You are on page 1of 137

Copyright

by

Abdulhamid Alsousy

2023

1
The Thesis Committee for Abdulhamid Alsousy
Certifies that this is the approved version of the following Thesis:

A Compositional Reservoir Simulation Study to Evaluate Impacts of

Captured CO2 Composition, Miscibility, and Injection Strategy on CO2-

EOR and Sequestration in a Carbonate Oil Reservoir

APPROVED BY
SUPERVISING COMMITTEE:

Kamy Sepehrnoori, Supervisor

Mojdeh Delshad, Co-Supervisor

2
A Compositional Reservoir Simulation Study to Evaluate Impacts of

Captured CO2 Composition, Miscibility, and Injection Strategy on CO2-

EOR and Sequestration in a Carbonate Oil Reservoir

by

Abdulhamid Alsousy

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 2023

3
Dedication

‫بسم هللا الرحمن الرحيم‬

To my family and my son Sultan, for their continuous support and encouragement.

4
Acknowledgements

I would like to express my heartfelt gratitude to Dr. Kamy Sepehrnoori, and Dr.

Mojdeh Delshad for giving me the opportunity to work under their guidance and

supervision. Their invaluable insight and continuous encouragement throughout have been

instrumental to this work.

I would like to extend my thanks to Dr. Marcos Machado for his support and advice.

Dr. Machado offered his time and expertise in CMG, assisting me in establishing a robust

numerical model and its numerical fine-tuning.

Last but not least, I am thankful to my sponsor Saudi Aramco, which made it

possible for me to join UT Austin’s Graduate program.

5
Abstract

A Compositional Reservoir Simulation Study to Evaluate Impacts of

Captured CO2 Composition, Miscibility, and Injection Strategy on CO2-

EOR and Sequestration in a Carbonate Oil Reservoir

Abdulhamid Alsousy, M.S.E

The University of Texas at Austin, 2023

Supervisor: Kamy Sepehrnoori

Co-Supervisor: Mojdeh Delshad

As the global energy demand rises, concerns regarding the increasing carbon levels

deepen. Pushing the international community to pour their time and resources into

exploring all avenues that bear potential to aid the decarbonization efforts. The

decarbonization efforts attempt to either reduce carbon dioxide emissions or to capture


carbon dioxide from the atmosphere. The oil and gas industry’s role falls into the first

category. Where captured CO2 is sequestered into geological stable formations as part of

carbon capture, utilization, and storage (CCUS) or carbon capture and storage (CCS)

projects. CCUS and CCS technologies hold the keys to decarbonization, possessing a large

capacity capable of storing over 8000 GtCO2, utilizing oil and gas reservoirs, saline

aquifers, and coal beds to discard CO2. In addition, the sequestration in geological

6
structures is long-term, with minimal risk of reintroducing the stored gas back to the

surface.

This work investigates two scenarios, one in which the reservoir undergoes a

tertiary production and another where the reservoir has reached the abandonment stage of

its life cycle. The analyses are carried out by employing a historically matched numerical

model of a real carbonate reservoir to explore CO2 storage implications on the reservoir’s

performance (EOR) and the efficiency of the injected gas storage in the subsurface. For a

holistic evaluation, the numerical model accounts for relative permeability hysteresis,

phase trapping, geochemistry, and thermodynamics. Various analyses are conducted to

establish the recommended gas blend injected, the importance of miscibility, and the

manner of injection (WAG or gas flood). The results showcased how miscible injection

outperforms immiscible in CO2-EOR and sequestration efficiency. Furthermore, gas flood

is recommended over WAG, especially when recycling produced gases is possible to store

larger volumes of carbon dioxide.

7
Table of Contents
List of Tables .....................................................................................................................11

List of Figures ....................................................................................................................12

Chapter 1: Introduction .....................................................................................................18

1.1: Background ........................................................................................................18

1.2: Research Objective ............................................................................................20

Chapter 2: Literature Review ............................................................................................22

2.1: Enhanced Oil Recovery (EOR) .........................................................................22

2.1.1: CO2-EOR ............................................................................................25

2.1.2: Surfactant EOR (SEOR) .....................................................................29

2.1.3: Foam EOR ..........................................................................................32

2.1.4: Polymer EOR ......................................................................................35

2.2: Carbon Capture, Utilization and Storage (CCUS) and Carbon Capture and
Storage (CCS) .....................................................................................................37

2.2.1: CO2 Capture ........................................................................................39

2.2.2: CO2 Transportation .............................................................................43

2.2.3: CO2 Storage and Utilization ...............................................................44

2.2.4: CCUS and CCS Projects .....................................................................46

2.2.5: CCUS and CCS Key Mechanisms ......................................................48

2.2.6: CCUS and CCS Numerical Simulation ..............................................51

Chapter 3: Methodology and Assumptions.......................................................................54

3.1 Numerical Simulation Model..............................................................................54

3.1.1: Field Model Description .....................................................................54

3.1.2: Assumptions and Considerations ........................................................59


8
3.1.2.1: Minimum Miscibility Pressure (MMP) ..................................59

3.1.2.2: Relative Permeability Hysteresis Model.................................61

3.1.2.3: CO2 Solubility in Water ..........................................................68

3.1.2.4: Thermal Modeling ..................................................................69

3.1.2.5: Equation of State (EOS) .........................................................70

3.1.2.6: Geochemistry ..........................................................................71

3.2: Approach............................................................................................................75

3.2.1: Injection Gas Composition Analysis ..................................................77

3.2.2: Depleted Reservoir Analysis ..............................................................79

3.2.3: Partially Depleted Reservoir Analysis ................................................81

Chapter 4: Results and Discussions ..................................................................................87

4.1: Composition of Injected CO2.............................................................................87

4.1.1: Single Impurity ...................................................................................87

4.1.2: Dual Impurity ......................................................................................89

4.1.3: Triple Impurity ....................................................................................95

4.2: Depleted Reservoir Analysis .............................................................................99

4.2.1: Miscible Injections ..............................................................................99

4.2.2: Immiscible Injection .........................................................................103

4.2.3: Miscible Thermal Modeling .............................................................106

4.3: Tertiary Production Analysis ...........................................................................108

4.3.1: Water-Alternating-Gas (WAG) ........................................................108

4.3.2: Continuous Gas Injection..................................................................115

9
Chapter 5: Conclusions and Recommendations .............................................................119

5.1: Conclusions and Recommendations ................................................................119

5.2: Recommendations for Future Work ................................................................121

References ........................................................................................................................123

Vita...................................................................................................................................137

10
List of Tables

Table 2.1: Common Geochemical Reactions. ................................................................28

Table 2.2: SEOR Screening Criteria (Sheng, 2015). .....................................................32

Table 3.1: Reservoir Model Properties. ........................................................................55

Table 3.2: Reservoir Oil Composition. ..........................................................................60

Table 3.3: Oil Properties. ...............................................................................................61

Table 3.4: Examples of EOS Models. ............................................................................70

Table 3.5: Modeled Geochemical Reactions. ................................................................72

Table 3.6: Different Injection Gas Composition Cases. ................................................79

Table 3.7: WAG Analysis Sensitivity Parameters. ........................................................85

Table 4.1: Single Impurity Cases. ..................................................................................87

Table 4.2: Dual Impurity Cases. ....................................................................................90

Table 4.3: Triple Impurity Cases. ..................................................................................95

Table 4.4: Carbon Dioxide Inventory at Different Injection Rates in Miscible

Flooding at 100% HCPV Injected. .............................................................102

Table 4.5: Carbon Dioxide Inventory at Different Injection Rates in Immiscible

Flooding at 100% HCPV Injected. .............................................................104


Table 3.7: WAG Analysis Sensitivity Parameters. ......................................................109

11
List of Figures

Figure 1.1: Cumulative Historical Global CO2 Emissions and Temperature Rise

Showcasing Multiple Scenarios of Forecasted Future Emissions (IPCC,

2021). ............................................................................................................19

Figure 2.1: Global Production Contribution from Different EOR Projects Forecast

(IEA, 2022b). ................................................................................................25

Figure 2.2: Number of EOR Projects Globally (IEA, 2022c). ........................................25

Figure 2.3: Surfactant Structure (Lake et al., 2014). .......................................................31

Figure 2.4: Foam Mobility vs Quality of Three Mediums and 0.1% Aerosol Foam

(Khan, 1965). ................................................................................................33

Figure 2.5: Apparent Foam Viscosity vs. Foam Quality in a Bentheimer Sandstone

(Tang et al., 2019). ........................................................................................34

Figure 2.6: Waterflooding and viscous fingering. (Left) Mobility Ratio > 1.0.

(Right) Polymer flood with Mobility Ratio < 1.0. (Olajire, 2014). ..............36

Figure 2.7: Commercial CCUS Facilities (Operating and Developing) Globally

(IEA, 2021a). ................................................................................................38

Figure 2.8: Stationary CO2 Emission Sources (NETL, 2015). ........................................40


Figure 2.9: Global Annual CO2 Emissions from Energy Combustion and Industrial

Processes (IEA, 2023). ..................................................................................40

Figure 2.10: Global CO2 Sources and Magnitude (IEA, 2021b). ......................................41

Figure 2.11: Global CCUS and CCS Formation Candidates (IEA, 2021b). .....................46

Figure 2:12: Commercial CCS Facilities (Operational and Under Development)

(GCCSI, 2023). .............................................................................................47

Figure 2.13: Global CCUS Locations in Oil and Gas Fields (OGCSI, 2023). ..................47

12
Figure 2.14: Capillary Pressure Curves (Peters, 2012). ....................................................50

Figure 2.15: CO2 Trapping Contribution of Trapping Mechanisms. (Left) Carbonate

and Sandstone Reservoir. (Right) Mafic and Ultramafic Rocks

(NASEM, 2019). ...........................................................................................51

Figure 3.1: Model Spatial Property Distribution. (a) Porosity (b) Permeability in md.

The model illustration is scaled by a factor of 3 on the z-axis. ....................55

Figure 3.2: Reservoir Model Vertical Grid Size in ft. The model illustration is scaled

by a factor of 3 on the z-axis. ........................................................................56

Figure 3.3: Rock Permeability vs Porosity Correlation...................................................57

Figure 3.4: Petrophysical Properties of Reservoir Rocks. (a) oil and water relative

permeabilities. (b) gas and water relative permeabilities. (c) capillary

pressure curves. .............................................................................................58

Figure 3.5: Capillary Trapping of Non-wetting Phase in a Porous Medium. Each

row reaches further points on the drainage curve than the row before it

(higher Snw), while the second column re-saturates via a screening curve

(lowering Snw), showing the resultant trapping of the non-wetting phase.

(Lake et al., 2014). ........................................................................................62

Figure 3.6: Hysteresis Models Performance on Historically Matched Case. ..................64

Figure 3.7: Hysteresis Loops in a Three-phase Oil/Water/Gas Flow Highlighting

Gas Relative Permeability Reduction (Larsen and Skauge, 1998). ..............65

Figure 3.8: Activity Coefficient vs. Ionic Strength Logarithm of Activity Models

(Rickard, 2012). ............................................................................................73

Figure 3.9: Original History Matched Case Vs. Hysteresis Model. ................................75

Figure 3.10: Reservoir Model and Five-Spot Well Pattern (Producers and Injectors)

Location. .......................................................................................................76
13
Figure 3.11: Gravity Effect in Gas Flood, Waterflood, and WAG (Samba and

Elsharafi, 2018). ............................................................................................82

Figure 3.12: Base Case Primary and Secondary Production Stages. Average

Reservoir Pressure, Gas Oil ratio (GOR), Oil RF, and Water Cut (WC). ....83

Figure 4.1: Single Impurity Impact on Carbon Dioxide Retention Compared to A

Pure Stream. ..................................................................................................88

Figure 4.2: Single Impurity Impact on Oil RF Compared to A Pure Stream. .................89

Figure 4.3: Dual Impurity (N2 + C1 and H2S) Impact on Carbon Dioxide Retention

Compared to A Pure Stream. (NH-N) Increasing concentration of N2 +

C1. (NH-H) Increasing concentration of H2S. (NH-E) Impurities have

equal concentrations......................................................................................91

Figure 4.4: Dual Impurity (N2 + C1 and C2) Impact on Carbon Dioxide Retention

Compared to A Pure Stream. (NC-N) Increasing concentration of N2 +

C1. (NC-C) Increasing concentration of C2. (NC-E) Impurities have

equal concentrations......................................................................................92

Figure 4.5: Dual Impurity (H2S and C2) Impact on Carbon Dioxide Retention

Compared to A Pure Stream. (HC-H) Increasing concentration of H2S.

(HC-C) Increasing concentration of C2. (HC-E) Impurities have equal

concentrations. ..............................................................................................92

Figure 4.6: Dual Impurity (N2 + C1 and H2S) Impact on Oil RF Compared to A Pure

Stream. (NH-N) Increasing concentration of N2 + C1. (NH-H) Increasing

concentration of H2S. (NH-E) Impurities have equal concentrations. ..........93

Figure 4.7: Dual Impurity (N2 + C1 and C2) Impact on Oil RF Compared to A Pure

Stream. (NC-N) Increasing concentration of N2 + C1. (NC-C) Increasing

concentration of C2. (NC-E) Impurities have equal concentrations. ............94


14
Figure 4.8: Dual Impurity (H2S and C2) Impact on Oil RF Compared to A Pure

Stream. (HC-H) Increasing concentration of H2S. (HC-C) Increasing

concentration of C2. (HC-E) Impurities have equal concentrations. ............94

Figure 4.9: Triple Impurity (H2S, N2 + C1 and C2) Impact on Carbon Dioxide

Retention Compared to A Pure Stream. (C) Increasing concentration of

C2. (H) Increasing concentration of H2S. (N) Increasing concentration of

N2 and C1.......................................................................................................96

Figure 4.10: Triple Impurity (H2S, N2 + C1 and C2) Impact on Carbon Dioxide

Retention Compared to A Pure Stream. (E) All impurities exist at equal

concentrations. ..............................................................................................96

Figure 4.11: Triple Impurity (H2S, N2 + C1 and C2) Impact on Oil RF Compared to A

Pure Stream. (C) Increasing concentration of C2. (H) Increasing

concentration of H2S. (N) Increasing concentration of N2 + C1. ..................97

Figure 4.12: Triple Impurity (H2S, N2 and C1 + C2) Impact on Oil RF Compared to A

Pure Stream. (E) All impurities exist at equal concentrations. .....................98

Figure 4.13: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection

Rates for Isothermal Miscible Injection Conditions. (a) RT on the left

axis and RF on the right axis (b) CO2 Retention in Mt...............................100

Figure 4.14: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection

Rates for Isothermal Miscible Injection Conditions. (a) RT on the left

axis and RF on the right axis (b) CO2 Retention in Mt...............................101

Figure 4.15: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection

Rates for Isothermal Immiscible Injection Conditions. (a) RT on the left

axis and RF on the right axis (b) CO2 Retention in Mt...............................104

15
Figure 4.16: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection

Rates Under Miscible Conditions. Marked at the Injected Amounts

(MtCO2) of Carbon Dioxide by its Counter Immiscible Flooding. (a) RT

on the left axis and RF on the right axis (b) CO2 Retention in Mt. ............105

Figure 4.17: Miscible Isothermal model (I) vs. Miscible non-Isothermal model (T) in

Terms of Oil RF (RF) and Carbon Dioxide Retention (RT) at different

injection rates. (a) RT on the left axis and RF on the right axis (b) CO2

Retention in Mt. ..........................................................................................107

Figure 4.18: Impact of Different Injection Temperature on Oil RF (RF) and CO2

Retention (RT) in Miscible non-Isothermal model (T). RT on the left

axis and RF on the right axis. The injectant temperature is listed by the

case name in °F. ..........................................................................................108

Figure 4.19: WAG Ratio and Cycle Duration Impact on Oil RF. The number of

samples shown by count along with their average (Avg), minimum

(min), maximum (max), P90, P50 (median) and P10. The x-axis on the

top figure indicates the cycle duration in years, while in the bottom

figure, the x-axis indicates Gas/Water ratio. ...............................................111

Figure 4.20: WAG Ratio and Cycle Duration Impact on CO2 Retention. The number

of samples shown by count along with their average (Avg), minimum

(min), maximum (max), P90, P50 (median) and P10. The x-axis on the

top figure indicates the cycle duration in years, while in the bottom

figure, the x-axis indicates Gas/Water ratio. ...............................................112

16
Figure 4.21: Impact of Gas Injection Rate at Constant WAG Ratio on Oil RF. The

average (Avg), P90, P50 (median), and P10 are shown for each gas rate

at a certain WAG ratio. The first row of the x-axis indicates Gas/Water

ratio, and the second row is the injection gas rate (rbbl/d). ........................113

Figure 4.22: Impact of Gas Injection Rate at Constant WAG Ratio on CO2 Retention.

The average (Avg), P90, P50 (median) and P10 are shown for each gas

rate at a certain WAG ratio. The first row of x-axis indicates Gas/Water

ratio, and the second row is the injection gas rate in (rbbl/d). ....................114

Figure 4.23: Continuous CO2 Flood Performance in Terms of EOR (RF) and

Sequestration (RT). (a) RT on the left axis and RF on the right axis (b)

CO2 Retention in Mt. ..................................................................................115

Figure 4.24: Average Oil RF in WAG Process at Different Gas Injection Rates

(MMSCFD). The number of samples shown by count along with their

average (Avg), P90, P50 (median) and P10. x-axis indicates the gas

injection rate (MMSCFD). ..........................................................................117

Figure 4.25: Average CO2 Retention in WAG Process at Different Gas Injection

Rates (MMSCFD). The number of samples shown by count along with

their average (Avg), P90, P50 (median) and P10. x-axis indicates the gas

injection rate (MMSCFD). ..........................................................................118

17
Chapter 1: Introduction

1.1: BACKGROUND

Rising greenhouse gases (GHG) concentrations in the atmosphere have gained

more attention in the last few years, garnering decarbonization technology more resources

and support. IPCC (2023) reports that most of the GHG comprise mainly of carbon dioxide

(410 ppm) and methane (1.866 ppb), raising the global temperature by 1.09°C above the

average temperatures of the 1850 – 1900 period. The increase in global temperature was

observed to be unequal across the globe, with most of it taking place over landmass (1.59

°C), while oceans exhibited half of that of the landmass (0.88°C). Figure 1.1 highlights the

historical carbon dioxide emissions since 1850, the temperature increase, and multiple

forecasts for the future. Due to the increasing temperatures and uneven distribution of

global warming, extreme weather events are predicted to increase in intensity and

frequency, which led to more emphasis on decarbonization efforts.

Decarbonization seeks to lower carbon dioxide concentrations in the atmosphere or

to maintain it at current levels, at the least, to address climate concerns. Approximately 37

giga tons of carbon dioxide (GtCO2) are emitted annually by fossil fuels combustion for

energy generation and industrial processes (IEA, 2023), overstating the importance of
decarbonization efforts. Typically, decarbonization functions through (1) capturing

anthropogenic CO2 from the atmosphere and large stationary emitters, such as power

plants, (2) enhancing the efficiency of industrial processes to become less CO2 intensive,

(3) transitioning to renewable energy sources to reduce the energy sector’s reliance on

fossil fuels as a primary energy source. The main challenge facing decarbonization is

finding a solution that is able to manage carbon dioxide in a net-zero or negative emissions

18
manner, is economically feasible, scalable promptly, and does not significantly harm life

quality, especially of those in vulnerable populations.

Figure 1.1: Cumulative Historical Global CO2 Emissions and Temperature Rise
Showcasing Multiple Scenarios of Forecasted Future Emissions (IPCC,
2021).

As the world strives to transition to a low-carbon economy and to limit the rise in

global temperatures, geological sequestration has emerged as a critical and indispensable

tool in achieving this ambitious goal. Geological sequestration offers a robust workflow to

combat carbon emissions across various sectors, with its capability to dispose of large

volumes of carbon dioxide to mitigate global warming (IEA, 2021b). Geological

sequestration encompasses a comprehensive approach to managing carbon dioxide

19
emissions from industrial facilities and power plants by capturing and securely storing it,

preventing it from being released into the atmosphere. Two variations of geological

sequestration are available: carbon capture, utilization, and storage (CCUS) and carbon

capture and storage (CCS). Both models of geological sequestration are similar in every

aspect but for the main objective of storage: CCUS aims to accomplish a goal by the

injection that is complemented by the storage efforts (e.g., EOR), while CCS only pursues

to dispose of CO2.

CCUS and CCS play a multifaceted and transformative role in decarbonizing

various sectors. For instance, carbon dioxide emissions from hard-to-abate sectors, such as

heavy industries (e.g., cement, steel, and petrochemicals) and energy sectors, are difficult

to reduce without negatively impacting life quality. However, by implementing CCUS and

CCS, these sectors would be able to function normally while the generated emissions are

captured at the source and prevented from reaching the atmosphere. Hence, introducing

stability to these sectors rather than pressing them to undergo harsh modifications to adapt

to a net-zero era. Consequently, CCUS and CCS technologies are capable of pushing the

needle towards negative emissions by disposing of carbon dioxide captured from the

atmosphere. Moreover, CCUS repurposes the captured CO2 to serve other goals, allowing

it to contribute to economic growth.

1.2: RESEARCH OBJECTIVE

This research aims to further our understanding of CCUS's potential in the oil and

gas industry. Historically, carbon dioxide has been used to enhance oil and gas production,

but at relatively small quantity compared to CCUS. The main concern was to use the least

volume possible of CO2 to improve hydrocarbon production, with no incentive to ensure

the entrapment of CO2 in the subsurface. However, increased awareness and the rise of

20
geological sequestration changed the narrative to attempt to store CO2 in the subsurface in

EOR projects. This trend benefits decarbonization and stabilizes the energy sector by

increasing hydrocarbon production. Therefore, this research investigates various aspects of

CO2 sequestration into an oil reservoir to evaluate its impact on enhancing oil recovery and

the efficiency of carbon dioxide entrapment.

21
Chapter 2: Literature Review

2.1: ENHANCED OIL RECOVERY (EOR)

Enhanced Oil Recovery (EOR) is the process of oil extraction from a reservoir that

could not be produced otherwise. Hydrocarbon oil and gas production is classified into

three stages: primary, secondary, and tertiary (EOR). The Primary stage depends on the

reservoir’s natural driving mechanism (e.g., gas cap, water influx, gravity drainage) to

produce hydrocarbons, with the aid of artificial lift equipment. Secondary recovery comes

afterward, in which water or gas is injected to drive hydrocarbons toward producers.

Primary and secondary production stages achieve a recovery factor (RF) of 35% to 45%

(primary ~5% - 20%, secondary ~15% - 25%) (Fink, 2021; Lake et al., 2014; Sheng, 2013;

Zitha et al., 2011; Chen et al., 2018; Bello et al., 2023). Then, tertiary stage comes into

play, in which EOR techniques are employed once the oil rate declines to the economic

limits during the primary and secondary production stages. EOR techniques can increase

the RF by 5% to 15% on average (Zitha et al., 2011; Lake et al., 2014; Islam et al., 2022;

Chen et al., 2018).

Current estimates of proven oil reserves globally range between 1,250 – 1,750

billion barrels, which is an underestimate, as the value continuously increases with new
discoveries and technological advances. At current production levels, this estimate predicts

the energy sector to be sustainable for the next 50 years (OPEC, 2021; Xu and Bell, 2022;

Worldometer, 2016). Therefore, EOR techniques that extend the volume of recoverable

hydrocarbons to 60% and beyond are highly crucial (Islam et al., 2022; Zitha et al., 2011;

Lake et al., 2014). Until new economically feasible technologies arise to further the volume

of recoverable hydrocarbons, EOR will remain extremely crucial to maximize oil RF. Even

22
then, EOR would retain its relevance and importance as it would remain capable of

extending the RF, further sustaining the world economy.

EOR involves the injection of substances into the reservoir, which are not naturally

found in the reservoir to promote hydrocarbon production (Lake et al., 2014). EOR

attempts to increase sweep efficiency (E) on microscopic (Ed) and macroscopic (Ev) levels:

microscopic denotes the volume of oil displaced by the injected fluid, while macroscopic

level, also known as volumetric sweep, refers to the area of the reservoir that has been in

contact with the injection fluid (Bello et al., 2023; Martin and Colpitts, 1996). Various

EOR methods are implemented in oil and gas fields to increase the volumes of produced

hydrocarbons. The primary categories of EOR techniques are:

1) Chemical EOR (CEOR): In this method, chemicals such as surfactants, polymers,

or alkaline substances are injected, individually or in combination, into a reservoir

to change the properties of the oil, making it easier to extract. The most common

chemical EOR methods are surfactant and polymer flooding (DOE, 2010). CEOR

increases RF by 5% - 15% via wettability alteration, mobility reduction, or capillary

forces reduction (Lake et al., 2014).

2) Thermal EOR (TEOR): In this method, heat is used to reduce the oil’s viscosity,

making it easier to extract. In general, TEOR is used in heavy oil reservoirs (API <

20) to overcome its high viscosity (Bello et al., 2023). The most common thermal

EOR methods are steam injection, in which steam is injected into the reservoir to

heat the oil, and in-situ combustion, wherein oxygen or air is injected into the

reservoir to burn some of the oil and generate heat (DOE, 2010). Other forms of

thermal EOR exist, such as SAGD and EM heating, offering a wide range for RF

from 5% - 65%, depending on which method is used. However, TEOR technologies

are costly and pose an environmental concern (Lake et al.,2014).


23
3) Solvent EOR: In this method, gases such as nitrogen (N2), carbon dioxide (CO2),

or hydrocarbon gas are injected into the reservoir to displace the oil. This method

is also known as gas flooding or gas EOR. The flooding could be miscible, in which

the injected gas mixes with oil in the absence of a distinct interface, or immiscible,

where the gas exists as a separate phase. Solvent EOR enables the recovery of an

additional 5% - 15% of the original oil in place (OOIP), on average (Lake et al.,

2014).

4) Microbial EOR (MEOR): In this method, either microorganisms or nutrients are

injected to grow the indigenous microorganisms in the reservoir. Consequently, oil

recovery is improved by altering the properties of the oil or the reservoir and

generating byproducts. For instance, microorganisms can break down complex

hydrocarbons, reduce viscosity, and produce gases that can assist in displacing the

oil, such as N2, H2, H2S, and CH4. MEOR is often used after TEOR and/or CEOR,

if significant oil is left behind, increasing RF by more than 20% of the OOIP. Still,

its effectiveness depends on the type of reservoir, the specific MEOR technique

used, and the effectiveness of the microorganisms employed (Pandey and Tiwari,

2022).

Over 40% of the EOR production in the U.S. is from thermal EOR, followed by

solvent EOR and Chemical EOR, as illustrated in Figure 2.1 (IEA, 2022b; Lake et al.,

2019; DOE, 2010). Each EOR method has its advantages and disadvantages, and the choice

of EOR method depends on the specific characteristics of each reservoir and its fluids, as

well as the associated economics and environmental evaluations. In this work, the focus is

on solvent EOR and specifically on CO2-EOR, as it pertains to the decarbonization of the

atmosphere and EOR.

24
Figure 2.1: Global Production Contribution from Different EOR Projects Forecast (IEA,
2022b).

Figure 2.2: Number of EOR Projects Globally (IEA, 2022c).

2.1.1: CO2-EOR

The process of injecting CO2 into an oil reservoir to enhance oil recovery is not a

novel technique. The technology was first introduced and patented by Wharton et al. in

1952, and the first field implementation was conducted in the United States in 1964

25
(Stalkup, 1978). Initially, EOR was not common, even though it was proven to work

technically, due to its unattractive economics inflicted by low oil prices and high injection

cost. Once oil prices started rising in the 1970s, interest in EOR increased and led to the

investigation of various EOR methods, which reported carbon dioxide EOR to be one of

the most influential EOR approaches (Sheng, 2013; Stalkup, 1978). Therefore, once oil

prices increased, improving the economics of CO2 flooding, by 1972, carbon dioxide

flooding became commercially viable and widely implemented (Chen et al., 2018; Stalkup,

1978). Today, CO2 flooding remains popular, however it is tied to regions with natural

carbon dioxide sources (Wang et al., 2017). Some of the pioneering commercial projects

are the Permian Basin’s SACROC unit in Texas (1972) and Saudi Arabia’s Uthmaniyah

field (1990) (Chen et al., 2018).

Carbon dioxide is injected into a reservoir, generally, for solvent EOR to pressurize

and drive mobile hydrocarbon toward producers. Moreover, CO2 can free immobile or

unproducible oil by changing its properties, such as viscosity, altering the capillary number

(Lake et al., 2019; Delshad et al., 1987). For a successful solvent EOR, it is paramount to

understand the in-situ chemical interactions, solubility, relative permeabilities, and

wettability of the reservoir of interest. Moreover, it is possible to use CO2-EOR in

combination with other forms of EOR, like foam, for better volumetric sweep efficiency.

CO2 is permeable through a permeable medium via four main mechanisms: gravity

forces, capillary forces, viscous forces, and diffusion. The dominant mechanism is

contingent on the reservoir’s properties but is typically diffusion, especially in tight and

fractured reservoirs (Eide et al., 2013; Song et al., 2020). Moreover, the displacement of

CO2 in a reservoir is via solution gas drive, miscible drive, or immiscible drive (Lake,

1989). In addition, there is a myriad of forms CO2 that could be injected into a reservoir,

such as gas, super-critical fluid (sc), dissolved (carbonated water (CW)), or water-
26
alternating-gas (WAG). Each injection approach has advantages and disadvantages; for

instance, pure gas injection could inject large volumes of CO2 into a reservoir, unlike CW,

which is limited by the volume dissolvable in the carrying water.

CO2 injection is classified as a miscible or an immiscible flood. For an injection

scheme to be miscible, the reservoir pressure must be above a certain pressure, which

allows CO2 to mix with oil as a homogenous phase with no interface, known as the

minimum miscibility pressure (MMP). If the reservoir is below MMP, then the flood is

immiscible, in which both phases exist separately with an interface in between. MMP is

influenced by the composition of the injected CO2 feed, reservoir pressure and temperature,

and oil composition. For example, carbon dioxide has a higher solubility in oil than in water

or brine (Chen et al., 2022; Lake, 1989). Generally, miscible injection is more desirable

since it is more efficient and, theoretically, capable of achieving 100% RF under fully

miscible flow. (Chen et al., 2022; Lake et al., 2014; Chen et al., 2018). Furthermore,

miscible CO2 injection was proven to be more successful than CW and immiscible flood

in field applications in terms of RF, economics, and environmentally (Chen et al., 2018;

Fath and Pouranfard, 2014). Miscibility is further discussed in Chapter 3.

A major aspect investigated in a CO2-EOR project is the in-situ chemical

interactions due to the injection of carbon dioxide (geochemistry). Introducing CO2 to a

brine forms carbonic acid (H2CO3) and causes the solution pH to decrease. The alteration

of the environment’s pH, in which the fluids reside, changes the interactions between them

and surrounding minerals. This has been confirmed in field projects in various types of

reservoirs with different minerals (Han et al., 2010; Mito et al., 2008; Kharaka et al., 2006;

Emberley et al., 2004; Klusman, 2003; Bowker and Shuler, 1991). Experiments by

Rosenbauer et al. (2005) on a limestone-water-scCO2 system proved that the alteration is

not limited to reservoir fluids, but also impacts the reservoir’s porosity and permeability.
27
This occurs due to the geochemical reaction causing mineral dissolution and carbonates

precipitation. The dissolution increases porosity and opens flow paths, while precipitated

minerals migrate and accumulate in flow paths, plugging them. Hence, understanding a

reservoir's mineralogy and fluids is vital to accurately predict the impact of their

interactions with CO2 accurately. Generally, carbon dioxide chemical interactions are

important in carbonate reservoirs since their mineralogical making (e.g., calcite (CaCO3))

is highly reactive to CO2. On the other hand, sandstone reservoirs are less reactive to CO2

than carbonate reservoirs; nonetheless, pH alteration could lead to precipitation of minerals

present within sandstone reservoirs. Subsequently, the mineral dissolution in sandstone

reservoirs is uniform due to the slow reaction rate, while in carbonate reservoirs the

dissolution is non-uniform due to the high reaction rate, possibly creating wormholes (Song

et al., 2020; Nightingale et al., 2009; Wellman et al., 2003). Table 2.1 chemical reactions

are some of the common mineral interactions in carbonate reservoirs, including mineral

dissolution (e.g., calcite and anhydrite) and mineral precipitation (e.g., dawsonite) (Liu et

al., 2011; Han et al., 2010).

Table 2.1: Common Geochemical Reactions.

Reaction Chemical Formula


Calcite 𝐶𝑎𝐶𝑂3 + 𝐻 + ↔ 𝐶𝑎2+ + 𝐻𝐶𝑂3−
Dolomite 𝐶𝑎𝑀𝑔(𝐶𝑂3 )2 + 2𝐻 + ↔ 𝐶𝑎2+ + 𝑀𝑔2+ + 2𝐻𝐶𝑂3−
Anhydrite 𝐶𝑎𝑆𝑂4 ↔ 𝐶𝑎2+ + 𝑆𝑂42−
Carbonic Acid 𝐶𝑂2 + 𝐻2 𝑂 ↔ 𝐻 + + 𝐻𝐶𝑂3−
Dawsonite 𝑁𝑎𝐴𝑙(𝐶𝑂3 )(𝑂𝐻) + 3𝐻 + ↔ 𝑁𝑎+ + 𝐻𝐶𝑂3− + 𝐴𝑙 3+ + 2𝐻2 𝑂
Aqueous reaction 𝐶𝑂32− + 𝐻 + ↔ 𝐻𝐶𝑂3−
Aqueous reaction 𝑂𝐻 − + 𝐻 + ↔ 𝐻2 𝑂

28
2.1.2: Surfactant EOR (SEOR)

Surfactant EOR (SEOR) is a technique that utilizes the injection of surface-active

agents, known as surfactants, to enhance oil recovery. Surfactants are chemical compounds

that possess both hydrophilic (water-liking) and hydrophobic (oil-liking) properties that

allow them to reduce capillary forces, a major contributor to oil entrapment, and alter

wettability. SEOR has been under development for over half a century and dates to the

1960s, when researchers first began studying the behavior of surfactants in oil-water

systems (Taber, 1958). One of the earliest examples of surfactant EOR can be traced back

to a field trial conducted in the 1970s in the Permian Basin of West Texas. The trial

demonstrated that a surfactant formulation could increase oil recovery by as much as 25%

compared to conventional water flooding (Economides, 2013). Moreover, recent

advancements in surfactant EOR have focused on developing more effective and

environmentally friendly surfactant formulations, in which biodegradable surfactants are

derived from natural sources such as plants and bacteria. These surfactants are less toxic

than traditional surfactants and can be broken down more easily in the environment (Atta

et al., 2021).

In a reservoir, capillary forces trap the oil, while viscous forces mobilize it (Lake

et al., 2014). Surfactants reduce the capillary forces trapping the oil by minimizing the

interfacial tension (IFT) between oil and water to ultralow values. The relationship between

capillary pressure (𝑃𝑐 ) and IFT (𝜎) across an interface with a curvature radius 𝑅, can be

expressed by the simple form of Laplace’s equation in eqn. 2.1. Moreover, the relationship

between capillary and viscous forces is represented by the local capillary number (𝑁𝑣𝑐 ).

𝑁𝑣𝑐 is a dimensionless measure of viscous forces to capillary forces, as shown in eqn. 2.2,

where 𝑢 is the displacing fluid’s Darcy velocity and µ is its viscosity. As capillary forces

are reduced by surfactants, viscous forces take over at a critical 𝑁𝑣𝑐 , after which the oil
29
flows more efficiently (Sheng, 2015). This phenomenon was determined to result from a

correlation between a fluid's residual saturation and 𝑁𝑣𝑐 , described as the capillary

desaturation curve (CDC).

2𝜎 cos 𝜃
𝑃𝑐 = 𝑃𝑛𝑤 − 𝑃𝑤 = (2.1)
𝑅
µ𝑢
𝑁𝑣𝑐 = (2.2)
𝜎

Surfactants consist of a hydrophobic tail (non-polar hydrocarbon chain) and a

hydrophilic head (polar), as shown in Figure 2.3. They function by adsorbing onto

interfaces to lower IFT and alter rock wettability. Based on polar configuration, surfactants

are categorized into four classes:

1) Anionic: the head is negatively charged and typically used in sandstone reservoirs.

Most SEORs utilize anionic surfactants, as they tend to be the most effective in

sandstone reservoirs due to their low adsorption (Lake et al., 2014; Sheng, 2015;

Atta et al., 2021).

2) Cationic: the head is positively charged, and typically used in carbonate reservoirs.

Cationic surfactants are the least likely to be used in sandstone reservoirs, to avoid

excessive adsorption (Atta et al., 2021). The main objective of surfactants in

carbonate reservoirs is to change wettability from oil-wet to water-wet, since they

are more likely to be oil-wet (Sheng, 2015; Jarrahian et al., 2012).

3) Nonionic: the head is not charged, hence less sensitive to electrolytes in aqueous

solutions. The main mechanisms which control their performance are van der

Waals and hydrogen bonding (Atta et al., 2021).

4) Amphoteric: The head is both negatively and positively charged simultaneously.

Another type changes its charge between positive and negative based on the
30
aqueous solution pH (Lake et al., 2014). This type of surfactants is not commonly

used due to its high cost (Yarveicy and Haghtalab, 2018). Amphoteric surfactants

are called zwitterionic, as well.

Figure 2.3: Surfactant Structure (Lake et al., 2014).

The surfactant EOR methods typically involve three stages: injection, flow, and

production. During the injection stage, surfactants are injected into the reservoir through

injection wells, mixed with the injection water or brine solution (Lake, 1989). The

surfactant solution is designed to be compatible with the formation water, reservoir rock,

oil composition, clay content, and reservoir conditions to reduce the interfacial tension

between the oil and water (Atta et al., 2021; Yarveicy and Haghtalab, 2018; Sheng, 2015;

Lake et al., 2014; Nelson, 1982). In the flow stage, the surfactant solution flows through

the reservoir, contacting the oil and reducing the interfacial tension. This allows the oil to

detach from the reservoir’s rock and flow more easily through the porous medium toward
the production wells. Finally, in the production stage, the oil is produced through

production wells, along with the surfactant solution and any formation water that may have

been displaced during the process, hence the interest of developing green-biodegradable

surfactants (Atta et al., 2021; Lake, 1989).

The effectiveness of SEOR can be affected by several factors, including the

surfactant type, concentration, injection rate, reservoir rock properties, pH, salinity, clay

content, and the presence of other fluids in the reservoir (Atta et al., 2021; Yarveicy and

Haghtalab, 2018; Sheng, 2015; Lake et al., 2014; Nelson, 1982). The process is often used
31
in combination with other EOR methods, such as water flooding, polymer flooding (SP) or

alkaline and polymer (ASP) to maximize oil recovery (Lake, 1989). Overall, surfactant

EOR is a complex process that requires careful planning and execution to achieve

maximum oil recovery, which has been shown to be effective in various oil reservoirs

(sandstone, carbonate, and fractured reservoirs). Sheng (2015) surveyed field

implementations of SEOR and provided a screening criterion based on them, some of

which are in Table 2.2.

Table 2.2: SEOR Screening Criteria (Sheng, 2015).

Lithology K, md Tres, °C Water Salinity, ppm Clay µo, cP So, %

Sandstone >10 <93.3 <50,000 Low <35 >30

2.1.3: Foam EOR

Foam EOR is a complex and expensive process that requires significant expertise

and resources. It was first considered to be a promising EOR technique in the 1960s (Sheng,

2013). Yet, some of the earliest patents, suggesting the use of detergents to enhance oil

recovery, were in the late 1920s (Taber, 1958). The process involves the injection of a

foam-like substance into an oil reservoir to improve oil recovery. The foam displaces oil

and reduces gas mobility to improve the overall efficiency of oil recovery (Bello et al.,

2023). Thus, increasing the portion of the reservoir volume exposed to gas is essential, as

it will boost oil production (Lake et al., 2014).

Foam comprises surfactants (foaming agents) that reduce surface tension, as in

SEOR, and foam introducing gas mobility control. When the surfactants are mixed with a

gas, such as nitrogen or carbon dioxide, and water/brine it promotes bubble formation and

creates a foamy substance; although foam could be generated without a foaming agent, the

32
bubbles would be unstable and easily breakable. Foam could be generated in-situ, prior to

injections (pre-generated), or as surfactant and gas slugs (surfactant-alternating-gas (SAG))

(Bello et al., 2023). Foam is classified based on bubble diameter and quality; foam quality

(𝛤) refers to the percentage of gas in foam and is represented by eqn. 2.3, where 𝑉𝐺 and 𝑉𝐿

are the gas and liquid volumes, respectively. Figure 2.4 displays the relationship between

foam quality and mobility in three different mediums of different permeabilities. The

effective viscosity of foam in all mediums is similar at the low end of quality and

quadruples at the other extreme, making foam’s viscosity over 500 times that of the air

used in the foams. More recent work by Tang et al. (2019) measured the apparent viscosity

of the foam at various foam qualities, and found that there is an optimum quality, which

beyond the foam viscosity decreases, as illustrated by Figure 2.5.

𝑉𝐺
𝛤= ∗ 100 (2.3)
𝑉𝐺 + 𝑉𝐿

Figure 2.4: Foam Mobility vs Quality of Three Mediums and 0.1% Aerosol Foam
(Khan, 1965).

33
Figure 2.5: Apparent Foam Viscosity vs. Foam Quality in a Bentheimer Sandstone
(Tang et al., 2019).

Foam EOR can overcome a problem which solvent EOR, TEOR and SEOR suffer

from, which is poor macroscopic efficiency (volumetric sweep efficiency). The problem is

typically caused by reservoir heterogeneity and low viscosity of injectant EOR agents

(Lake et al., 2014). Therefore, it is particularly effective in reservoirs with low to medium

permeability and heterogeneous permeability distribution, where gas override and viscous

fingering are prone to happen, leaving behind hydrocarbons. Foam EOR introduces
mobility control, defined by eqn. 2.4, by increasing the viscosity of the gas to lower its

mobility to values close to that of the oil (mobility ratio ≤ 1). The extent of reduction is

impacted by the bubble size; the bigger the bubbles in the foam, the higher the reduction

(Lake et al., 2014). Ideally, gas mobility would be reduced below that of the oil phase, to

achieve a uniform injection front. Other mechanisms could be capitalized on to limit gas

mobility such as capillary forces reduction, via IFT, to change the capillary number 𝑁𝑣𝑐

(Bello et al., 2023). That is why foam has been utilized along with CO2 in some gas EOR

34
projects, further improving its performance and for CO2 to act as a stabilizer for the foam

(Wang et al., 2017).

𝑘
𝑀= (2.4)
µ

The process of creating foam for EOR requires a thorough understanding of the

reservoir characteristics, including the composition of the oil, the rock properties, and the

fluid flow behavior. The foam must be designed to meet the specific needs of the reservoir,
taking into account factors such as temperature, pressure, salinity, surfactant concentration,

wettability and the presence of other chemicals (Farzaneh and Soharbi, 2013; Geng et al.,

2021). Thus, further testing and experimentation are needed, to further the knowledge

concerning foam, and to correct previous studies that have been conducted at ambient

conditions rather than reservoir conditions (Bello et al., 2023). In addition, the introduction

process of foam into the reservoir is crucial and needs to be tailored, since each method’s

performance differs. Furthermore, monitoring and control systems to manage the injection

rate, pressure, and other parameters are crucial for an effective EOR (Sheng, 2013; Bello

et al., 2023).

2.1.4: Polymer EOR

Polymer EOR utilizes long-chain polymer molecules to improve oil recovery. The

polymers are water-soluble, viscoelastic with pseudoplastic and shear thickening behavior

under shear stress (Sheng, 2013). These long molecules impact the injected water, in which

they are dissolved, and lower its mobility via viscosity increment. This results in reducing

the water/oil mobility ratio, hence stabilizing the displacement, and improving water flood

performance and volumetric sweep, as illustrated in Figure 2.6 (Gbadamosi et al., 2022;

Lake et al., 2014; Olajire, 2014). According to a survey of experimental work on polymer
35
floods, Gbadamosi et al. (2022) reported an oil recovery factor of 17% - 98%. The large

range is caused by the usage of different core samples of different properties, different

polymers, and a combination of associated chemicals (ASP, SP), highlighting the

importance of choosing the best choice for polymer injection scheme.

Figure 2.6: Waterflooding and viscous fingering. (Left) Mobility Ratio > 1.0. (Right)
Polymer flood with Mobility Ratio < 1.0. (Olajire, 2014).

Polymers typically are classified into two categories: organic (biopolymers) and

synthetic. Some of the most common polymers used are hydrolyzed polyacrylamide

(HPAM), copolymers consisting of acrylic acid (AA) and acrylamide (AMD) (Gbadamosi

et al., 2022; Beteta et al., 2020; Lake et al., 2014; Sheng, 2013). They are exceptionally

good at managing water and that makes them a great EOR method for mature and declining

oil reservoirs. Moreover, by incorporating other chemicals, such as alkali, the sweep

efficiency is further improved by freeing capillary trapped hydrocarbons (Olajire, 2014).

In addition, polymer flooding demands less volumes of water, leading to a reduced water

cut in production wells (Gbadamosi et al., 2022). However, treating produced water from

polymer flooding is not simple (Olajire, 2014).

36
Polymers have been employed for a myriad of applications, such as improving well

performance, water mobility control, fracturing fluids, and EOR. In addition, polymers

could be used to assist other CEOR techniques as mentioned in the previous sections. Due

to its versatility, especially for heavy oils with high mobility ratios. There have been more

pilots deployed recently in North America (Wang et al., 2017).

2.2: CARBON CAPTURE, UTILIZATION AND STORAGE (CCUS) AND CARBON CAPTURE
AND STORAGE (CCS)

The Process of capturing carbon (CO2) from the atmosphere, utilizing the carbon

dioxide and storing it, is coined as carbon capture, utilization and storage (CCUS). Another

variation of CCUS exists and is called carbon capture and storage (CCS). Both CCUS and

CCS technologies’ primary goal is to reduce carbon dioxide emissions to mitigate climate

change and achieve net-zero. The processes are designed to stabilize CO2 levels by

capturing it from the atmosphere or from industrial facilities with a high carbon footprint,

such as power plants, factories, and other sources, and then either storing it underground

or using it for industrial purposes.

CCUS and CCS sequestration projects have the potential to capture 90% of CO2

emitted by fossil fuels (WCA, 2018). The captured carbon dioxide could be utilized in
enhancing oil recovery as part of CO2-EOR, as aforementioned, making sequestration

projects more economically feasible and hydrocarbon energy environmentally friendly. In

addition to assisting in stabilizing the energy sector by availing more oil, the carbon dioxide

sequestered is expected to remain downhole for a millennia with a low possibility of being

released back to the surface (Hepburn et al., 2019). Moreover, governments have started

subsidizing CCS and CCUS efforts to incentivize companies to pursue CCS and CCUS by

lessening the financial burden, making hydrocarbons a greener energy.

37
The major difference between CCUS and CCS is in their goals and their eligible

storage candidates. CCUS aims to achieve an additional goal besides storing CO2, which

in the petroleum industry is to enhance oil and gas recovery. On the other hand, CCS only

pursues carbon dioxide storage in a geological trap such as depleted reservoirs, deep saline

formations, unminable coal seams, and salt caverns. Otherwise, both CCS and CCUS are

generally similar.

CCUS and CCS have recently gained more attention as the increased CO2

concentrations in the atmosphere were taken more seriously. Prior attempts and efforts

have failed to gain either CCUS or CCS traction. However, due to recent commitments to

achieve net-zero, a notable increase in CCUS projects has been observed, as illustrated by

Figure 2.7 (IEA, 2021a).

Figure 2.7: Commercial CCUS Facilities (Operating and Developing) Globally (IEA,
2021a).

38
2.2.1: CO2 Capture

Carbon dioxide emissions are produced by natural causes (e.g., volcanic activity,

forest fires, organic matter decay) or of human origin (anthropogenic). According to the

national energy technology laboratory (2015) National Carbon Sequestration Database and

Geographic Information System (NATCARB), more than 6350 large stationary sources of

CO2 were identified in the US and Canada and mapped in Figure 2.8. These facilities were

found to generate over 3 billion metric tons of carbon dioxide (GtCO2). It is only logical to

assume that the number of stationary sources has increased since the annual global

emissions have continuously risen, as illustrated in Figure 2.9, from 35.56 GtCO2 to 37.49

GtCO2. Moreover, the implementation of capture projects could be prioritized at areas with

high carbon intensity that have been recognized in Figure 2.10. In other words, there are

ample stationary sources that could be capitalized on for capture projects to fuel CO2

sequestration efforts. These projects would enable sequestration endeavors to target

anthropogenic carbon dioxide, which currently represents 30% of stored CO2 (Islam et al.,

2022). Furthermore, these efforts have extended the global capture capacity to over 40

metric tons of CO2 (MtCO2) per year (IEA, 2021b). By 2030 more than 200 new projects

are expected to come online, as indicated by Figure 2.7, and are expected to raise the

capacity of CO2 capture to over 200 MtCO2/year (IEA, 2021a). However, this capture

capacity is insufficient to meet the net-zero ambitions by 2050 (1.2 GtCO2/year), and much

work is needed in this area (IEA, 2022a). Without sufficient CO2 capture capacity, CCS

and CCUS would fall apart, as there would be nothing to store or repurpose.

39
Figure 2.8: Stationary CO2 Emission Sources (NETL, 2015).

Figure 2.9: Global Annual CO2 Emissions from Energy Combustion and Industrial
Processes (IEA, 2023).
40
Figure 2.10: Global CO2 Sources and Magnitude (IEA, 2021b).

Most of the emissions are byproducts of essential industries, such as the energy

sector, refineries, and steel and cement production (IEA, 2021b). Therefore, it is difficult

to curb emission levels without technological advances or lowering life quality. The

available solution currently is to offset carbon levels in the atmosphere, on a large scale,

through capture technology. There is a myriad of techniques available to capture carbon,

which can be broadly classified into three categories:

1) Post-combustion: The primary method used in power plants to capture CO2 after

fossil fuels have been burned. This category remains the most implemented for

capture (IEA, 2020).

2) Pre-combustion: Mainly used in industrial processes and involves chemical

procedures to extract CO2 prior to combustion. The process generates synthetic gas

via gasification of fuel. The synthetic gas consists mainly of hydrogen and carbon

monoxide, which are later used to produce carbon dioxide (Mohammad et al.,

2020).

41
3) Oxy-fuel combustion: Burns fossil fuels in a pure oxygen environment instead of

air, leading to a high-purity CO2 byproduct. Subsequently, there is no need for a

chemical sorbent process for capture, however the method remains highly

expensive due to the specialized equipment associated with it (Elhenawy et al.,

2020; IEA, 2020; Ahmed and Zahid, 2019).

Establishing the stage at which carbon dioxide would be captured is important.

Based on which stage the capturing is planned to happen, the technologies chosen need to

be tailored to accommodate the feed gas (flue gas) composition, desired CO2 volume

capture, capture efficiency, water content, resultant CO2 purity, and the presence of other

chemicals (Islam et al., 2022; Mohammad et al., 2020).

The leading capture technologies presently are post-combustion, and its most

common applications are Absorption, adsorption, and separation by membranes (Chao et

al., 2021; Islam et al., 2022). Absorption uses a separation column containing a solvent to

capture carbon dioxide from the flue gas, and then it is isolated from it by a stripper. In the

membrane process, the gas is dehydrated and then flowed through a membrane, which

enables the separation of carbon dioxide from the flue gas by a vacuum pump. Adsorption

processes utilize solid adsorbent in the columns, with two types of adsorption: pressure-

swing and vacuum swing. Absorption is responsible for 57% of captured CO2, followed by

adsorption at 14%, membranes at 8%, and 21% split between other alternatives. Absorption

has a more significant share because it is a mature technology used in petrochemical

processes that happens to be the most economical, as well (Islam et al., 2022; Chao et al.,

2021).

New capturing technologies are frequently brought to the stage, but they tend to

face problems in scalability. A promising technique is direct air capture (DAC), in which

carbon dioxide is captured out of the air. The concept is familiar and has been proven to
42
capture CO2 from gases with relatively high concentrations of it, in post-combustion sites

(Azarabadi and Lackner, 2020). However, DAC still requires more research to become

more efficient in the extraction process. Until then, DAC is incapable of large-scale

deployment for its unattractive economics (Hanna et al., 2021). A currently discussed

solution to overcome this challenge is to offer governmental subsidies for each MtCO2

removed from the atmosphere. By subsidizing efforts to further develop capture

technology, breakthroughs are bound to happen sooner, making the venture more

economical.

2.2.2: CO2 Transportation

The transportation of carbon dioxide to the injection site of either a CCS or CCUS

project is highly subjective. The location of sequestration project, its proximity to a capture

site, and the transportation means available are highly influential for transportation

economic evaluation. For brevity, this section will glance over available forms of

transportation of carbon dioxide.

Both CCS and CCUS require similar infrastructure to transport the carbon dioxide

delivered to them. There are a few options available for CO2 transportation, such as low-

pressure (trucks), mid-pressure (tankers and rail), and high-pressure (pipelines). Both

onshore and offshore sequestration projects favor pipeline transportation for its large,

continuous, and economical supply (Islam et al., 2022). Other forms carry smaller volumes

and provide an intermittent supply. Nevertheless, pipeline networks are not extensive, and

the use of other forms of transportation is inevitable and more economical than establishing

a new pipeline.

Regardless of the choice of transportation method, there are precautions that need

to be taken into consideration. Carbon dioxide is a dangerous poisonous gas that poses

43
safety concerns to individuals in its vicinity. Pressurizing carbon dioxide adds another risk,

making it possibly explosive. In addition, the carrier vessel and CO2 must meet certain

specifications to avoid corrosion, scaling, or leakages. Generally, the supply chain required

to transport CO2 is extremely complex and expensive (DOE, 2022).

2.2.3: CO2 Storage and Utilization

Captured carbon dioxide could be repurposed to serve a multitude of purposes. Its

applications are incredibly diverse, and there are applications for it in every industry; some

of the applications are:

1) Food and beverage applications: Carbon dioxide is used in food packaging,

preservatives, carbonated drinks, and refrigeration.

2) Medical applications: Carbon dioxide is used in respiratory treatments, anesthesia

gas, cryotherapy, and surgical operations.

3) Industrial applications: Carbon dioxide is used in chemical generation (e.g.,

methane, ethane, urea), fire extinguishers, plant growth stimulation, creating light

rigid materials (carbon fiber), and carbon batteries.

All these applications encourage capture technology and motivate it to develop.

Especially when captured carbon dioxide is sellable as a commodity in multiple markets

with established demand.

In regards to the oil and gas industry, carbon dioxide is mainly used for EOR

projects. Nevertheless, with the advent of emission awareness, CO2-EOR gained a new

purpose to sequester as much carbon dioxide as possible as part of CCUS workflow. On

the other hand, although CCS offers no beneficial value to the oil and gas industry, the

mantle of CCS has been taken by it. As the experts of the subsurface, an environment of

44
high pressure and temperature with complex flow dynamics, whom have studied it for over

a century.

CCUS, in the oil and gas industry, targets hydrocarbon-bearing structures to

improve their production. In contrast, CCS targets any geological structures capable of

withholding CO2, such as saline aquifers and salt caverns. Carbon dioxide sequestration in

depleted or partially depleted reservoirs is generally preferred. It allows the oil and gas

industry to avoid the abandonment stage and the associated complex and costly exit

strategies by transforming the fields into a carbon sequestration project.

A survey of possible structures capable of containing carbon dioxide shows an

abundance of candidates all over the globe, with an estimated storage capacity of 8,000 to

55,000 Gt (IEA, 2021b). Most of the storage capacity (~75%), mapped in Figure 2.11, is

onshore in depleted oil and gas reservoirs and saline aquifer formations. Oil reservoirs are

more preferential than saline aquifers as they tend to have better seals (caprocks) and due

to the higher solubility of carbon dioxide in oil (Chen et al., 2022; Lake et al., 2019; Lake,

1989). In addition, Hepburn et al. (2019) evaluation of the potential carbon dioxide

utilization in EOR, by 2050, ranged from 100 – 1,800 MtCO2/yr. However, Hepburn et al.

economic analysis, based on 2015 material and CO2 cost and revenues, indicated that

CCUS would require governmental subsidies of $45 - $60 per MtCO2 to break even. In

contrast, carbon dioxide from natural sources tends to be less than 50% of the cost of

anthropogenic CO2 (Lake et al., 2019). As a result, anthropogenic CO2 accounts for 30%

of sequestered carbon dioxide (Islam et al., 2022).

45
Figure 2.11: Global CCUS and CCS Formation Candidates (IEA, 2021b).

2.2.4: CCUS and CCS Projects

Some of the earliest recognized large-scale CCUS and CCS projects are the

SACROC unit (1972) of the Permian basin and the Sleipner offshore gas field in Norway

(1996), respectively (Han et al., 2010; Ma et al., 2021). Over 175 million MtCO2 were

injected into SACROC, storing ~60% of it by 2009, while Sleipner stored 20 million

MtCO2 in a saline aquifer (Hovorka et al., 2021; Chadwick et al., 2010). According to

GCCSI (2023), there are 30 commercial-scale CCS projects as of 2023, most of which are

in the US and acquire CO2 from non-energy plant sources.

The main challenges facing CCS and CCUS projects stem from the low capture

capacity, carbon dioxide transportation and the projects' economics. If those obstacles are

tackled, there is an abundance of storage capacity, many times over the required volume,

as illustrated by Figure 2.11. Fortunately, as more capture facilities have been established,

the global capture capacity has grown to ~45 MtCO2/year (IEA, 2021b). By 2030, more

than 200 new projects are expected to come online, raising the capacity of CO2 capture to

46
over 200 MtCO2/year. The increased volume will fuel all sequestration efforts globally,

illustrated by Figures 2.12 and 2.13.

Figure 2:12: Commercial CCS Facilities (Operational and Under Development) (GCCSI,
2023).

Figure 2.13: Global CCUS Locations in Oil and Gas Fields (OGCSI, 2023).
47
2.2.5: CCUS and CCS Key Mechanisms

CCS and CCUS attempt to sequester as much as possible carbon dioxide in the

subsurface. To achieve this goal, it is not sufficient to solely inject it underground. The

trapping structure’s petrophysics must be evaluated to ensure its capability of withholding

CO2. Otherwise, carbon dioxide could seep into other formations contaminating them or

possibly leaking back to the surface. In addition, if the injection is done without proper

planning, a good portion of it might be produced back to the surface, along with reservoir

fluids. Therefore, understanding the trapping mechanisms of CO2 in a reservoir is crucial.

In the subsurface, four main mechanisms control CO2 trapping, which are structural

trapping, solubility trapping, mineral trapping, and residual trapping (Chen et al., 2022;

Cao et al., 2020; Han et al., 2010; Delshad et al., 2010).

Structural trapping is the primary mechanism of carbon dioxide sequestration. In

the absence of structural trapping, storage of CO2 would not be possible due to the lack of

a confining structure. A structural trap refers to any geological trap or barrier capable of

holding carbon dioxide in place due to the presence of faults, seals, or stratigraphic

variation. In this mechanism, the trapping is governed by buoyancy and capillary forces,

and the trapped carbon dioxide is mobile. Buoyancy forces drive the fluid to flow, while

capillary forces prevent it. For the trapping to be possible, the formation must be overlain

by a caprock (ultra-low permeability) to prevent gases from escaping, by ensuring capillary

forces outweigh buoyance forces.

Solubility trapping is vital for carbon dioxide storage. CO2 is trapped by dissolving

into in-situ fluids (oil and gas). The solubility of a substance is a measure of how much of

it is dissolvable within a certain solvent, and is affected by temperature, pressure, salinity,

and chemical makeup of both the solvent and the solute. For example, carbon dioxide

solubility in oil is much higher than in water or brine (Chen et al., 2022). Solubility plays
48
an important role in the propagation of carbon dioxide and plume migration inside the

reservoir (Kumar et al., 2005). However, the trapped volume is dynamic, and changes as

reservoir conditions change.

Residual trapping stores CO2 as an immobile gas within the reservoir. When CO2

is injected into a reservoir, it migrates through the pore space of the rock, displacing the

fluids that were originally present. However, some of the CO2 can become trapped within

the pores, especially smaller ones, and crevices of the rock, rendering it immobile.

Subsequently, residual trapping plays an important role in carbon dioxide propagation and

plume migration, as well (Kumar et al., 2005). The trapping is governed by capillary forces,

which is why it is also known as capillary trapping. Hence, residual trapping magnitude

depends on the rock’s properties, such as pore size, fluid affinity, and relative permeability

of all phases (Peters, 2012). Moreover, residual trapping gives rise to hysteresis,

magnifying the magnitude of trapping (Spiteri et al., 2005). The impact of capillary

pressure (𝑃𝑐 ) required to trap or mobilize fluids within a porous medium is highlighted in

Figure 2.14, where higher capillary pressure (work) is needed for drainage than for an

imbibition cycle to displace the same saturation. Figure 2.14 shows the capillary pressure

curves for a sample initially saturated with water, which gas is injected into (drainage), the

gas is unable to penetrate the medium until a certain 𝑃𝑐 is reached (entry capillary pressure),

allowing it to overcome the capillary forces and displace the water. Conversely, in

imbibition the gas is extracted from the medium, the residual gas saturation would differ

from the initial gas saturation due to hysteresis and trapping (Peters, 2012). As a result, the

relative permeabilities curves of the phases are affected, since they are dependent on their

respective saturations.

49
Figure 2.14: Capillary Pressure Curves (Peters, 2012).

Mineral trapping of carbon dioxide is the most stable and safe storage method;

however, it takes years to take effect (Zhang and Depaolo, 2017). Mineral trapping occurs

when acid CO2 (gas, supercritical, or dissolved) interacts with minerals present in the

reservoir, leading to the formation of stable carbonates, like calcite. However, carbon

dioxide must dissolve into brine first for it to interact with minerals. Certain formations

exhibit fast mineralization due to their richness in calcium, magnesium, or iron such as

mafic and ultramafic rocks. An example of a field application of mafic rocks for mineral

trapping is the Carbfix project in Iceland. Carbfix used CO2 as an injectant into basaltic

lava, and mineralization was confirmed to happen in less than a year (Snæbjörnsdóttir et

al., 2017; Gunnarsson et al., 2018; Matter et al., 2016). However, that is not the case for a

typical oil and gas field, and because of that mineral trapping is insignificant relative to

other trapping mechanisms, especially short-term (NASEM, 2019). In addition, due to the

long duration it takes for mineral trapping, in oil and gas fields, to take place on a large

enough scale, it is often overlooked, as shown in Figure 2.15.

50
Figure 2.15: CO2 Trapping Contribution of Trapping Mechanisms. (Left) Carbonate and
Sandstone Reservoir. (Right) Mafic and Ultramafic Rocks (NASEM, 2019).

2.2.6: CCUS and CCS Numerical Simulation

Proper planning of CCS and CCUS projects requires holistic analysis incorporating

all aspects and stages of the project. Starting from the first stage of carbon capture and

ending with the injection of carbon dioxide into the reservoir. Such type of an analysis

ensures the success and robustness of the pursuit. However, the plan needs to incorporate

appropriate assumptions and deliverables, otherwise failure is inevitable. The focus of this

work is regarding the injection of carbon dioxide into a carbonate oil reservoir.
There has been a lot of work done in the area of modeling the subsurface, to assist

decision making and risk abatement. Continuously striving to lower the risks associated

with hydrocarbon endeavors, numerical modeling became a standard tool in the oil and gas

industry. Encouraged by advances in the computing industry, boosting the reliance of

modeling capabilities, complexities, and accuracy at a reduced runtime. As a result,

numerical modeling permeated into all aspects of the oil and gas industry, from facility

design to hydrocarbon’s PVT.

51
Modeling could be done via analytical models or numerical models. Analytical

models are relatively simplistic and provide closed-form solutions. The Buckley-Leverett

model is a classic analytical model used to predict two-phase flow in porous media. On the

other hand, numerical models provide more detailed and accurate predictions of fluid

behavior. Finite difference, finite element, and discrete element methods are commonly

used in numerical modeling, which solve flow, heat transfer and chemical equations in a

discretized structure of the subsurface. Therefore, numerical models are complex and time

demanding, while analytical models are fast and simple. Nevertheless, numerical models

are mostly utilized due to their capability to accommodate complex scenarios.

Carbon dioxide sequestration has two forms, like CO2-EOR, miscible and

immiscible floods. Thus, experience acquired from modeling CO2-EOR was employed as

the basis of sequestration projects. Since sequestration projects are concerned with carbon

dioxide retention, more research has been dedicated towards understanding its trapping

mechanisms, and how to properly model it. The knowledge gained is valuable, however

most of the modeling studies relied on synthetic models (Sun et al., 2020). Although

synthetic models are important in isolating the impact of a phenomena from outside

variables, it should not be relied on for project planning. Synthetic modeling offers possibly

misleading results as it does not account for multiple phenomena interacting with each

other due to its simplified nature. In addition, current geochemistry evaluation focuses on

saline aquifers, and relies on CO2-EOR knowledge for hydrocarbon reservoirs, which does

not consider long-term geochemistry (Sun et al., 2020). Hence, adding to the inaccuracy of

model prediction. An accurate carbon sequestration model must incorporate the major

phenomena involved, such as the main trapping mechanisms discussed previously.

Major findings from the literature for modeling sequestration projects within

partially depleted oil reservoirs are:


52
1) Recovery factor increases as the injection rate of carbon dioxide increases in

miscible flood (Safi et al., 2016; Ghulami et al., 2015). This might be due to the

higher injection pressure required to enable higher injection rate. The higher

injection pressure enhances miscibility and overcomes capillary forces enabling

CO2 to contact a larger volume of oil.

2) Solubility of CO2 decreases with increasing salinity and temperature, but increases

as pressure increases (Sun et al., 2020; Kutsienyo et al., 2019; Ampomah et al.,

2015; Lake et al., 2014).

3) Hysteresis plays an important role in carbon dioxide storage, increasing the

contribution of residual trapping to the overall trapped CO2 (Doughty and Myer,

2009; Juanes et al., 2006; Spiteri et al., 2005; Kumar et al., 2005). Otherwise, part

of the CO2, which was supposed to be trapped, would exaggerate the dissolved

carbon dioxide by the simulator and overestimate mineral precipitation.

Multiple studies oversimplify the numerical modeling neglecting major trapping

mechanisms (Ampomah, 2016; Lashgari et al., 2019).

53
Chapter 3: Methodology and Assumptions

This chapter elaborates on the analysis of the simulation results. The chapter

discusses the reservoir simulator, the numerical model utilized, and its imposed

assumptions.

3.1 NUMERICAL SIMULATION MODEL

The analysis in this work employs numerical models and a compositional reservoir

simulator software developed by the Computer Modeling Group (CMG) called GEM.

GEM models the fluid flow and transport in the subsurface. Moreover, it is capable of

modeling multi-phase and multi-component phase behavior, and geochemical modeling,

making it an ideal tool for simulating CO2-EOR and sequestration projects. GEM allows

for a myriad of assumptions and correlations for the simulation process, some of which

will be discussed in this chapter, which lead to the decision to use GEM over other

software. For example, alternative simulation software has some limitations, such as lack

of compositional modeling (IMEX), or geochemistry (ECLIPSE).

3.1.1: Field Model Description

The numerical model used is based on an actual reservoir to emulate real-life

reservoir behavior rather than an idealistic synthetic model. The model used is a sector

model of a large, tight carbonate field. The reservoir is shallow and has a relatively low

temperature of 130 °F and contains a superlight oil (API > 40). A carbonate reservoir was

chosen for the model, as ~60% of proven reserves are contained within them (Chen et al.,

2018; Sheng, 2013). The general properties of the reservoir model are described by Table

3.1. The reservoir exhibits high heterogeneity and mud stritations (thin layers) are

prevalent, as shown by the thin tighter layers in the porosity and permeability distributions

54
in Figure 3.1. To accommodate the inherent heterogeneity of the subsurface, the spatial

properties of the model have been based on field measurements. To further mimic real

reservoir’s structure, the geometry of the layers within the model were not assumed to be

flat but rather have varying thickness, depths, and pinch-outs, as illustrated by Figure 3.2.

Table 3.1: Reservoir Model Properties.

Property Value
Avg. Porosity 6.35%
Avg. Permeability 62 md
Initial Pressure 3120 psi
kv/kh 0.005
HCPV 3.47x108 ft3
Avg. Thickness 510 ft
Length 3830 ft
Width 3775 ft

(a) (b)

Figure 3.1: Model Spatial Property Distribution. (a) Porosity (b) Permeability in md.
The model illustration is scaled by a factor of 3 on the z-axis.

55
Figure 3.2: Reservoir Model Vertical Grid Size in ft. The model illustration is scaled by
a factor of 3 on the z-axis.

The reservoir model consists mainly of limestone, the carbonates of which are

icehouse carbonates that are known to be some of the most heterogeneous carbonates

(Isdiken, 2013). In accordance with the reservoir’s geology, four rock types were

established in the model. Figure 3.3 shows the porosity permeability correlation of three

of the four rocks, while the fourth rock, is not shown due to not having sufficient data

points. To accommodate all the complexities of the reservoir, a 32,674 corner-point grid
model is required, with the average dimensions of each grid being 123ft x 122ft x 15ft

(x,y,z). Additionally, the relative permeabilities of each rock type and capillary pressure

curve are showcased in Figure 3.4.

56
Figure 3.3: Rock Permeability vs Porosity Correlation.

57
1

Relative Permeability
0.8
0.6
0.4
0.2
0
0.2 0.4 0.6 0.8 1
Sw
krw1 krw2 krw3 krw4
kro1 kro2 kro3 kro4 (a)
1
Relative Permeability

0.8
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8
Sg
krw1 krw2 krw3 krw4
krg1 krg2 krg3 krg4
(b)
Capillary Pressure (Pcow),

0.75

0.5
psi

0.25

0
0.2 0.4 0.6 0.8 1
Sw
Pc1 Pc2 Pc3 Pc4
(c)

Figure 3.4: Petrophysical Properties of Reservoir Rocks. (a) oil and water relative
permeabilities. (b) gas and water relative permeabilities. (c) capillary
pressure curves.
58
3.1.2: Assumptions and Considerations

For the modeling endeavor, assumptions are integrated to enable the analysis of a

complex system or to isolate a phenomenon and analyze its behavior separately. Identifying

and acknowledging the assumptions made are crucial to understanding the behavior of the

system under consideration. Throughout this section, the key assumptions employed in the

model are discussed, aiming to provide a comprehensive understanding of the model and

the context in which it operates.

3.1.2.1: Minimum Miscibility Pressure (MMP)

Minimum miscibility pressure (MMP) evaluation is a crucial element in solvent

EOR, as it determines miscibility. Miscible solvent EOR is more desirable due to its

superior performance when it comes to oil recovery (Chen et al., 2018; Fath and

Pouranfard, 2014; Lake et al., 2014). However, in terms of carbon dioxide storage within

the subsurface, there is still some uncertainty. Thus, miscibility must be taken into account

when evaluating hybrid CO2-EOR and sequestration projects in oil reservoirs.

Minimum miscibility pressure is a function of reservoir pressure, temperature, and

the compositions of oil and injected carbon dioxide stream (Lake et al., 2014; Lake, 1989;

Stalkup, 1978). The presence of any impurities in the CO2 gas injected (e.g., CO, H2S,
NOx, water, etc.) could alter MMP (Lake et al., 2014; Li and Yan, 2009). Consequently,

compositional modeling is paramount for dynamic modeling of miscibility. The

composition of the oil used in the field is shown in Table 3.2. It is common to create

pseudo-components combining multiple components to reduce the number of components

modeled and ease the simulation process, which is why all components are typically

grouped beyond C6 (Lake et al., 2014). Moreover, components of close molecular weight

and in small concentrations could also be grouped to enable GEM to easily track them,

59
which is the case in this work where nitrogen and methane and C5 and C6 were grouped

as two pseudo-components.

Table 3.2: Reservoir Oil Composition.

Component Mol%
N2 2.17
CO2 0.25
C1 27.31
C2 10.99
C3 11.75
C4 7.34
C5 4.53
C6 3.63
C7+ 32.03

Experimental measurement of MMP provides a reliable estimate. Some of the

common measurements are slim tube, rising bubble, and vanishing interfacial tension, with

slim tube test being the most common test (Chen et al., 2022). A major drawback of the

experimental measurements are the long time and the high expense it demands (Chen et

al., 2022). Therefore, many models, numerical and analytical, were developed to

circumvent the costly and time intensive experiments.


Literature of experimental work establishes that MMP pressure increases with

increasing temperature and decreases with increasing pressure (Chen et al., 2022; Lake,

1989). In addition, salinity has been found to increase MMP as it increases (Kutsienyo et

al., 2019). Furthermore, the composition of the injected carbon dioxide was observed to

alter MMP pressure. Merely by introducing impurities to the CO2 feed, the MMP is

affected, at the same pressure, temperature, salinity and oil composition. For instance,

60
hydrogen sulfide (H2S) is one of the few known gases to lower the MMP. Conversely,

nitrogen and methane increase MMP (Chen et al., 2018; Trivedi et al., 2007).

Utilizing CMG-WINPROP, a fluid property characterization tool, analysis of fluid

behavior under different conditions is possible. In addition, WINPROP can be used to tune

the equation of state (EOS) and to assess miscibility pressures calculations. Nevertheless,

it was used to confirm the accuracy of CMG’s characterization of the oil properties. The

prediction and accuracy of properties are shown in Table 3.3, and it agrees with the

experimental measurements. In this work, both miscible and immiscible injection schemes

will be explored.

Table 3.3: Oil Properties.

Property Model Value Actual Value

Oil API 47 45

Oil Viscosity 0.4 cp 0.4 cp

Bubble Point Pressure 1819 psi 1792 psi

Solution GOR 900 SCF/STB 849 SCF/STB

MMP 1900 psi 2000 psi

3.1.2.2: Relative Permeability Hysteresis Model

Residual trapping (capillary trapping) is one of the major carbon dioxide trapping

mechanisms in the subsurface that gives rise to hysteresis. Hysteresis influences fluid flow

in porous media based on its previous saturation history by altering the relative

permeability of the present phases. For example, during imbibition, part of the non-wetting

phase gets trapped, causing the saturation of the wetting phase to be less than it is on the

drainage curve at the same pressure, especially in cyclic drainage/imbibition processes

61
(Peters, 2012; Ghomian, 2008; Larsen and Skauge, 1998). This phenomenon is more

evident in cyclic injection of wetting and non-wetting phases, as illustrated by Figure 3.5.

Generally, relative permeability of a phase is a function of its saturation as well as

the porous media’s wettability and saturation history (Larsen and Skauge, 1998). The

alteration in relative permeability based on the saturation history is what is known as

hysteresis. The magnitude of hysteresis’s impact is influenced by the present fluids’

characteristics and the porous media’s petrophysical properties (Lake et al., 2014; Peters,

2012; Larsen and Skauge, 1998). According to Land (1968), the non-wetting phase starts

by invading the larger pores, and as its saturation increases, it starts invading smaller pores,

in which it will be trapped once its saturation decreases, as illustrated by Figure 3.5.

Figure 3.5: Capillary Trapping of Non-wetting Phase in a Porous Medium. Each row
reaches further points on the drainage curve than the row before it (higher
Snw), while the second column re-saturates via a screening curve (lowering
Snw), showing the resultant trapping of the non-wetting phase. (Lake et al.,
2014).

62
In two-phase fluid systems, only one independent saturation exists, allowing

relative permeabilities to move in two directions along the drainage or imbibition curves.

Nevertheless, this assumption is not applicable in three-phase systems (Larsen and Skauge,

1998). In three-phase systems, the number of independent saturations increases (at least 2),

opening countless new directions for the system to follow (Larsen and Skauge, 1998).

Therefore, accurately modeling the three-phase flow system’s relative permeabilities is

essential in modeling the entrapment of carbon dioxide in the subsurface by hysteresis.

Additionally, incorporating hysteresis, in accordance with reality, improves the oil RF

prediction by reducing the gas phase’s mobility and replacing the trapped oil with gas,

reducing oil’s residual saturation and mobilizing it (Beygi et al., 2015). Furthermore,

Hosseini et al. (2012) report that gas hysteresis is more significant in water-wet systems

than water or oil. Therefore, only gas hysteresis is assumed in this study, along with a

modifier for oil residual saturation as a function of gas saturation.

Various workflows and models exist to model hysteresis, but GEM contains a few

of the common models, such as Carlson’s (1981), Land’s (1968), Larsen and Skauge’s

(1998) three-phase model, and Carlson’s three-phase model. Other models are also offered,

but they have been disregarded due to their simplicity. Although Carlson’s and Land’s

models are proposed to simulate a three-phase system, they are inaccurate because they

were developed for two-phase flow systems (GEM User Manual, 2022.10; Kutsienyo et

al., 2019; Larsen and Skauge, 1998). That is consistent with the findings of Skauge and

Larsen (1994), wherein it was observed that the residual oil saturation in three-phase flow

is considerably lower than in two-phase systems. The deviation of the two-phase flow

models prediction from the three-phase flow models can be observed by the mass of

trapped CO2 in Figure 3.6, which highlights the historically matched case run with different

hysteresis models. Conversely, Figure 3.6 shows Carlson’s three-phase results, which
63
adapts Larsen and Skauge’s work to account for three-phase flow, and Larsen and Skauge’s

models coupled with Land’s have a similar performance. Therefore, Larsen and Skauge’s

model was chosen for the analysis of this work, especially since it has been developed to

function under cyclic processes such as WAG.

Figure 3.6: Hysteresis Models Performance on Historically Matched Case.

Larsen and Skauge’s model accounts for the relative permeability of present phases

and their saturation history, unlike two-phase models (Kutsienyo et al., 2019; Larson and

Skauge, 1998). In addition, the three-phase model does not operate under the assumption

of reversible scanning curves to estimate relative permeabilities, instead it creates

secondary curves for a more accurate representation of the fluid flow in the model

(Kutsienyo et al., 2019). The relative permeabilities are cycle-dependent in this model, that

is that the relative permeabilities are altered after each cycle.

The relative permeability model starts off by following the primary drainage or

imbibition curve’s relative permeability, as shown by curve G1 in Figure 3.7’s loop 1. Any
64
subsequent transitions between imbibition and drainage are modeled by screening curves.

Different scanning curves are employed depending on the gas saturation, increasing or

decreasing, both of which use the primary gas relative permeability, such as G1 in Figure

3.7, along with Land’s trapping function, as described by eqn. 3.1 and 3.2, where C is

Land’s constant which relates the initial gas saturation (Sgi) and the trapped gas saturation

(Sgr). Accordingly, in a waterflood the primary imbibition and drainage curves are

sufficient, due to lack of a third phase (gas).

Figure 3.7: Hysteresis Loops in a Three-phase Oil/Water/Gas Flow Highlighting Gas


Relative Permeability Reduction (Larsen and Skauge, 1998).

1 1
𝐶= − (3.1)
𝑆𝑔𝑟 𝑆𝑔𝑖

(𝑆𝑔ℎ − 𝑆𝑔𝑐𝑟𝑖𝑡 )
𝑆𝑔𝑟 = 𝑆𝑔𝑐𝑟𝑖𝑡 + (3.2)
1 + 𝐶 ∗ (𝑆𝑔ℎ − 𝑆𝑔𝑐𝑟𝑖𝑡 )

Drainage secondary curves (increasing gas saturation) are calculated as follows:

65
𝑑𝑟𝑎𝑖𝑛 (3.3)
𝑘𝑟𝑔 (𝑆𝑔 ) = 𝑘𝑟𝑔 (𝑆𝑔𝑓 )

1 2 4
𝑆𝑔𝑓 = 𝑆𝑔𝑐𝑟𝑖𝑡 + [(𝑆𝑔 − 𝑆𝑔𝑟 ) + √(𝑆𝑔 − 𝑆𝑔𝑟 ) + (𝑆𝑔 − 𝑆𝑔𝑟 )] (3.4)
2 𝐶

𝑑𝑟𝑎𝑖𝑛 𝑖𝑛𝑝𝑢𝑡 𝑖𝑛𝑝𝑢𝑡 𝑆𝑤𝑐𝑜𝑛 𝛼


[𝑘𝑟𝑔 (𝑆𝑔 )] = {[𝑘𝑟𝑔 (𝑆𝑔 ) − 𝑘𝑟𝑔 (𝑆𝑔𝑠𝑡𝑎𝑟𝑡 )] 𝑖𝑚𝑏
∗ ( 𝑠𝑡𝑎𝑟𝑡 ) } + [𝑘𝑟𝑔 (𝑆𝑔𝑠𝑡𝑎𝑟𝑡 )] (3.5)
𝑛 𝑆𝑤 𝑛
𝑛−1

where:
• 𝑑𝑟𝑎𝑖𝑛
𝐾𝑟𝑔 (𝑆𝑔 ) = calculated 𝑘𝑟 for the secondary drainage curve at gas saturation
𝑖𝑛𝑝𝑢𝑡
• 𝐾𝑟𝑔 (𝑥) = 𝑘𝑟 input at gas saturation 𝑥

• 𝑖𝑚𝑏
𝐾𝑟𝑔 (𝑥) = 𝑘𝑟 at the gas saturation x

• 𝑆𝑔𝑠𝑡𝑎𝑟𝑡 = gas saturation at the start of the secondary drainage

• 𝑆𝑔𝑓 = free gas saturation

• 𝑆𝑤𝑐𝑜𝑛 = connate water saturation

• 𝑆𝑤𝑠𝑡𝑎𝑟𝑡 = water saturation at the start of the secondary drainage curve

• α = reduction exponent of the gas phase’s relative permeability in the

presence of mobile water

• n = hysteresis loop number

The first set of brackets, in eqn. 3.5, transforms the primary drainage curve G1 to
be zero at gas saturation 𝑆𝑔𝑠𝑡𝑎𝑟𝑡 . The second set reduces the gas relative permeability in

presence of mobile water, indicating that the relative permeability of the gas is a function

of both the gas and water’s saturation history. On the other hand, any secondary imbibition

curves are dealt with by transforming the saturations based on eqns. 3.1, 3.2 and 3.3, such

that the imbibition gas relative permeability is calculated as follows:

66
𝑡𝑟𝑎𝑛𝑠
(𝑆𝑔𝑟 )𝑛 = (𝑆𝑔𝑟 )𝑛 − (𝑆𝑔𝑒𝑛𝑑 )𝑛−1 (3.6)

𝑡𝑟𝑎𝑛𝑠
(𝑆𝑔𝑖 ) = (𝑆𝑔𝑖 )𝑛 − (𝑆𝑔𝑒𝑛𝑑 )𝑛−1 (3.7)
𝑛

1 1
(𝐶 𝑡𝑟𝑎𝑛𝑠 )𝑛 = ( 𝑡𝑟𝑎𝑛𝑠 ) − ( 𝑡𝑟𝑎𝑛𝑠 ) (3.8)
𝑆𝑔𝑟 𝑆𝑔𝑖
𝑛 𝑛

1 2 4
(𝑆𝑔𝑓 )𝑛 = 𝑆𝑔𝑐𝑟𝑖𝑡 + [(𝑆𝑔 − 𝑆𝑔𝑟 ) + √(𝑆𝑔 − 𝑆𝑔𝑟 ) + 𝑡𝑟𝑎𝑛𝑠 (𝑆𝑔 − 𝑆𝑔𝑟 )] (3.9)
2 𝐶

𝑡𝑟𝑎𝑛𝑠
(𝑆𝑔𝑓 ) = (𝑆𝑔𝑓 )𝑛 + (𝑆𝑔𝑒𝑛𝑑 )𝑛−1 (3.10)
𝑛

𝑖𝑚𝑏 𝑑𝑟𝑎𝑖𝑛 𝑡𝑟𝑎𝑛𝑠


[𝑘𝑟𝑔 (𝑆𝑔 )]𝑛 = [𝑘𝑟𝑔 (𝑆𝑔𝑓 )] (3.11)
𝑛

The gas phase trapping induced by hysteresis impacts the oil residual saturation

(Sor). The gas displaces the oil and competes with it to exist within the reservoir’s pores.

As a result, the Sor value is reduced by the gas invading the pores, mobilizing trapped oil,

and increasing the oil recovery. The modified residual oil saturation is modeled as dictated

by eqn. 3.12, where a is an input parameter used to modify 𝑆𝑜𝑟 .

𝑚𝑜𝑑 (3.12)
𝑆𝑜𝑟 = 𝑆𝑜𝑟 − 𝑎(𝑆𝑔 − 𝑆𝑔𝑓 )

To apply this workflow to the model of interest in this study, relative permeabilities

and Land’s constant C or maximum residual gas saturation (Sgrmax) are required as

simulation inputs. In addition, since the model consists of 4 rock types, the inputs are

needed for all of them separately. Regarding the trapping function input, experimental data

for the field is available for one rock type, which is Sgrmax = 0.39. Therefore, C was
calculated based on Sgrmax as input for 𝑆𝑔𝑟 in eqn. 3.1, and the value was assumed to be the

same for the other rock types. Consequently, each rock type will have its own S grmax since
67
the calculation utilizes the residual saturation of gas from the primary imbibition curve. In

terms of 𝛼 and 𝑎, the values were determined based on analogous work with the values of

1.69 and 1.19, respectively.

3.1.2.3: CO2 Solubility in Water

Henry’s law is used to model the solubility of the components present in the

aqueous phase. The model is implemented to observe CO2 solubility into the aqueous phase

and its contribution to sequestration as a function of temperature, pressure, and salinity; in

this work, salinity is assumed to be constant to reduce the computational complexity of the

model. The general form of Henry’s law is illustrated by eqn. 3.13, where 𝑓𝑖 is the fugacity

of component 𝑖 at a specific composition within the aqueous phase 𝑥𝑖 and 𝐻𝑖 is the

corresponding Henry’s constant. Henry’s constant is calculated according to eqn. 3.14 as


a function of pressure (𝑃: pressure, 𝑃𝑟𝑒𝑓 : reference pressure), temperature (𝑇) and molar

volume (𝑣 ∞ : molar volume at infinite dilution of the component).

𝑓𝑖 = 𝐻𝑖 𝑥𝑖 (3.13)

𝑟𝑒𝑓 𝑣𝑖∞ ∗ (𝑃 − 𝑃𝑟𝑒𝑓 )


ln(𝐻𝑖 ) = ln(𝐻𝑖 ) + (3.14)
𝑅𝑇

GEM offers only Henry’s model to account for solubility. However, it allows for

alternative correlations to modify Henry’s constants (e.g., Harvey’s model and modified

Harvey’s model). GEM also allows for manual input instead of resorting to any correlation,

to accommodate empirical data. Henry’s law was used in this work for all components of

the aqueous phase, but for carbon dioxide.

In regard to carbon dioxide, the solubility modeling is enhanced by incorporating

Harvey’s correlation (1996) in Henry’s constant calculation. As a result, more complex

68
temperature-dependent correlations are implemented for 𝑣 ∞ , in eqn. 3.14, and the salting-

out phenomenon is accounted for. Incorporating Harvey’s work in the calculation of 𝐻 is


done by modifying the reference Henry constant (𝐻 𝑟𝑒𝑓 ) and setting 𝑃𝑟𝑒𝑓 to the solvent

(water) vapor pressure (𝑃1𝑠 ) as a function of temperature (𝑇 ∗ : reduced temperature).

Equation 3.15 shows Harvey’s correlation used to calculate 𝐻 𝑟𝑒𝑓 , where A, B, and C are

constants introduced to constrain the correlation’s behavior allowing it to function under

high temperatures (~680 °F for CO2) (Harvey, 1996).

𝐴 𝐵(1 − 𝑇 ∗ )0.355
ln(𝐻 𝑟𝑒𝑓 ) = ln(𝑃1𝑠 ) + + + 𝐶 exp(1 − 𝑇 ∗ ) (𝑇 ∗ )−0.41 (3.15)
𝑇∗ 𝑇∗
𝑇
𝑇∗ = (3.16)
𝑇𝑐

To further enhance the accuracy of modeling the components dissolving within the

aqueous phase, the aqueous phase needs to be adequately modeled. Therefore, well

established empirical correlations were applied to account for the aqueous medium’s

density (Rowe and Chou, 1970) and viscosity (Kestin et al., 1981).

3.1.2.4: Thermal Modeling

Two versions of the model are used, a thermal version and an isothermal version.

Most of the analysis is conducted under the assumption of isothermal conditions, due to

the high computational time associated with thermal modeling. However, few thermal

models were run to evaluate the impact of thermal interaction on the results, to mitigate its

computational complexity. In the Results chapter, the distinction will be made clearly,

establishing thermal and isothermal models.

69
3.1.2.5: Equation of State (EOS)

Equation of state (EOS) is a correlation that relates temperature, pressure and molar

volume to each other (Lake, 1989). One of the earliest and simplest forms of EOS is the

ideal gas law which was introduced by Boyle in 1663. Since then, many complex EOSs

were introduced, some of which are shown in Table 3.4. The cubic EOS are the most

common as they are more fit for modeling three-phase systems, and van der Waals is one

of its earliest examples. The general form of cubic EOS is expressed in a pressure-explicit

form, as shown by eqn. 3.17, where θ, η, δ, and ε are based on experimental data and are

specific to each EOS. The performance and limitations of each model differ; thus, the

choice of EOS needs to be based upon the desired application.

Table 3.4: Examples of EOS Models.

EOS Year

van der Waals 1873

Clausius 1880

Berthelot 1900

Redlich-Kwong 1949

Redlich-Kwong-Soave 1972

Lee-Erbar-Edmister 1973

Peng-Robinson 1976

Patel-Teja 1982

3P1T 1987

𝑅𝑇 𝜃(𝑉 − 𝜂)
𝑃= − (3.17)
𝑉 − 𝑏 (𝑉 − 𝑏)(𝑉 2 + 𝛿𝑉 + 𝜀)

70
Redlich-Kwong-Soave (SRK) and Peng-Robinson (PR) EOS models have been

utilized the most in modeling gas EOR processes, due to their reliability in modeling gases

and mixtures (Lake et al., 2014). GEM recognizes their reliability and accuracy and only

offers SRK and PR EOS in their compositional modeling software GEM. PR EOS was

chosen due to its ability to provide more accurate predictions for binary mixtures of CO2

and impurities commonly found in carbon dioxide streams, such as nitrogen, methane, and

ethane (Li and Yan, 2009).

3.1.2.6: Geochemistry

Reservoir fluids exist in equilibrium due to the extended period the fluids resided

in the reservoir. Any possible interaction that could have happened had already subsided

due to the long period the fluids were entrapped within the reservoir’s structure. Once the

in-situ equilibrium is disturbed, through production or injection, fluid-fluid, and fluid-solid

interactions are inevitable. Thus, in waterflood processes, the injected water is first tested

for its compatibility with the reservoir rock mineralogy, in-situ fluids, and microorganisms.

In this work, the geochemical interaction caused by CO2 injection is evaluated.

Geochemistry is important for carbon dioxide sequestration as it dictates the extent

of mineral trapping. However, geochemistry impact is often considered minor and

neglected (Beygi, 2016). The solid species often modeled for carbonate reservoirs are

calcite, dolomite, and anhydrite, as the most common minerals in carbonates. This work

only considers calcite, because it represents over 60% of the reservoir considered by

volume, while dolomite and anhydrite make up less than 10% each, and the rest is split

between different minerals and clays. In addition, calcite reaction to carbon dioxide is

extremely fast, while the other mineral reactions are much slower in comparison, some of

which requiring >100 years (NASEM, 2019). Therefore, the reactions implemented in this

71
work, highlighted in Table 3.5, include only calcite and common intra-aqueous reactions

related to calcite (Kutsienyo et al., 2019).

Table 3.5: Modeled Geochemical Reactions.

Reaction Reaction Chemical Formula

Calcite 𝐶𝑎𝐶𝑂3 + 𝐻 + ↔ 𝐶𝑎2+ + 𝐻𝐶𝑂3−

Carbonic acid 𝐶𝑂2 + 𝐻2 𝑂 ↔ 𝐻 + + 𝐻𝐶𝑂3−

Aqueous reaction 𝐶𝑂32− + 𝐻 + ↔ 𝐻𝐶𝑂3−

Aqueous reaction 𝑂𝐻 − + 𝐻 + ↔ 𝐻2 𝑂

To emulate geochemical reactions in the subsurface, geochemical activity modeling

is required, which utilizes chemical thermodynamics and kinetics. GEM avails a few

activity models to handle the geochemical interactions within the aqueous phase: Ideal,

Debye-Huckel, B-dot, Pitzer, and Pitzer2. The Ideal model assumes all component activity

coefficients to be equal to unity. Whereas the other models have their own values rather

than assuming unity. By default, all the models, other than the ideal, refer to Wolery (1992)

and Parkhurst and Appelo (2013) work for activity coefficients, unless the user specifies

otherwise (GEM User Manual 2022.10). The most accurate model is the Pitzer2 and is

appropriate for high salinity solutions, especially high ionic strength solutions (Pitzer and

Kim, 1974; GEM User Manual 2022.10). Due to the exhaustive and rigorous workflows

of the Pitzer models, they are computationally expensive, especially Pitzer2 with its

increased accuracy. Therefore, the B-dot model was chosen for the analysis as it is an

extension of the Debye-Huckel model. Equations 3.18 and 3.19 highlight the general

forms of Debye-Huckel and B-dot, respectively. Where 𝛾𝑖 is the activity coefficient of ion

𝑖 and 𝐴, 𝐵̇, åi and 𝑧𝑖 are constants for ion 𝑖, and 𝐼 is the solution’s ionic strength. In addition,
72
both models perform closely but start diverging close to the ionic strength of 0.74 molal

(log ionic strength ~ -0.3 molal), as illustrated by Figure 3.8. The model is generally

accurate up to 300 °C and ionic strength of 3 molal (Rickard, 2012).

𝐴𝑧𝑖2 √𝐼
log(𝛾𝑖 ) = − (3.18)
1 + åi 𝐵̇ √𝐼

𝐴𝑧𝑖2 √𝐼
log(𝛾𝑖 ) = − + 𝐵̇𝑖 𝐼 (3.19)
1 + åi 𝐵̇ √𝐼

Figure 3.8: Activity Coefficient vs. Ionic Strength Logarithm of Activity Models
(Rickard, 2012).

The aqueous phase chemical reactions equilibrium in GEM are governed by eqn.
3.20, where K eq is the chemical equilibrium constant, Q is the activity product, and the

exponent 𝑣𝑖 equal to the stoichiometric coefficient. The chemical equilibrium (K eq ) is a

user input, and the user could elect to define it based on Wolery’s or Phreeqc’s databases.
In addition, the user can define log (K eq ) as a constant or as a function of temperature by

defining the ai coefficients in eqn. 3.21. The approach implemented in this work is the

73
latter using Wolery’s database, to give the thermal models more integrity in reflecting

reality.
𝑛𝑐𝑡
𝑣
K eq = Q , where 𝑄 = ∏ 𝛼𝑘𝑖 (3.20)
𝑖=1

log (K eq ) = a0 + a1 𝑇 + a2 𝑇 2 + a3 𝑇 3 + a4 𝑇 4 (3.21)

Mineral precipitation and dissolution reactions rate are calculated by the transition

state theory (TST), which is expressed by GEM as in eqn. 3.22, 23 and 24.

𝑛𝑐𝑡 𝜉 Ϛ
Q Q
r = sign [1 − ( ̂ S𝜔 [k 0 + ∑ 𝐾𝑖 𝑎𝑖𝜔𝑖 ] |1 − (
)] A ) | (3.22)
K eq K eq
𝑖=1

𝐴̂0 𝑁𝑚
𝐴̂ = (3.23)
𝑁𝑚0

𝐸𝑎 1 1
𝑘0 = 𝑘0∗ exp [− ( − )] (3.24)
𝑅 𝑇 𝑇∗

Where:

• 𝑆𝜔 = water saturation

• ̂ = reactive surface area at the current time step


A

• ̂ 0 = initial reactive surface area


A

• 𝑁𝑚 = number of moles of the mineral at the current time step

• 𝑁𝑚 = initial number of moles of the mineral

• 𝑘0 = rate constant at current temperature T

• 𝑘0∗ = rate constant at reference temperature T*

• 𝐸𝑎 = reaction activation energy

• 𝑅 = gas constant

74
Wolery and Phreeqc databases could be used to obtain reactions’ 𝐸𝑎 , 𝐴̂, 𝑘0∗ and 𝐾𝑒𝑞 ,

and the former was implemented in this work.

3.2: APPROACH

The model described in the previous section has been applied for the analysis. To

ensure the accuracy of the model performance, a simulation model that was history-

matched on an actual CO2-EOR field performance was chosen. However, the history-

matched model originally did not incorporate relative permeability hysteresis,

geochemistry, phase trapping, or thermal interaction. Since this evaluation aims to assess

EOR and CO2 sequestration, the model was updated to include these mechanisms.

Afterward, to ensure the model’s integrity is preserved, it was compared against the

original history-matched version. Figure 3.9 illustrates that the new model is still in good

agreement with the original case, indicating its continuing reliability.

Figure 3.9: Original History Matched Case Vs. Hysteresis Model.

75
Afterward, all the wells in the original case were replaced by one set of 5-spot

injection pattern, as shown in Figure 3.10. The wells were replaced to simplify the problem

and to maximize the distance between the injector and producers, delaying gas

breakthrough. Furthermore, the original wells were not operational simultaneously but

were put on production on varying years, as incremental and maintenance wells with

operational downtime. The producers of the 5-spot injection introduced were deliberately

not placed at the edges of the model to avoid any numerical complications. The new wells

(producers and injector) have the same properties as the original history-matched case: all

the wells are stimulated, vertical, and cased-hole completion. Each new well’s performance

and skin were based on the average of the original wells within its vicinity.

Figure 3.10: Reservoir Model and Five-Spot Well Pattern (Producers and Injectors)
Location.

76
The different analyses conducted are:

1) Impact of CO2 composition on oil RF and CO2 retention

2) CO2 injection in a depleted oil reservoir for EOR and sequestration

3) CO2 injection in a partially depleted oil reservoir for EOR and sequestration. Both

continuous CO2 injection and water alternating gas (WAG) are studied.

Each scenario is elaborated on in the following sections.

3.2.1: Injection Gas Composition Analysis

This evaluation was conducted to optimize the depleted and partially depleted

analysis. The analysis aims to find the desired composition of the carbon dioxide feed

injected, improving both EOR and CO2 sequestration. Usually, the injected CO2 feed is not

pure but contains some impurities such as N2, C1, C2, and H2S. The purity of CO2 relies on

the capture technology or the source of CO2 (Islam et al., 2022; Lake et al., 2019). For

instance, McElmo and Bravo domes, natural sources of CO2, provide a highly pure carbon

dioxide (> 99 mol%) (Lake et al., 2019). Additionally, based on Li and Yan's (2009)

investigation of EOS performance in modeling binary mixture’s vapor-liquid equilibrium

of CO2 and common impurities, it is found that the EOS predictions deviate significantly

from experimental data once the impurities exceed 30 mol% of the gas. Therefore, in this

analysis, impurities are limited to not exceed 30 mol% of the injected gas.

The analysis is conducted as miscible injection since miscible injection triumphs

immiscible in terms of EOR (Chen et al., 2022; Chen et al., 2018; Lake et al., 2014; Fath

and Pouranfard, 2014). Moreover, the carbon dioxide would mix with the oil phase,

impacting its properties under miscible conditions. MMP is estimated to be 2000 psi

experimentally, which is supported by Sebastian et al. correlation (1985) and Sinha et al.

(2021) work. Furthermore, MMP estimate by GEM is in agreement with the estimate, as

77
aforementioned. To further ensure miscibility, the reservoir pressure was kept above the

MMP for the whole duration of the simulation.

The simulation lasts 10 years of continuous miscible gas injection, where the

injection rate is maintained at a constant value for a fair comparison. The model

investigated various compositions of the injected gas consisting of CO2, N2, C1, C2, and

H2S. In this analysis, all possible combinations were tested for Single-impurity, dual-

impurity, and triple-impurity, as stated in Table 3.6. The study seeks to understand the

influence of each individual impurity, and which one dominates when more than one

impurity exists. Since the main objective of this evaluation is to observe the impact on oil

recovery and sequestration, the model is isothermal and does not account for mineralization

(geochemistry), however all other assumptions discussed in Chapter 3.1.2 are

implemented.

78
Table 3.6: Different Injection Gas Composition Cases.

Case Composition
%Impurity 1 %Impurity 2 %Impurity 3 %CO2
1 99
10 90
Single Impurity
20 80
30 70
5 5 90
15 5 80
25 5 70
Dual Impurity 1 1 98
5 5 90
10 10 80
15 15 70
10 5 5 90
15 5 5 80
20 5 5 70
Triple Impurity
1 1 1 97
5 5 5 85
10 10 10 70

3.2.2: Depleted Reservoir Analysis

Depleted reservoirs are evaluated in this work due to their potential to contribute to

the atmosphere's decarbonization. A depleted reservoir, in this context, is a reservoir that

has reached its abandonment stage with no significant production potential. The oil and gas

industry has an incentive to pursue sequestration projects in depleted reservoirs rather than

going through complex and expensive exit strategies. Turning depleted fields into a carbon

dioxide sink would improve the carbon footprint of the industry, making oil and gas a

greener primary energy source. The extended injection of CO2 would pressurize the

79
reservoir and require pressure relief wells (i.e., producers) to prevent the reservoir from

over pressurizing. leading to additional hydrocarbon recovery at a negligible carbon

footprint.

To purely observe the magnitude of the gas injection on the reservoir, the reservoir

was initialized at residual oil saturation. By initializing the model at residual oil saturation,

the reservoir is assumed to be fully swept, netting a 78.5% RF. The evaluation was

conducted as miscible and immiscible injection by changing the initialization pressure of

the model, however the pressure differential was fixed at 500 psi between the initialization

pressure and the minimum bottomhole pressure (BHP) constraint on the producers. This

fixation of differential pressure was done to isolate the effect of the injection on production

and not to introduce another variable into the study. The models are initialized at 2500 and

1500 psi for miscible and immiscible cases, respectively. The production wells are limited

only by the minimum BHP of the producers, for proper pressure relief. In addition, the

producers are prevented from reaching the frac pressure of 4200 psi by limiting the

injection BHP to 4100 psi.

The results of the scenarios investigated are compared based on the incremental oil

RF produced via the relief wells (producers) and the retention efficiency of carbon dioxide.

The retention efficiency takes into consideration the in-situ CO2 originally present in the

reservoir as well as the injected gas and is calculated as expressed by eqn. 3.25. Moreover,

the incremental oil RF is based on the original volume present in the model, listed in Table

3.1.

𝑃𝑟𝑜𝑑𝑢𝑐𝑒𝑑 𝐶𝑂2
𝐶𝑂2 𝑅𝑒𝑡𝑒𝑛𝑡𝑖𝑜𝑛 𝐸𝑓𝑓𝑒𝑐𝑖𝑒𝑛𝑐𝑦 = (1 − ) ∗ 100 (3.25)
𝐼𝑛𝑗𝑒𝑐𝑡𝑒𝑑 𝐶02

80
For any simulation to be accepted, it needs to meet certain criteria. The criteria are

put in place to ensure a fair and realistic comparison. The criteria are:

1) Reservoir Pressure must not reach 4100 psi, to avoid inducing fractures.

2) Injection Rate must be constant for the entirety of the simulation duration.

3) Low material balance to avoid inaccurate results (<0.001%).

Due to the complex grid system of the model (e.g., pinch-outs, discontinuities,

multiple rock types, high heterogeneity, etc.), thermal modeling adds much complexity to

the numerical simulation and makes each run computationally expensive. Nevertheless,

few thermal models have been simulated to determine the importance and need for non-

isothermal modeling. The gas stream injected in this study refers to the optimum

composition from the previous section’s results (90% CO2, 5% N2 and C1, 5% C2).

3.2.3: Partially Depleted Reservoir Analysis

Partially depleted oil reservoirs offer the oil and gas industry opportunities to

enhance oil production and to lower its carbon footprint. Partially depleted oil reservoirs

go into tertiary production stage once primary and secondary stages are not capable of

producing more oil, while leaving oil behind. As discussed previously, one of the most

common EOR methods currently is solvent EOR, specifically CO2-EOR. The petroleum

industry has been attempting to increase the efficiency of carbon dioxide injection to

produce as much oil as possible with the least possible amount of carbon dioxide, for

economical purposes. However, as part of the decarbonization efforts, the volume of

injected carbon dioxide is much larger. Therefore, it is essential to examine the impact of

the increased volumes injected to avoid releasing CO2 back into the atmosphere through

producers.

81
Carbon dioxide injection in this analysis is conducted as continuous gas injection,

like the depleted reservoir case, and as water-alternating-gas (WAG). Generally, WAG has

a better performance compared to gas flooding, as the water impedes the gas, negatively

impacting its mobility (Larsen and Skauge, 1998). Hence, gas breakthrough is delayed, and

the sweep is enhanced, as illustrated by Figure 3.11. Consequently, both forms of injection

are studied to compare their performance in terms of trapping/retention of CO2 in the

subsurface.

Figure 3.11: Gravity Effect in Gas Flood, Waterflood, and WAG (Samba and Elsharafi,
2018).

In preparation for the analysis, a base case simulation model was established. The

base case underwent primary production, ending with an oil RF of 10%. During the primary

production phase, the producers were constrained to not exceed an oil rate of 150 bbls/d,

GOR of 1500 SCF/STB, WC of 95%, and minimum BHP of 1500 psi. The oil rate

constraint was based on the history-matched case well performance, while GOR, WC, and

BHP were designed to enable the reservoir to reach 10% RF without allowing a significant
82
volume of gas to evolve from the oil phase. At the end of the primary production stage, oil

production rate starts declining and a waterflood commences. The waterflood is maintained

at a voidage replacement ratio (VRR) of 1.2 to pressurize the reservoir and redissolve the

evolved gas back into the oil phase. Once the average reservoir pressure reached 2500 psi,

VRR was switched to 1.0, to maintain the pressure above MMP by 500 psi, conforming

with the depleted case, in preparation for the upcoming tertiary stage. The secondary stage

was kept operational until it could not produce oil, with the producer operating under the

same constraints as the primary stage. The base case performance is demonstrated in

Figure 3.12, and based on it the tertiary stage was set to commence in 2055, when oil RF

starts to plateau at 51.4%. The extended waterflood was meant to produce as oil much as

possible without EOR, to observe the gas flood enhancement. The tertiary stage utilizes

the gas composition recommended by the results of gas composition analysis (90% CO2,

5% N2 and C1, 5% C2).

Figure 3.12: Base Case Primary and Secondary Production Stages. Average Reservoir
Pressure, Gas Oil ratio (GOR), Oil RF, and Water Cut (WC).

83
The tertiary stage analysis compares continuous gas injection and WAG in terms

of EOR and sequestration. The producers are based on the history-matched case for both

studies, as it was for the primary and secondary production. However, the minimum BHP

was increased to 2000 psi to ensure miscibility, and the WC constraint was increased to

99%. In terms of injection, multiple scenarios were investigated as long as it did not lead

to over-pressurizing the reservoir to the frac pressure.

In WAG, various combinations of injection rate, wag ratio, and cycle period (slug

size) are conducted to study their influence on EOR and sequestration. To ensure the WAG

ratio is accurate, the fluid rates are specified at bottom hole conditions. In WAG, the water

volume injected could become problematic, due to the limitations imposed on the

producers. If the water volume introduced into the reservoir is too high, it will increase the

WC of the producing wells, causing them to shut-in once it reaches 99%; the shut-in period

lasts for a year in an attempt to reduce the WC. However, as WAG persists, more water

will be introduced to the model, leading to oversaturating and pressurizing the model. The

limited size of the model used is partly responsible for the rapid oversaturation. Therefore,

low injection rates are used in WAG, listed in Table 3.7, relative to the abandoned reservoir

scenario. As a result of the slow injection rate, some simulation cases were not able to

achieve a 100% HCPV of fluid injected. Hence, WAG was run for 50 years to enable all

cases to achieve a 30% HCPV of fluids injected, based on Lake et al. (2019) survey of

CO2-EOR and sequestration projects. The 30% HCPV injected is not only gas, but includes

water as well, since the PV has been utilized.

84
Table 3.7: WAG Analysis Sensitivity Parameters.

Cycle Duration, WAG Ratio, Gas Injection Rate, Gas Injection

yr w:g rbbl/day Rate, MMSCFD

0.25 1:4 1000 1.5

0.5 1:3 2000 3

1 1:2 3000 4.5

2 1:1 4000 6.2

- 2:1 5000 7.8

- 3:1 6000 9.5

- 4:1 8000 13.8

In regard to continuous gas flood, the injection is miscible. Furthermore, since the

production wells are based on the history-matched case, the injection rates are less than

that of the depleted reservoir scenario to avoid reaching frac pressure. The maximum

injection rate that would not result in over pressurizing the reservoir is 10 MMSCFD.

Therefore, the rates utilized in this study match those of the WAG analysis in Table 3.7 at

surface conditions, for a fair comparison. Similarly, the simulations are also carried out up

to 30% HCPV, the cutoff used for WAG study. Furthermore, as the injection is continuous
at a constant rate in gas flood, unlike WAG, the gas flood cases were able to reach 30%

HCPV injected much faster (13 – 30 years).

All the cases were compared based on CO2 retention efficiency and EOR. The

retention efficiency takes into consideration the in-situ CO2 originally found in the

reservoir, as well as the injected gas as described by eqn. 3.25. In terms of EOR, the cases

are compared based on the incremental oil RF produced. This analysis forgoes

85
geochemistry and thermal modeling. Thus, mineralization and intra-aqueous reactions are

disabled due to the long simulation times, amplified by the model’s complexity, and

numerical issues we faced. Consequently, the carbon dioxide that is supposed to

mineralize, will contribute to the other mechanisms, inflating them. Nevertheless,

mineralization is not a major contributor in the short-term (NASEM, 2019).

86
Chapter 4: Results and Discussions

In this chapter, the results of the analyses are displayed and discussed. The results

are presented based on each scenario described in the methodology Chapter 3.2.

4.1: COMPOSITION OF INJECTED CO2

In this analysis, the impact of the composition of the injected gas on oil RF and

carbon dioxide sequestration is explored. The evaluation assesses all possible combinations

and concentrations of an injected gas consisting of CO2, N2, C1, C2, and H2S, while limiting

the impurities not to exceed 30 mol% of the injected stream.

4.1.1: Single Impurity

This section investigates the impact of single impurity on CO2-EOR and carbon

dioxide sequestration performance. A total of 12 numerical experiments were conducted

to understand how the impurity affects the performance relative to a 100% CO2 stream.

The experiments carried out are listed in Table 4.1.

Table 4.1: Single Impurity Cases.

Case Injected CO2 Composition

%Impurity %CO2

1 99

10 90
Single Impurity
20 80

30 70

87
Figure 4.1 highlights the performance of the different gas compositions in terms of

carbon dioxide storage efficiency. It is observed that the presence of any of the impurities

had a negative impact on the retention of carbon dioxide. Hydrogen sulfide (H2S) had the

slightest influence with a negligible effect, which is not surprising, as H 2S is known to

enhance the miscibility of CO2 in oil (Chen et al., 2018; Trivedi et al., 2007; Lake, 1989).

C2 also has a minor impact but is slightly more negative than H2S. Lastly, N2 + C1 are found

to be detrimental to carbon dioxide retention.

Figure 4.1: Single Impurity Impact on Carbon Dioxide Retention Compared to A Pure
Stream.

Regarding oil RF and CO2-EOR, Figure 4.2 illustrates the performance of the

different gas compositions. Both N2 + C1 and C2 had a positive impact, increasing the oil

RF above that of a pure CO2 stream, with N2 + C1 having the largest increment. On the

other hand, H2S was the only impurity to negatively impact the oil RF.

88
Figure 4.2: Single Impurity Impact on Oil RF Compared to A Pure Stream.

4.1.2: Dual Impurity

This section investigates the impact of dual impurity on the performance of CO2-

EOR and carbon dioxide sequestration. A total of 36 numerical experiments were

conducted to understand how the contaminants affect the performance, relative to a 100%

CO2 stream. In addition, this section seeks to understand the impact of the impurities, when

two co-exist, if they would negate or amplify each other’s effect, and if one contaminant
would dominate. Therefore, the numerical simulations evaluated various CO2

compositions with equal and unequal concentrations, as listed in Table 4.2.

89
Table 4.2: Dual Impurity Cases.

Case Injected CO2 Composition


%Impurity 1 %Impurity 2 %CO2
5 5 90
15 5 80
25 5 70
Dual Impurity
5 5 90
10 10 80
15 15 70

Figures 4.3, 4.4, and 4.5 showcase the results of the analysis of dual impurity on

the volumes of carbon dioxide sequestered compared to a pure CO2 stream. Consistent with

the previous section (single impurity), all the cases exhibit a decline in the amount of

carbon dioxide sequestered. Yet, the magnitude of decrease differs based on the contents

of the feed gas. In all scenarios where N2 + C1 impurity is present, the case exhibits the

lowest volume of carbon dioxide sequestered. However, the presence of other impurities,

along with N2 + C1, decreased its negative effect on carbon dioxide retention, compared to

the single impurity analysis. Furthermore, N2 + C1 are found to be the least dominant when

it comes to carbon sequestration, while H2S is the most dominant. This is observed in the
cases where N2 + C1 exist in equal concentration with another impurity; the performance

leans toward the case where the other impurity is present at higher mol%. On the other

hand, H2S is the least damaging to CO2 retention, similar to the single impurity analysis,

with minor reduction due to the presence of other impurities in the stream. In the cases

where H2S and C2 co-exist, the performance is similar, with H2S being slightly better.

Generally, the impact of the impurities on CO2 retention could be summarized as follows:

90
1) H2S is the least damaging to CO2 retention, and its effect dominates over other

impurities.

2) C2 decreases CO2 retention slightly more than H2S (~0.25%), and its effect

dominates over N2 + C1 but not quite as much as H2S.

3) N2 + C1 decrease carbon dioxide retention the most and have the least dominant

effect.

Figure 4.3: Dual Impurity (N2 + C1 and H2S) Impact on Carbon Dioxide Retention
Compared to A Pure Stream. (NH-N) Increasing concentration of N2 + C1.
(NH-H) Increasing concentration of H2S. (NH-E) Impurities have equal
concentrations.

91
Figure 4.4: Dual Impurity (N2 + C1 and C2) Impact on Carbon Dioxide Retention
Compared to A Pure Stream. (NC-N) Increasing concentration of N2 + C1.
(NC-C) Increasing concentration of C2. (NC-E) Impurities have equal
concentrations.

Figure 4.5: Dual Impurity (H2S and C2) Impact on Carbon Dioxide Retention Compared
to A Pure Stream. (HC-H) Increasing concentration of H2S. (HC-C)
Increasing concentration of C2. (HC-E) Impurities have equal
concentrations.

92
Figures 4.6, 4.7, and 4.8 demonstrate the analysis results on CO2-EOR, compared

to a pure CO2 stream. Like the single impurity section, N2 + C1, and C2 drive the oil RF

higher than that of a pure stream, while H2S decreases it. Hence, the best combination

consists of N2 + C1, and C2, adding an additional 10% to the oil RF. When comparing the

cases wherein the impurities exist in equal mol%, it is found that the most dominant

impurities are C2, H2S, and N2 + C1, respectively. However, N2 + C1 are not as

overwhelmed by the other impurity as in carbon dioxide retention.

Figure 4.6: Dual Impurity (N2 + C1 and H2S) Impact on Oil RF Compared to A Pure
Stream. (NH-N) Increasing concentration of N2 + C1. (NH-H) Increasing
concentration of H2S. (NH-E) Impurities have equal concentrations.

93
Figure 4.7: Dual Impurity (N2 + C1 and C2) Impact on Oil RF Compared to A Pure
Stream. (NC-N) Increasing concentration of N2 + C1. (NC-C) Increasing
concentration of C2. (NC-E) Impurities have equal concentrations.

Figure 4.8: Dual Impurity (H2S and C2) Impact on Oil RF Compared to A Pure Stream.
(HC-H) Increasing concentration of H2S. (HC-C) Increasing concentration
of C2. (HC-E) Impurities have equal concentrations.

94
4.1.3: Triple Impurity

This section investigates the impact of triple impurity on the performance of CO2-

EOR and carbon dioxide sequestration. A total of 12 numerical experiments were

conducted to understand how the contaminants affect the performance, relative to a 100%

CO2 stream. This section explores the behavior when all the impurities are present in the

injected stream, to evaluate the impact on the results and compare with the dual impurity

cases. The experiments conducted in this investigation are listed in Table 4.3.

Table 4.3: Triple Impurity Cases.

Case Injected CO2 Composition


%Impurity 1 %Impurity 2 %Impurity 3 %CO2
10 5 5 90
15 5 5 80
20 5 5 70
Triple Impurity
1 1 1 97
5 5 5 85
10 10 10 70

Figures 4.9 and 4.10 contain the analysis results of triple impurity effect on carbon

retention. The results in Figure 4.9, where one impurity has higher concentrations than the

rest, are in line with the dual impurity analysis. The least reduction in carbon retention

comes from cases rich in H2S, C2, and N2 + C1, respectively. In the cases where all

impurities are present at equal concentrations, Figure 4.10, the stream seems to result in

outcomes closest to the scenario in which the gas stream is C2 rich, as can be inferred from

Figures 4.9 and 4.10 at 70 mol%.

95
Figure 4.9: Triple Impurity (H2S, N2 + C1 and C2) Impact on Carbon Dioxide Retention
Compared to A Pure Stream. (C) Increasing concentration of C2. (H)
Increasing concentration of H2S. (N) Increasing concentration of N2 and C1.

Figure 4.10: Triple Impurity (H2S, N2 + C1 and C2) Impact on Carbon Dioxide Retention
Compared to A Pure Stream. (E) All impurities exist at equal
concentrations.

96
Figures 4.11 and 4.12 illustrate the results of the analysis of triple impurity effect

on CO2-EOR. Once more, the results in Figure 4.11 are consistent with the dual impurity

analysis. Increasing the mol% of N2 + C1 and C2 have a positive effect on the oil recovery,

with N2 + C1 offering the largest increment. On the other hand, H2S reduces oil recovery

as its concentration increases. When it comes to the cases of impurities in equal

concentrations, Figure 4.12, the stream results in a worse oil RF than the cases in Figure

4.11 due to the presence of H2S. However, as the impurities’ concentrations increase, the

negative influence is lessened as N2 + C1 and C2 counteract the effect of H2S.

Figure 4.11: Triple Impurity (H2S, N2 + C1 and C2) Impact on Oil RF Compared to A
Pure Stream. (C) Increasing concentration of C2. (H) Increasing
concentration of H2S. (N) Increasing concentration of N2 + C1.

97
Figure 4.12: Triple Impurity (H2S, N2 and C1 + C2) Impact on Oil RF Compared to A
Pure Stream. (E) All impurities exist at equal concentrations.

Based on the results of this section and the previous ones, it is consistently found

that:

1) All impurities negatively affect carbon retention, with the least impactful being

H2S, C2, and N2 + C1, respectively. Furthermore, the influence of H2S is the

strongest in terms of carbon dioxide retention.

2) N2 + C1 and C2 promote oil RF, while H2S is detrimental to it, with N2 + C1 offering
the largest increase.

Combining C2 and N2 + C1 results in the best mixture, enhancing EOR without

compromising carbon retention. Therefore, the recommended gas composition is 90 mol%

CO2, 5 mol% C2, and 5 mol% N2 + C1. The purity of the stream is 90 mol% due to the

deterioration imposed on carbon dioxide retention in return for a minor increase in oil RF.

In addition, a highly pure stream (≥ 99 mol%) is not recommended because it does not

significantly benefit sequestration compared to a 90% blend, while resulting in a lower RF.

98
Furthermore, it could be economically expensive to provide a highly pure anthropogenic

CO2 (Islam et al., 2022, IEA, 2020; Mohammad et al., 2020).

4.2: DEPLETED RESERVOIR ANALYSIS

Analysis of a reservoir that is no longer able to produce and has reached the

abandonment stage of its life cycle, is presented in this section. The reservoir is at its

residual oil saturation of 16.35%. The evaluation investigates the implications of

miscibility and injection rate on CO2-EOR and storage under a continuous gas flood.

Normally, this type of injection would be considered a CCS project, however the pressure

relief wells utilized end up producing oil, so it could be regarded as CCUS, as well. The

simulation cases presented are mainly isothermal, with a few thermal cases, and account

for geochemistry. Furthermore, any breakdown of carbon dioxide inventory within the

reservoir is not accurate. Rather the reported values are an approximation as CMG’s Result

module faces some issues reporting it in three-phase oil/CO2/water systems; nonetheless,

this does not affect the simulation accuracy (CMG support, personal communication,

2023).

4.2.1: Miscible Injections

This section evaluates miscible injection of CO2 for CO2-EOR and sequestration at

different operating conditions. Figure 4.13 shows the performance, in terms of EOR (RF)

and Storage (RT), during a 30-year gas flood. The results indicate that a higher injection

rate is preferential for EOR, increasing the oil RF by 700% from the lowest injection rate

(5 MMSCFD) to the highest (30 MMSCFD). Conversely, higher injection rates accelerated

gas breakthroughs, reducing CO2 retention in the subsurface. Nevertheless, higher injection

rate cases sequestered larger volumes of carbon dioxide at a quicker pace but at lower

99
retention efficiency. Thus, the capability to capture CO2 from produced reservoir fluids for

recycling becomes more crucial as the injection rates rise.

Figure 4.13: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection Rates
for Isothermal Miscible Injection Conditions. (a) RT on the left axis and RF
on the right axis (b) CO2 Retention in Mt.

Evaluating the results through a different paradigm, all the simulations were

assessed at the same HCPV injection for a fair comparison, as is typically done in CO2-

EOR projects. The simulations presented in Figure 4.13 injected 61% - 344% HCPV of

carbon dioxide, which could skew the comparisons in favor of the higher injection

volumes. To establish a reasonable cutoff for HCPV injected, a survey by Lake et al. (2019)

100
of CO2-EOR and sequestration projects was utilized. The survey reports that CO2-EOR

projects inject 25% - 60% HCPV, with a median value of 30%. These values could be

attributed to economics, equipment integrity, and availability of CO2, among other reasons.

Hence, all the simulation cases were compared up to 100% HCPV, Figure 4.14, for the

analysis to cover the entire range and be more comprehensive. By doing so, the period of

gas floods ranged from 8 - 48 years until 100% HCPV injection was attained.

Figure 4.14: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection Rates
for Isothermal Miscible Injection Conditions. (a) RT on the left axis and RF
on the right axis (b) CO2 Retention in Mt.

The results from Figure 4.14 agree with the temporal evaluation in Figure 4.13,

where the oil RF is enhanced by increased injection rates. At the same time, Figure 4.14

101
contradicts the temporal analysis, indicating that the higher injection rates are more

beneficial for the retention efficiency of CO2, as well as the cumulative volume stored.

Additionally, storage is achieved in a shorter period, where the 5 MMSCFD case took 48

years to reach 100% HCPV, while the 30 MMSCFD case achieved it in 8 years. This

outcome is justified by: (1) increased solubility of CO2 in reservoir fluids due to increased

pressure, (2) faster mineralization due to the plume contacting a larger part of the reservoir

and the abundance of CO2, (3) higher driving force of the injectant, (4) diminishing phase

trapping due to the increased pressure. An issue that needs to be addressed is the surface

equipment footprint required to compress and inject the gas, to avoid producing more CO2

than is disposed of. Table 4.4 shows the forms carbon dioxide is stored within the reservoir.

All the categories, in which CO2 exists in, increase with higher injection rate but for the

Trapped CO2. This can be attributed to the elevated pressure reducing the trapped volume.

Table 4.4: Carbon Dioxide Inventory at Different Injection Rates in Miscible Flooding
at 100% HCPV Injected.

Carbon Dioxide Inventory, MtCO2


Case
Dissolved Mineralized Trapped Structural Total

5 MMSCFD 94,861 1,833 50,539 1,137,451 1,284,684

15 MMSCFD 125,077 1990 113,713 1,401,327 1,642,107

30 MMSCFD 153,421 2247 84,213 1,754,307 1,994,188

The result of the comparisons concludes that higher injection rates are beneficial

for both EOR and storage efforts under miscible conditions, transforming reservoirs

reaching their abandonment stage into an opportunity that could be capitalized on for

decarbonization and oil production rather than a liability.

102
4.2.2: Immiscible Injection

It is well established that immiscible gas injection is not efficient and

underperforms in comparison to miscible gas injection (Chen et al., 2018; Fath and

Pouranfard, 2014; Lake et al., 2014). Nevertheless, immiscible injection, in regard to

carbon dioxide sequestration, is still to be fully understood. Thus, this section evaluates

immiscible injection on CO2-EOR and sequestration of CO2 at different operating

conditions. Figure 4.15 shows EOR (RF) and Storage (RT) performance during a gas

flood. Similar to miscible flooding, this evaluation is based on HCPV injected rather than

a temporal comparison. The results highlight that the EOR is much worse than in the

miscible case, as anticipated, with the highest additional RF being 0.26% compared to

2.68% in the miscible flood. However, the carbon dioxide retention efficiency is less

sensitive to the injection rate, unlike miscible injection. Consequently, immiscible flooding

results in higher retention efficiency, even at lower injection rates, than in miscible

injection. In terms of the volumes of CO2 stored, higher injection rates result in larger

volumes sequestered. Compared to the miscible case, immiscible flooding has smaller

amounts of CO2 stored in MtCO2, due to the lower pressure of the reservoir, allowing a

lesser amount of carbon dioxide to fill the pore volume (lower compression of CO2). Table

4.5 illustrates the forms carbon dioxide exists in the model at the 100% HCPV injected.

All the categories, in which CO2 exists in, increase with higher injection, unlike miscible

flooding, due to the much lower pressure of the reservoir leading to the entrapment of

carbon dioxide.

103
Figure 4.15: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection Rates
for Isothermal Immiscible Injection Conditions. (a) RT on the left axis and
RF on the right axis (b) CO2 Retention in Mt.

Table 4.5: Carbon Dioxide Inventory at Different Injection Rates in Immiscible


Flooding at 100% HCPV Injected.

Carbon Dioxide Inventory, MtCO2


Case
Dissolved Mineralized Trapped Structural Total

5 MMSCFD 92,305 1,967 379,030 281,261 754,563

15 MMSCFD 122,440 2168 495,557 382,896 1,003,061

30 MMSCFD 133,109 2180 570,052 455,608 1,160,949


104
Another level of confirmation is needed to ensure that the results reported above

are accurate. Figure 4.16 shows the miscible flood performance at the same amount of

CO2 injected in the immiscible case to achieve 100% HCPV. Now, it is clear that miscible

flooding is, in fact, better than immiscible flooding in every aspect, where immiscible flood

averages 50% CO2 retention and an oil RF below 0.3%.

Figure 4.16: Oil RF (RF) and Carbon Dioxide Retention (RT) at Different Injection Rates
Under Miscible Conditions. Marked at the Injected Amounts (MtCO2) of
Carbon Dioxide by its Counter Immiscible Flooding. (a) RT on the left axis
and RF on the right axis (b) CO2 Retention in Mt.

105
4.2.3: Miscible Thermal Modeling

As per the previous sections, miscible flooding has been established to be superior

to immiscible flood, therefore this section works under miscible conditions. The

comparison between isothermal and non-isothermal models is conducted at the extreme

cases (5 and 30 MMSCFD) to amplify any possible deviations and examine the reliability

of isothermal modeling. Figure 4.17 compares both cases, where the injected CO2 stream

temperature is equal to the reservoir temperature of 130 °F. This is done for a fair

comparison between both models, since the isothermal model operates at the same

reservoir and injected temperature of 130 °F. The results show that isothermal modeling

results in better RF and CO2 retention than non-isothermal modeling, with the deviation

increasing at the higher rate. Nevertheless, both models gave comparable oil recovery

factor and carbon retention with minor differences of ~0.25% in RF and ~1% in RT, which

is reasonable because fluid properties are the same, since the pressure and temperature are

similar. The negligible differences between the models do not justify the added

computational expense associated with non-isothermal modeling, when the injectant fluid

temperature is at the same or close to the reservoir temperature.

106
Figure 4.17: Miscible Isothermal model (I) vs. Miscible non-Isothermal model (T) in
Terms of Oil RF (RF) and Carbon Dioxide Retention (RT) at different
injection rates. (a) RT on the left axis and RF on the right axis (b) CO2
Retention in Mt.

Furthermore, the injection rate of 30 MMSCFD case was simulated with different
injection temperatures. Figure 4.18 showcases the impact on carbon dioxide retention and

oil RF. The results demonstrate an improved oil RF as the injected gas temperature is

lowered, however at the expense of lower carbon dioxide retention. The increase in oil RF

at lower temperatures is expected, reducing MMP and enhancing the mixing of carbon

dioxide with oil and its solubility in water (Lake, 1989; Stalkup, 1978). Although the

107
solubility of CO2 is enhanced, the improved production results in producing bigger

volumes of CO2 as well.

Figure 4.18: Impact of Different Injection Temperature on Oil RF (RF) and CO2
Retention (RT) in Miscible non-Isothermal model (T). RT on the left axis
and RF on the right axis. The injectant temperature is listed by the case
name in °F.

4.3: TERTIARY PRODUCTION ANALYSIS

In this section, the tertiary production stage is studied. The evaluation compares the

performance of CO2-EOR and storage under a continuous gas flood and WAG. The models

presented are isothermal, with no geochemistry modeled, due to numerical issues we faced.

4.3.1: Water-Alternating-Gas (WAG)

This section looks at the performance of WAG under different operational

strategies. The analysis explored the influence of different injection rates, WAG ratios, and

cycle duration (slug size) on CO2 EOR and storage. A total of 84 cases are conducted to

cover all possible permutations of the parameters in Table 3.7.


108
Table 3.7: WAG Analysis Sensitivity Parameters.

Gas Injection Rate, Gas Injection


Cycle Duration, yr WAG Ratio, g:w
rbbl/day Rate, MMSCFD

0.25 1:4 1000 1.5

0.5 1:3 2000 3

1 1:2 3000 4.5

2 1:1 4000 6.2

2:1 5000 7.8

3:1 6000 9.5

4:1 8000 13.8

Figures 4.19 and 4.20 illustrate the influence of cycle duration and WAG ratio on

oil RF and carbon dioxide retention within the reservoir. Cycle duration is found to have

no significant impact on the oil RF, with the largest drop being 0.2% when the cycle time

is less than one year. The same behavior is observed in carbon dioxide retention, Figure

4.20, however the biggest drop in retention occurs when the cycles are less than six months.

On the other hand, the WAG ratio is found to be highly influential, increasing the oil RF

as the gas volume exceeds that of the water. Unfortunately, as the WAG ratio enhances the
oil RF, it reduces carbon dioxide retention. Different injection rates were investigated at

the same WAG ratios to further understand the impacts of WAG ratio. Figures 4.21 and

4.22 highlight the results of this sensitivity, illustrating that higher gas injection rates lead

to improved oil RF and carbon dioxide retention. The results observed from this analysis

are consistent with Braga’s findings (2021) in his investigation of recovery factor in WAG

109
processes. Braga’s work examined various parameters in his research (WAG ratio, cycle

duration, reservoir thickness, injection rate, etc.) concluding the following:

1) cycle duration has the least influence on oil RF.

2) WAG ratio and increased injection rates increase the oil RF.

Based on the results, a WAG ratio of 1:1, 1:2, or 2:1 is recommended, as they

optimize oil RF without harming carbon dioxide retention significantly. Additionally,

based on surface equipment limitations, higher injection rates are advised to promote both

EOR and sequestration efforts. In terms of cycle length, it is not crucial, but it is advised

that it does not fall below six months, to avoid a 3% decline in retention.

110
Figure 4.19: WAG Ratio and Cycle Duration Impact on Oil RF. The number of samples
shown by count along with their average (Avg), minimum (min), maximum
(max), P90, P50 (median) and P10. The x-axis on the top figure indicates
the cycle duration in years, while in the bottom figure, the x-axis indicates
Gas/Water ratio.

111
Figure 4.20: WAG Ratio and Cycle Duration Impact on CO2 Retention. The number of
samples shown by count along with their average (Avg), minimum (min),
maximum (max), P90, P50 (median) and P10. The x-axis on the top figure
indicates the cycle duration in years, while in the bottom figure, the x-axis
indicates Gas/Water ratio.

112
Figure 4.21: Impact of Gas Injection Rate at Constant WAG Ratio on Oil RF. The average (Avg), P90, P50 (median), and P10
are shown for each gas rate at a certain WAG ratio. The first row of the x-axis indicates Gas/Water ratio, and the
second row is the injection gas rate (rbbl/d).

113
Figure 4.22: Impact of Gas Injection Rate at Constant WAG Ratio on CO2 Retention. The average (Avg), P90, P50 (median)
and P10 are shown for each gas rate at a certain WAG ratio. The first row of x-axis indicates Gas/Water ratio, and
the second row is the injection gas rate in (rbbl/d).

114
4.3.2: Continuous Gas Injection

This section investigates miscible gas flood at different injection rates, to

understand the performance in EOR and sequestration. Figure 4.23 compares the

performance of all the models up to 30% HCPV, similar to the WAG simulations. Parallel

to the depleted reservoir scenario’s miscible injection, it is found that higher injection rates

result in improved EOR and sequestration performance.

Figure 4.23: Continuous CO2 Flood Performance in Terms of EOR (RF) and
Sequestration (RT). (a) RT on the left axis and RF on the right axis (b) CO2
Retention in Mt.

115
The results of the WAG simulations are shown by Figures 4.24 and 4.25 based on

the average surface injection rate. Comparing Figures 4.23, 4.24, and 4.25, the additional

oil recovery is observed to be further enhanced by continuous gas flood over WAG, even

when considering the best WAG ratios (0.5, 1, and 2). The only exception is at the injection

rate of 9.5 MMSCFD, where WAG overcomes the continuous gas flood oil RF. Therefore,

a full-field model could add more insights into the effect of higher injection, which is not

possible in the utilized sector model. In terms of carbon dioxide retention, it is observed

that WAG results in higher retention efficiency than the gas flood. The best case in gas

flood (10 MMSCFD) nets an 81% in CO2 retention, while the worst case in WAG reaches

83.8%. Nevertheless, WAG stored volume of CO2 ranges from 500,000 – 965,000 MtCO2,

whereas the continuous gas flood stores a larger volume of CO2, as shown in Figure 4.23,

within the same HCPV injected. Hence, continuous gas injection is recommended if the

field has capability to recycle produced CO2, especially due to a better EOR performance.

116
Figure 4.24: Average Oil RF in WAG Process at Different Gas Injection Rates
(MMSCFD). The number of samples shown by count along with their
average (Avg), P90, P50 (median) and P10. x-axis indicates the gas
injection rate (MMSCFD).

117
Figure 4.25: Average CO2 Retention in WAG Process at Different Gas Injection Rates
(MMSCFD). The number of samples shown by count along with their
average (Avg), P90, P50 (median) and P10. x-axis indicates the gas
injection rate (MMSCFD).

118
Chapter 5: Conclusions and Recommendations

The fight against climate change is of high significance, and action is paramount to

curb future emissions. The release of greenhouse gases like carbon dioxide into the

atmosphere intensifies the greenhouse effect, leading to global warming. Therefore, the

exploration of all avenues bearing the potential to contribute to combating climate change

is a must. To address this pressing issue, the adoption of Carbon Capture, Utilization, and

Storage (CCUS) technologies becomes crucial. The role CCUS plays in the transition to a

low-carbon future is pivotal, due to its sufficient storage capacity and its ability to

significantly reduce carbon emissions from industries and infrastructures.

5.1: CONCLUSIONS AND RECOMMENDATIONS

This analysis investigated different aspects of CCUS in carbonates oil reservoir.

The results of the analyses are summarized as follows:

1) Gas composition analysis:

a. All contaminants experimented with (N2+C1, C2, H2S) were found to negatively

affect CO2 retention. Whereas N2+C1 and C2 presence improves oil recovery,

and H2S decreases it.


b. In gas mixtures consisting of more than one impurity, H2S dominates CO2

retention, while C2 dominates oil recovery.

2) Depleted oil reservoir analysis

a. Miscible gas flood outperforms immiscible gas flood in enhancing both oil

recovery and CO2 retention. Additionally, miscible flood was observed to be

more efficient than immiscible, storing larger amounts of carbon dioxide within

the same PV. Moreover, higher injection rates below the frac gradient are

preferential, improving CO2 retention and oil RF.


119
b. Thermal modeling of the injection highlights improvements in oil production,

but at lower storage, as the injection gas temperature decreases. The impact of

cold CO2 injection in the higher temperature reservoir on geomechanics and

potential fracture initiation was not considered in this research. However, if the

injected fluid is at the same temperature as the reservoir, the outcome of the

simulation is similar, with negligible deviation. Hence, thermal modeling is not

needed when the injected gas is at the reservoir’s temperature.

c. Geochemical reactions and CO2 mineralization contribution to carbon dioxide

trapping is negligible in the short-term and raises the computational complexity

of the simulation. Thus, disregarding geochemistry would not affect the

accuracy of short-term EOR and sequestration project simulations.

3) Partially depleted oil reservoir analysis

a. In WAG injection strategy, the cycle duration (slug size) has negligible effects

on the oil recovery and sequestration efforts, while WAG ratio is highly

influential.

b. Higher injection rates of the gas phase, at the same WAG ratio, improve oil

recovery as well as CO2 retention. However, as the WAG ratio increases in

favor of the gas, the oil recovery is further enhanced, but at reduced retention

compared to a lower WAG ratio.

c. Miscible continuous gas flood shows similar behavior to the depleted reservoir

scenario, where the oil recovery factor and CO2 retention are improved as the

injection is increased.

d. Continuous gas flood resulted in higher oil recovery than WAG, at the expense

of reduced CO2 retention. This result was not expected and needs to be further

explored in a full field model or a larger reservoir model.


120
Based on the analyses conducted in this work, the following is recommended:

1) The optimum gas mixture for CCUS, co-optimizing oil recovery and

sequestration, consists of 5% N2+C1, 5% C2 and 90% CO2.

2) Implementing CCUS in depleted oil fields, reaching abandonment stage of their

lifecycle, as a carbon dioxide sink offers additional oil recovery.

3) Miscible gas flood is superior to immiscible injection in both EOR and

sequestration.

4) Injection of cooler gas to further enhance miscible gas flood performance, while

staying out of hydrate forming conditions and fracture initiation due to thermal

stresses.

5) Thermal simulation is not necessary when the injected gas is close to the

reservoir’s temperature.

6) WAG performance is optimized at WAG ratios of 1:2, 1:1, and 2:1, and cycle

periods longer than 6 months.

5.2: RECOMMENDATIONS FOR FUTURE WORK

The numerical model utilized in this work contains geological and compositional

data of an actual carbonate reservoir considered for a CO2-EOR project. In addition, the

incorporation of various phenomena into the model, such as relative permeability

hysteresis, CO2 solubility in water, and geochemical reactions, offering a more insightful

understanding of the behavior of an actual reservoir. However, future work is needed to

add more depth to the results of this thesis.

One goal for future works is to expand on thermal modeling to analyze hydrate

formation. However, a challenge facing hydrate formation analysis is the incapability of

CMG-GEM to model phase solidifying (e.g., hydrate). Therefore, other methods need to

121
be explored for this purpose. For example, by modeling phase viscosity and density it is

possible to study injectivity, which in turn might be useful to infer hydrate formation.

Another goal for future investigation is to evaluate reservoir geomechanics. In this

analysis, only rock compressibility was incorporated as a function of pressure. However,

further studies should research geomechanics in more depth due to CO2 sequestration,

especially in reservoirs with faults. Moreover, geochemistry is important for such analysis

as it would alter the rock properties via precipitation and dissolution. Furthermore, the

analysis could be utilized to study injectivity behavior, as well.

Finally, a comprehensive analysis of CCUS and CCS economics is of great

significance. If geological sequestration was found to be excessively costly, all technical

evaluations would lose their practicality and purpose. A comprehensive analysis should

account for gas supply, compression, transportation, and injection, as well as wellbore

integrity and a surveillance program of carbon dioxide in the subsurface. This analysis

would assist governments in determining an appropriate subsidy system to encourage

companies to pursue CCS and CCUS.

122
References

Ahmed, U. and Zahid, U. (2019). Techno-Economic Assessment of Future Generation


IGCC Processes with Control on Greenhouse Gas Emissions. Computer Aided
Chemical Engineering, 46, 529-534. Elsevier. https://doi.org/10.1016/B978-0-12-
818634-3.50089-8

Alvarado, V. and Manrique, E. (2010). Enhanced oil recovery: An update review. Energies,
3(9), 1529-1575. https://doi.org/10.3390/en3091529

Ampomah, W., Balch, R. S., Grigg, R. B., Cather, M., Will, R. A., and Lee, S. Y. (2016).
Optimization of CO2-EOR Process in Partially Depleted Oil Reservoirs. Presented
at the SPE Western Regional Meeting, SPE, Anchorage, Alaska, USA, SPE-
180376-MS. https://doi.org/10.2118/180376-MS

Ampomah, W., Balch, R. S., Grigg, R. B., Dai, Z., and Pan, F. (2015). Compositional
Simulation of CO2 Storage Capacity in Depleted Oil Reservoirs. Presented at the
Carbon Management Technology Conference, CMTC, Sugar Land, Texas, CMTC-
439476-MS. https://doi.org/10.7122/439476-MS

Atta, D. Y., Negash, B. M., Yekeen, N., and Habte, A. D. (2021). A State-of-The-Art
Review on The Application of Natural Surfactants in Enhanced Oil Recovery.
Journal of Molecular Liquids, 321, 114888.
https://doi.org/10.1016/j.molliq.2020.114888

Azarabadi, H., and Lackner, K. S. (2020). Postcombustion Capture or Direct Air Capture
in Decarbonizing US Natural Gas Power? Environ. Sci. & Technol., 54(8), 5102–
5111. https://doi.org/10.1021/acs.est.0c00161

Bello, A., Ivanova, A., and Cheremisin, A. (2023). Foam EOR as an Optimization
Technique for Gas EOR: A Comprehensive Review of Laboratory and Field
Implementations. Energies, 16(2), 972. https://doi.org/10.3390/en16020972

Beteta, A., Nurmi, L., Rosati, L., Hanski, S., McIver, K., Sorbie, K., and Toivonen, S.
(2020). Polymer Chemical Structure and its Impact on EOR Performance.
Presented at the SPE Improved Oil Recovery Conference, SPE, Virtual,
D011S009R003. https://doi.org/10.2118/200441-MS
123
Beygi, M. (2016). Development of Compositional Three-phase Relative Permeability and
Hysteresis Models and Their Application to EOR Processes. PhD Dissertation,
University of Texas at Austin.

Beygi, M. R., Delshad, M., Pudugramam, V. S., Pope, G. A., and Wheeler, M. F. (2015).
Novel Three-Phase Compositional Relative Permeability and Three-Phase
Hysteresis Models. SPE J., 20(01), 21–34. https://doi.org/10.2118/165324-PA

Bowker, K. and Shuler, P. J. (1991). Carbon Dioxyde Injection and Resultant Alteration of
the Weber Sandstone, Rangely Field, Colorado. AAPG Bulletin, 75(9), 1489–1499.

Braga, C. T. S. (2021). Impacto Da Segregação Gravitacional Na Recuperação De Óleo


No Caso De Injeção Wag Em Cenário Típico Do Pré-Sal [Influence of Gravity
Segregation on Oil Recovery for Wag Injection in A Typical Pre-Salt Case]. MS
Thesis, Pontifícia Universidade Católica Do Rio De Janeiro.
https://doi.org/10.17771/PUCRio.acad.55925

Cao, C., Liu, H., Hou, Z., Mehmood, F., Liao, J., and Feng, W. (2020). A Review of CO2
Storage in View of Safety and Cost-Effectiveness. Energies, 13(3), 600.
https://doi.org/10.3390/en13030600

Carlson, F. (1981). Simulation of Relative Permeability Hysteresis to the Non-Wetting


Phase. SPE 10157.

Chadwick, A., Williams, G., Delepine, N., Clochard, V., Labat, K., Sturton, S.,
Buddensiek, M.-L., Dillen, M., Nickel, M., Lima, A. L., Arts, R., Neele, F., and
Rossi, G. (2010). Quantitative Analysis of Time-Lapse Seismic Monitoring Data at
The Sleipner CO2 Storage Operation. The Leading Edge, 29(2), 170–177.
https://doi.org/10.1190/1.3304820

Chao, C., Deng, Y., Dewil, R., Baeyens, J., and Fan, X. (2021). Post-Combustion Carbon
Capture. Renewable and Sustainable Energy Reviews, 138, 110490.
https://doi.org/10.1016/j.rser.2020.110490

Chen, P., Kalam, M. Z., Al Kindi, S. A., Abolhag, Y. H., and Shtepani, E. (2018).
Maximize EOR Potential by Optimizing Miscible CO2 Injection Parameters in
Carbonate Reservoirs. Presented at the Offshore Technology Conference, OTC,
124
Houston, Texas, USA, D021S024R002. https://doi.org/10.4043/28640-MS.
https://doi.org/10.4043/28640-MS

Chen, Z., Zhou, Y., and Li, H. (2022). A Review of Phase Behavior Mechanisms of CO2
EOR and Storage in Subsurface Formations. Ind. & Eng. Chem. Res., 61(29),
10298–10318. https://doi.org/10.1021/acs.iecr.2c00204

Christiansen, R. L., and Haines, H. K. (1987). Rapid Measurement of Minimum Miscibility


Pressure with The Rising-Bubble Apparatus. SPE Res. Eng., 2(4), 523–527. SPE-
13114-PA. https://doi.org/10.2118/13114-PA

Computer Modeling Group, (2022). GEM User Manual (version 2022.10). Computer
Modeling Group Ltd

Delshad, M.., Wheeler, M. F., and Kong, X. (2010). A Critical Assessment of CO2 Injection
Strategies in Saline Aquifers. Presented at the SPE Western Regional Meeting,
SPE, Anaheim, California, USA, SPE-132442-MS.
https://doi.org/10.2118/132442-MS

Delshad, M., Delshad, M., Pope, G. A., and Lake, L. W. (1987). Two- and Three-Phase
Relative Permeabilities of Micellar Fluids. SPE Formation Evaluation, 2(3), 327–
337. https://doi.org/10.2118/13581-PA

Doughty, C., and Myer, L. R. (2009). Scoping calculations on leakage of CO2 in geologic
storage: The impact of Overburden Permeability, Phase Trapping, and Dissolution.
In B. J. McPherson and E. T. Sundquist (Eds.), Geophysical Monograph Series,
183, 217–237. American Geophysical Union.
https://doi.org/10.1029/2005GM000343

Economides, M. J. (2013). Petroleum Production Systems (2nd ed.). Prentice Hall.

Eide, Ø., Fernø, M. A., Karpyn, Z., Haugen, Å., and Graue, A. (2013). CO2 Injections for
Enhanced Oil Recovery Visualized with an Industrial CT-scanner. Presented at the
IOR 2013 - 17th European Symposium on Improved Oil Recovery, Saint
Petersburg, Russia. https://doi.org/10.3997/2214-4609.20142641

125
Elhenawy, S. E. M., Khraisheh, M., AlMomani, F., and Walker, G. (2020). Metal-Organic
Frameworks as a Platform for CO2 Capture and Chemical Processes: Adsorption,
Membrane Separation, Catalytic-Conversion, and Electrochemical Reduction of
CO2. Catalysts, 10(11), 1293. https://doi.org/10.3390/catal10111293

Emberley, S., Hutcheon, I., Shevalier, M., Durocher, K., Gunter, W. D., and Perkins, E. H.
(2004). Geochemical Monitoring of Fluid-Rock Interaction and CO2 Storage at The
Weyburn CO2-Injection Enhanced Oil Recovery Site, Saskatchewan, Canada.
Energy, 29(9–10), 1393–1401. https://doi.org/10.1016/j.energy.2004.03.073

Farzaneh, S. A., and Sohrabi, M. (2013). A Review of The Status of Foam Applications in
Enhanced Oil Recovery. Presented at the EAGE Annual Conference & Exhibition
incorporating SPE Europec, SPE, London, UK, SPE-164917-MS.
https://doi.org/10.2118/164917-MS

Fath, A., and Pouranfard, A. (2014). Evaluation of Miscible and Immiscible CO2 Injection
in One of The Iranian Oil Fields. Egyptian Journal of Petroleum, 23(3), 255–270.
https://doi.org/10.1016/j.ejpe.2014.08.002

Fink, J. K. (2021). Petroleum Engineer’s Guide to Oil Field Chemicals and Fluids (3rd
ed.). Elsevier Science & Technology.

Gbadamosi, A., Patil, S., Kamal, M. S., Adewunmi, A. A., Yusuff, A. S., Agi, A., and
Oseh, J. (2022). Application of Polymers for Chemical Enhanced Oil Recovery: A
Review. Polymers, 14(7), 1433. https://doi.org/10.3390/polym14071433

Geng, D., Li, J., Li, H., and Huang, W. (2021). Effects of Particle Combinations with
Different Wettability on Foam Structure and Stability. Front. Mater., 8, 699187.
https://doi.org/10.3389/fmats.2021.699187

Ghomian, Y. (2008). Reservoir Simulation Studies for Coupled CO2 Sequestration and
enhanced Oil Recovery. PhD Dissertation. University of Texas.

Ghulami, M., Sasaki, K., Sugai, Y., and Nguele, R. (2015). Numerical Simulation Study
on Gas Miscibility of an Oil Field Located in Afghanistan. Presented at the 21st
Formation Evaluation Symposium of Japan, 13-14 October, Fukuoka, Japan.

126
Global CCS Institute (GCCSI). (2023). Facilities Database. https://co2re.co/FacilityData

Gunnarsson, I., Aradóttir, E. S., Oelkers, E. H., Clark, D. E., Arnarson, M. Þ., Sigfússon,
B., Snæbjörnsdóttir, S. Ó., Matter, J. M., Stute, M., Júlíusson, B. M., and Gíslason,
S. R. (2018). The Rapid and Cost-Effective Capture and Subsurface Mineral
Storage of Carbon and Sulfur at The Carbfix2 Site. Int. J. Greenhouse Gas Control,
79, 117–126. https://doi.org/10.1016/j.ijggc.2018.08.014

Han, W. S., McPherson, B. J., Lichtner, P. C., and Wang, F. P. (2010). Evaluation of
Trapping Mechanisms in Geologic CO2 Sequestration: Case Study of SACROC
Northern Platform, A 35-Year CO2 Injection Site. American Journal of Science,
310(4), 282–324. https://doi.org/10.2475/04.2010.03

Hanna, R., Abdulla, A., Xu, Y., and Victor, D. G. (2021). Emergency Deployment of Direct
Air Capture as A Response to The Climate Crisis. Nature Communications, 12(1),
368. https://doi.org/10.1038/s41467-020-20437-0

Harvey, A. (1996). Semiempirical Correlation for Henry’s Constants Over Large


Temperature Ranges. AlChE J., 42(5): 1491-1494.
https://doi.org/10.1002/aic.690420531

Hepburn, C., Adlen, E., Beddington, J., Carter, E. A., Fuss, S., Mac Dowell, N., Minx, J.
C., Smith, P., and Williams, C. K. (2019). The Technological and Economic
Prospects for CO2 Utilization and Removal. Nature, 575, 87–97.
https://doi.org/10.1038/s41586-019-1681-6

Hoseini, J., Masoudi, R., Mousavi Mirkalaei, S. M., Ataei, A., and Demiral, B. (2011).
Investigating the Effect of Hysteresis Modelling on Numerical Simulation of
Immiscible WAG Injection. Presented at the International Petroleum Technology
Conference, IPTC, Bangkok, Thailand, IPTC-15055-MS.
https://doi.org/10.2523/IPTC-15055-MS

Hovorka, S., Smyth, R., Romanak, K., Yang, C., Nicot, J., Hardage, B., Sava, D. (2021).
SACROC Research Report. Austin: Bureau of Economic Geology.
https://www.beg.utexas.edu/gccc/research/sacroc

127
IPCC (2023). Climate Change 2023 Synthesis Report.
https://www.ipcc.ch/report/ar6/syr/downloads/report/IPCC_AR6_SYR_FullVolu
me.pdf

IPCC (2021). Climate Change 2021: The Physical Science Basis.


https://www.ipcc.ch/report/ar6/wg1/downloads/report/IPCC_AR6_WGI_FullRep
ort_small.pdf

International Energy Agency (IEA) (2021a). Carbon Capture in 2021: Off and Running or
Another False Start?. https://www.iea.org/commentaries/carbon-capture-in-2021-
off-and-running-or-another-false-start

International Energy Agency (IEA) (2022a). Carbon Capture, Utilisation and Storage.
https://www.iea.org/reports/carbon-capture-utilisation-and-storage-2

International Energy Agency (IEA) (2022b). EOR Production in The New Policies
Scenario, 2000-2040. https://www.iea.org/data-and-statistics/charts/eor-
production-in-the-new-policies-scenario-2000-2040

International Energy Agency (IEA) (2023). Global CO2 Emissions from Energy
Combustion and Industrial Processes, 1900-2022. https://www.iea.org/data-and-
statistics/charts/global-CO2-emissions-from-energy-combustion-and-industrial-
processes-1900-2022

International Energy Agency (IEA) (2022c). Number of EOR Projects in Operation


Globally, 1971-2017. https://www.iea.org/data-and-statistics/charts/number-of-
eor-projects-in-operation-globally-1971-2017

International Energy Agency (IEA) (2021b). The World Has Vast Capacity to Store CO2:
Net Zero Means We’ll Need it. https://www.iea.org/commentaries/the-world-has-
vast-capacity-to-store-co2-net-zero-means-we-ll-need-it

International Energy Agency (IEA) (2020). Timely Advances in Carbon Capture,


Utilisation and Storage. https://www.iea.org/reports/the-role-of-ccus-in-low-
carbon-power-systems/timely-advances-in-carbon-capture-utilisation-and-storage

128
Isdiken, B., (2013). Integrated Geological and Petrophysical Investigation on Carbonate
Rocks of the Middle Early to Late Early Canyon High Frequency Sequence in the
Northern Platform Area of the SACROC Unit. MS Thesis. University of Texas.

Islam, R., Sohel, R. N., and Hasan, F. (2022). Role of Oil and Gas Industry in Meeting
Climate Goals Through Carbon Capture, Storage and Utilization CCUS. Presented
at the SPE Western Regional Meeting, SPE, Bakersfield, California, USA,
D011S005R002. https://doi.org/10.2118/209311-MS

Jarrahian, Kh., Seiedi, O., Sheykhan, M., Sefti, M. V., and Ayatollahi, Sh. (2012).
Wettability Alteration of Carbonate Rocks by Surfactants: A Mechanistic Study.
Colloids and Surfaces A: Physicochem. Eng. Asp., 410, 1–10.
https://doi.org/10.1016/j.colsurfa.2012.06.007

Juanes, R., Spiteri, E. J., Orr, F. M., and Blunt, M. J. (2006). Impact of Relative
Permeability Hysteresis on Geological CO2 Storage. Water Resour. Res., 42(12).
https://doi.org/10.1029/2005WR004806

Kestin, J., Khalifa, H. E. and Correia, R. J. (1981). Tables of the Dynamic and Kinematic
Viscosity of Aqueous NaCl Solutions in the Temperature Range 20-150 C and
Pressure Range 0.1-35 MPa. J. Phys. Chem. Ref. Data, 10, 71-87.

Khabibullin, R., Emadi Ltd, A., Abu Grin, Z., Oskui Ltd, R., Alkan, H., Grivet, M., and
Elgridi, K. (2017). Investigation of CO2 Application for Enhanced Oil Recovery in
a North African Field—A New Approach to EOS Development. Presented at the
IOR 2017 - 19th European Symposium on Improved Oil Recovery, Stavanger,
Norway. https://doi.org/10.3997/2214-4609.201700276

Khan, S. (1965). The flow of Foam Through Porous Media. MS Thesis, Stanford
University.

Kharaka, Y. K., Cole, D. R., Thordsen, J. J., Kakouros, E., and Nance, H. S. (2006). Gas–
Water–Rock Interactions in Sedimentary Basins: CO2 Sequestration in The Frio
Formation, Texas, USA. Journal of Geochemical Exploration, 89(1–3), 183–186.
https://doi.org/10.1016/j.gexplo.2005.11.077

129
Klusman, R. W. (2003). A Geochemical Perspective and Assessment of Leakage Potential
for A Mature Carbon Dioxide–Enhanced Oil Recovery Project and as A Prototype
for Carbon Dioxide Sequestration; Rangely Field, Colorado. AAPG Bulletin, 87(9),
1485–1507. https://doi.org/10.1306/04220302032

Kumar, A., Ozah, R., Noh, M., Pope, G. A., Bryant, S., Sepehrnoori, K., and Lake, L. W.
(2005). Reservoir Simulation of CO2 Storage in Deep Saline Aquifers. SPE J.,
10(03), 336–348. https://doi.org/10.2118/89343-PA

Kutsienyo, E. J., Ampomah, W., Sun, Q., Balch, R. S., You, J., Aggrey, W. N., and Cather,
M. (2019). Evaluation of CO2-EOR Performance and Storage Mechanisms in an
Active Partially Depleted Oil Reservoir. Presented at the SPE Europec at 81st
EAGE Conference and Exhibition, SPE, London, England, UK, D041S008R009.
https://doi.org/10.2118/195534-MS

Lake, L. (1989). Enhanced oil recovery. Prentice Hall.

Lake, L. W., Lotfollahi, M., and Bryant, S. L. (2019). CO2 Enhanced Oil Recovery
Experience and its Messages for CO2 Storage. Science of Carbon Storage in Deep
Saline Formations, 15–31. Elsevier. https://doi.org/10.1016/B978-0-12-812752-
0.00002-2

Lake, L. W., Johns, R. T., Rossen, W. R., and Pope, G. A. (2014). Fundamentals of
Enhanced Oil Recovery (2nd ed.). SPE

Land, C. (1968). Calculation of Imbibition Relative Permeability for Two and Three-Phase
Flow from Rock Properties. SPE J., 8(2), 149-156. https://doi.org/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. DOI:10.2118/38456-PA

Lashgari, H. R., Sun, A., Zhang, T., Pope, G. A., and Lake, L. W. (2019). Evaluation of
Carbon Dioxide Storage and Miscible Gas EOR in Shale Oil Reservoirs. Fuel, 241,
1223–1235. https://doi.org/10.1016/j.fuel.2018.11.076

130
Li, H., and Yan, J. (2009). Evaluating Cubic Equations of State for Calculation of Vapor–
Liquid Equilibrium of CO2 and CO2-Mixtures for CO2 Capture and Storage
Processes. Applied Energy, 86(6), 826–836.
https://doi.org/10.1016/j.apenergy.2008.05.018

Lian, L., Qin, J., Yang, S., Yang, Y., Li, S., and Chen, X. (2014). An Improved Viscosity
Model for CO2-Crude System. Petrol. Explor. Develop., 41(5), 648–653.
https://doi.org/10.1016/S1876-3804(14)60077-X

Liu, F., Lu, P., Zhu, C., and Xiao, Y. (2011). Coupled Reactive Flow and Transport
Modeling of CO2 Sequestration in The Mt. Simon Sandstone Formation, Midwest
U.S.A. International Journal of Greenhouse Gas Control, 5(2), 294–307.
https://doi.org/10.1016/j.ijggc.2010.08.008

Ma, J., Li, L., Wang, H., Du, Y., Ma, J., Zhang, X., and Wang, Z. (2022). Carbon Capture
and Storage: History and The Road Ahead. Engineering, 14, 33–43.
https://doi.org/10.1016/j.eng.2021.11.024

Martin, F. D., and Colpitts, R. M. (1996). Reservoir Engineering. in Standard Handbook


of Petroleum and Natural Gas Engineering. Elsevier. https://doi.org/10.1016/B978-
088415643-7/50009-1

Matter, J. M., Stute, M., Snæbjörnsdottir, S. Ó., Oelkers, E. H., Gislason, S. R., Aradottir,
E. S., Sigfusson, B., Gunnarsson, I., Sigurdardottir, H., Gunnlaugsson, E.,
Axelsson, G., Alfredsson, H. A., Wolff-Boenisch, D., Mesfin, K., Taya, D. F. de la
R., Hall, J., Dideriksen, K., and Broecker, W. S. (2016). Rapid Carbon
Mineralization for Permanent Disposal of Anthropogenic Carbon Dioxide
Emissions. Science, 352(6291), 1312–1314.
https://doi.org/10.1126/science.aad8132

Mito, S., Xue, Z., and Ohsumi, T. (2008). Case Study of Geochemical Reactions at The
Nagaoka CO2 Injection Site, Japan. International Journal of Greenhouse Gas
Control, 2(3), 309–318. https://doi.org/10.1016/j.ijggc.2008.04.007

Mohammad, M., Isaifan, R. J., Weldu, Y. W., Rahman, M. A., and Ghamdi, S. G. A.
(2020). Progress on Carbon Dioxide Capture, Storage and Utilisation. International
Journal of Global Warming, 20(2), 124.
https://doi.org/10.1504/IJGW.2020.105386
131
National Academies of Sciences Engineering and Medicine (NASEM) (2019). Negative
Emissions Technologies and Reliable Sequestration: A Research Agenda. National
Academies Press, Washington, D.C. https://doi.org/10.17226/25259

National Energy Technology Laboratory (NETL). 2015. Carbon Storage Atlas V (5th ed.).
https://www.netl.doe.gov/sites/default/files/2018-10/ATLAS-V-2015.pdf

Nelson, R. C. (1982). The Salinity-Requirement Diagram–A Useful Tool in Chemical


Flooding Research and Development. SPE J., 22(2), 259–270.
https://doi.org/10.2118/8824-PA

Nightingale, M., Johnson, G., Shevalier, M., Hutcheon, I., Perkins, E., and Mayer, B.
(2009). Impact of Injected CO2 on Reservoir Mineralogy During CO2-EOR. Energy
Procedia, 1(1), 3399–3406. https://doi.org/10.1016/j.egypro.2009.02.129

Oil and Gas Climate Initiative (OGCI) (2023). CO2 Storage Resource Catalogue.
https://www.ogci.com/co2-storage-resource-catalogue/

Olajire, A. A. (2014). Review of ASP EOR (Alkaline Surfactant Polymer Enhanced Oil
Recovery) Technology in the Petroleum Industry: Prospects and challenges.
Energy, 77, 963–982. https://doi.org/10.1016/j.energy.2014.09.005

OPEC (2021). OPEC Share of World Crude Oil Reserves.


https://www.opec.org/opec_web/en/data_graphs/330.htm

Pandey, L., Tiwari, P. (2022). Microbial Enhanced Oil Recovery Principles and Potential.
Springer. https://doi.org/10.1007/978-981-16-5465-7

Parkhurst, D. and Appelo, C. (2013). Description of Input and Examples for PHREEQC
Version 3 – A Computer Program for Speciation, Batch-Reaction, One-
Dimensional Transport, and Inverse Geochemical Calculations. U.S. Geological
Survey Techniques and Methods, Book 6.

Peters, E. (2012). Advanced Petrophysics Volume 2. Live Oak Book Company.

132
Pitzer, K. S., and Kim, J. J. (1974). Thermodynamics of Electrolytes. IV. Activity and
Osmotic Coefficients for Mixed Electrolytes. Journal of the American Chemical
Society, 96(18), 5701–5707. https://doi.org/10.1021/ja00825a004

Rickard, D. (2012). Chapter 2 – Sulfur Chemistry in Aqueous Solutions. Developments in


Sedimentology, 65, 31–83. Elsevier. https://doi.org/10.1016/B978-0-444-52989-
3.00002-7

Rosenbauer, R. J., Koksalan, T., and Palandri, J. L. (2005). Experimental Investigation of


CO2–Brine–Rock Interactions at Elevated Temperature and Pressure: Implications
for CO2 Sequestration in Deep-Saline Aquifers. Fuel Processing Technology,
86(14), 1581–1597. https://doi.org/10.1016/j.fuproc.2005.01.011

Rowe, A. M. and Chou, J. C. S. (1970). Pressure-Volume-Temperature-Concentration


Relation of Aqueous NaCl Solutions. J. Chem. Eng. Data, 15, 61-66.

Safi, R., Agarwal, R. K., and Banerjee, S. (2016). Numerical Simulation and Optimization
of CO2 Utilization for Enhanced Oil Recovery from Depleted Reservoirs. Chemical
Engineering Science, 144, 30–38. https://doi.org/10.1016/j.ces.2016.01.021

Samba, M. A., and Elsharafi, M. O. (2018). Literature Review of Water Alternation Gas
Injection. Journal of Earth Energy Engineering, 7(2), 33–45.
https://doi.org/10.25299/jeee.2018.vol7(2).2117

Sebastian, H. M., Wenger, R. S., and Renner, T. A. (1985). Correlation of Minimum


Miscibility Pressure for Impure CO2 Streams. Journal of Petroleum Technology,
37(11), 2076–2082. https://doi.org/10.2118/12648-PA

Sheng, J. J. (2013). Enhanced Oil Recovery Field Case Studies. Elsevier.


https://doi.org/10.1016/C2010-0-67974-0

Sheng, J. J. (2015). Status of surfactant EOR technology. Petroleum, 1(2), 97–105.


https://doi.org/10.1016/j.petlm.2015.07.003

133
Sinha, U., Dindoruk, B., and Soliman, M. (2021). Prediction of CO2 Minimum Miscibility
Pressure Using an Augmented Machine-Learning-Based Model. SPE J., 26(04),
1666–1678. https://doi.org/10.2118/200326-PA

Skauge, A. and Larsen, J. (1994). Three-Phase Permeabilities and Trapped Gas


Measurements Related to WAG Processes. International Journal of The Society of
Core Analysis, 9421, 227-236.

Snæbjörnsdóttir, S. Ó., Oelkers, E. H., Mesfin, K., Aradóttir, E. S., Dideriksen, K.,
Gunnarsson, I., Gunnlaugsson, E., Matter, J. M., Stute, M., and Gislason, S. R.
(2017). The Chemistry and Saturation States of Subsurface Fluids During The
Insitu Mineralisation of CO2 and H2S at The Carbfix Site in SW-Iceland.
International Journal of Greenhouse Gas Control, 58, 87–102.
https://doi.org/10.1016/j.ijggc.2017.01.007

Song, Z., Song, Y., Li, Y., Bai, B., Song, K., and Hou, J. (2020). A Critical Review of CO2
Enhanced Oil Recovery in Tight Oil Reservoirs of North America and China. Fuel,
276, 118006. https://doi.org/10.1016/j.fuel.2020.118006

Spiteri, E. J., Juanes, R., Blunt, M. J., and Orr, F. M. (2005). Relative Permeability
Hysteresis: Trapping Models and Application to Geological CO2 Sequestration.
Presented at the SPE Annual Technical Conference and Exhibition, SPE, Dallas,
Texas, SPE-96448-MS. https://doi.org/10.2118/96448-MS

Stalkup, F. I. (1978). Carbon Dioxide Miscible Flooding: Past, Present, And Outlook for
the Future. Journal of Petroleum Technology, 30(8), 1102–1112.
https://doi.org/10.2118/7042-PA

Sun, Q., Ampomah, W., Kutsienyo, E., Appold, M., Adu-Gyamfi, B., Dai, Z., and
Soltanian, M. R. (2020). Assessment of CO2 Trapping Mechanisms in Partially
Depleted Oil-Bearing Sands. Fuel, 278, 118356.
https://doi.org/10.1016/j.fuel.2020.118356

Taber, J. J. (1958). The Injection of Detergent Slugs in Water Floods. Transactions of the
AIME, 213(1), 186–192. https://doi.org/10.2118/1007-G

134
Tang, J., Vincent-Bonnieu, S., and Rossen, W. R. (2019). Experimental Investigation of
the Effect of Oil on Steady-State Foam Flow in Porous Media. SPE Journal, 24(01),
140–157. https://doi.org/10.2118/194015-PA

Trivedi, J. J., Babadagli, T., Lavoie, R. G., and Nimchuk, D. (2007). Acid Gas
Sequestration During Tertiary Oil Recovery: Optimal Injection Strategies and
Importance of Operational Parameters. Journal of Canadian Petroleum Technology,
46(3), 60–68. https://doi.org/10.2118/07-03-06

U.S. Department of Energy (DOE) (2010). Enhanced Oil Recovery.


https://www.energy.gov/fecm/enhanced-oil-recovery

U.S. Department of Energy (DOE) (2022). Carbon Capture, Transport, and Storage:
Supply Chain Deep Dive Assessment. https://www.energy.gov/fecm/carbon-
capture-transport-and-storage-supply-chain-review-deep-dive-assessment

Wang, L., Tian, Y., Yu, X., Wang, C., Yao, B., Wang, S., Winterfeld, P. H., Wang, X.,
Yang, Z., Wang, Y., Cui, J., and Wu, Y.-S. (2017). Advances in
Improved/Enhanced Oil Recovery Technologies for Tight and Shale Reservoirs.
Fuel, 210, 425–445. https://doi.org/10.1016/j.fuel.2017.08.095

Wellman, T. P., Grigg, R. B., McPherson, B. J., Svec, R. K., and Lichtner, P. C. (2003).
Evaluation of CO2-Brine-Reservoir Rock Interaction with Laboratory Flow Tests
and Reactive Transport Modeling. Presented at the International Symposium on
Oilfield Chemistry, SPE, Houston, Texas, SPE-80228-MS.
https://doi.org/10.2118/80228-MS

Wolery, T. (1992). EQ3/6, A Software Package for Geochemical Modeling of Aqueous


Systems: Package Overview and Installation Guide (Version 7.0). Lawrence
Livermore National Laboratory.

World Coal Association. (WCA) (2018). Carbon Capture, Use and Storage.
https://www.worldcoal.org/clean-coal-technologies/clean-coal-2/

Worldometer (2016). Oil Left in The World. https://www.worldometers.info/oil/#oil-


reserves

135
Xu, C. and Bell, L. (2022). Global Oil and Gas Reserves Increase in 2022. Oil and Gas
Journal. https://www.ogj.com/exploration-
development/reserves/article/14286688/global-oil-and-gas-reserves-increase-in-
2022

Yarveicy, H., and Haghtalab, A. (2018). Effect of Amphoteric Surfactant on Phase


Behavior of Hydrocarbon-Electrolyte-Water System-An Application in Enhanced
Oil Recovery. Journal of Dispersion Science and Technology, 39(4), 522–530.
https://doi.org/10.1080/01932691.2017.1332525

Zhang, S., and DePaolo, D. J. (2017). Rates of CO2 Mineralization in Geological Carbon
Storage. Acc. Chem. Res. 50(9), 2075–2084. doi: 10.1021/acs.accounts.7b00334

Zitha, P.L., Felder, R.D., Zornes, D.R., Brown, K.G., and Mohanty, K.K. (2011).
Increasing Hydrocarbon Recovery Factors.
https://www.spe.org/en/industry/increasing-hydrocarbon-recovery-factors/

136
Vita

Abdulhamid Alsousy is a Saudi petroleum engineer. He gained his bachelor’s

degree in Petroleum Engineering from the University of Oklahoma in 2016. Then,

Abdulhamid joined Saudi Aramco as a reservoir engineer for 4 years, and 1 year as a

production engineer, on onshore and offshore fields. Afterwards, Abdulhamid embarked

to pursue his Master’s degree in Petroleum Engineering at UT Austin.

Email: ana.sousy@yahoo.com

This dissertation was typed by Abdulhamid Alsousy.

137

You might also like