate reservoirs are a major proportion in which the process is applied. The
Permian basin has been an important showcase for application of CO2 flooding in the
US.

8.2.1 Model description and properties

A three dimensional layer-cake model is constructed with layered permeability


and porosity fields. All reservoir properties are based on the Permian Basin setting.
Reservoir gridblock size is 30 ft, 30 ft and 10 ft in the X, Y and Z directions with the
reservoir discretized into 11x11x8 gridblocks. The model is constructed as a quarter of
a 10-acre five spot pattern with one injector and one producer. The formation depth
is 5000 ft with the temperature 180◦ F . The average permeability is 10 mD and the
average porosity is 15%. The following correlation by Lucia and Fogg (1990) is applied
to correlate permeability and porosity.

k(in mD) = 7.38X106 Xφ6.72 (8.1)

The permeability and porosity fields can be seen in Figures 8.1 and 8.2. The reservoir
fluid is same as the one used in Chapter 6 (see Section 6.2). A constant initial saturation
of 0.4 is assumed in the reservoir. The relative permeabilities are average values from
the data from Oak (1990) for an intermediate wet sample and modeled using the Corey
115
function. A summary of all the reservoir properties is shown in Table 8.1 and the relative
permeability parameters are given in Table 8.2. Three cases are constructed, the first
being WAG, the second FAWAG and the third low tension gas flooding. The models are
run for a period of 7 years with the half cycle length of WAG being 180 days leading to
a total of 7 cycles. The injection rate of CO2 is limited to 1% HCPV/month resulting
in 70.2 Mscf/day for the injector. The WAG ratio is assumed to be 1.0 equivalent to a
water injection rate of 40 STB/day for the first two cycles which later increases to 80
STB/day. The surfactant injected in the FAWAG and low tension gas flooding cases is
at a concentration of 5000 ppm. The injector is rate controlled while the producer is
pressure controlled with an initial BHP of 2250 psia gradually dropping to 750 psia in
the final cycle.

8.2.2 Results and discussion

The results from all the three models are summarized in Table 8.3. The recycling
fraction plotted in Figure 8.3 shows that the FAWAG model and the low tension gas
flooding (LTGF) models have much lower recycling ratios. The BHP of the injector
from the three models is shown in Figure 8.4. As expected, the FAWAG model has a
higher BHP compared to the WAG model due to foam which impedes the gas flow and
increases the injection pressure. The LTGF model has a lower BHP compared to the
other two models as the surfactant improves the relative permeability of all the three
phases at higher capillary number closer to the injection well. The recovery from the
WAG model is around 40% for around 95% of the pore volume injected. The oil recovery
plot in Figure 8.5 shows that the FAWAG model deviates from the WAG model after the
first cycle when foam becomes active. The FAWAG model has higher recovery at lower
injected fluid volumes compared to the WAG model. The LTGF model has even higher
recovery compared to the FAWAG model but since it involves injecting more efficient
surfactants for ultralow IFT, the economics need to be considered before a decision on

116
the development process is made.

8.3 Sandstones (Gulf Coast)

CO2 flooding has been on the rise in the Gulf Coast sandstone fields for example
Hastings field, Little Creek field, Weeks Island among others. The construction of a CO2
pipeline network (“Green Pipeline” Davis et al., 2011) is instrumental for large scale
CO2 flooding in the region.

8.3.1 Model description and properties

A three dimensional layer-cake model is constructed with stochastic permeability


and porosity fields. All reservoir properties are based on the Gulf Coast sandstone
reservoirs setting. Reservoir gridblock size is set to 30 ft, 30 ft and 10 ft in the X,
Y and Z directions with the reservoir discretized into 11x11x8 gridblocks. The model
is constructed as a quarter of a 10-acre five spot pattern with one injector and one
producer. The formation depth is set to 6140ft with the temperature 220◦ F . The average
permeability is 100 mD and the average porosity is 19%. The porosity-permeability
relationship is defined to be an exponential relationship (similar to Senocak et al., 2008).

k(in mD) = 3.387X10−4 Xe58.7φ (8.2)

The permeability and porosity fields, generated stochastically with a dimensionless cor-
relation lengths of 0.5 and 0.3 in the X and Y directions, can be seen in Figures 8.6 and
8.7. The reservoir fluid is same as the one used in Chapter 6 (see Section 6.2). A constant
initial water saturation of 0.4 is assumed in the reservoir with pressure at the saturation
point. The relative permeabilities are average values from the data from Oak (1990) for
an intermediate wet sample and modeled using the Corey function. A summary of all the
reservoir properties is shown in Table 8.4 and the relative permeability parameters are
the same as those used in the earlier section in Table 8.2. Three cases are constructed, the
117
first being WAG, the second FAWAG and the third low tension gas flooding. The models
are run for a period of 7 years with the half cycle length of WAG being 180 days leading
to a total of 7 cycles. The injection rate of CO2 is limited to 1% HCPV/month resulting
in 111.4 Mscf/day for the injector. The WAG ratio is assumed to be 1 equivalent to a
water injection rate of 64 STB/day for the first two cycles which later increases to 128
STB/day. The surfactant injected in the FAWAG and low tension gas flooding models
is at a concentration of 5000 ppm. The injector is rate controlled while the producer is
pressure controlled with an initial BHP of 3000 psia gradually dropping to 1000 psia in
the final cycle.

8.3.2 Results and discussion

The results from all the three models are summarized in Table 8.5. The recycling
fraction plotted in Figure 8.8 shows that the FAWAG model and the low tension gas
flooding (LTGF) cases have much lower recycling ratios. The BHP of the injector from
the three simulations is shown in Figure 8.9. As expected, the FAWAG case has a higher
injection BHP compared to the WAG case due to foam with higher effective viscosity
compared to the gas. The LTGF model has a lower injection BHP compared to the
other two simulations as the surfactant improves the relative permeability of all the
three phases at higher capillary number closer to the injection well. The recovery from
the WAG model is around 50% for around 100% of the pore volume injected. The oil
recovery plot in Figure 8.10 shows that the FAWAG case deviates from the WAG model
after the first cycle when foam becomes active. The FAWAG case has higher recovery at
lower injected fluid volumes compared to the WAG simulation. The LTGF case has even
higher recovery compared to the FAWAG but since it involves injecting costly low IFT
surfactant, the economics need to be considered before a decision on the development
process is made.

118
8.4 Summary and Conclusions

The carbonate and the sandstone simulation cases show that foam can give addi-
tional benefit over a WAG injection scheme. In either of the models, an adverse perme-
ability profile and a high heterogeneity was used resulting in foam giving good incremental
recovery. One-on-one comparison between the simulation cases shows that the sandstone
reservoirs give higher recovery compared to the carbonate ones. In the sandstone reser-
voirs, though the permeability and porosity fields are heterogeneous, they are stochastic
and hence have an innate conforming ability leading to a later breakthrough as shown in
the effluent volume fraction in Figures 8.11 and 8.12.

119
Table 8.1: Summary of reservoir properties - carbonate reservoir case

Table 8.2: Summary of relative permeability end points - carbonate reservoir case

Table 8.3: Comparison of results - carbonate reservoir case

120
Table 8.4: Summary of reservoir properties - sandstone reservoir case

Table 8.5: Comparison of results - sandstone reservoir case

121
Figure 8.1: Permeability in the X-direction (mD) - carbonate case

Figure 8.2: Porosity distribution in the carbonate case

122
Figure 8.3: Recycling fraction comparison for the carbonate reservoir cases

Figure 8.4: Comparison of the injector bottom-hole pressure for the carbonate reservoir
cases

123
Figure 8.5: Oil recovery (fraction of OOIP) comparison for the carbonate reservoir cases

Figure 8.6: Permeability in the X-direction (mD) - sandstone reservoir case

124
Figure 8.7: Porosity distribution in the sandstone reservoir case

Figure 8.8: Recycling fraction comparison for the sandstone reservoir cases

125
Figure 8.9: Comparison of the injector bottom-hole pressure for the sandstone reservoir
cases

Figure 8.10: Oil recovery (fraction of OOIP) comparison for the sandstone reservoir cases

126
Figure 8.11: Effluent volume fraction for the carbonate reservoir case with WAG

Figure 8.12: Effluent volume fraction for the sandstone reservoir with WAG

127
Chapter 9

Summary, Conclusions and Recommendations

9.1 Summary and Conclusions

This study investigated two mobility control processes for gas flooding namely
WAG and foam. New petrophysical and foam models have been developed and imple-
mented in DOE-CO2 , the in-house compositional gas flooding simulator at The University
of Texas at Austin. These models help further our understanding of these EOR processes
and help us make a better analysis of the development mechanism for reservoirs suitable
for gas flooding. CO2 flooding is the focus of this study as it is the most popular EOR
process in terms of proportion of oil production through EOR processes. The following
are the main conclusions from this study:

• A new three-phase relative permeability model (UTKR3P) incorporating hysteresis


and compositional dependency has been developed, tested against experimental
data for different wettability conditions, and implemented into a compositional
reservoir simulator (Beygi, 2014).

• UTKR3P has been validated using multi-cyclic three-phase water-alternating-gas


experimental data for mixed wet/oil wet rocks from Oak (1990).

• The hysteresis model (UTHYST) developed, is an extension of the Land trapping


model but with a dynamic Land’s coefficient. The model overcomes some of the
limitations of published three-phase hysteresis models for non-water wet rocks and
mitigates some of the complexity and difficulties associated with commonly applied
models in numerical reservoir simulators (Beygi et al., 2013).
128
• The local equilibrium approximation of the population balance based foam model
by Chen et al. (2010) has been implemented into DOE-CO2 , the in-house compo-
sitional gas flooding simulator.

• Experimental validation for N2 and CO2 foams with simulation data shows that
the local equilibrium approximation works well for modeling foam texture (Naderi
et al., 2013).

• Simulations with layer cake models for WAG and simultaneous injection schemes
have been performed to test the local equilibrium foam implementation.

Field-Scale Case Study - Field C:

• Field scale simulations were performed for Field-C to compare the EOR processes
of CO2 flooding and surfactant flooding and suggest a best possible development
scheme for the field.

• A comparison of continuous CO2 flood injection with a WAG flood shows that the
WAG scheme has lower recycling and higher recovery with much less quantity of
gas injected.

• Sensitivities were performed for the WAG process with respect to production con-
straints, cycle length, and hysteresis. The simulation results were not very sensitive
to the changes in the injection rate constraints. Simulations show that a shorter
half cycle length for the WAG process yields a slightly higher recovery but may im-
pose operational difficulties. A stronger hysteresis on the gas phase leads to higher
trapping and higher oil recovery though the difference in the recovery was small in
this case.

• The surfactant flood simulation shows that it has a slower response due to the lower
injectivities compared with a gas flooding scheme. The incremental recovery with
129
a WAG-CO2 flood was higher and faster than with a surfactant flood. But the
ultimate decision has to be taken after an economic analysis and consideration of
operational factors.

Field-Scale Case Study - Field L:

• Simulations were performed for Field-L to test the impact of foam as a mobility
control process. Even though the pilot area selected for the simulations has minimal
heterogeneity, the foam assisted WAG (FAWAG) scheme has lower CO2 utilization
factor and lower recycling ratio compared to the WAG scheme.

• The mechanistic foam model is compared with the empirical foam model in DOE-
CO2 . Results show that assumptions made to arrive at an equivalent empirical
model from the mechanistic foam model are not adequate.

• Simulations also show that low tension gas flooding has significant benefits over
a simple FAWAG process. This simulation makes an important assumption of
the way the surfactant functions are handled in the simulator. In low tension
gas flooding, there are two different surfactants are injected in the water slug of
the WAG scheme. Technically, these two surfactants are completely different in
their formulation and are injected in different phases of production. The foaming
surfactant is injected in the initial phases and the low IFT surfactant during the
later stages. In the current capability of DOE-CO2 , the surfactant has been treated
as a tracer for the purposes of foam simulation and it is not possible to switch over
from one surfactant functionality to the other during the simulation.

• Simulations to benchmark the foam modeling in DOE-CO2 with a commercial sim-


ulator (STARS) show that the results give very similar trend with some differences.

130
Field applications - carbonate and sandstone reservoir replica models:

• The carbonate and the sandstone reservoir models replicating the Permian Basin
carbonates and the Gulf Coast sandstones respectively show that foam can give
additional benefit over a WAG injection scheme. In either of the simulation cases,
an adverse permeability profile and a high heterogeneity resulted in foam giving
good incremental recovery. Similar wettability for the carbonate and the sandstone
reservoirs have been assumed and the same relative permeability curves have been
used. The difference in the rock type manifests in the distribution of permeability.

• One-on-one comparison between the simulations shows that the sandstone reser-
voirs give higher recovery compared to the carbonate case. In the sandstone
reservoirs, though the permeability and porosity fields are heterogeneous, they are
stochastic and hence have an innate conforming ability leading to a later break-
through.

9.2 Recommendations

The following work(s) could be taken up to further our understanding of mobility


control processes in CO2 flooding:

• Application of the cycle dependent relative permeability model (UTKR3P) to the


field scale model could show the real impact of hysteresis and trapping due to flow
reversals.

• For the simulation of low tension gas flooding process, DOE-CO2 could be modified
to handle two different types of surfactants, one as a dilute surfactant for foam
generation and one as an agent for residual oil reduction.

131
Appendix: Input Files

A1: Input File for Chapter 4: Case 3 - UTKR3P Relperm Model


with UTHYST Hysteresis Model
**********************************************************************
CC FIELD: SYNTHETIC
CC CASE : 2D CASE WITH HYSTERESIS
CC INJECTION SCHEME : WAG CO2 INJECTION
CC********************************************************************
CC LAYERCAKE MODEL
CC
CC RESERVOIR PROPERTIES
CC GRID BLOCKS : 60*1*10
CC GRID BLOCK SIZE : 50*100*20
CC MODEL DEPTH : 6000*FT
CC POROSITY : 0.2
CC PERMEABILITY X : 1010 mD
CC KV/KH : 0.25
CC TEMPERATURE : 220 F
CC INITIAL PRESSURE : 1500 PSIA
CC INIT. WATER SAT. : 0.45
CC RELPERM MODEL : UTKR3P
CC HYSTERESIS MODEL : UTHYST
CC INJECTION : WAG CO2 INJECTION
CC TIME PERIOD : 0-60 DAYS
CC
CC********************************************************************
CC FILE CREATED BY : VENKATESWARAN S PUDUGRAMAM
CC DATE CREATED : MAY 2013
CC
CC
CC
CC********************************************************************
CC
CC..+....1....+....2....+....3....+....4....+....5....+....6...+...7..
CC****************CONSTANTS AND VALUES FOR MODELS*********************
CC
CC IGEOM, NBM
23 0
CC

132
CC NXM, NYM, NZM, NPM, NCM, NWM, IXYZ, POSTPROCESSOR
60 1 10 3 8 2 1 2
CC
CC NS1M,NBWM,NPRM,NPFM,NCTM,NHSM, NPRPERMM, NS3OUT1,NS3OUT2
1 101 15 1 1 1000 3 3500 3500
CC
CC IC2,NCMT,NREG,NTAB
2161
CC
CC EE, NZPM
0.5761 3
CC
CC TOLP, TOLVOL, QLIM
0.5D-1 0.5D-4 1.0D-30
CC
CC NUMAX,INQUA,INCON,INVEL,NOSWTM,NUMPRE,NUMOUT,NUMPVT,NST,IFLIP
10 3 3 5 10 30 12 30 20 0
CC*********************************************************************
CC CASE NAME WITH FORMAT ( 17A4, A2 ) OF TOTAL 70 COLUMNS.
*—-HEADER
UTKR3P Relperm Model with UTHYST Hysteresis Model
CC
CC NUMBER OF COMPONENTS.
*——–NC
8
CC COMPONENT NAMES WITH FORMAT ( 1X, A8 ), NC CARDS.
CC..+..8
*—-NAME
CO2
N2-C1
C2-C3
C4-C6
C7+
C10+
C14+
C20+
CC BLACK OIL OPTION, SALINITY, AQUIFER OPTION
CC (0:OFF,1:ON) PPM (0:OFF,1:ON)
*—–IBOST SLNTY IAQUIF
0 0. 0
CC CRITICAL PRESS. (PSI), TEMP. (R) AND VOL. (CU FT/LB-MOLE),
CC MOLECULAR WT. (LB/LB-MOLE), ACENTRIC FACTOR, PARACHOR. NC CARDS.
*——–PC TC VC WT OM PARACH VSP
1069.87 547.56 1.603 44.010 0.2250 78.00 -0.05435

133
631.54 319.44 1.795 17.570 0.0145 69.90 -0.19268
657.00 835.84 3.452 36.510 0.1279 132.75 -0.1242
491.36 1050.45 5.199 69.850 0.2423 229.61 -0.04714
409.01 1196.58 6.291 104.380 0.3533 298.42 0.12961
310.05 1359.53 9.330 154.130 0.5003 423.52 0.14725
250.43 1359.53 16.898 222.780 0.6836 587.09 0.13847
187.67 2020.83 29.737 516.410 0.8328 1518.16 0.0593
CC
CC NC CARDS.
*—-PARAA PARAB
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
CC
CC BINARY INTERACTION COEFFICIENTS, CIJ. NC CARDS.
*—–DELTA
0.0000
0.1110 0.0000
0.1200 0.0109 0.0000
0.1750 0.0112 0.0000 0.0000
0.0750 0.0109 0.0002 0.0035 0.0000
0.0500 0.0102 0.0000 0.0035 0.0000 0.0000
0.0700 0.0102 0.0000 0.0035 0.0005 0.0000 0.0000
0.0750 0.0102 0.0000 0.0035 0.0005 0.0000 0.0000 0.0000
CC
CC BINARY INTERACTION COEFFICIENTS, DIJ. NC CARDS.
*——-DIJ
0.000
0.000 0.000
0.000 0.000 0.000
0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
CC
CC MAXIMUM NUMBER OF PHASES ( 3 OR 4 ), IVISC AND ISINGL (0 OR 1)
*——–NP IVISC IVISC COEF ISINGL ISOLU
31110

134
CC
CC VISCOSITY COEFFICIENT
*—–COEF1,COEF2,COEF3,COEF4,COEF5
0.1023 0.023364 0.058533 -0.040758 0.0093324
CC IEOS: 1, IPEM: 0, 1 OR 2
CC
*——IEOS IPEM ISTAM IEST IVSP KI
1 1 -1 0 0 0
CC
CC ITERATION TOLERANCES FOR PERSCHKE’S FLASH ROUTINES.
*—-TOLFLA TOLFLM TOLPD TOLSAM TOLSAS TOLSUM
1.E-08 1.E-08 1.E-08 1.E-08 1.E-08 1.E-08
CC
CC MAXIMUM NUMBER OF ITERATIONS FOR PERSCHKE’S FLASH ROUTINES.
*—-MAXFLA MAXFLM MAXPD MAXSAM MAXSAS MAXSANR
1000 1000 1000 1000 1000 20
CC
CC ITERATION TOLERANCE AND MAX NUMBER OF ITERATIONS FOR VECTOR-FLASH.
*—-IVCFL TOLVFL MAXVFL
0 1.E-8 20
CC
CC SWITCHING PARAMETERS FOR PERSCHKE’S FLASH ROUTINES.
*—-SWIPCC SWIPSA
.000001 0.000001
CC
CC PHASE IDENTIFICATION PARAMETERS FOR PERSCHKE’S FLASH ROUTINES.
*——IOIL ITRK DMSLIM IZGAS IDEN
1 2 20. 0 0
CC
CC TRACER FLAG(0:OFF,1:ON), ASPHALTENE FLAG(0:OFF,1:ON)
*—-IFLAGT IASPR
00
CC
CC********************************************************************
CC *
CC OUTPUT OPTIONS *
CC *
CC********************************************************************
CC
CC
CC HISTORY PRINTING PARAMETER FOR <<HISTORY.CPR>>.
*— NHSSKIP NSTSKIP IPV
200 40 0
CC

135
CC REFERENCE CONCENTRATION, CONC0, USED FOR EFFLUENT CONCENTRATION (NC)
*—–CONC0
1. 1. 1. 1. 1. 1. 1. 1.
CC
CC NUMBER OF PRINTS FOR <<.TAB>> (ALSO FOR TRAPPING & ASPHALTENE DATA)
*——-NPR
15
CC
CC TIME(DAYS) AND FLAGS ( 0 OR 1 ) FOR <<TABLE.CPR>>. NPR CARDS.
*——-TPR MPRP MPRSAT MPRIND MPROMFR MPRPMFR MPRPRO MPRATES
2. 1 1 1 1 1 1 1
60. 1 1 1 1 1 1 1
140. 1 1 1 1 1 1 1
220. 1 1 1 1 1 1 1
300. 1 1 1 1 1 1 1
380. 1 1 1 1 1 1 1
460. 1 1 1 1 1 1 1
540. 1 1 1 1 1 1 1
600. 1 1 1 1 1 1 1
680. 1 1 1 1 1 1 1
760. 1 1 1 1 1 1 1
840. 1 1 1 1 1 1 1
920. 1 1 1 1 1 1 1
1000. 1 1 1 1 1 1 1
1080. 1 1 1 1 1 1 1
CC
CC NUMBER OF PRINTS FOR <<PROFILE.CPR>>.
*——-NPF
0
CC
CC NUMBER OF PRINTS FOR <<CONTOUR.CPR>>.
*——-NCT
0
CC
CC NPRPERM
*——-
1
CC
CC
*——- IPRPERMG IPRPERMPH
33
CC
CC********************************************************************
CC *

136
CC RESERVOIR AND WELL DATA *
CC *
CC********************************************************************
CC
CC A FLAG FOR RESERVOIR GEOMETRY:
CC 1-D: 11(Y), 12(X), 13(Z), 2-D: 21(XY), 22(YZ), 23(XZ), 3-D: 31
*—–IGEOM INUG
23 0
CC
CC NUMBER OF GRID BLOCKS IN X, Y, AND Z.
*——–NX NY NZ
60 1 10
CC
CC NUMBER OF WELLS AND WELLBORE MODELS: 1-) Babu and Odeh 2-) Peaceman
*——–NW IWM
21
CC
CC WELLBORE RATIUS (FT). NW WELLBORE RADIUS SHOULD BE LISTED BELOW.
*——–RW: (NW)
2*0.25
CC
CC WELL LOCATIONS. NW CARDS.
*——-LXW LYW IDIR LZWF LZWL
1 1 3 1 10
60 1 3 1 10
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable) FOR GRID BLOCK SIZE IN X-DIRECTION.
*——-MDX
0
CC
CC CONSTANT GRID BLOCK SIZE IN X-DIRECTION (FT).
*——–DX
50.0
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable ) FOR GRID BLOCK SIZE IN Y-DIRECTION.
*——-MDY
0
CC
CC CONSTANT GRID BLOCK SIZE IN Y-DIRECTION (FT).
*——–DY
100.0
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable ) FOR GRID BLOCK SIZE IN Z-DIRECTION.
*——-MDZ

137
0
CC
CC VARIABLE GRID BLOCK SIZE IN Z-DIRECTION (FT).
*——–DZ(I)
20.0
CC
CC A FLAG FOR FORMATION DEPTH.
*——–MD
0
CC
CC DEPTH (FT) OF THE MIDDLE OF THE MOST UPPER LAYER.
*———D (THETAX THETAY FOR DIPPING)
6000.
CC
CC A FLAG ( 0 -> Constant OR 1-> Variable ) FOR FORMATION POROSITY.
*——MPOR MNTG MTRANZ MMOD
00000
CC
CC HETEROGENEOUS POROSITY (FRACTION) AT FORMATION REFERENCE PRESSURE
*—-PORSTD
0.20
CC
CC A FLAG ( 0-> Constant OR 1-> Variable ) FOR PERMEABILITY IN X-DIRECTION.
*—-MPERMX
0
CC
CC HETEROGENEOUS PERMEABILITY (MD) IN X-DIRECTION.
*—–PERMX
1010.
CC
CC A FLAG ( 0-> Constant OR 1 -> Variable) FOR PERMEABILITY IN Y-DIRECTION.
*—-MPERMY
4
CC
CC
*–FACTY
1.0
CC
CC FLAG ( 0-> Constant OR 1-> Variable ) FOR PERMEABILITY IN Z-DIRECTION.
*—-MPERMZ
4
CC
CC
* –FACTZ

138
5.0
CC
CC FORMATION COMPRESSIBILITY (1/PSI) AND REFERENCE PRESSURE (PSI).
*——–CF PF
5.0e-6 14.7
CC H2O COMPRESSIBILITY (1/PSI), REFERENCE PRESSURE (PSI) AND
CC MOLAR DENSITY (LB-MOLE/CU FT).
*——–CW PW DENMWS
3.30e-6 14.70 3.467
CC
CC WATER MOLECULAR WT. (LBM/LBM-MOLE) AND VISCOSITY (CP).
*——-WTW VISCW VISCG VISCO DENG DENO
18. 0.700 4*0.0
CC
CC FORMATION TEMPERATURE (F).
*—–TEMPF
220.0
CC
CC STANDARD TEMPERATURE (F) AND STANDARD PRESSURE (PSI).
*—–TFSTD PSTD
60. 14.7
CC
CC A FLAG ( 1, 2, 3 OR 4 ) FOR NUMERICAL DISPERSION CONTROL.
*—-IUPSTW
1
CC
CC ITC ( 0 : NO 2ND ORDER TIME, 1 : 2ND ORDER TIME ON )
*—-ITC
0
CC RESTART OPTIONS.
CC ISTART ( 1 OR 2 ), ISTORE ( 0 OR 1 ).
*—-ISTART ISTORE
11
CC
CC A FLAG ( 0 OR 1 ) FOR AUTOMATIC TIME-STEP SELECTION ( = 1 ).
*——-MDT
1
CC
CC A FLAG ( 0 OR 1 ) FOR PHYSICAL DISPERSION CALCULATION.
*—–MDISP
0
CC
CC FLAGS FOR REL. PERM, CAP. PRESS., AND HYSTERESIS MODELS
*—–IPERM ICPRES ICAP IRPERM IRTYPE IHYST

139
700001
CC
CC 2-PHASE RESIDUAL SATURATIONS
*– SR1W2 SR1W3 SR2W1 SR2W3 SR3W1 SR3W2
0.195 0.195 0.232 0.232 0.33 0.33
CC
CC 2-PHASE END-POINT RELATIVE PERMEABILITIES
*– PR1W2 PR1W3 PR2W1 PR2W3 PR3W1 PR3W2
0.95 0.95 0.82 0.82 0.56 0.56
CC
CC ’C1’s FOR 2-PHASE RELATIVE PERMEABILITY CURVETURES
*– C11W2 C11W3 C12W1 C12W3 C13W1 C13W2
2.82 2.82 3.90 3.90 3.16 3.16
CC
CC ’C2’s FOR 2-PHASE RELATIVE PERMEABILITY CURVETURES
*– C21W2 C21W3 C22W1 C22W3 C23W1 C23W2
0.0 0.0 0.0 0.0 0.0 0.0
CC
CC CRITICAL AND CONNATE SATURATIONS
*– S1C S2C S3C
0.1 0.1 0.05
CC
CC PARAMETER FOR CORRELATING THE 2-PHASE AND 3-PHASE RESIDUALS
*– BSR1 BSR2 BSR3 ISR2P ISR3P
0.0 0.0 0.0 0 1
CC
CC OPTION FOR EXPLICIT THREE-PHASE RELATIVE PERMEABILITY PARAMETERS
*– IPR3P IC13P IC23P
000 000 000
CC
CC
*—– IHYSTPH1 IHYSTPH2 IHYSTPH3
004
CC
CC
*—– SATTOLHYST ILAND
0.01 0
CC
CC
*—- HYSTB1 HYSTB2 HYSTB3
0.0 0.0 1.0
CC
CC
*—- HYSTA1 HYSTA2 HYSTA3

140
0.0 0.0 1.05
CC
CC FLAGS FOR PRESSURE EQUATION SOLVER
*– IPRESS IPREC METHSL OMEGA
4 4 1 1.0
CC
CC ITERATIVE PRESSURE SOLVER PARAMETERS.
*—–ITMAX LEVLIT IDGTS NS1 NS2 ZETA
5000 1 1 5 1000000 1.E-07
CC
CC INITIAL TIME (DAYS).
*———T
0.
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL PRESSURE.
*——–MP
0
CC
CC VARIABLE INITIAL PRESSURE (PSIA).
*———P
1500.
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL WATER SATURATION.
*——MSAT
0
CC
CC VARIABLE INITIAL WATER SATURATION (FRACTION).
*——-SAT
0.45
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL OVERALL COMPOSITION.
*—–MOMFR
0
CC
CC CONSTANT INITIAL COMPOSITION (MOLE FRACTION), NC CARDS.
*——OMFR
0.0114 0.1454 0.1751 0.125 0.1261 0.1398 0.1237 0.1535
CC
CC********************************************************************
CC *
CC RECURRENT DATA *
CC *
CC********************************************************************
CC

141
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
15.0 0.0001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
0.5 0.001 0.7 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 0.0 1
CC
CC OVERALL COMPONENT
*——-KC Z1
1 1.0
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
30.0 0.001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.001 0.7 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 1.0 0

142
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
45.0 0.001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.001 0.7 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 0.0 1
CC
CC OVERALL COMPONENT
*——-KC Z1
1 1.0
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
60.0 0.0001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.0001 0.7 0.7 .001 .7

143
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 1.0 0
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC END OF INPUT.
*——–TM DT NWELLS GORLIM WORLIM —————-
-1. -1. -1 -1.E10 -1.E10
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
75.0 0.00001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.0001 0.1 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 0.0 1
CC
CC OVERALL COMPONENT
*——-KC Z1
1 1.0
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2

144
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
90.0 0.0001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.0001 0.7 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
12
CC
CC CONSTANT PRESSURE INJECTION WELL
*— PBHC FWMLC NCOMP
2000. 1.0 0
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
1000.
CC
CC END OF INPUT.
*——–TM DT NWELLS GORLIM WORLIM —————-
-1. -1. -1 -1.E10 -1.E10

145
A2: Input File for Chapter 5: Simultaneous Injection of Gas and
Water with Surfactant
**********************************************************************
CC FIELD: LAYERCAKE MODEL
CC CASE : SIMULT. INJECTION OF FOAM
CC INJECTION SCHEME : SIMULTANEOUS INJECTION OF FOAM
CC********************************************************************
CC LAYERCAKE MODEL
CC
CC RESERVOIR PROPERTIES
CC GRID BLOCKS : 60*1*10
CC GRID BLOCK SIZE : 20*50*20
CC MODEL DEPTH : 6000*FT
CC POROSITY : 0.2
CC PERMEABILITY X : VARYING LAYERING
CC KV/KH : 0.1
CC TEMPERATURE : 220 F
CC INITIAL PRESSURE : 6000 PSIA
CC INIT. WATER SAT. : 0.45
CC RELPERM MODEL : UTKR3P
CC FOAM MODEL : MECHANISTIC FOAM MODEL (IFOAM3)
CC INJECTION :SIMULT. INJECTION OF CO2 AND WATER WITH SURFACTANT
CC TIME PERIOD : 360 DAYS
CC
CC********************************************************************
CC FILE CREATED BY : VENKATESWARAN S PUDUGRAMAM
CC DATE CREATED : MAY 2013
CC
CC
CC
CC********************************************************************
CC
CC..+....1....+....2....+....3....+....4....+....5....+....6...+...7..
CC****************CONSTANTS AND VALUES FOR MODELS*********************
CC
CC IGEOM, NBM
23 0
CC
CC NXM, NYM, NZM, NPM, NCM, NWM, IXYZ, POSTPROCESSOR
60 1 10 3 8 2 1 2
CC
CC NS1M,NBWM,NPRM,NPFM,NCTM,NHSM, NPRPERMM, NS3OUT1,NS3OUT2
1 101 15 1 1 1000 3 3500 3500

146
CC
CC IC2,NCMT,NREG,NTAB
2161
CC
CC EE, NZPM
0.5761 3
CC
CC TOLP, TOLVOL, QLIM
0.5D-1 0.5D-4 1.0D-30
CC
CC NUMAX,INQUA,INCON,INVEL,NOSWTM,NUMPRE,NUMOUT,NUMPVT,NST,IFLIP
10 3 3 5 10 30 12 30 20 0
CC*********************************************************************
CC CASE NAME WITH FORMAT ( 17A4, A2 ) OF TOTAL 70 COLUMNS.
*—-HEADER
Simultaneous Injection of Gas and Water with Surfactant
CC
CC NUMBER OF COMPONENTS.
*——–NC
8
CC COMPONENT NAMES WITH FORMAT ( 1X, A8 ), NC CARDS.
CC..+..8
*—-NAME
CO2
N2-C1
C2-C3
C4-C6
C7+
C10+
C14+
C20+
CC BLACK OIL OPTION, SALINITY, AQUIFER OPTION
CC (0:OFF,1:ON) PPM (0:OFF,1:ON)
*—–IBOST SLNTY IAQUIF
0 0. 0
CC CRITICAL PRESS. (PSI), TEMP. (R) AND VOL. (CU FT/LB-MOLE),
CC MOLECULAR WT. (LB/LB-MOLE), ACENTRIC FACTOR, PARACHOR. NC CARDS.
*——–PC TC VC WT OM PARACH VSP
1069.87 547.56 1.603 44.010 0.2250 78.00 -0.05435
631.54 319.44 1.795 17.570 0.0145 69.90 -0.19268
657.00 614.01 3.452 36.510 0.1279 132.75 -0.1242
491.36 835.84 5.199 69.850 0.2423 229.61 -0.04714
409.01 1050.45 6.291 104.380 0.3533 298.42 0.12961
310.05 1196.58 9.330 154.130 0.5003 423.52 0.14725

147
250.43 1359.53 16.898 222.780 0.6836 587.09 0.13847
187.67 2020.83 29.737 516.410 0.8328 1518.16 0.0593
CC
CC NC CARDS.
*—-PARAA PARAB
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
0.457235529 0.077796074
CC
CC BINARY INTERACTION COEFFICIENTS, CIJ. NC CARDS.
*—–DELTA
0.0000
0.1110 0.0000
0.1200 0.0109 0.0000
0.1750 0.0112 0.0000 0.0000
0.0750 0.0109 0.0002 0.0035 0.0000
0.0500 0.0102 0.0000 0.0035 0.0000 0.0000
0.0700 0.0102 0.0000 0.0035 0.0005 0.0000 0.0000
0.0750 0.0102 0.0000 0.0035 0.0005 0.0000 0.0000 0.0000
CC
CC BINARY INTERACTION COEFFICIENTS, DIJ. NC CARDS.
*——-DIJ
0.000
0.000 0.000
0.000 0.000 0.000
0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000 0.000
0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
CC
CC MAXIMUM NUMBER OF PHASES ( 3 OR 4 ), IVISC AND ISINGL (0 OR 1)
*——–NP IVISC IVISC COEF ISINGL ISOLU
31110
CC
CC VISCOSITY COEFFICIENT
*—–COEF1,COEF2,COEF3,COEF4,COEF5
0.1023 0.023364 0.058533 -0.040758 0.0093324
CC IEOS: 1, IPEM: 0, 1 OR 2

148
CC
*——IEOS IPEM ISTAM IEST IVSP KI
1 1 -1 0 1 0
CC
CC ITERATION TOLERANCES FOR PERSCHKE’S FLASH ROUTINES.
*—-TOLFLA TOLFLM TOLPD TOLSAM TOLSAS TOLSUM
1.E-08 1.E-08 1.E-08 1.E-08 1.E-08 1.E-08
CC
CC MAXIMUM NUMBER OF ITERATIONS FOR PERSCHKE’S FLASH ROUTINES.
*—-MAXFLA MAXFLM MAXPD MAXSAM MAXSAS MAXSANR
1000 1000 1000 1000 1000 20
CC
CC ITERATION TOLERANCE AND MAX NUMBER OF ITERATIONS FOR VECTOR-FLASH.
*—-IVCFL TOLVFL MAXVFL
0 1.E-8 20
CC
CC SWITCHING PARAMETERS FOR PERSCHKE’S FLASH ROUTINES.
*—-SWIPCC SWIPSA
.000001 0.000001
CC
CC PHASE IDENTIFICATION PARAMETERS FOR PERSCHKE’S FLASH ROUTINES.
*——IOIL ITRK DMSLIM IZGAS IDEN
1 2 20. 0 0
CC
CC TRACER FLAG(0:OFF,1:ON), ASPHALTENE FLAG(0:OFF,1:ON)
*—-IFLAGT IASPR
10
CC
CC NUMBER OF TRACER USED
*——NCTR
1
CC..+..8.
CC NAME OF TRACER ( 1X, A8 )
*—NAMET
TRACFOAM
CC
CC
*—-ITYPET IUNIT IPGT RKT BI RLAMDA ICAPFL
1 1 0 0. 0. 0. 0
CC
CC MOL. DIFFUSION COEFF. FOR TRACER IN EACH PHASE, NP CARDS
*—-DIFUNT
0.0
0.0

149
0.0
CC
CC FLOWING FRACTION AND MASS TRANSFER COEFF., NP CARDS
*——-FFL FFH CM
0. 1. 0.
0. 1. 0.
0. 1. 0.
CC
CC DEFINE SALINITY FOR TRACER ADSORPTION
*—-NMONCT NDIVCT BETASE CSE1
1 1 0. 0.
CC
CC PARAMETERS FOR TRACER ADSORPTION
*—–A1ADT A2ADT BADT IADIRV
0.1 0. 0.05 1
CC
CC POLYMER OPTION ( 0 : OFF, 1 : ON )
*—-IFPLYT
0
CC
CC DILUTE SURFACTANT OPTION ( 0 : OFF, 1 : ON )
*—–IFOSW
0
CC
CC FOAM OPTION (0:OFF, 1:Resistance Table, 2:Pc* Model)
*—–IFOAM
3
CC
CC TRACER NUMBER FOR FOAM
*—–NFOAM
1
CC
CC PARAMETERS FOR FOAM GENERATION CRITERIA
*— AKFP0, AKFM0, AKFM20, ANFS, PCSM, CF3SR, ALPHF, XTRP, SOLIM, IFOAM3P
1.97E11 3.048 0.0 2.83E10 1.45 90.0 1.93E-11 0. 1 1
CC
CC
*— IFO3LE, BETAF, ITRAPFGC
10 0
CC
CC********************************************************************
CC *
CC OUTPUT OPTIONS *
CC *

150
CC********************************************************************
CC
CC
CC HISTORY PRINTING PARAMETER FOR <<HISTORY.CPR>>.
*— NHSSKIP NSTSKIP IPV
30 40 0
CC
CC REFERENCE CONCENTRATION, CONC0, USED FOR EFFLUENT CONCENTRATION (NC)
*—–CONC0
1. 1. 1. 1. 1. 1. 1. 1.
CC
CC NUMBER OF PRINTS FOR <<.TAB>> (ALSO FOR TRAPPING & ASPHALTENE DATA)
*——-NPR
15
CC
CC TIME(DAYS) AND FLAGS ( 0 OR 1 ) FOR <<TABLE.CPR>>. NPR CARDS.
*——-TPR MPRP MPRSAT MPRIND MPROMFR MPRPMFR MPRPRO MPRATES
2. 1 1 1 1 1 1 1
30. 1 1 1 1 1 1 1
60. 1 1 1 1 1 1 1
90. 1 1 1 1 1 1 1
120. 1 1 1 1 1 1 1
150. 1 1 1 1 1 1 1
180. 1 1 1 1 1 1 1
210. 1 1 1 1 1 1 1
240. 1 1 1 1 1 1 1
270. 1 1 1 1 1 1 1
300. 1 1 1 1 1 1 1
315. 1 1 1 1 1 1 1
330. 1 1 1 1 1 1 1
345. 1 1 1 1 1 1 1
360. 1 1 1 1 1 1 1
CC
CC NUMBER OF PRINTS FOR <<PROFILE.CPR>>.
*——-NPF
0
CC
CC NUMBER OF PRINTS FOR <<CONTOUR.CPR>>.
*——-NCT
0
CC
CC NPRPERM
*——-
0

151
CC
CC********************************************************************
CC *
CC RESERVOIR AND WELL DATA *
CC *
CC********************************************************************
CC
CC A FLAG FOR RESERVOIR GEOMETRY:
CC 1-D: 11(Y), 12(X), 13(Z), 2-D: 21(XY), 22(YZ), 23(XZ), 3-D: 31
*—–IGEOM INUG
23 0
CC
CC NUMBER OF GRID BLOCKS IN X, Y, AND Z.
*——–NX NY NZ
60 1 10
CC
CC NUMBER OF WELLS AND WELLBORE MODELS: 1-) Babu and Odeh 2-) Peaceman
*——–NW IWM
21
CC
CC WELLBORE RATIUS (FT). NW WELLBORE RADIUS SHOULD BE LISTED BELOW.
*——–RW: (NW)
2*0.25
CC
CC WELL LOCATIONS. NW CARDS.
*——-LXW LYW IDIR LZWF LZWL
1 1 3 1 10
60 1 3 1 10
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable) FOR GRID BLOCK SIZE IN X-DIRECTION.
*——-MDX
0
CC
CC CONSTANT GRID BLOCK SIZE IN X-DIRECTION (FT).
*——–DX
20.0
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable ) FOR GRID BLOCK SIZE IN Y-DIRECTION.
*——-MDY
0
CC
CC CONSTANT GRID BLOCK SIZE IN Y-DIRECTION (FT).
*——–DY
50.0

152
CC
CC A FLAG ( 0 -> Constant OR 1 -> Variable ) FOR GRID BLOCK SIZE IN Z-DIRECTION.
*——-MDZ
0
CC
CC VARIABLE GRID BLOCK SIZE IN Z-DIRECTION (FT).
*——–DZ(I)
20.0
CC
CC A FLAG FOR FORMATION DEPTH.
*——–MD
0
CC
CC DEPTH (FT) OF THE MIDDLE OF THE MOST UPPER LAYER.
*———D (THETAX THETAY FOR DIPPING)
6000.
CC
CC A FLAG ( 0 -> Constant OR 1-> Variable ) FOR FORMATION POROSITY.
*——MPOR MNTG MTRANZ MMOD
00000
CC
CC HETEROGENEOUS POROSITY (FRACTION) AT FORMATION REFERENCE PRESSURE
*—-PORSTD
0.20
CC
CC A FLAG ( 0-> Constant OR 1-> Variable ) FOR PERMEABILITY IN X-DIRECTION.
*—-MPERMX
2
CC
CC HETEROGENEOUS PERMEABILITY (MD) IN X-DIRECTION.
*—–PERMX
100. 80.0 40.0 30.0 25.0 30.0 30.0 35.0 70.0 40.0
CC
CC A FLAG ( 0-> Constant OR 1 -> Variable) FOR PERMEABILITY IN Y-DIRECTION.
*—-MPERMY
4
CC
CC
*–FACTY
1.0
CC
CC FLAG ( 0-> Constant OR 1-> Variable ) FOR PERMEABILITY IN Z-DIRECTION.
*—-MPERMZ
4

153
CC
CC
* –FACTZ
0.1
CC
CC FORMATION COMPRESSIBILITY (1/PSI) AND REFERENCE PRESSURE (PSI).
*——–CF PF
5.0e-6 14.7
CC H2O COMPRESSIBILITY (1/PSI), REFERENCE PRESSURE (PSI) AND
CC MOLAR DENSITY (LB-MOLE/CU FT).
*——–CW PW DENMWS
3.30e-6 14.70 3.467
CC
CC WATER MOLECULAR WT. (LBM/LBM-MOLE) AND VISCOSITY (CP).
*——-WTW VISCW VISCG VISCO DENG DENO
18. 0.700 4*0.0
CC
CC FORMATION TEMPERATURE (F).
*—–TEMPF
220.0
CC
CC STANDARD TEMPERATURE (F) AND STANDARD PRESSURE (PSI).
*—–TFSTD PSTD
60. 14.7
CC
CC A FLAG ( 1, 2, 3 OR 4 ) FOR NUMERICAL DISPERSION CONTROL.
*—-IUPSTW
1
CC
CC ITC ( 0 : NO 2ND ORDER TIME, 1 : 2ND ORDER TIME ON )
*—-ITC
0
CC RESTART OPTIONS.
CC ISTART ( 1 OR 2 ), ISTORE ( 0 OR 1 ).
*—-ISTART ISTORE
11
CC
CC A FLAG ( 0 OR 1 ) FOR AUTOMATIC TIME-STEP SELECTION ( = 1 ).
*——-MDT
1
CC
CC A FLAG ( 0 OR 1 ) FOR PHYSICAL DISPERSION CALCULATION.
*—–MDISP
0

154
CC
CC FLAGS FOR REL. PERM, CAP. PRESS., AND HYSTERESIS MODELS
*—–IPERM ICPRES ICAP IRPERM IRTYPE IHYST
710000
CC CAP. PRESS. PARAMETERS
CC
*——-EPC CPC
2.0 0.5
CC INTERFACIAL TENSIONS (DYNES/CM).
CC
* — RIFTWO RIFTWG
30 30
CC
CC 2-PHASE RESIDUAL SATURATIONS
*– SR1W2 SR1W3 SR2W1 SR2W3 SR3W1 SR3W2
0.00 0.00 0.00 0.00 0.30 0.30
CC
CC 2-PHASE END-POINT RELATIVE PERMEABILITIES
*– PR1W2 PR1W3 PR2W1 PR2W3 PR3W1 PR3W2
0.20 0.20 0.70 0.70 0.90 0.90
CC
CC ’C1’s FOR 2-PHASE RELATIVE PERMEABILITY CURVETURES
*– C11W2 C11W3 C12W1 C12W3 C13W1 C13W2
1.5 1.5 2.5 2.5 3.5 3.5
CC
CC ’C2’s FOR 2-PHASE RELATIVE PERMEABILITY CURVETURES
*– C21W2 C21W3 C22W1 C22W3 C23W1 C23W2
0.0 0.0 0.0 0.0 0.0 0.0
CC
CC CRITICAL AND CONNATE SATURATIONS
*– S1C S2C S3C
0.1 0.1 0.05
CC
CC PARAMETER FOR CORRELATING THE 2-PHASE AND 3-PHASE RESIDUALS
*– BSR1 BSR2 BSR3 ISR2P ISR3P
1.0 1.0 1.0 0 1
CC
CC OPTION FOR EXPLICIT THREE-PHASE RELATIVE PERMEABILITY PARAMETERS
*– IPR3P IC13P IC23P
000 000 000
CC
CC FLAGS FOR PRESSURE EQUATION SOLVER
*—-IPRESS IPREC METHSL OMEGA
4 4 1 1.0

155
CC
CC ITERATIVE PRESSURE SOLVER PARAMETERS.
*—–ITMAX LEVLIT IDGTS NS1 NS2 ZETA
100 1 1 5 1000000 1.E-07
CC
CC INITIAL TIME (DAYS).
*———T
0.
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL PRESSURE.
*——–MP
0
CC
CC VARIABLE INITIAL PRESSURE (PSIA).
*———P
6000.
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL WATER SATURATION.
*——MSAT
0
CC
CC VARIABLE INITIAL WATER SATURATION (FRACTION).
*——-SAT
0.45
CC
CC A FLAG ( 0 OR 1 ) FOR INITIAL OVERALL COMPOSITION.
*—–MOMFR
0
CC
CC CONSTANT INITIAL COMPOSITION (MOLE FRACTION), NC CARDS.
*——OMFR
0.0114 0.1454 0.1751 0.125 0.1261 0.1398 0.1237 0.1535
CC
CC NO. OF TRACER INITIALLY PRESENTED
*—–NCINT
0
CC
CC********************************************************************
CC *
CC RECURRENT DATA *
CC *
CC********************************************************************
CC
CC

156
CC MAXIMUM TIME (DAYS), TIME STEP (DAYS) AND WELL DATA.
*—-TM DT NWELLS GORLIM WORLIM
360.0 0.001 2 -1.E+20 -1.E+10
CC
CC PARAMETERS FOR TIME STEP SELECTORS.
*—–DTMAX DTMIN DSLIM DPLIM DVLIM DMFACT
1.0 0.001 0.7 0.7 .001 .7
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
14
CC
CC CONSTANT RATE INJECTION WELL
*— WATER RATE(STB/D) GAS RATE(MSCF/D) NCOMP ISWITCH PBHC(PSI)
200.0 250.0 1 1 6500
CC
CC OVERALL COMPONENT
*——-KC Z1
1 1.0
CC
CC NUMBER OF TRACER INJECTED
*—-NCOMPT
1
CC
CC TRACER NUMBER AND CONCENTRATION OF TRACER INJECTED
*——KCTR Z1TR
1 5000
CC
CC WELL NO. AND WELL TYPE
*——–LW IQTYPE
2 -2
CC
CC CONSTANT PRESSURE PRODUCTION WELL
*— PBHC
600.
CC
CC END OF INPUT.
*——–TM DT NWELLS GORLIM WORLIM —————-
-1. -1. -1 -1.E10 -1.E10

157
References

Acs, G., Doleschall, S., and Farkas, E. 1985. General Purpose Compositional Model.
SPE J. 25 (4): 543-553. SPE-10515-PA. doi:10.2118/10515-PA
Ahmadloo, F., Ashgari, K., and Yadali Jamaloei, B. 2009. Experimental and Theoretical
Studies of Three-Phase Relative Permeability. Paper SPE 124538 presented at the
2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana,
USA, 4-7 October. doi:10.2118/124538-MS
Aziz, K., and Settari, K. 1979. Petroleum Reservoir Simulation. London: Applied
Science Publisher.
Baker, L. 1988. Three-phase relative permeability correlations. Paper SPE 17639 pre-
sented at the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, Oklahoma,
17-20 April. doi:10.2118/17369-MS
Balbinski, E. F., Fishlock, T. P., Goodyear, S. G., and Jones, P. I. R. 1999. Key character-
istics of three-phase oil relative permeability formulations for improved oil recovery
predictions. Petroleum Geoscience 5 (4): 339-346. doi:10.1144/petgeo.5.4.339
Basry, A. Al, Hajeri, S. Al, Saadawi, H., Aryani, F. Al, Obeidi, A., Negahban, S., and
Yafei, G. Al. 2011. Lessons Learned from the First Miscible CO2 -EOR Pilot
Project in Heterogeneous Carbonate Oil Reservoir in Abu Dhabi, UAE. Paper
SPE 142665 presented at the SPE Middle East Oil and Gas Show and Conference,
Manama, Bahrain, 25-28 September. doi:10.2118/142665-MS
Beliveau, D. 1987. Midale CO2 Flood Pilot. J. Cdn. Pet. Tech. 26 (6): 66-69.
doi:10.2118/87-06-05
Bertin, H. J., Quintard, M. Y., and Castanier, L. M. 1998. Development of a Bubble-
Population Correlation for Foam-Flow Modeling in Porous Media. SPE J. 3 (4):
356-362. SPE-52596-PA. doi:10.2118/52596-PA
Beygi, M. R., Delshad, M., Pudugramam, V. S., Pope, G. A., and Wheeler, M. F. 2013.
A New Approach to Model Hysteresis and Its Impact on CO2 -EOR Processes
with Mobility Control Strategies. Paper SPE 165324 presented at the SPE West-

158
ern Regional and AAPG Pacific Section Joint Technical Conference, Monterey,
California, USA, 19-25 April. doi:10.2118/165324-MS
Biu Victor, T. 2011. Simulation of CO2 Enhanced Oil Recovery in Heterogeneous Models.
Paper SPE 150807 presented at the Nigeria Annual International Conference and
Exhibition, Abuja, Nigeria, 30 July - 3 August. doi:10.2118/150807-MS
Blunt, M. J. 2000. An Empirical Model for Three-Phase Relative Permeability. SPE J.
5 (4): 435-445. SPE-67950-PA. doi:10.2118/67950-PA
Brinkman, F. P., Kane, T. V, Mccullough, R. F. L., and Miertschin, J. W. 1998. Use of
Full-Field Simulation to Design a Miscible CO2 Flood. Paper SPE 39629 presented
at the SPE/DOE Improved Oil Recovery Synposium, Tulsa, Oklahoma, 19-22
April. doi:10.2118/39629-MS
Carlson, F. 1981. Simulation of Relative Permeability Hysteresis to the Nonwetting
Phase. Paper SPE 10157 presented at the 56th Annual Fall Technical Conference
and Exhibition, San Antonio, Texas, USA, 5-7 October. doi:10.2118/10157-MS
Chang, S., Owusu, L., French, S. B., and Kovarik, F. S. 1990. The Effect of Microscopic
Heterogeneity on CO2 -Foam Mobility: Part 2-Mechanistic Foam Simulation. Pa-
per SPE 20191 presented at the SPE/DOE Seventh Symposium on Enhanced Oil
Recovery, Tulsa, Oklahoma, 22-25 April. doi:10.2118/20191-MS
Chang, Y.-B. 1990. Development and application of an equation of state compositional
simulator. PhD dissertation, The University of Texas at Austin.
Chen, Q., Gerritsen, M. G., and Kovscek, A. R. 2010. Modeling Foam Displacement With
the Local-Equilibrium Approximation?: Theory and Experimental Verification.
SPE J. 15 (1): 171-183. SPE-116735-PA. doi:10.2118/116735-PA
Cheng, L., Reme, A. B., Shan, D., Coombe, D. A., and W.R., R. 2000. Simulating
Foam Processes at High and Low Foam Qualities. Paper SPE 59287 presented
at SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, 3-5 April.
doi:10.2118/59287-MS
Chou, S. I. 1990. Percolation Theory of Foam in Porous Media. Paper SPE 20239
presented at the SPE/DOE Seventh Symposium on Enhanced Oil Recovery, Tulsa,
Oklahoma, 22-25 April. doi:10.2118/20239-MS
Christensen, J. R., Stenby, E. H., and Skauge, A. 1998. Compositional and Relative Per-
meability Hysteresis Effects on Near-Miscible WAG. Paper SPE 39627 presented
159
at the SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, 19-22
April. doi:10.2118/39627-MS
Corey, A., Rathjens, C., Henderson, J., and Wyllie, M. 1956. Three-phase relative
permeability. J. Pet Tech 8 (11): 63-65. SPE-737-G. doi:10.2118/737-G
Davis, D., Scott, M., Roberson, K., and Robinson, A. 2011. Large Scale CO2 Flood Be-
gins Along Texas Gulf Coast. Paper SPE 144961 presented at the SPE Enhanced
Oil Recovery Conference, Kuala Lumpur, Malaysia, 19-21 July. doi:10.2118/144961-
MS
Dawson, A. G., Jackson, D. D., and Buskirk, D. L. 1989. Impact of Solvent Injection
Strategy and Reservoir Description on Hydrocarbon Miscible EOR for the Prud-
hoe Bay Unit, Alaska. Paper SPE 19657 presented at the Annual Technical Con-
ference and Exhibition, San Antonio, Texas, 8-11 October. doi:10.2118/19657-MS
Delshad, M., and Pope, G. 1989. Comparison of the three-phase oil relative permeability
models. Transport in Porous Media 4 : 59-83. doi:10.1007/BF00134742
Delshad, M., Pope, G. A., and Lake, L. W. 1987. Two- and Three-Phase Relative
Permeabilities of Micellar Fluids. SPE Form Eval 2 (3): 327-337. SPE-13581-
PA. doi:10.2118/13581-PA
Dria, D., Pope, G., and Sepehrnoori, K. 1993. Three-phase Gas/Oil/Brine Relative
Permeabilities measured under CO2 Flooding Conditions. SPE Res Eng 8 (2):
143-150. SPE-20184-PA. doi:10.2118/20184-PA
Egermann, P., Vizika, O., Dallet, L., Requin, C., and Sonier, F. 2000. Hysteresis in
Three-Phase Flow: Experiments, Modeling and Reservoir Simulations. Paper
SPE 65127 presented at the SPE European Petroleum Conference, Paris, France,
24-25 October. doi:10.2118/65127-MS
Element, D. J., Masters, J. H. K., Sargent, N. C., Jayasekara, A. J., and Goodyear,
S. G. 2003. Assessment of three-phase relative permeability models using lab-
oratory hysteresis data. Paper SPE 84903 presented at the SPE International
Improved Oil Recovery Conference, Kuala Lumpur, Malaysia, 20-21 October.
doi:10.2118/84903-MS
Falls, A., Hirasaki, G., Patzek, T., Gauglitz, P., Miller, D., and Ratulowski, T. 1988. De-
velopment of a Mechanistic Foam Simulator: The Population Balance and Genera-

160
tion by Snap-Off. SPE Res Eng 3 (3): 884-892. SPE-14961-PA. doi:10.2118/14961
-PA
Farajzadeh, R., Andrianov, A., Krastev, R., Hirasaki, G. J., and Rossen, W. R. 2012.
Foam-oil interaction in porous media: implications for foam assisted enhanced oil
recovery. Advances in colloid and interface science 183-184 : 1-13. doi:10.1016/
j.cis.2012.07.002
Fayers, F. 1989. Extension of Stone’s method 1 and conditions for real character-
istics in three-phase flow. SPE Res Eng 4 (4): 437-445. SPE-16965-PA.
doi:10.2118/16965-PA
Fayers, F., and Foakes, A. 2000. An Improved Three Phase Flow Model Incorporating
Compositional Variance. Paper SPE 59313 presented at the SPE/DOE Improved
Oil Recovery Symposium, Tulsa, Oklahoma, USA, 3-5 April. doi:10.2118/59313-
MS
Fayers, F., and Matthews, J. 1984. Estimation of Normalized Stone’s Methods for Es-
timating Three-Phase Relative Permeabilities. SPE J. 24 (2): 224-232. SPE-
11277-PA. doi:10.2118/11277-PA
Fergui, O., Quintard, M., Bertin, H., and Defives, D. 1995. Transient Foam Flow in
Porous Media: Experiments and Simulation. In Proceedings of the 8th European
IOR Symposium, Vienna, Austria.
Fisher, A., Foulser, R., and Goodyear, S. 1990. Mathematical modeling of foam flooding.
Paper SPE 20195 presented at the SPE/DOE Seventh Symposium on Enhanced
Oil Recovery, Tulsa, Oklahoma, 22-25 April. doi:10.2118/20195-MS
Friedmann, F., Hughes, T., Smith, M., Hild, G., Wilson, A., and Davies, S. 1997. De-
velopment and Testing of a New Foam-Gel Technology to Improve Conformance
of the Rangely CO2 Flood. Paper SPE 38837 presented at the SPE Annual
Technical Conference and Exhibition, San Antonio, Texas, USA, 5-8 October.
doi:10.2118/38837-MS
Friedmann, F., Chen, W., and Gauglitz, P. 1991. Experimental and Simulation Study
of High-Temperature Foam Displacement in Porous Media. SPE Res Eng 6 (1):
37-45. SPE-17357-PA. doi:10.2118/17357-PA
Gai, X. 2004. A coupled Geomechanics and Reservoir Flow Model on Parallel Computers.
PhD dissertation, The University of Texas at Austin.
161
Genetti, D., Whitaker, C., Smith, D., and Price, L. 2003. Applying Improved Recovery
Processes and Effective Reservoir Management to Maximize Oil Recovery at Salt
Creek. Paper SPE 81458 presented at the Middle East Oil Show and Conference,
Bahrain, 5-8 April. doi:10.2118/81458-MS
Ghomian, Y., Pope, G., and Sepehrnoori, K. 2008. Hysteresis and field-scale optimization
of WAG injection for coupled CO2 -EOR and sequestration. Paper SPE 110639
presented at the SPE/DOE Improved Oil Recovery Symposium , Tulsa, Okla-
homa, 19-23 April. doi:10.2118/110639-MS
Ghomian, Y., Urun, M., and Pope, G. A. 2008. Investigation of Economic Incentives for
CO2 Sequestration. Paper SPE 116717 presented at the SPE Annual Technical
Conference and Exhibition, Denver, Colorado, USA, 21-24 September. doi:10.2118
/116717-MS
Goodyear, S., Hawkyard, I., Masters, J. H. K., Woods, C. L., Jayasekera, A. J., and
Balbinski, D. E. 2003. Subsurface Issues for CO2 Flooding of UKCS Reser-
voirs. Chemical Engineering Research and Design 81 (3): 315-325. doi:10.1205/
02638760360596865
Harpole, K., and Hallenbeck, L. 1996. East Vacuum Grayburg San Andres Unit CO2
Flood Ten Year Performance Review: Evolution of a Reservoir Management
Strategy and Results of WAG Optimization. Paper SPE 36710 presented at
the SPE Annual Technical Conference and Exhibition, Denver, Colorado, US.
doi:10.2118/36710-MS
Hervey, J. R., and Iakovakis, A. C. 1991. Performance Review of a Miscible CO2 tertiary
project: Rangely Weber Sand Unit, Colorado. SPE Res Eng 6 (2): 163-168.
SPE-19653-PA. doi:10.2118/19653-PA
Hill, W., Tinney, T., Young, L., and Stark, K. 1994. CO2 Operating plan, South Welch
Unit, Dawson County, Texas. Paper SPE 27676 presented at the SPE Permian
Basin Oil and Gas Recovery Conference, Midland, Texas, USA, 16-18 March.
doi:10.2118/27676-MS
Hoiland, R., Joyner, H., and Stalder, J. 1986. Case History of a Successful Rocky Moun-
tain Pilot CO2 Flood. Paper SPE 14939 presented at the SPE/DOE Fifth Sympo-
sium on Enhanced Oil Recovery, Tulsa, Oklahoma, 20-23 April. doi:10.2118/14939-
MS

162
Holtz, M. H. 2008. Summary of Gulf Coast Sandstone CO2 EOR Flooding Application
and Response. SPE 113368 presented at the SPE/DOE Improved Oil Recovery
Symposium, Tulsa, Oklahoma, USA, 19-23 April. doi:10.2118/113368-MS
Honarpour, M., Nagarajan, N. R., Grijalba, A. C., Valle, M., and Adesoye, K. 2010.
Rock-Fluid Characterization for Miscible CO2 Injection: Residual Oil Zone, Semi-
nole Field, Permian Basin. Paper SPE 133089 presented at the SPE Annual Tech-
nical Conference and Exhibition, Florence, Italy, 19-22 September. doi:10.2118/133089-
MS
Hsu, C., Morell, J., and Falls, A. 1997. Field-Scale CO2 Flood Simulations and Their
Impact on the Performance of the Wasson Denver Unit. SPE Res Eng 12 (1):
4-11. SPE-29116-PA. doi:10.2118/29116-PA
Hustad, O. 2000. A Coupled Model for Three-Phase Capillary Pressure and Relative Per-
meability. Paper SPE 63150 presented at the SPE Annual Technical Conference
and Exhibition, Dallas, Texas, USA, 1-4 October. doi:10.2118/63150-MS
Hustad, O. S., and Holt, T. 1992. Gravity Stable Displacement of Oil by Hydrocar-
bon Gas After Waterflooding. Paper SPE 24116 presented at the SPE/DOE
Eighth Symposium of Enhanced Oil Recovery, Tulsa, Oklahoma, USA, 22-24
April. doi:10.2118/24116-MS
Islam, M. R., and Farouq Ali, S. M. 1990. Numerical simulation of foam flow in porous
media. J. Cdn. Pet. Tech. 29 (47-51).
Jerauld, G. 1997. General Three-Phase Relative Permeability Model for Prudhoe Bay.
SPE Res Eng 12 (4): 255-263. SPE-36178-PA. doi:10.2118/36178-PA
Jeschke, P., Schoeling, L., and Hemmings, J. 2000. CO2 flood potential of califor-
nia oil reservoirs and possible CO2 sources. Paper SPE 63305 presented at the
SPE/AAPG Western Regional Meeting, Long Beach, California, USA, 19-23 June.
doi:10.2118/63305-MS
Johnston, J. 1988. Weeks Island gravity stable CO2 pilot. Paper SPE 17351 presented
at the SPE/DOE Enhanced Oil Recovery Symposium, Tulsa, Oklahoma, 17-20
April. doi:10.2118/17351-MS
Khatib, Z., Hirasaki, G., and Falls, A. 1988. Effects of Capillary Pressure on Coalescence
and Phase Mobilities in Foams Flowing Through Porous Media. SPE Res Eng 3
(3): 919-926. SPE-15442-PA. doi:10.2118/15442-PA
163
Killough, J. E. 1976. Reservoir Simulation With History-Dependent Saturation Func-
tions. SPE J. 16 (1): 37-48. SPE-5106-PA. doi:10.2118/5106-PA
Koottungal, L. 2010, April. 2010 worldwide EOR survey. Oil and Gas Journal.
Kovscek, A. R., Patzek, T. W., and Radke, C. J. 1995. A Mechanistic Population Balance
Model for Transient and Steady-State Foam Flow in Boise Sandstone. Chemical
Engineering Science 50 (23): 3783-3799. doi:10.1016/0009-2509(95)00199-F
Land, C. 1968. Calculation of Imbibition Relative Permeability for Two- and Three-
Phase Flow From Rock Properties. SPE J. 8 (2): 149-156. SPE-1942-PA.
doi:10.2118/1942-PA
Larsen, J., and Skauge, A. 1998. Methodology for numerical simulation with cycle-
dependent relative permeabilities. SPE J. 3 (2): 163-173. SPE-38456-PA.
doi:10.2118/38456 -PA
Marfoe, C., and Kazemi, H. 1987. Numerical Simulation of Foam Flow in Porous Me-
dia. Paper SPE 16709 presented at the SPE Annual Technical Conference and
Exhibition, Dallas, Texas, USA, 27-30 September. doi:10.2118/16709-MS
Masalmeh, S., and Wei, L. 2010. Impact of Relative Permeability Hysteresis, IFT depen-
dent and Three Phase Models on the Performance of Gas Based EOR Processes.
Paper SPE 138203 presented at the Abu Dhabi International Exhibition and Con-
ference, Abu Dhabi, UAE, 1-4 November. doi:10.2118/138203-MS
Masoner, L., and Wackowski, R. 1995. Rangely Weber Sand Unit CO2 Project Update.
SPE Res Eng 10 (3): 203-207. SPE-27755-PA. doi:10.2118/27755-PA
Merchant, D. 2010. Life Beyond 80: A Look at Conventional WAG Recovery Beyond
80% HCPV Injected in CO2 Tertiary Floods. Paper SPE 139516 presented at
the SPE International Conference on CO2 Capture, Storage and Utilization, New
Orleans, Louisiana, USA, 10-12 November. doi:10.2118/139516-MS
Mohammadi, S. S., Coombe, D. A., and Stevenson, V. M. 1993. Test of Steam-foam
Process for Mobility Control in South Casper Creek Reservoir. J. Cdn. Pet.
Tech. 32 (10): 49-54. doi:10.2118/93-10-06
Mollaei, A., and Delshad, M. 2011. A Novel Forecasting Tool for Water Alternating
Gas (WAG) Floods. Paper SPE 148742 presented at the SPE Eastern Regional
Meeting, Columbus, Ohio, USA, 17-19 August. doi:10.2118/148742-MS

164
Naderi Beni, A., Varavei, A., Delshad, M., and Farajzadeh, R. 2013. Modeling Gas
Solubility in Water for Foam Propagation in Porous Media. Paper SPE 163579
presented at the SPE Reservoir Simulation Symposium, The Woodlands, Texas,
USA, 18-20 February. doi:10.2118/163579-MS
Nghiem, L. X., and Coombe, D. A. 1997. Modeling Asphaltene Precipitation During
Primary Depletion. SPE J. 2 (2): 170-176. doi:10.2118/36106-PA
Oak, M. J. 1990. Three-phase relative permeability of water-wet Berea. Paper SPE 20183
presented at the SPE/DOE Seventh Symposium on Enhanced Oil Recovery, Tulsa,
Oklahoma, 22-25 April. doi:10.2118/20183-MS
Oak, M. J. 1991. Three-phase relative permeability of intermediate-wet Berea sand-
stone. Paper SPE 22599 presented at the SPE Annual Technical Conference and
Exhibition, Dallas, Texas, USA, 8-9 October. doi:10.2118/22599-MS
Oudinot, A. Y., Koperna, G. J., Z.G., P., Liu, N., Heath, J., Wells, A., . . . Wilson, T.
2009. CO2 Injection Performance in the Fruitland Coal Fairway, San Juan Basin:
Results of a Field Pilot. Paper SPE 127073 presented at the SPE International
Conference on CO2 Capture, Storage and Utilization, San Diego, California, USA,
2-4 November. doi:10.2118/127073-MS
Patzek, T., and Myhill, N. 1989. Simulation of the Bishop Steam Foam Pilot. Paper SPE
18786 presented at the SPE California Regional Meeting, Bakersfield, California,
USA, 5-7 April. doi:10.2118/18786-MS
Patzek, T. W. 1966. Description of Foam Flow in Porous Media by the Population
Balance Method. Paper SPE 16321 submitted and published unsolicited with
permission from author.
Pejic, D., and Maini, B. B. 2003. Three-Phase Relative Permeability of Petroleum
Reservoirs. Paper SPE 81021 presented at the SPE Latin American and Car-
ribean Petroleum Engineering Conference, Trinidad, West Indies, 27-30 April.
doi:10.2118 /81021-MS
Perry, G. E. 1982. Weeks Island “ S” Sand Reservoir B Gravity Stable Miscible CO2 Dis-
placement, Iberia Parish, Louisiana. Paper SPE 10695 presented at the SPE/DOE
Third Joint Symposium on Enhanced Oil Recovery, Tulsa, Oklahoma, 4-7 April.
doi:10.2118/10695-MS

165
Robinson, D., and Peng, D. 1978. The characterization of the heptanes and heavier
fractions for the GPA Peng-Robinson programs. Gas Processors Association.
Rosman, A., Riyadi, S., and Kifli, A. 2011. Oil Recovery Optimization by Immiscible
WAG in Offshore Mature Field?: Dulang Case Study. Paper SPE 144531 pre-
sented at the SPE Enhanced Oil Recovery Conference, Kuala Lumpur, Malaysia,
19-21 July. doi:10.2118/144531-MS
Rossen, W. 1996. Foams in Enhanced Oil Recovery. In Foams: Theory, Measurement
and Applications. New York: Marcel Dekker.
Rossen, W. R., Zeilinger, S. C., Shi, J., and Lim, M. T. 1994. Mechanistic Simulation
of Foam Processes in Porous Media. Paper SPE 28940 presented at the SPE
Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 25-
28 September. doi:10.2118/28940-MS
Sahin, S., Kalfa, U., and Celebioglu, D. 2008. Bati Raman Field Immiscible CO2 Ap-
plication - Status Quo and Future Plans. SPE Res Eval & Eng 11 (4): 778-791.
SPE-106575-PA. doi:10.2118/106575-PA
Senocak, D., Pennell, S., Gibson, C., and Hughes, R. 2008. Effective Use of Heterogeneity
Measures in the Evaluation of a Mature CO2 Flood. Paper SPE 39660 presented
at the SPE/DOE Improved Oil Recovery Symposium, Tulsa, Oklahoma, USA,
19-23 April. doi:10.2118/113977-MS
Shahverdi, H., Sohrabi, M., Fatemi, M., Jamiolahmady, M., Irelan, S., and Robertson,
G. 2011. Evaluation of Three-Phase Relative Permeability Models for WAG In-
jection Using Water-Wet and Mixed-Wet Core Flood Experiments. Paper SPE
143030 presented at the SPE EUROPEC/EAGE Annual Conference and Exhibi-
tion, Vienna, Austria, 23-26 May. doi:10.2118/143030-MS
Skjaeveland, S., and Kleppe, J. 1992. Recent Advances in Improved Oil Recovery Meth-
ods for North Sea Sandstone Reservoirs. Stavanger: SPOR Monograph Series,
Norwegian Petroleum Directorate.
Spiteri, E, Juanes, R., Blunt, M., and Orr, F. 2005. Relative-Permeability Hysteresis:
Trapping Models and Application to Geological CO2 Sequestration. Paper SPE
96448 presented at the SPE Annual Technical Conference and Exhibition, Dallas,
Texas, USA, 9-12 October. doi:10.2118/96448-MS

166
Spiteri, EJ, and Juanes, R. 2004. Impact of relative permeability hysteresis on the numer-
ical simulation of WAG injection. Paper SPE 89921 presented at the SPE Annual
Technical Conference and Exhibition, Houston, Texas, USA, 26-29 September.
doi:10.2118/89921-MS
Stein, M., Frey, D., Walker, R., and Pariani, G. 1992. Slaughter Estate Unit CO2 flood:
Comparison between Pilot and Field-Scale Performance. J. Pet Tech 44 (9):
1026-1032. SPE-19375-PA. doi:10.2118/19375-PA
Stone, H. 1973. Estimation of three-phase relative permeability and residual oil data. J.
Pet Tech 12 (4): 52-61.
Stone, H. L. 1970. Probability Model for Estimating Three-Phase Relative Permeability.
J. Pet Tech 22 (2): 214-218. SPE-2116-PA. doi:10.2118/2116-PA
Surguchev, L. M., Korbel, R., Haugen, S., and Krakstad, O. S. 1992. Screening of
WAG Injection Strategies for Heterogeneous Reservoirs. Paper SPE 25075 pre-
sented at the European Petroleum Conference, Cannes, France, 16-18 November.
doi:10.2118/25075-MS
Sweatman, R. E., Parker, M. E., and Crookshank, S. L. 2009. Industry Experience
With CO2 -Enhanced Oil Recovery Technology. Paper SPE 126446 presented at
the SPE International Conference on CO2 Capture, Storage and Utilization, San
Diego, California, USA, 2-4 November. doi:10.2118/126446-MS
Turek, E. A., Metcalfe, R., Yarborough, L., and Robinson Jr., R. 1984. Phase Equi-
libria in CO2 -Multicomponent Hydrocarbon Systems: Experimental Data and
an Improved Prediction Technique. SPE J. 24 (3): 308-324. SPE-9231-PA.
doi:10.2118/9231-PA
Vassenden, F., and Holt, T. 1998. Experimental foundation for relative permeability
modeling of foam. Paper SPE 39660 presented at the SPE/DOE Improved Oil
Recovery Symposium, Tulsa, Oklahoma, USA, 19-22 April. doi:10.2118/39660-
MS
Vikas. 2002. Simulation of CO2 Sequestration. MS thesis, The University of Texas at
Austin.
Wassmuth, F., Green, K., and Hodgins, L. 2005. Conformance Control for Miscible CO2
Floods in Fractured Carbonates. Paper presented at the 6th Canadian Interna-

167
tional Petroleum Conference, Calgary, Alberta, Canada, 7-9 June. doi:10.2118/
2005-243
Yuan, C., and Pope, G. 2012. A New Method To Model Relative Permeability in
Compositional Simulators To Avoid Discontinuous Changes Caused by Phase-
Identification Problems. SPE J. 17 (4): 1221-1230. SPE-142093-PA. doi:10.2118/142093-
PA
Zhou, Z. H., and Rossen, W. R. 1995. Applying Fractional-Flow Theory to Foam Pro-
cesses at the “ Limiting Capillary Pressure”. SPE Advanced Technology Series 3
(1): 154-162. SPE-24180-PA. doi:10.2118/24180-PA

168

You might also like