You are on page 1of 171

Subscriber access provided by UNIV OF WESTERN ONTARIO

Review
A Literature Review of CO2-, Natural Gas-, and Water-Based
Fluids for Enhanced Oil Recovery in Unconventional Reservoirs.
Lauren C. Burrows, Foad Haeri, Patricia Cvetic, Sean Sanguinito, Fan
Shi, Deepak Tapriyal, Angela Lea Goodman, and Robert M. Enick
Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.9b03658 • Publication Date (Web): 25 Mar 2020
Downloaded from pubs.acs.org on March 26, 2020

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 170 Energy & Fuels

1
2
3
4
5
6
7
8
A Literature Review of CO2, Natural Gas, and
9
10
11
12
Water-Based Fluids for Enhanced Oil Recovery in
13
14
15
16
Unconventional Reservoirs.
17
18
19
20
21
Lauren C. Burrows,a,b Foad Haeri,a,c Patricia Cvetic,a,c Sean Sanguinito,a,c Fan Shi, a,c Deepak
22
23 Tapriyal,a,c Angela Goodman,a,* Robert M. Enick d,*
24
25
26 AUTHOR ADDRESS
27
28 a
United States Department of Energy, National Energy Technology Laboratory, P.O. Box
29
30
31 10940, Pittsburgh, PA 15236, United States
32
b
33 Oak Ridge Institute of Science and Technology, 100 ORAU Way, Oak Ridge, TN 37830, United
34
35 States
36
37 c
38 Leidos Research Support Team, 626 Cochrans Mill Road, Pittsburgh, PA 15236, United States
39
d
40 Department of Chemical and Petroleum Engineering, University of Pittsburgh, Pittsburgh, PA
41
42 15261, United States
43
44
45
46
47 ABSTRACT. Primary oil recovery from fractured unconventional formations, such as
48
49 shale or tight sands, is typically less than 10%. The development of an economically viable
50
51
enhanced oil recovery (EOR) technique applicable to unconventional liquid reservoirs (ULRs)
52
53
54 would lead to tremendous increases in domestic oil production. Although injection techniques such
55
56 as waterflooding and CO2 EOR have proven profitable in conventional formations for decades,
57
58
59
60 ACS Paragon Plus Environment
1
Energy & Fuels Page 2 of 170

1
2
3 EOR in ULRs presents a far more difficult challenge. The extremely low permeability and mixed
4
5
6 wettability of unconventional formations are the foremost obstacles to success. Because of the
7
8 challenges associated with water-based EOR techniques (a.k.a. chemical EOR) in shale, several
9
10 non-aqueous injection fluids have been considered, including CO2, natural gas and (to a lesser
11
12
13
degree) nitrogen. All these fluids have significantly lower viscosities than water, allowing them to
14
15 more easily penetrate shale nanopores. Unlike water, they also each possess some degree of
16
17 miscibility with oil, which enables the gas to extract oil through a combination of mechanisms.
18
19
Based on laboratory-scale experimentation, CO2 and rich natural gas (methane-rich natural gas
20
21
22 containing high concentrations of ethane, propane and butane) are the most promising EOR fluids.
23
24 The interpretation of results from field tests in the Bakken and Eagle Ford Formations have been
25
26 complicated by interference of frac-hits or well-bashing caused by hydraulic fracturing at nearby
27
28
29 wells. In this review we cover mechanisms, laboratory experiments, numerical simulations and
30
31 field tests involving high-pressure CO2, natural gas, ethane, nitrogen and water.
32
33 1.0 Introduction to enhanced oil recovery in unconventional liquid reservoirs.
34
35
36
Advances in hydraulic fracturing and horizontal drilling of unconventional liquid reservoirs
37
38 (ULRs) have enabled the United States to become a top oil producer.1 Still, only a small percentage
39
40 of the oil present in unconventional reservoirs—less than 10%—can be accessed using current
41
42
technologies.2 After drilling and fracturing a well, oil production typically declines to less than
43
44
45 20% of the initial rate within one year. The lack of options for extending the life of the well forces
46
47 operators to drill additional wells to maintain production.3-4 Because drilling is both expensive and
48
49 carries environmental risks, technologies to enhance oil recovery from mature wells are needed.5-
50
51
52
6
If methods can be developed to recover more oil from unconventional oil wells, a second “shale
53
54 revolution” could be possible.7-8 Unfortunately, techniques that improve oil production in
55
56
57
58
59
60 ACS Paragon Plus Environment
2
Page 3 of 170 Energy & Fuels

1
2
3 conventional formations do not work well in unconventional formations. In conventional
4
5
6 formations, once the reservoir pressure declines and primary production slows, techniques such as
7
8 waterflooding and miscible gas injection can recover a large percentage of the remaining oil. 9
9
10 Miscible gases such as carbon dioxide (CO2), can easily flow through pores and extract
11
12
13
hydrocarbons throughout the reservoir. In contrast, the average permeability of unconventional
14
15 formations is several orders of magnitude lower than that of conventional formations. Therefore,
16
17 injected fluids only flow easily through hydraulic fractures (and possibly natural fractures) and the
18
19
extraction of oil is limited to portions of the formation near fracture surfaces.
20
21
22 Injection of low-viscosity, oil-miscible gases is considered a promising technique for EOR
23
24 in ULRs,4, 10 but it is expected to operate by different mechanisms than those present during EOR
25
26 in conventional reservoirs. Although the injected fluid can flow into the formation via high-
27
28
29 permeability hydraulic fractures, the injected fluid must then diffuse from fractures into oil-rich
30
31 pores in shale or tight formations. Then, through a combination of mechanisms, oil can be driven
32
33 into the high-permeability fractures and produced. Several gases have been considered for EOR in
34
35
36
unconventional reservoirs based on their physical properties, availability and cost—most notably,
37
38 CO2, natural gas, and nitrogen (each of which has also been employed for EOR in conventional
39
40 formations). The abilities of these gases to mobilize oil in shale formations have been, and
41
42
continue to be, examined in computational simulations, laboratory experiments and field tests.
43
44
45 Injection of low salinity brine and aqueous surfactant solutions to promote hydrocarbon mobility
46
47 is another approach being considered.
48
49 Reviews of EOR in ULRs have been presented by several teams.1, 11-17 Reviews have been
50
51
52 presented that focus on high-pressures gases including natural gas, CO2 and nitrogen.12, 16-18

53
54 Selection criteria for EOR in ULRs using miscible gases based on laboratory and field results have
55
56
57
58
59
60 ACS Paragon Plus Environment
3
Energy & Fuels Page 4 of 170

1
2
3 been presented,19 along with an editorial review of those criteria.20 Although reviews of
4
5
6 nanoparticle EOR have also been published; they contain EOR strategies and experimental results
7
8 that are more relevant to conventional formations than to unconventional formations.21-22 A review
9
10 of water-based alkaline-surfactant-polymer (ASP) fluids for EOR in conventional and
11
12
13
unconventional formations was also completed, again with most of the information pertaining to
14
15 conventional formations.23 Reviews of EOR in ULR field projects have also been presented.13-14,
16
17 24-26
18
19
The focus of this review is on techniques to recover hydrocarbons from tight and shale
20
21
22 formations that have already been fractured and yielded primary oil production. A comprehensive
23
24 overview of research activities related to EOR in unconventional liquid reservoirs (ULR) that have
25
26 been conducted in the laboratory and in the field is provided. Emphasis is placed on the selection
27
28
29 of high-pressure gases, including CO2, nitrogen, rich natural gas, lean natural gas, and methane
30
31 that can be injected into a fractured shale formation to promote the recovery of additional oil.
32
33 Although propane and ethane have been shown to be exceptional solvents in laboratory-scale
34
35
36
testing, these valuable alkanes are less likely to be considered for field-scale tests than CO2 and
37
38 natural gas due to their high cost. The mode of injection—whether continuous injection; huff-n-
39
40 puff injection into the same well used for production; or multiple-well cyclic injection into a
41
42
dedicated injector that motivates production from a separate producer—will also be examined.
43
44
45 Although the primary emphasis of this report will be related to high-pressure gas injection, an
46
47 overview of water-based chemical EOR strategies for ULR will also be presented. For example,
48
49 there is a wealth of information related to the use of low salinity water (LSW) or the addition of
50
51
52 surfactants to hydraulic fracturing fluid in order to promote oil recovery during flowback via
53
54 wettability alteration and/or interfacial tension reduction. Therefore, the use of LSW or surfactant
55
56
57
58
59
60 ACS Paragon Plus Environment
4
Page 5 of 170 Energy & Fuels

1
2
3 solutions for EOR in previously fractured formations may be a viable path forward for aqueous
4
5
6 EOR in ULR. Gaps in the literature and paths forward will be discussed.
7
8 This review will not delve into efforts toward enhanced gas recovery from unconventional
9
10 gas reservoirs that produce little, if any, oil, 27-30 such as the fields in the Barnett,31 the Qusaiba in
11
12
13
Saudi Arabia 32, and the Longmaxi in China.33 30, 34 35
Mechanical approaches to improving oil
14
36-40
15 recovery such as improving the fracturing process, optimizing wellbore geometry and/or
16
17 spacing and stability,41-42 improving perforation strategies,43 controlling proppant placement,44
18
19
modeling fracture patterns,45 developing choke management strategies,46 improving post-
20
21
22 fracturing operations and fluid cleanup,47-48 refracturing,49-50 production data analysis,51 and
23
24 improving well monitoring technology,52 are also beyond the scope of this report.
25
26 Although this report is not the first literature review of EOR in ULRs, it is intended to be
27
28
29 quite comprehensive. An overview of the attributes of unconventional formations will be provided.
30
31 The properties of various gases (e.g. ethane, CO2, natural gas, methane, and nitrogen) and the
32
33 mechanisms associated with their injection will be presented, including the different modes of
34
35
36
fluid injection (continuous or huff-n-puff). The results from numerous laboratory-scale high
37
38 pressure gas injection experiments, and 31 field tests will be reviewed. Suggestions for improving
39
40 oil recovery in future EOR trials based on field trials, laboratory experiments, and insights gleaned
41
42
from modeling and simulation studies will be outlined. In an effort to put forth a fully
43
44
45 comprehensive review of the literature to date, we have included non-peer-reviewed conference
46
47 papers in this review article. Of the 383 publications cited in this article, 190 are non-peer-reviewed
48
49 papers, most of which are related to conference presentations. These non-peer-reviewed papers
50
51
52 contain important information related to the current state-of-the-art related to EOR in
53
54 unconventional formations. However, readers are therefore advised to recognize that the findings
55
56
57
58
59
60 ACS Paragon Plus Environment
5
Energy & Fuels Page 6 of 170

1
2
3 and conclusions of these non-peer-reviewed papers have not been subjected to the scrutiny of
4
5
6 multiple reviewers.
7
8 The first experimental and simulation studies investigating the concept of using CO2 as an
9
10 EOR agent for unconventional liquid reservoirs (ULR) were recorded between 2006 to 2012. 53-60
11
12
13
Since then, there has been an explosive growth in research on EOR in ULRs. Most of the
14
15 experimental and simulation efforts have focused on the Bakken and Eagle Ford Formations
16
17 (Figure 1). It is often stressed that results of EOR in ULRs are specific to the shale formation
18
19
studied.61-62
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40 Figure 1. Experiment types and formations studied in research related to EOR in ULRs.Based on
41
literature published between 2008 and Feb 2020.
42
43
44
45 1.1 Differences between conventional and unconventional liquid reservoirs. EOR
46
47 strategies in conventional formations depend on fluid flow through a porous rock matrix. In tight
48
49
and shale formations, rock permeabilities are so low that flow through the matrix is impractical or
50
51
52 impossible—the injection pressure needed to establish flow through an unfractured shale or tight
53
54 formation would be incredibly high (or the flow rate at a reasonable pressure drop would be
55
56
57
58
59
60 ACS Paragon Plus Environment
6
Page 7 of 170 Energy & Fuels

1
2
3 extremely low). Figure 2 (top) illustrates the permeability differences between unconventional and
4
5
6 conventional formations. In general, conventional formations have permeabilities between 100 and
7
8 0.1 mD.14 In contrast, shale formations have permeability values below 0.001 mD—between 100
9
10 and 10,000 times lower than those of conventional formations. For example, oil-producing shale
11
12
13
reservoirs such as Eagle Ford, Wolfcamp, Barnett, and Upper Bakken can be considered
14
15 “extremely tight”, while the Middle Bakken is considered a “tight” formation.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 2. Comparison of permeabilities and pore radii in conventional and unconventional oil
39 reservoirs. Permeability ranges from Canadian Society of Unconventional Resources.
40
41
http://www.csur.com. Pore size ranges from Loucks et al.63
42
43
44 Many characteristics of shale contribute to its low permeability, including nanoscale pore
45
46 sizes, low porosity, low pore connectivity. First, pore radii in conventional reservoirs are generally
47
48
49 greater than 2 microns (μm), (Figure 2, micropore, mesopore, and macropores);64 whereas typical
50
51 pore radii in shales are at least 20-400 times smaller, ranging from 5 nm to 100 nanometers (nm).63,
52
53 65
As shown in Figure 2, pore throat sizes for major oil-producing shale reservoirs—the Eagle
54
55
Ford, Wolfcamp and Barnett Shales—all fall in the nanopore range. The porosity, or the percentage
56
57
58
59
60 ACS Paragon Plus Environment
7
Energy & Fuels Page 8 of 170

1
2
3 of space occupied by pores, in unconventional reservoirs is between 5-15%, compared to 20-30%
4
5
6 in conventional reservoirs.66
7
8 Pores in conventional reservoirs tend to be water-wet (adhering water and releasing oil) to
9
10 intermediate-wet, while pores in ULRs tend to be intermediate-wet to oil-wet (adhering oil).63, 67
11
12
13
For example, the Eagle Ford Formation is characterized as intermediate-wet,68 while the Permian
14
15 Basin and Bakken Formation are between intermediate and oil-wet.69, 70
In conventional
16
17 reservoirs, water can flow through the pores and push oil towards the production well; but in
18
19
unconventional reservoirs, because the accessible, oil-containing pores in shales are oil-wet,
20
21
22 waterflooding is not an efficient method of enhancing oil recovery. Miscible gas injection is a
23
24 more promising approach. The injection of high-pressure gases, such as CO2 and natural gas that
25
26 have substantial solubility in oil and the ability to extract light hydrocarbons from oil, may have
27
28
29 easier access to oil-rich porous network of unconventional formations.
30
31 1.2 Mechanism of high-pressure gas EOR in ULRs. The mechanism of high-pressure
32
33 gas EOR in ULRs is different from the mechanism of EOR in conventional sandstones and
34
35
36
carbonates. In conventional reservoirs, miscible gases such as CO2 can enter the permeable matrix
37
38 and flow through pores (Figure 3, top).71 CO2 can dissolve in the oil, and the hydrocarbons within
39
40 the crude oil can be extracted into the CO2-rich phase. In this manner, multiple-contact miscibility
41
42
between the oil and CO2 can be established in the porous medium (even though these fluids are
43
44
45 not “first-contact” miscible when simply combined in a vessel).72 This multiple-contact, pressure-
46
47 driven oil displacement mechanism is sometimes called a “flushing” mechanism.73 When this
48
49 process is conducted at an injection pressure high enough that essentially all of the crude oil
50
51
52 residing in the high-permeability porous medium is recovered, the pressure is referred to as the
53
54 minimum miscibility pressure (MMP). CO2 pressures above the MMP may make CO2 a stronger
55
56
57
58
59
60 ACS Paragon Plus Environment
8
Page 9 of 170 Energy & Fuels

1
2
3 solvent in general, but no further improvement in oil recovery occurs because, in a conventional
4
5
6 formation, oil recovery is complete at the MMP.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 Figure 3. Oil recovery mechanisms of unconventional and unconventional reservoirs. The yellow
52 and orange lines represent lower and higher molecular weight hydrocarbons, respectively.
53 Relative pore sizes are not drawn to scale.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
9
Energy & Fuels Page 10 of 170

1
2
3 The mechanism of oil recovery in unconventional reservoirs has been the subject of many
4
5
6 experiments and simulations. Here, we introduce a proposed mechanism of miscible gas EOR in
7
8 fractured ULRs as a summary of results of several research groups.71, 74-80 In the mechanism
9
10 described here, CO2 is used as an example of a miscible gas, but other miscible gases such as
11
12
13
natural gas or ethane would operate by similar mechanisms. The mechanism of CO2 EOR in
14
15 fractured shale reservoirs can be divided into four steps (Figure 3, bottom):
16
17 Step 1: Injection. Driven by high injection pressures at the wellhead, CO2 flows rapidly
18
19
through the network of fractures to the matrix-fracture interface. CO2 does not yet permeate the
20
21
22 shale matrix. Although the injection pressure is typically maintained below the fracture pressure,
23
24 the rapid injection of a high-pressure CO2 may cause injection induced fracturing or the slight
25
26 opening of natural fractures.
27
28
29 Step 2: Early soaking period. The high injection pressures at the wellhead forces CO2 to
30
31 permeate pores (and microfractures) close to the matrix-fracture interface. A relatively low
32
33 interfacial tension (IFT) between CO2-rich and oil-rich phases will be established. CO2 dissolves
34
35
36
into the oil in these pores, and causes oil swelling and oil viscosity reduction.81 As the oil in the
37
38 pores swells, the pressure increases slightly in the pores, creating a local gradient where oil moves
39
40 into the fractures. In addition, the exposure of CO2 to the matrix-fracture interface may cause a
41
42
favorable shift in wettability from oil-wet (hydrophobic) to water-wet (hydrophilic) during this
43
44
45 step and the next step.76 Pressure-driven CO2 permeation may also drive some oil in an undesirable
46
47 direction into the matrix and away from the fracture. During the early soaking period, little
48
49 molecular diffusion occurs.
50
51
52 Step 3: Late soaking period. The soaking period continues as the injection pressure
53
54 diminishes. CO2 moves deeper into the pores by molecular diffusion, a process driven by the
55
56
57
58
59
60 ACS Paragon Plus Environment
10
Page 11 of 170 Energy & Fuels

1
2
3 concentration gradient between oil and CO2. Because of the higher concentration of oil in the pores
4
5
6 (and higher concentration of CO2 in the fractures), CO2 migrates into pores, and oil migrates into
7
8 fractures. Relative to the convection-based processes that occur during EOR in conventional
9
10 reservoirs, diffusion-based migration of CO2 within unconventional formations is slow and
11
12
13
commonly requires long soaking periods in the field—although fast oil recovery can be achieved
14
15 using small samples in the laboratory. Because diffusion coefficients are related to molecular
16
17 weight, lower molecular weight injection fluids such as CO2 and methane may be more effective
18
19
at diffusing into pores.82 In addition, lower molecular weight hydrocarbons are mobilized
20
21
22 preferentially over high molecular weight hydrocarbons, due to their higher diffusion
23
24 coefficients.73, 80 The vaporization of the lighter hydrocarbons in the oil becomes significant.81 The
25
26 CO2 solubilities of heavy hydrocarbons increases at higher pressures, so pressures above the MMP
27
28
29 can increase oil recovery. In the late soaking period, oil swelling and viscosity reduction continue
30
31 to facilitate extraction of oil from pores to fractures. The injected CO2 also contributes to pressure
32
33 support within the formation.
34
35
36
Step 4: Production. Pressure is reduced at the well head, enabling the oil (and CO2) in the
37
38 fractures to flow towards the well bore. If the pressure within the high-permeability fractures falls
39
40 below the bubble point of the mixture, solution gas drive can help to move the oil toward the
41
42
production well. Some CO2 remains stored in the shale pores. Due to relative permeability
43
44
45 hysteresis, gas relative permeability during this production step is lower than it is (at the same gas
46
47 saturation) during the injection step, facilitating the movement of the oil phase.83
48
49 Of all the mechanisms that may be working in conjunction to promote oil recovery, there
50
51
52 appears to be a consensus that diffusion, vaporization, oil swelling, oil viscosity reduction, and
53
54 pressure support may be the most important. However, these effects have been more thoroughly
55
56
57
58
59
60 ACS Paragon Plus Environment
11
Energy & Fuels Page 12 of 170

1
2
3 studied than the others. Therefore, none of the remaining effects, such as IFT reduction, wettability
4
5
6 alteration, solution gas drive, relative permeability hysteresis, or injection induced fracturing
7
8 should be discounted.
9
10 1.3 Modes of injection in EOR in ULRs. It is important to classify the modes by which
11
12
13
gases can be injected into previously fractured formation for EOR because these designations
14
15 appear in many descriptions of laboratory studies, computer models and field tests. There are some
16
17 variance in the use of these terms, especially cyclic and huff-n-puff, but in this report, we will use
18
19
the following designations (regardless of differences that may appear in the original papers that
20
21
22 are cited): “continuous” (CO2 is injected into one well and produced from a nearby well),
23
24 “multiple-well cyclic” (CO2 is injected into a well, allowed to soak, and produced from a separate
25
26 nearby well), and “huff-n-puff” (CO2 is injected, allowed to soak, and produced from the same
27
28
29 well).
30
31 Continuous: CO2 is injected into one well and produced from a nearby well. In this
32
33 scenario, CO2 is continuously injected into a well, and CO2, oil, and brine are produced from one
34
35
36
or more nearby production wells (Figure 4A). Typically, continuous tests are only considered in
37
38 formations with high enough permeability for CO2 to move by convective flow through the matrix.
39
40 In this case, the EOR process is analogous to EOR in a conventional formation, but with diffusion
41
42
playing a more prominent role. Further, very good inter-well communication may indicate that
43
44
45 high-permeability flow paths that cause conformance control problem may exist, in which case a
46
47 significant portion of the CO2 flows to the neighboring well rather than diffusing into the
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
12
Page 13 of 170 Energy & Fuels

1
2
3 formation.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 4. Modes of fluid injection in EOR.
29
30
31
32
Multiple-well cyclic: CO2 is injected, allowed to soak, and produced from a nearby well.
33
34 In the rarely reported case of multiple-well cyclic injection, the CO2 is injected into a dedicated
35
36 injection well (Figure 4B). CO2 is then given a long period of time to “soak” into a fractured shale
37
38
matrix by diffusive flow, and to strip oil from the matrix-fracture interfaces. Afterward, a nearby
39
40
41 dedicated production well is opened for production and the CO2, oil, and brine production follows.
42
43 Multiple-well cyclic injection is rarely used because of the difficulty of ensuring proper
44
45 communication between the injection and production wells in fractured shale reservoirs. In some
46
47
48 cases, multiple-well cyclic CO2 injection may be more effective than huff-n-puff in individual
49
50 wells when it is conducted in an asynchronous manner with multiple parallel wells that provides
51
52 beneficial inter-well interference.84
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
13
Energy & Fuels Page 14 of 170

1
2
3 Huff-n-puff: CO2 is injected, allowed to soak, and produced from the same well. Huff-n-
4
5
6 puff injection is a very commonly implemented strategy that is analogous to the cyclic mode, and
7
8 is also referred to as cyclic or intermittent injection. During huff-n-puff injection, both injection
9
10 and production occur in the same well because the very low permeability of the fractured shale
11
12
13
causes poor inter-well communication (Figure 4C). Huff-n-puff can be considered as a stripping
14
15 process relying on a fracture network, whether natural or induced, and is akin to diffusive flow. A
16
17 long soak time is necessary to enable CO2 to permeate the matrix by diffusive flow. Huff-n-puff
18
19
injection is much more commonly implemented than multiple-well cyclic injection. (It should be
20
21
22 noted that the “multiple-well cyclic” strategy illustrated in Figure 4B is so uncommon that the term
23
24 “cyclic” is most commonly associated with the term “huff-n-puff”.)
25
26 2.0 Comparison of fluids being considered for EOR in ULRs. Aqueous liquids and
27
28
29 high-pressure gases have both been considered for EOR in ULRs. Water-based technologies
30
31 employ fresh water or low salinity water (LSW), produced water, or mixtures thereof, and
32
33 chemical EOR involves dissolved additives (e.g. alkaline compounds, surfactants, polymers),
34
35
36
emulsions of solvents, and dispersions of nanoparticles that may improve oil recovery. Water-
37
38 based technologies enjoy economic and safety advantages over the waterless high-pressure gas
39
40 alternatives for EOR and will be discussed in detail in Section 5. High-pressure gases such as
41
42
nitrogen, methane, ethane, and CO2 have oil recovery advantages because they can both apply
43
44
45 pressure to the reservoir, and their low viscosities allow them to more readily penetrate
46
47 unconventional reservoirs. Further, high-pressure gases can recover oil via mechanisms related to
48
49 the dissolution of the gas into the oil, or the dissolution of the lighter oil components into the high-
50
51
52 pressure gas. Such “solvent” attributes are not found in water. Most high-pressure gases have
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
14
Page 15 of 170 Energy & Fuels

1
2
3 higher cost and have more challenging logistics than the use of water or chemical EOR. Table 1
4
5
6 compiles the properties of CO2, methane, ethane and nitrogen that are relevant to EOR.
7
8
9
10 Table 1. Attributes of fluids considered for EOR in ULRs.
11
12
13 Advantageous for oil recovery Moderate effect on oil recovery Disadvantageous for oil recovery
14
15 entry Attribute CO2 Ethane Methane Nitrogen Water
16 1 Kinetic diameter (nm) 85-86 0.330 0.444 0.376 0.364 0.265
17 2 Molecular weight (g/mol) 44.0 30.0 16.0 28.0 18.0
18 3 Critical temperature, °Ca 31.0 32.2 −82.6 −147.0 374.0
4 Critical pressure, MPaa 7.38 4.90 4.61 3.40 22.06
19
5 Critical Density (g/cm3) a 0.467 0.207 0.166 0.313 0.322
20 6 Viscosity (cP), 100 °C, 20.7 MPa a 0.039 0.038 0.018 0.025 0.287
21 7 MMP, Bakken oil (MPa),110 °C 87 17.4 9.26 31.1 101.4 -b
22 8 Solubility in oil High Moderate Moderate Low Very Low
23 9 Swells oil? Significant Significant Yes No No
24 10 Reduces oil viscosity? Significant Yes Yes No No
25 11 Reduces gas-oil IFT? Significant Significant Yes Yes No
26 12 Adsorbs on shale? Yes Yes No No Yes
27 13 Induces asphaltene precipitation? Significant Significant Yes Yes No
28 14 Contaminant if present in produced oil/gas? Yes No No Yes Yes
15 Forms acid in water? Yes No No No No
29
16 Cost (USD per pound)c 0.05 0.24 0.19 d 0.05 <0.01
30
31
32
a
Tc, Pc, ,  values obtained from NIST Chem Webbook data. b Information was not available. c
33 Rough estimates of gas prices are provided, which can fluctuate greatly. The cost for these fluids
34 out of the gas supplier plant. Truck transport costs an additional $0.0015/lb per mile. Additional
35 costs are associated with service companies pumping these fluids into the well. d Liquefied natural
36 gas.
37
38
39
40 2.1 Physical properties of EOR fluids. Molecular size. The molecular size of a gas relates
41
42 to its ability to penetrate nanoscale shale pores. Consider the kinetic diameter of a gas, which is a
43
44
measure of the size of a molecule as it travels along a straight path based on its collision with other
45
46
47 gas molecules. In terms of kinetic diameter, CO2 (0.330 nm) is the smallest of the high-pressure
48
49 gases considered for EOR, including ethane (0.444 nm), methane (0.376 nm), or nitrogen (0.364
50
51 nm) (Table 1, entry 1). Water, with a kinetic diameter of 0.265 nm, is smaller than CO2. The small
52
53
54 size of these molecules is an important aspect of their ability to gain access to the oil-containing
55
56 nanopores of tight formations and shales. As shown in Figure 2, these kinetic diameters are in the
57
58
59
60 ACS Paragon Plus Environment
15
Energy & Fuels Page 16 of 170

1
2
3 picopore (<1 nm) range, which affords them with the possibility of entering the smallest of shale
4
5
6 pores in extremely tight unconventional formations. Because the dimensions of the pores within
7
8 the shale matrix can be comparable to the mean free path of these gases at reservoir conditions,
9
10 Knudsen diffusion can occur during EOR. Diffusivity for Knudsen diffusion is inversely
11
12
13
proportional to the square root of the molecular weight of the gas. Because the molecular weight
14
15 of methane (16) is much smaller than that of CO2 (44) (See Table 1, entry 2), methane could be
16
17 expected to have a diffusion coefficient that is (44/16)0.5 or 1.66 times greater than that of CO2.82
18
19
Supercritical temperature and pressure. The temperature and pressure at which
20
21
22 supercritical conditions are achieved is an important factor in choosing an EOR fluid. CO2 has a
23
24 relatively low critical point of 31 °C and a critical pressure of 7.38 MPa (Table 1, entries 3 and 4).
25
26 This allows the density and solvent strength of CO2 to be manipulated at any temperature via
27
28
29 pressure changes. Higher pressures lead to higher densities approaching liquid-like values and
30
31 greater solvent strength. The critical temperature (32 °C) and critical pressure (4.90 MPa) of ethane
32
33 are similar to those of CO2. Methane and nitrogen both have very low critical temperatures and
34
35
36
are therefore, well within the supercritical region at EOR conditions. These high-pressure gases
37
38 cannot reach liquid-like density values at elevated pressures like CO2 and ethane can.
39
40 Supercritical Density. The ability of a high-pressure gas to dissolve oil is roughly
41
42
correlated to the density of that fluid (Table 1, entry 5). Denser fluids behave as better solvents.
43
44
45 CO2 has the highest critical density of the high-pressure gases considered. Ethane has very similar
46
47 critical temperature and pressure as CO2 but has a lower molecular weight than CO2. As a result,
48
49 ethane has a lower density than CO2 at the same conditions. Methane is the lowest-density high-
50
51
52 pressure gas considered for use as an EOR fluid for ULR. Nitrogen is lower in density than CO2
53
54 at the same temperature and pressure.
55
56
57
58
59
60 ACS Paragon Plus Environment
16
Page 17 of 170 Energy & Fuels

1
2
3 The density of a high-pressure gas is strongly dependent on temperature and pressure. This
4
5
6 temperature and pressure solubility dependence is especially apparent for CO2.88 CO2 is a good
7
8 solvent for oil at high pressures (~12-14 MPa) and a very poor solvent for oil at low pressures.89
9
10 CO2 is a dense fluid at high pressure more akin to a liquid with high solubility in oil whereas CO2
11
12
13
is lower density at lower pressures and has far lower solubility in oil. This dramatic difference
14
15 enables CO2-oil separations to be performed simply by depressurization and phase separation of
16
17 the vapor-phase CO2-rich fluid and the hydrocarbon-rich liquid phase. Conveniently, the solvent
18
19
strength of CO2 increases with fluid density over the temperature range associated with EOR in
20
21
22 both conventional and unconventional formations. Therefore, CO2 can dissolve oil under high
23
24 formation pressures and can be easily separated from oil under near-atmospheric pressures after
25
26 production. Because formation temperatures are so far above the methane critical temperature,
27
28
29 methane does not exhibit the same advantageous changes in density (and solvent strength) that
30
31 CO2 does.
32
33 Viscosity. The viscosity of EOR gases is a function of temperature and pressure. Table 1
34
35
36
compiles viscosities at 100 °C and 20.7 MPa (Table 1, entry 6). The viscosity of high-pressure
37
38 CO2 (0.039 cP) is about seven times lower than that of water (0.287 cP, Table 1). The viscosity of
39
40 ethane is comparable to that of CO2 at the same temperature and pressure. Methane is roughly two
41
42
times less viscous than CO2. In general, the low viscosity of these high-pressure gases may
43
44
45 promote EOR in ULRs by facilitating the flow through hydraulic fractures and natural fracture
46
47 networks and allowing the gases to access more of the shale matrix. On the other hand, there is
48
49 no direct relation between injectivity and viscosity. Further, the low viscosity of these fluids can
50
51
52 also result in gravity override or viscous fingering, which are disadvantageous for oil recovery. At
53
54 the pore scale, viscosity does not have a significant effect on oil recovery.
55
56
57
58
59
60 ACS Paragon Plus Environment
17
Energy & Fuels Page 18 of 170

1
2
3 2.2 Interactions of EOR fluids with oil. Gas-oil minimum miscibility pressure (MMP).
4
5
6 In conventional reservoirs, the MMP represents the pressure at which the maximum oil recovery
7
8 can be attained. In contrast, in EOR in ULRs, oil recovery is incomplete at the MMP. Still, the
9
10 MMP measurements provide a useful qualitative measure of how “strong” a high-pressure gas is
11
12
13
as a solvent for the crude oil. For example, the MMP of CO2, methane, lean natural gas, rich
14
15 natural gas, and nitrogen can be compared at formation temperature to determine their relative
16
17 solvent strengths; the gas with the lowest MMP is the best solvent. The oil-gas MMP’s of high-
18
19
pressure gases are listed in Table 1, entry 7. Based upon these values, ethane is a stronger solvent
20
21
22 for oil than CO2.87, 73 Methane is less miscible with oil than CO2. This disadvantage for methane
23
24 can be significantly mitigated for natural gases that are primarily composed of methane but are
25
26 rich in ethane, propane and butane. The MMP associated with nitrogen is nearly six times higher
27
28
29 than that associated with CO2 and more than three times higher than that of methane,78 indicating
30
31 that it is the least miscible with oil of the four gases listed. Methods to measure or estimate the
32
33 MMP include the slim tube test,60, 85 pressure-composition (Px) pseudo-binary diagrams, the rising
34
35
36
bubble method, 90-92 vanishing interfacial tension, 93, 94 capillary rise vanishing interfacial tension,
37
73, 95, 94
38 and the diminishing interface method.96, 97
39
40 Solubility in oil. The ability of a gas to dissolve in oil positively influences oil recovery.
41
42
Gas dissolved in oil can cause oil swelling and oil viscosity reduction, both of which help oil
43
44
45 migrate from pores to a fracture. Pressure-composition experiments clearly indicate that
46
47 substantial amounts of CO2 can dissolve in oil (Table 1, entry 8). Although methane is certainly
48
49 less soluble in oil than CO2, methane exhibits an appreciable solubility in oil.98-99 For example, the
50
51
52 solubility of methane and CO2 both reach 6 Mscf/bbl, however, this solubility is reached at 27.6
53
54 MPa for CO2, and at 48.3 MPa for methane.89, 90 This can also be illustrated by measurements of
55
56
57
58
59
60 ACS Paragon Plus Environment
18
Page 19 of 170 Energy & Fuels

1
2
3 the gas:oil ratio (GOR) for oil saturated with CO2 (or other gases) as a function of pressure. As
4
5
6 shown in Figure 5, the ratio of CO2 in oil (purple) increases rapidly as pressure increases, while
7
8 the ratio of other gases (enriched natural gas, natural gas and nitrogen) in oil all remain low.92
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 Figure 5. Measured gas/oil ratios of live oil with different high-pressure gases. Data from Li et
28 al., 2017. 92
29
30 Oil swelling due to dissolved high-pressure gas. Oil swelling due to dissolved gas creates
31
32
33
a localized pressure gradient which causes oil to move from pores to fractures. Therefore, gases
34
35 that cause more oil swelling are good candidates for EOR. Habibi et al. presented an excellent
36
37 visual comparison of the differences between CO2 and nitrogen regarding oil swelling due to
38
39 dissolution of the high-pressure gas. 91, 92 Oil was placed in a windowed vessel, followed by
40
41
42 addition of CO2 that initially elevated the pressure to 13.8 MPa (Figure 6). After 5 hours, the oil
43
44 phase was 1.32 times larger than the original oil and the pressure of the closed system fell to 11.8
45
46 MPa. When the swelling experiment was repeated with nitrogen, no detectable oil swelling
47
48
49 occurred, even at 20.7 MPa (Figure 7). Dissolution of methane in oil can also cause oil swelling.98
50
51 Light hydrocarbon gases such as ethane and propane also induce significant degrees of oil
52
53 swelling. 100
54
55
56
57
58
59
60 ACS Paragon Plus Environment
19
Energy & Fuels Page 20 of 170

1
2
3
4
5
6
7
8
9
10
11
12
13 Figure 6. CO2-oil interface at reservoir temperature. (a) Oil (235 mL, 0.78 mol) in a cell at
14
15
atmospheric pressure and no CO2, at reservoir temperature, 50 °C. (b) CO2 (6.4 mol) injected to
16 increase the pressure from 0 to 13.8 MPa (c) After 5 hr, further CO2 dissolution into the oil phase
17 causes oil to swell. Reprinted from Habibi, A.; Yassin, M. R.; Dehghanpour, H.; Bryan, D.,
18 Experimental investigation of CO2-oil interactions in tight rocks: A Montney case study. Fuel
19 2017, 203, 853-867 with permission from Elsevier.
20
21
22
23
24
25
26
27
28
29
30
31
32
33 Figure 7. Nitrogen -oil interface at reservior temperature. (a) Oil (295 mL, 0.98 mol) in a cell at
34 atmospheric pressure and no N2, at reservoir temperature, 50 °C. (b) Nitrogen (1.46 mol) is injected
35 to increase the pressure to 13.8 MPa. (c) Additional nitrogen (0.65 mol) is injected to increase
36 pressure from 13.8 MPa to 20.7 MPa. Reprinted from Habibi, A.; Yassin, M. R.; Dehghanpour,
37 H.; Bryan, D., Experimental investigation of CO2-oil interactions in tight rocks: A Montney case
38
39
study. Fuel 2017, 203, 853-867 with permission from Elsevier.
40
41
42
43 Reduction in oil viscosity due to dissolved gas. The dissolution of some high-pressure gases
44
45
will also reduce oil viscosity. The reduction of oil viscosity facilitates its displacement from pores
46
47
48 and fractures. This viscosity-reducing effect is more prevalent in CO2 than with other gases (Table
49
50 1, entry 10). Figure 8 shows the effect of dissolved CO2 and four other gases on the viscosity of a
51
52 live Bakken oil.92 As the saturation pressure is increased, the viscosity of oil-CO2 mixture (purple)
53
54
55 decreases more rapidly than the viscosity of oil-natural gas (red) or oil-enriched natural gas (blue)
56
57
58
59
60 ACS Paragon Plus Environment
20
Page 21 of 170 Energy & Fuels

1
2
3 mixtures. The experiments showed either no change or slight viscosity increases for nitrogen-based
4
5
6 systems (green). Therefore, the benefits of oil viscosity reduction for EOR will likely be most
7
8 pronounced with CO2 injection.
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 8. Measured viscosities of live oil with different high-pressure gases. Data from Li et al.,
29 2017.92
30
31 IFT reduction due to dissolved gas. Dissolution of high-pressure gases in oil decreases the
32
33
34
IFT between the oil and the high-pressure gas (Table 1, entry 11). The reduction in IFT with
35
36 increasing pressure is most dramatic in the gas phase. As the pressure is increased further and the
37
38 CO2 enters the less compressible liquid phase (above 6.7 mPa), the decrease in IFT with increasing
39
40 pressure continues but becomes less significant. The reduction in IFT will promote oil recovery if
41
42
43 the fluids are flowing through the porous media or fractures, however the IFT reduction is not an
44
45 important parameter in the low permeability portions of the unconventional porous media where
46
47 CO2 transport is accomplished primarily by diffusion processes. Reductions in IFT are less
48
49
50 significant for gases such methane or nitrogen. For example, at 313 K and 5 MPa, the IFT values
51
52 for the CO2-decane, methane-decane and nitrogen-decane systems are 8, 13, and 19 mN/m,
53
54 respectively.101
55
56
57
58
59
60 ACS Paragon Plus Environment
21
Energy & Fuels Page 22 of 170

1
2
3 2.3 Interactions between EOR fluids and the shale reservoir. Adsorption within shale.
4
5
6 CO2 can be adsorbed by the kerogen102-103 or clays within the shale.104, 105 The strong preferential
7
8 adsorption of CO2 over methane is the primary mechanism behind CO2 enhanced natural gas
9
10 recovery from unconventional gas reservoirs.28, 102, 106-109
Experimental 28, 102, 110, 104
and
11
12 23, 107, 108, 109
13
computational techniques have been used to study CO2 adsorption in shale. CO2
14
15 adsorbs to shale approximately 3.7-7 times more strongly than methane.28, 110 One study reported
16
17 that that up to 17 mg CO2/g can be adsorbed to Lower Bakken shale at a pressure of 40 MPa.111
18
19
CO2 adsorption is decreased when water is present, as water can block access of CO2 to shale.112
20
21
22 Although retention of CO2 by the shale formation via adsorption is beneficial for carbon storage,
23
24 it also reduces the amount of the injected CO2 that can mobilize oil via the mechanisms previously
25
26 described. Ethane also adsorbs to shale.113 Methane and nitrogen are not expected to be adsorbed
27
28
29 in shale (Table 1, entry 12).102
30
31 Asphaltene precipitation. Most crude oils contain a mixture of paraffins (linear and
32
33 branched hydrocarbons), naphthenes (hydrocarbons with cyclic groups) and aromatics
34
35
36
(hydrocarbons that contain aromatic groups such as benzene and naphthalene). If an oil is rich in
37
38 aromatics, mixing the crude oil with a strong aromatic solvent such as toluene or xylene results in
39
40 a single, liquid phase. However, mixing the crude oil with a weaker, paraffinic, solvent such as
41
42
pentane will result in a portion of the asphaltenes precipitating and forming a highly viscous
43
44
45 deposit. Although CO2 is a significantly weaker solvent than pentane for hydrocarbons, it has
46
47 sufficient solvent strength to both dissolve in the oil and also extract the lighter components of
48
49 crude oil in the CO2-rich phase, causing asphaltenes to precipitate (Table 1, entry 13). These
50
51
52 deposits can be problematic whether they occur within the formation or within production
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
22
Page 23 of 170 Energy & Fuels

1
2
3 wellbores. Because methane can dissolve in oil and become miscible with light hydrocarbons at
4
5
6 elevated pressure, methane is likely to induce some asphaltene precipitation.
7
8 Produced gas contamination. The non-flammability of CO2 is commonly recognized as
9
10 one of its most appealing safety features. However, the non-flammability of CO2 means it is a
11
12
13
contaminant for produced natural gas (Table 1, entry 14). During the puff phase, the produced gas
14
15 stream will contain natural gas and injected CO2. When the gas is first produced, it is richer in
16
17 CO2. Later in the puff phase, the produced gas mixture becomes richer in natural gas. Because CO2
18
19
has no heating value, the produced gas mixture could be directed to a facility capable of removing
20
21
22 CO2 from the produced gas. A more likely scenario, however, is the venting of the early, CO2-
23
24 rich produced gas mixture to the atmosphere until the subsequently mixture becomes rich enough
25
26 in natural gas to allow it to be either flared or directed to a gas processing plant. In either case,
27
28
29 significant economic costs and environmental impacts associated with the “puff” production of the
30
31 CO2-rich natural gas would complicate the use of CO2 as an EOR solvent for ULRs. The non-
32
33 flammable nature of nitrogen, another fluid considered for EOR in ULRs, will also result in
34
35
36
contamination issues of the natural gas produced along with the oil from an ULR. However, if
37
38 natural gas is selected as the EOR solvent, there would be no concerns about the production of the
39
40 natural gas that was originally injected during the huff cycle along with the natural gas that was
41
42
originally present in the formation. The natural gas would emerge as a single phase that would
43
44
45 either be flared, sent to a gas processing plant, or sold.
46
47 Acid formation. When high-pressure CO2 is in contact with water or brine, CO2 dissolves
48
49 in the aqueous phase and reacts with water to form carbonic acid. The pH of brine is reduced to ~
50
51
52 3 due to the formation of carbonic acid. This acidity can influence the mineral characteristics of
53
54 the formation. For example, although calcium carbonate is only very slightly soluble in water
55
56
57
58
59
60 ACS Paragon Plus Environment
23
Energy & Fuels Page 24 of 170

1
2
3 (0.00015 mol/L at 25 °C), its solubility in acidic water can reach a level of more than 0.03 mol/liter
4
5
6 at a CO2 pressure of 6.2 MPa when the pH is ~3.114 In some cases, the dissolution of calcium
7
8 carbonate may be beneficial in that it enhances injectivity or increase sweep efficiency. The
9
10 formation of acidic brine within the pores of shale is also likely to occur as the injected CO2
11
12
13
diffuses into the formation brine, and the impact of this acid on the mineralogy of the shale and
14
15 the production of oil must be assessed.112 Methane, ethane, and nitrogen do not undergo any
16
17 chemical changes upon exposure to water or brine (Table 1, entry 15).
18
19
2.4 Other considerations. Cost. There are no reports of pure liquid propane, butane, LPG
20
21
22 (liquefied petroleum gas, propane-butane mixtures), Y-grade solvents, mineral oil or energized
23
24 mineral oil being used for laboratory or field studies of EOR in ULR because they are more
25
26 expensive than the gases previously discussed. The cost for the gases shown in Table 1 are roughly
27
28
29 $0.05/lb for CO2, $0.24/lb for ethane, $0.19/lb for LNG (liquefied natural gas), $0.05/lb for
30
31 nitrogen, and <$0.01/lb for water—plus an additional $0.0015/lb per mile of truck transport (Table
32
33 1, entry 16). Additional costs are associated with service companies pumping these fluids into the
34
35
36
well. A study by Hoffman indicated that in case of a high oil price ($80/bbl) and a low gas price
37
38 ($5/Mscf), the natural gas injection process for oil recovery in a four-section area of the Elm
39
40 Coulee field in eastern Montana is of favorable net present value and a high rate of return. 60
41
42
Availability of high-pressure gas. Natural gas can be an attractive fluid to consider for
43
44
45 EOR, especially if natural gas is being produced nearby and there is not a local market for its sale.
46
47 Rather than flaring the gas, it could be used for EOR in ULRs. Even if the natural gas can be
48
49 gathered and sold (wellhead prices can vary from $0.05-0.25/lb), one could consider diverting it
50
51
52 for EOR injection instead. The supplies of ethane derived from the wet gas produced from shale
53
54 could be the primary source of ethane for EOR in unconventional fields, especially if a nearby
55
56
57
58
59
60 ACS Paragon Plus Environment
24
Page 25 of 170 Energy & Fuels

1
2
3 market (e.g. a cracker plant that transforms ethane into ethylene, which is then polymerized to
4
5
6 make polyethylene resin) is not present and sources of CO2 are not convenient.115-116 Ethane could
7
8 be used as an EOR solvent in unconventional formations in regions where CO2 supply pipelines
9
10 are not accessible.116-117 It is possible to separate nitrogen from air using a small processing plant
11
12
13
in the field, which would render nitrogen readily available. Ultimately, the availability of each gas
14
15 is dependent on the location of the project and the equipment and infrastructure available.
16
17 Environmental considerations. Using anthropogenic CO2 for EOR in ULR has the benefit
18
19
of enabling geologic storage of a portion of the injected CO2 because a significant portion of the
20
21
22 injected CO2 would be retained by the formation. This would help reduce net CO2 emissions and
23
24 curb the increasing levels of CO2 in the atmosphere. It should be noted, however, that CO2 streams
25
26 containing natural gas are often vented to the atmosphere until the stream contains pure natural
27
28
29 gas. This process would likely negate any environmental benefit of CO2 EOR in ULRs. The use
30
31 of natural gas as an EOR injection gas would not lead to deliberate venting of greenhouse gases.
32
33 3.0 Laboratory experiments related to high-pressure gas EOR in ULRs. The various
34
35
36
injection modes used in the field (described in Figure 4, Section 1.4) can be modeled in laboratory
37
38 experiments using cylindrical core samples. Continuous injection involves injection of high-
39
40 pressure gases through one end of the core (Figure 9A). The core is confined in a core holder using
41
42
a Teflon sleeve upon which an overburden pressure is applied. The sleeve prevents the flow of
43
44
45 high-pressure fluids around the perimeter of the core and forces the fluids to move only through
46
47 the core. Typically, continuous injection using confined cores is performed using core samples
48
49 from conventional reservoirs (>0.1 mD). Due to the low permeability of shale, few, if any,
50
51
52 successful results of continuous oil production from confined shale cores have been reported.
53
54 There have been a few reports of continuous injection into higher permeability “tight” cores (0.1-
55
56
57
58
59
60 ACS Paragon Plus Environment
25
Energy & Fuels Page 26 of 170

1
2
3 0.001 mD).56, 118, 119, 120 Researchers at Texas A&M attempted a continuous injection of CO2 into
4
5
6 a Bakken shale core.121 At an overburden pressure of 16.5 MPa at 116 °C, and a targeted flow rate
7
8 of 0.01 cm3/min, extremely high pressure drops occurred and steady state was never attained. This
9
10 result demonstrated that continuous injection of CO2 into shale is not possible in a reasonable time
11
12
13
frame.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Figure 9. Injection modes for laboratory core flooding experiments.
37
38
39
40
41 Laboratory-scale representation of multiple-well cyclic injection involves injection of fluid
42
43 into one end of a confined core, followed by a soak time, and subsequent production from the other
44
45 end of the core (Figure 9B). Again, because of the low permeability of shale, injection into one
46
47
48 end of the core and production in the other end is not realistic in shales. Therefore, multiple-well
49
50 cyclic injections in unconventional cores are also relatively uncommon. Reports of EOR
51
52 experiments using multiple-well cyclic injection usually involve cores with permeabilities well
53
54
55 above 0.1 mD.122-123
56
57
58
59
60 ACS Paragon Plus Environment
26
Page 27 of 170 Energy & Fuels

1
2
3 Confined core huff-n-puff injection involves injection into one end of a confined core,
4
5
6 followed by a soak time, and subsequent production from the same end of the core (Figure 9C). In
7
8 this setup, the face of the core being injected with fluid is meant to represent a fracture face. The
9
10 confined core huff-n-puff injection mode was demonstrated in the laboratory by Habibi et al..124-
11
12
13
125
(Note that “huff-n-puff” is called “cyclic” in the papers by Habibi et al.) Cores used in this
14
15 study had diameters of 3.8 cm and permeabilities between 0.30 to 7.53 mD. Although the authors
16
17 refer to these cores as “tight” sandstone, according to the classifications in Figure 1, they would
18
19
be better described as low or moderate permeability conventional cores. Because of the difficulty
20
21
22 of achieving flow of high-pressure gas through a confined core, confined core huff-n-puff is rarely
23
24 used in experiments involving cores from ULRs.
25
26 Continuous, multiple-well cyclic, and confined core huff-n-puff injections all employ
27
28
29 confined cores (Figure 9A, B and C, respectively). Huff-n-puff experiments can also be conducted
30
31 using an unconfined core that is immersed in high-pressure gas, allowing the gas to flow around
32
33 the core, utilizing the entire surface area of the core (Figure 9D). Immersed core huff-n-puff
34
35
36
extraction setups are designed to model a high-permeability fracture (the surrounding high-
37
38 pressure gas) and the low permeability shale matrix (the core). The majority of recent core flooding
39
40 experiments involving shale cores are performed using an immersed core huff-n-puff injection.
41
42
This injection mode is the most efficient way to extract oil from shale cores with permeabilities
43
44
45 below 0.001 mD. Oil recoveries approaching 100% OOIP are possible using an immersed core
46
47 huff-n-puff apparatus. Detailed descriptions of immersed-core huff-n-puff experimental set-ups
48
49 have been reported by the Energy and Environmental Research Center (EERC),76, 99, 108, 109 and the
50
51
52 Schechter Group at Texas A&M University. 126-127
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
27
Energy & Fuels Page 28 of 170

1
2
3 3.1 Core Preparation Procedures. When performing a core flooding experiment, it is
4
5
6 important to know the initial oil saturation of the core, so that the percentage of oil recovered
7
8 during EOR can be calculated. The choice of method for determining the initial saturation depends
9
10 on the permeability of the sample, whether the sample is from an oil-producing zone or an outcrop,
11
12
13
and what equipment is available to the researcher. If a core is from an outcrop, it must be saturated
14
15 with oil prior to the extraction experiments.
16
17 If a core has a high enough permeability to conduct continuous displacements (such as
18
19
those illustrated in Figure 9A), then the procedure is analogous to that commonly reported for core
20
21
22 flooding experiments for conventional formations. First, cores are cleaned, dried and weighed
23
24 (Figure 10A). Once the core is clean, dried and weighed, the permeability or porosity is often
25
26 determined by measuring injection of water or brine.78, 118, 128-130 128, 131-132 Oil is injected into the
27
28
29 core, resulting in the displacement of water (or brine). Injection rates may be set in a range of
30
31 0.005-0.1 mL/min.129, 133 This process typically involves injecting 3-10 pore volumes of oil, or
32
33 until no more brine is produced, at which point the core retains its irreducible water saturation. 118,
34
35 129-130, 132-134
36
The saturation of oil and irreducible water saturation in the core is determined by
37
38 material balance.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
28
Page 29 of 170 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 10. Methods of introducing oil to cores and measuring oil recovery associated with
31
extraction experiments. (The unconfined method by injecting fluid around a core, as shown in B
32
33
and C, is meant to simulate a hydraulic or natural fracture with injected fluid rapidly saturating the
34 high-permeability fractures and diffusive flow or imbibition flowing into the low-permeability
35 matrix.)
36
37
38 For unconventional cores with permeability values so low that continuous displacement is
39
40
41 not feasible, a different procedure is followed (Figure 10B). Cores are first cleaned using the Dean
42
43 Stark solvent extraction technique.59,113,114, 115, 116, 117 The cleaned cores are then placed in a vacuum
44
45 to remove air or residual solvent. Instead of solvent extraction, cores are sometimes simply dried
46
47
48
in a vacuum or oven.134, 135, 136 Although not commonly reported, after cleaning the core may be
49
50 immersed in high-pressure brine to introduce aqueous fluid into the pore network.136 One then
51
52 immerses the core in high-pressure oil for several weeks to saturate the core with oil. In general,
53
54
lower-permeability cores require longer soaking times and higher soaking pressures. For example,
55
56
57
58
59
60 ACS Paragon Plus Environment
29
Energy & Fuels Page 30 of 170

1
2
3 shale outcrop cores can require up to 2-4 months of soaking at elevated pressure (68.9 MPa).121
4
5
6 Others report shorter soak times and lower pressures (13-28 MPa, 2-7 days). 124,137,136 After the oil
7
8 saturation step, the cores are weighed and the EOR experiment is performed. One can measure the
9
10 amount of CO2 and oil exiting the extraction vessel by reducing the pressure of the effluent stream,
11
12
13
which causes the gaseous CO2 and liquid oil to separate by density. Then, the amounts of CO2 and
14
15 oil produced can be measured. The oil recovery at the end of the experiment can also be measured
16
17 by determining the weight loss of core.
18
19
There is an alternative method for determining oil recovery in a core that does not involve
20
21
22 the cleaning and saturation process. Researchers at the EERC perform extraction experiments
23
24 using oil-saturated unconventional cores from the Bakken Formation. The cores from the
25
26 formation are used “as-is” with no effort made to increase their oil content. The cores are weighed,
27
28
29 and the high-pressure extraction experiments performed (Figure 10C).71, 74, 87, 111, 138-139 At the end
30
31 of the extraction experiments, the cores are weighed to determine how much oil was recovered.
32
33 The cores are then ground into a powder and the remaining oil extracted with methylene chloride
34
35
36
and acetone. The mass of this remaining oil corresponds to the oil in the core after the EOR process
37
38 was complete. The sum of the oil extracted during the experiment, and the oil extracted after
39
40 grinding the core amounts to the oil initially present in the core. The percentage of oil extracted
41
42
during the experiment is then calculated. The advantage of this method is the ability to begin EOR
43
44
45 tests upon receipt of the cores without conducting the months-long core cleaning and saturations
46
47 procedures. A drawback of this procedure is the disintegration of the core, preventing it from being
48
49 re-used.
50
51
52 3.2 Effect of gas choice on oil recovery. Initial experiments to evaluate EOR in ULRs
53
54 focus on the choice of gas used for extraction. Gases investigated typically include CO2, methane,
55
56
57
58
59
60 ACS Paragon Plus Environment
30
Page 31 of 170 Energy & Fuels

1
2
3 natural gas (methane-rich fluid containing C2+ components, primarily C2-C4), ethane and
4
5
6 nitrogen. Experiments vary in the permeabilities of the cores. Experiments performed on tight or
7
8 shale cores (permeabilities <0.01 mD) are most relevant for EOR in ULRs. Here, we present all
9
10 reported experiments in which the oil recovery abilities of multiple high-pressure gases are
11
12
13
compared.
14
15 In 2015, Alharthy and colleagues reported experiments performed at the EERC on Middle
16
17 and Lower Bakken cores.76 The oil recovery abilities of CO2, methane, a methane/ethane blend,
18
19
and nitrogen were tested (Figure 11A). The core preparation procedure shown in Figure 10C was
20
21
22 used. Each core was subjected to experimental temperature and pressure meant to simulate the
23
24 original reservoir conditions (110 °C and 34.5 MPa) and maximum recovery times were set at 24
25
26 hours. The bulk of the oil recovery occurred within the first 8-10 hours, with minimal additional
27
28
29 recovery in the remaining time. The highest initial rates of oil recovery were observed using CO2.
30
31 A natural gas mixture (85:15 methane:ethane blend) and methane also gave fast recovery of oil.
32
33 After 24 h, all three of these gases performed well, extracting ~90% OOIP. On the other hand,
34
35
36
nitrogen only extracted of 32% OOIP. Clearly, CO2 and hydrocarbon gases performed best, while
37
38 nitrogen performed worst.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
31
Energy & Fuels Page 32 of 170

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48 Figure 11. (A) Oil recoveries using Middle Bakken cores.76 (B) 24-hour recoveries for the Middle
49 Bakken (B) and Lower Bakken (C) cores.87 (D) Oil recovery versus injection cycles for CO2,
50 methane and nitrogen for Wolfcamp Shale cores.137 (E) Oil recovery ratios of four sandstone cores
51 from the Ordos Basin with different injection fluids.133
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
32
Page 33 of 170 Energy & Fuels

1
2
3 In 2017, Jin et al. from the EERC performed a series of experiments to compare recovery
4
5
6 abilities of CO2, methane, ethane, and nitrogen utilizing core samples from the Upper, Middle, and
7
8 Lower Bakken.87 The core preparation procedure shown in Figure 10C was used. Recovery
9
10 performance was measured for each sample over a 24-hour period. Oil recoveries for Middle
11
12
13
Bakken and Lower Bakken cores are illustrated in Figure 11B and 11C, respectively. Ethane
14
15 exhibited the best results. CO2, a 85:15 methane:ethane blend, and methane performed
16
17 comparably—approaching or exceeding 90% recovery after only 8 hours in Middle Bakken cores
18
19
and then attaining ~95% oil recovery at 24 hours. Nitrogen performed poorly (approximately 26%
20
21
22 recovery after 24 hours). Oil recoveries for the Lower Bakken samples were not as high (just under
23
24 20% for methane, slightly higher than 25% for an 85:15 methane:ethane blend, and above 30% for
25
26 CO2) but nitrogen still fared the worst, with less than 10% recovery.87
27
28
29 Li et al. at Texas Technical University compared how different gases—CO2, methane, and
30
31 nitrogen—affect oil recovery in Wolfcamp Shale cores (Figure 11D).137 A compositional model
32
33 was built to analyze the laboratory results.140 The core preparation procedure shown in Figure 10B
34
35
36
was used, and extractions were performed at 40 °C and 13.8 MPa. As in other studies, CO2
37
38 performed best (between 60-70% recovery) while methane recovered 30-40%, and nitrogen
39
40 recovered 40-50%. Nitrogen fared slightly better than methane in this study, while still returning
41
42
lower oil recovery compared to CO2. Their simulation study that also considered a CO2-propane
43
44
45 mixture, a C1-C2-C4 mixture, a nitrogen-CO2 mixture, and a C1-CO2 mixture. Given the very
46
47 high solvent strength of propane and butane, it is not surprising that gas mixtures enriched in these
48
49 components were also effective EOR solvents, and that a produced gas (~2% N2, 57% C1, 15%
50
51
52 C2, 16% C3, 10% C4) was predicted to have slightly better oil recovery and economics than CO2.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
33
Energy & Fuels Page 34 of 170

1
2
3 In 2018, Chen et al. at the Chinese Academy of Sciences tested four different injection
4
5
6 fluids: water, “active water” (water containing a surfactant), CO2 and nitrogen.133 Sandstone cores
7
8 from the Ordos Basin with average permeabilities of 0.51 mD and average porosities of 14.6%
9
10 were used. Here, the core preparation procedure shown in Figure 10A was used. Although the
11
12
13
authors refer to these cores as “tight” sandstone, according to the classification system shown in
14
15 Figure 1, these cores would be better described as low-permeability conventional cores. Cores
16
17 were placed in a core holder within a nuclear magnetic resonance (NMR) spectrometer and
18
19
pressurized fluids were continuously injected at 10.0 MPa at 67 °C. CO2 flooding gave the highest
20
21
22 oil recovery, followed by nitrogen, active water, and water (Figure 11E). The presence of a
23
24 surfactant improved recovery over water alone by 10%. By monitoring the cores using NMR, the
25
26 authors showed that oil recovery from micro and sub-micropores was 60-70% higher during gas
27
28
29 flooding (CO2 and nitrogen) than during waterflooding (water and “active water”).
30
31 Based solely on technical performance, the most promising pure component fluids for EOR
32
33 in ULR are ethane and CO2 followed by the much less effective gases with very low critical points,
34
35
36
methane and nitrogen. (In the field, the actual choice of fluid is heavily limited by availability and
37
38 logistics.) In general, oil recovery using different gases increases as the miscibility between the
39
40 gas and oil, increases. This correlates to the gas with the lowest MMP being the most effective oil
41
42
recovery agent. Propane has an even lower MMP (3.8 MPa, 110 °C with Bakken crude oil) than
43
44
45 either CO2 or ethane, and would likely perform better in EOR experiments.141 Further, although
46
47 there are no reports of MMP or laboratory-scale EOR studies with pure butane, it is likely that this
48
49 fluid would exhibit MMP values even lower than those reported for propane. These findings affirm
50
51
52 that produced natural gas rich in C2-C4 components is a promising EOR fluid. The selection of
53
54 the best gas does not guarantee success, however. For any gas, even one with exceptional
55
56
57
58
59
60 ACS Paragon Plus Environment
34
Page 35 of 170 Energy & Fuels

1
2
3 performance in laboratory-scale tests, many factors can influence oil recovery results. For
4
5
6 example, the size of the shale sample, the ratio of surface area to sample volume, the identity of
7
8 the gas, the temperature and pressure of the experiment, and the duration of the experiment can
9
10 impact oil recovery.
11
12
13
3.3 CO2 core flooding experiments testing EOR in ULRs. Many articles and conference
14
15 proceedings have emerged in recent years which report investigations regarding CO2 extraction of
16
17 hydrocarbons from tight and shale cores. Experiments are designed to identify the effects of
18
19
injection pressure,73, 77, 121, 134, 136, 142 rock lithology,75, 111, 143-145 soak time and number of huff-n-
20
21
22 puff cycles,134, 121, 128 water saturation,118, 131, 136 and asphaltene deposition130, 135, 146-150 on oil
23
24 recovery. Table 2 lists all reports of CO2 core flooding experiments which may be relevant to CO2
25
26 EOR in unconventional reservoirs. Although many of the authors draw conclusions regarding EOR
27
28
29 from tight and shale reservoirs, the actual permeabilities of the cores tested vary widely. The
30
31 entries in Table 2 are listed in order of core permeability, with higher-permeability cores listed
32
33 first and lower-permeability cores listed last. In many cases, the permeability of the shale cores
34
35
36
tested is unknown. Reports describing cores as “shale cores” were listed in Table 2 as “extremely
37
38 tight”, even though the true permeability may be unknown. Experiments with higher-permeability
39
40 cores are easier to perform, but conclusions made based on these experiments may not be
41
42
applicable to tight and shale formations due to the differences in oil recovery mechanisms. As
43
44
45 expected, the manner by which the experiments are performed changes substantially depending on
46
47 core permeability. For example, experiments employing higher permeability cores (above 0.01
48
49 mD, entries 1-20) are performed using a variety of injection modes, including continuous,
50
51
52 multiple-well cyclic, and confined core huff-n-puff (defined in Figure 9A-C). These experiments
53
54 all involve a core that is confined around its perimeter, typically with an elastomeric Hassler
55
56
57
58
59
60 ACS Paragon Plus Environment
35
Energy & Fuels Page 36 of 170

1
2
3 sleeve, forcing CO2 to flow through the core. With extremely tight shale cores (entries 21-35), on
4
5
6 the other hand, injections are performed using an immersed core huff-n-puff setup (defined in
7
8 Figure 9D).
9
10 Aside from the injection modes, there are several other differences between experiments
11
12
13
using conventional cores (permeability above 0. 1 mD) and unconventional cores (permeability
14
15 below 0.1 mD). Shale cores used in immersed core huff-n-puff experiments are typically smaller
16
17 than higher-permeability cores used in continuous, multiple-well cyclic, or confined-core huff-n-
18
19
puff experiments. In addition, injection pressures are usually higher in experiments using shale
20
21
22 cores. Temperatures may be higher as well, but this parameter is usually chosen based upon the
23
24 reservoir temperature. Finally, the role of water is more commonly tested in cores with higher
25
26 permeabilities in which continuous flooding can occur. For each experiment, injection mode,
27
28
29 reservoir, core permeability, core size, injection pressure, injection temperature, the presence or
30
31 absence of water, and recovery factors are all included in Table 2. In our overview of CO2 core
32
33 flooding in unconventional reservoirs, we will focus our discussion on the reports which employ
34
35
36
extremely low-permeability cores (Table 2, entries 17-36). Our overview is organized by the
37
38 parameters being tested (e.g. effect of porosity, permeability, surface area, rock lithology, injection
39
40 pressure, etc. on oil recovery).
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
36
Page 37 of 170 Energy & Fuels

1
2
3 Table 2. CO2 core flooding experiments listed in order of core permeability.
4 High, 100-10 mD Moderate, 10-1 mD Low, 1-0.1 mD Tight, 0.1-0.01 mD Very Tight, 0.01-0.001 mD Extremely tight, 0.001-0.0001 mD
5
6 entry Injection Core reservoir core permeability core length core dia- Injection temp Water recovery factor reference
7 modea prepb (mD) (cm) meter (cm) pressure (MPa) (°C) (% OOIP)
1 cont, mwc A artificial 12.5-13.8 c 6.0-20.0 2.5 17.5-25.0 60 yes 57-93 Ding et al. 151
8
2 cont, mwc A West China 5.8 97.3 d 2.5 12.9 44 yes 39-48 Zhou et al.123
9 3 HnP-cnf A Northwest China 2.3 97.3 d 2.5 12.9 44 yes 24-38 Ma et al.152
10 4 HnP-cnf, cont. A Monterey 1.3 6.9 3.8 9.3 38 no 54, 93 Vega et al. 58
11 5 cont. A Pembina Cardium 0.8-1.7 20.3-25.4 5.1 7.2-14.0 27, 53 yes 63-87 Cao et al.130
12 6 cont. e
Qianjiang 0.73 n.r. 2.5 17.5, 35 40, 80 no 51, 56 Wang et al.153
13 7 mwc A Lucaogou 0.63 18 2.5 32 75 no 11 Wei et al.122
8 cont. A Pembina Cardium 0.40-2.78 15.2-17.8 5.1 12.0 53 yes 63-83 Li et al.131
14
9 HnP-cnf A Changqing 0.36, 0.81 6.1-6.6 2.5 5-16 61 yes 27-53 Qian et al.129
15 10 HnP-cnf A Montney 0.30-7.53 2.2-5.0 3.8 6.9-9.7 50-90 no 67-73 Habibi et al.124
16 11 mwc A Canadian Bakken 0.27-0.83 n.r. 3.8 7-15 63 yes 43-63 Song et al.128
17 12 cont A Ordos 0.20-1.61 5.4 2.5 10 67 yes 30-43 Chen et al.133
18 13 HnP-cnf B Eagle Ford 0.20-0.25 n.r. n.r. 8.3-13.8 40 yes 45, 68 Li et al.136
19 14 cont A Monteney 0.02, 1.3 7.3, 6.9 3.2, 3.8 7.6-9.0 45, 90 no 0-25 Kovscek et al. 56
15 cont A Bakken 0.016-0.132 15.2-17.8 5.1 10 56 yes 43-92 Han et al.118
20
16 cont B Bakken 0.0018 5.0 2.5 10 38 no 70 Zhang et al.120
21 17 mwc f
Wolfcamp n.r.g 5.1 2.5 31.2 71 no n.r. f Liu et al.154
22 18 HnP-Imm B Mancos, Eagle Ford n.r.h 5 3.8 10.3-24.1 35 no 33-85 Gamadi et al.134
23 19 HnP-Imm Ci Barnett n.r. g 3.3-4.0 2.5 11, 20.7 66 no 18-55 Tovar et al.126
24 20 HnP-Imm B Wolfcamp n.r.h n.r. 2.5 14.5-24.1 74 no 37-55 Tovar et al.121, 78
25 21 HnP-Imm B Wolfcamp n.r.h 3.8-5.8 2.5 10-24.1 77 no 1-49 Adel et al.77
22 HnP-Imm B Wolfcamp 0.0003-0.0005 5.1 3.8 8.3-16.5 40 no 40-70 Li et al.142
26
23 HnP-Imm B Wolfcamp 0.0002-0.00035 5.1 3.8 8.3-13.8 40 yes 40-68 Li et al.136
27 24 HnP-Imm B Wolfcamp 0.0003-0.0005 5.1 3.8 13.8 40 no 65 Li et al.137
28 25 HnP-Imm C Bakken n.r. j 3-4 1 34.5 110 no 40-95 Hawthorne et al.74
29 26 HnP-Imm C Bakken n.r.j 3-4 1 34.5 110 no 75-100 Sorensen et al.145
30 27 HnP-Imm C Bakken n.r., k 0.038 3-4 1.1 34.5 110 no 40, 95 Alharthy et al.76
31 28 HnP-Imm C Bakken 0.00075-0.10 3-4 1.1 34.5 110 no 18-95 Jin et al.143
29 HnP-Imm C Bakken 0.0001-0.067 3-4 1.1 34.5 110 no 58-100 Jin et al.139
32
30 HnP-Imm C Bakken n.r.l 4 1.1 34.5 110 no 15-65 Jin et al.111
33 31 HnP-Imm C Bakken 0.00075-0.05 3-4 1.1 34.5 110 no >90 Jin et al.87
34 32 HnP-Imm C Bakken n.r. j 3-4 1 17.2-51.7 110 no 27-100 Hawthorne et al.73
35 33 HnP-Imm B Not reported 0.0001 10.2 3.8 15.2 20 no 30 Meng et al.155
36 34 HnP-Imm B Jianghan 0.00035-0.00045 5.0 2.5 6-18 20 no 46, 68 Li et al.156
37 35 HnP-Imm B Eagle Ford 0.00029 3.8 5.0 13.8 40 no 61 Li et al 157
36 HnP-Imm C Bakken n.r. 4-5 1.1 34.5 110 no 12-100 Hawthorne et al. 79-80
38 a
Injection mode, as described in Figure 9. Cont. = continuous, mwc = multiple well cyclic, HnP-cnf = huff-n-puff, confined core, HnP-imm = huff-n-puff, immersed core. b Core preparation procedure, as
39 described in Figure 10. c Fractured cores. d 22 cores of total length 97.3 cm. e Cores were used as received. Oil content measured by NMR. f Amount of oil recovered reported in g, not a percentage. g
40 Permeability in the nanodarcy range.h Described as shale cores. i Preserved sidewall cores used as received. Cores were not ground to a powder at the end of the experiment. j Middle Bakken, 0.002-0.04
41 mD, Lower Bakken. k Lower Bakken. l Upper and Lower Bakken shale.
42
43
44
45 ACS Paragon Plus Environment
46 37
47
Energy & Fuels Page 38 of 170

1
2
3 Effect of porosity and permeability on oil recovery. In an early experiment testing oil
4
5
6 recovery in rocks of varying porosity, Hawthorne et al. performed core flooding experiments with
7
8 CO2 on three different sample types: (1) a conventional sandstone core with 25% porosity and
9
10 1000 mD permeability, (2) Middle Bakken siltstone with 4-8% porosity and 0.002 to 0.04 mD
11
12
13
permeability, and (3) Upper and Lower Bakken shale samples with ultralow permeability (Table
14
15 2, entry 25).74 CO2 was allowed to flow around small cores (1 cm diameter, 4 cm length) at 5000
16
17 psi. The high-permeability conventional core had all of its oil recovered within one hour. The low-
18
19
permeability middle Bakken core had all oil recovered in 3 hours. In the Upper and Lower Bakken
20
21
22 shale samples, however, complete hydrocarbon recovery required 96 hours of CO2 flow. Clearly,
23
24 oil was extracted more easily from higher-permeability, higher-porosity cores. Based upon these
25
26 results, the authors conclude that CO2 EOR is possible in the Bakken, but will require long shut-
27
28
29 in times, especially in the lowest permeability media.
30
31 Effect of exposed surface area on oil recovery. In a second, related experiment, Hawthorne
32
33 et al. crushed the shale samples to a 3 mm sieve, and CO2 was allowed to flow around the crushed
34
35
36
sample. In this experiment, 80% of hydrocarbons were recovered in only 2-4 hours. This result
37
38 suggests that the amount of exposed surface area of the rock is critical for hydrocarbon recovery,
39
40 especially for shale.71, 74 The size of the cores used for CO2 core flooding experiments also effects
41
42
the amount of surface area exposed to CO2—which in turn, effects oil recovery. Figure 12 shows
43
44
45 a compilation of oil recovery results after 24 h of CO2 huff-n-puff laboratory testing at near-
46
47 miscible and miscible pressure conditions.73-77, 126, 132, 134, 139 The y-axis represents the percent of
48
49 oil recovered from the sample and the x-axis represents the ratio of the volume (V) of the piece
50
51
52 of shale to the surface area (A) exposed to CO2 during the “huff” phase (V/A). If the core is
53
54 confined by a sleeve, then A is the area of one circular end of the core. The largest values of V/A
55
56
57
58
59
60 ACS Paragon Plus Environment
38
Page 39 of 170 Energy & Fuels

1
2
3 are associated with small pieces of shale immersed in CO2 and the smallest values of V/A with
4
5
6 larger, confined shale cores. In the U.S. Middle Bakken experiments, 90-100% of the oil was
7
8 recovered for the smallest pieces of shale immersed in CO2. For relatively large cores (3.8 cm in
9
10 diameter and 5.1 cm long), V/A increased to 0.7 and oil recovery dropped to 65-72%. For even
11
12
13
larger cores (3.81 cm diameter, 7.0 cm long confined cores), the ratio of V/A increased to 7 and
14
15 oil recovery was only 25%. The green dashed line is a qualitative line that illustrates the trend of
16
17 the highest reported oil recovery for the Bakken sample as a function of V/A. Results for shale
18
19
samples for other formations are also provided, but there was not a wide range of V/A values
20
21
22 available for other formations, therefore no line correlating data from other formation is provided.
23
24 Oil recovery decreases dramatically with increasing V/A, which is to be expected in a process
25
26 where CO2 diffusion is responsible for the recovery.
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 Figure 12. Compilation of oil recovery results after 24 hr of CO2 huff-n-puff laboratory testing at
52 near-miscible and miscible pressure conditions for Bakken, Mancos, Eagle Ford and other
53 unspecified shale.73-77, 126, 132, 134, 139 Oil recovery (%) is plotted versus core volume/core surface
54 area of shale exposed to CO2 (V/A). Percent oil recovery was compiled at 24 hours, regardless of
55 cycle length and number of cycles (e.g. one 24-h cycle or six 4-hour cycles).
56
57
58
59
60 ACS Paragon Plus Environment
39
Energy & Fuels Page 40 of 170

1
2
3 Effect of injection pressure on oil recovery. Many researchers have investigated whether
4
5
6 pressures above the MMP yield a significant amount of additional oil recovery during huff-n-puff
7
8 experiments. Experiments testing the effect of injection pressure on oil recovery by CO2 typically
9
10 involve first determining the CO2-oil MMP of the formation oil at reservoir temperature, and then
11
12
13
performing extraction experiments with CO2 below, near, and above the MMP. Results of six
14
15 different experiments performed by different research groups are discussed below and summarized
16
17 in Figure 13. 134, 142, 136, 73, 121, 77
18
19
At Texas Technical University, Gamadi et al. conducted a series of immersed core huff-n-
20
21
22 puff core tests using Mancos and Eagle Ford shale cores (Table 2, entry 18).134 The effect of
23
24 changing soaking period, soaking pressure, and numbers of cycles was tested. Oil recovery
25
26 increased with increasing pressure up to the MMP, but only modest improvements in oil recovery
27
28
29 were attained at pressures above the MMP (Figure 13A). However, shorter soak times could be
30
31 used when the CO2 was injected at high pressures above the MMP. For a specified total duration
32
33 (for example 48 hours) higher oil recovery was seen with shorter shut-in times and more cycles
34
35
36
(50% recovery for four 12-hour cycles), as opposed to longer shut-in times and fewer cycles (38%
37
38 for two 24-hour cycles).
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
40
Page 41 of 170 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29 Figure 13. Oil recovery with increasing pressure in shale core flooding experiments. CO2-oil
30 MMPs are indicated in blue. 134, 142, 136, 73, 121, 77
31
32
33
34
The effect of the injection pressure on EOR in Wolfcamp shale cores was studied by Li et
35
36 al. (Table 2, entry 22).142 The estimated MMP for the CO2−Wolfcamp crude oil system was about
37
38 11.2 MPa at 40 °C (Figure 13B). Wolfcamp shale cores were studied using huff-n-puff
39
40
experiments using cores with permeabilities between 300 and 500 nD. An injection pressure of
41
42
43 1.4 MPa higher than the MMP improved recovery in the shale core. Beyond that, oil recovery
44
45 leveled off with increasing pressures. A second set of experiments by Li et al. using extremely
46
47 tight Wolfcamp cores (permeabilities of 200 nD to 350 nD) showed recovery of 68% after 7 cycles
48
49
50 (Figure 13C, Table 2, entry 23).136 When the pressure was increased above the MMP, oil recovery
51
52 increased by about 9%. Pressure conditions further above the MMP did not significantly increase
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
41
Energy & Fuels Page 42 of 170

1
2
3 oil recovery. When the core flooding experiment was conducted with imbibed water (15 wt% KCl)
4
5
6 and Wolfcamp oil, the oil recovery was reduced to 45%.
7
8 Hawthorne et al. at the EERC studied the effect of pressure on oil recovery from shale
9
10 pores (Table 2, entry 32).73 Using an immersed core flooding apparatus, the oil recovery by CO2
11
12
13
extraction was measured at 17.2, 34.5 and 51.7 MPa. They found that higher pressures increased
14
15 oil recovery (Figure 13D). In particular, recovery of higher molecular weight hydrocarbons
16
17 increased with increasing injection pressure.
18
19
In the Schechter Group at Texas A&M University, Tovar et al. conducted a study of the
20
21
22 pressure-dependency of CO2 EOR in the immersed core huff-n-puff injection mode using
23
24 preserved sidewall cores from the Wolfcamp Formation (Table 2, entry 20).121 They found that oil
25
26 recovery increased with increasing pressures, even beyond the MMP (Figure 13E). The highest
27
28
29 recovery (~40%) was obtained at the highest pressure and the longest soak time (not shown). Adel
30
31 et al., in the same research group, performed huff-n-puff core floods that demonstrated a continued
32
33 increase in recovery as pressure increases beyond the MMP (Figure 13F, Table 2, entry 21).77
34
35
36
In general, researchers found that increasing injection pressures beyond the CO2-oil MMP
37
38 increased oil recovery. In some cases, recovery increased steadily with increasing pressure,73, 77,
39
40 121
while in other cases, oil recovery plateaued at some value above the MMP.134, 136, 142 Increased
41
42
recoveries at higher pressures may be due to the increased ability of CO2 to access smaller pores.
43
44 129
45 In general, during EOR in ULRs, oil recovery will likely be improved by increasing injection
46
47 pressures, however the magnitude of the increase may vary significantly.
48
49 Effect of soak time and number of cycles on huff-n-puff oil recovery. The goal of many core
50
51
52 flooding experiments is to identify conditions which will recover as much oil as possible in the
53
54 shortest amount of time. Researchers agree that increasing the amount of soak time in a huff-n-
55
56
57
58
59
60 ACS Paragon Plus Environment
42
Page 43 of 170 Energy & Fuels

1
2
3 puff experiment will increase oil recovery, but there is a point beyond which the oil recovery
4
5
6 plateaus and recovery in the field would become uneconomical. Many researchers also agree that
7
8 increasing the number of cycles in a huff-n-puff experiment increases recovery, but this requires a
9
10 greater volume of CO2. The process also affords diminishing returns in terms of oil recovery after
11
12
13
only a few cycles. There have also been studies that consider modifying the duration of the cycles
14
15 to optimize performance.
16
17 Gamadi et al. conducted CO2 core flooding experiments using Mancos and Eagle Ford
18
19
shale cores (Table 2, entry 18).134 Soaking period, soaking pressure, and numbers of cycles were
20
21
22 altered and recoveries of 33% to 85% were observed. For a specified total duration (for example
23
24 48 hours) higher oil recovery was seen with shorter shut-in times and more cycles (50% recovery
25
26 for four 12-hour cycles), as opposed to longer shut-in times and fewer cycles (38% for two 24-
27
28
29 hour cycles). Figure 14 shows results from four experiments in which the effect of cycle time was
30
31 tested. Four cycles were performed for each experiment, and the cycle times varied (6, 12, 24, and
32
33 48 h). The oil recovery did increase with increasing duration of cycles, however, better
34
35
36
efficiency—in terms of oil recovered per hour (y axis)—was achieved with shorter cycles.
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 Figure 14. Oil recovery efficiency versus duration of cycles. Data from Gamadi et al., 2014.134
55
56
57
58
59
60 ACS Paragon Plus Environment
43
Energy & Fuels Page 44 of 170

1
2
3 Tovar et al. also studied the effect of injection pressure on oil recovery (Table 2, entry
4
5
6 20).121 The highest recovery (~40%) was obtained at the highest injection pressure and the longest
7
8 soak time. The same researchers also demonstrated that when multiple cycles are used for huff-n-
9
10 puff, the early cycles yield the greatest amount of oil production.
11
12
13
With regard to the duration of the soak time, most papers report an increase in incremental
14
15 oil with increasing soak time, however this effect becomes much less significant once a certain
16
17 duration is reached. For example, oil recovery from 1.5 inch diameter Canadian Bakken core plugs
18
19
increased with soak times up to 6 hours, but little additional oil was recovered with a 9 hour soak
20
21
22 (Table 2, entry 11).128 However, when one considers a fixed total operational time for recovery,
23
24 greater oil recovery may be obtained with shorter soak times and shorter production times because
25
26 this allows for more cycles to be conducted.
27
28
29 Effect of shale lithology on oil recovery. Because ULRs such as the Bakken have complex
30
31 lithologies, researchers at the EERC performed detailed core analyses of six different Bakken wells
32
33 to characterize their porosity, permeability, water content, total organic carbon, mineralogy, and
34
35
36
natural fractures.75 Cores from the Three Forks, Lower Bakken shale, Middle Bakken, and Upper
37
38 Bakken shale were included in this study. Cores from the Middle Bakken and Three Forks
39
40 Formations showed rapid oil recovery, with 80% of oil being recovered within 8 hours of CO2
41
42
extraction. In contrast, the lower permeability Upper and Lower Bakken shales gave 54-67%
43
44
45 recoveries after 24 hours. Although the oil recoveries for the Upper and Lower Bakken shales
46
47 samples were significantly lower, the authors express that the observed oil recovery values are
48
49 encouraging. For reference, their report includes a detailed table outlining core properties such as
50
51
52 sample depth, TOC, mineralogical data, and oil recovery results. A subsequent report by
53
54 Hawthorne et al. described rock characterization for 51 different cores from the Bakken formation
55
56
57
58
59
60 ACS Paragon Plus Environment
44
Page 45 of 170 Energy & Fuels

1
2
3 (Table 2, entry 36).79 They found that oil recovery most depended on total organic carbon content
4
5
6 and pore throat radii. Many reports investigating the potential for CO2 EOR include complex rock
7
8 characterization data alongside oil recovery data. The task of interpreting this characterization data
9
10 to determine which properties most influence oil recovery is non-trivial. Jin analyzed the results
11
12
13
of core flood experiments performed on 21 different samples from the Bakken Petroleum System.
14
15 The percentage of oil recovery after seven hours was recorded for each sample, and the results
16
17 were subjected to a statistical analysis to discover which of five shale characteristics (porosity,
18
19
permeability, total organic content, pore throat radius and water saturation) most significantly
20
21
22 affected oil recovery. The authors concluded that total organic carbon (TOC) and pore-throat
23
24 radius most significantly impacted oil recovery during core flood experiments (Table 2, entry
25
26 29).139
27
28
29 Mineralogical data collected using x-ray diffraction (XRD) revealed that the two main
30
31 components of Upper and Lower Bakken shale are quartz and illite.111, 138 An analysis of the
32
33 kerogen content via pyrolysis showed that the majority of the kerogen was type II marine kerogen.
34
35
36
The authors report that 40% of kerogen in Bakken Shale is in the immature category, which
37
38 indicates that the kerogen has not transformed into oil. When combined with small pore sizes, the
39
40 high content of immature kerogen indicates that much of the oil may have lower mobility than
41
42
otherwise expected. Oil that is contained in these immature kerogen pores would be tightly
43
44
45 adsorbed and difficult to extract. Extraction experiments using the Upper and Lower Bakken Shale
46
47 cores described in this study afforded oil recovery rates between 15 and 65% (Table 2, entry 30).111
48
49 Asphaltene deposition during EOR in ULR. Asphaltene precipitation can occur during CO2
50
51
52 flooding when oil mixes with CO2, and the insoluble components are precipitated within the
53
54 reservoir. 95, 96 This process is known to reduce permeability in conventional oil reservoirs, and the
55
56
57
58
59
60 ACS Paragon Plus Environment
45
Energy & Fuels Page 46 of 170

1
2
3 same process is thought to occur in ULRs. Due to the small pore sizes in ULRs, asphaltene
4
5
6 plugging may be even more detrimental to permeability than in conventional reservoirs.150 Several
7
8 researchers have studied the effect of asphaltene precipitation on oil recovery in ULRs.
9
10 Shen studied the potential for asphaltene plugging by comparing the pore size distribution
11
12
13
in shale samples to the size of asphaltene aggregates formed upon exposure of oil to CO2 (or
14
15 methane).149 The measured pore sizes in Eagle Ford and Mancos Shale were mostly between 3 nm
16
17 and 50 nm. Pre-filtered crude oil was confined in a pressurized cell at 21 °C and CO2 was injected
18
19
into the cell. Asphaltene deposits were collected on nanomembranes. Most of the asphaltene
20
21
22 particles were between 30-200 nm in size. The large size of the asphaltene particles compared to
23
24 the small size of the shale pores indicates that the asphaltenes can plug shale pores during CO2
25
26 extraction. CO2 gave slightly more asphaltene deposits and larger particle sizes than methane.
27
28
29 Next, Shen and Sheng performed an experimental and numerical investigation of
30
31 permeability reduction due to asphaltene deposition during CO2 huff-n-puff injection with
32
33 Wolfcamp shale oil and Eagle Ford shale outcrop samples.135 They found a 27% permeability
34
35
36
reduction after the first cycle of CO2 huff-n-puff injection, a 49% permeability reduction after 6
37
38 cycles. The reduction in permeability was immediate after the first cycle; indicating that a fast
39
40 asphaltene deposition occurs, which can block pore pathways for hydrocarbon extraction.
41
42
Shen and Sheng also studied asphaltene precipitation and deposition-associated formation
43
44
45 damage.158 Eagle Ford outcrop shale cores and Wolfcamp oil were used in their experiments. The
46
47 reduction in pore size and core permeability upon CO2 exposure was measured. After 6 cycles of
48
49 CO2 soaking, the number of pores with diameters between 100 to 800 nm decreased, and the
50
51
52 number of pores with diameters smaller than 100 nm increased. When the same experiment was
53
54 performed using decane instead of Wolfcamp crude oil, no change in pore size distribution was
55
56
57
58
59
60 ACS Paragon Plus Environment
46
Page 47 of 170 Energy & Fuels

1
2
3 observed. Therefore, the change in pore size distribution was likely due to pore plugging and
4
5
6 asphaltene adsorption. After 6 cycles of CO2 soaking, the original permeability (0.126 mD) was
7
8 reduced by more than one-third due to asphaltene precipitation and deposition.
9
10 There are two mechanisms by which asphaltenes can reduce shale permeability: (1)
11
12
13
adsorption of high molecular weight asphaltenes to rock surfaces, and (2) by mechanical plugging
14
15 of pores by asphaltene aggregates suspended in oil. Reverse core flooding using n-heptane and
16
17 toluene was used to determine which of these two mechanisms predominates in ULRs.150 An Eagle
18
19
Ford shale core was saturated with Wolfcamp crude oil and immersed in CO2. The core was soaked
20
21
22 for 6 hours, and the pressure released slowly to produce CO2 and oil. After soaking for 6 cycles, a
23
24 0.9 cm slice of the rock was cut from the core. N-heptane flooding of this 0.9 cm slice was first
25
26 performed at 20.7 MPa to induce asphaltene precipitation within the pores. The permeability after
27
28
29 n-heptane flooding was 0.199 mD. A second flood with toluene removed asphaltenes from the
30
31 pores previously deposited within the rock during the n-heptane flood. The permeability after
32
33 toluene flooding was 0.333 mD, which the authors assumed was approximately equal to the
34
35
36
permeability of the core in the absence of asphaltenes. The authors concluded that the permeability
37
38 of the core was reduced from 0.333 mD to 0.199 mD due to asphaltene deposition during CO2
39
40 flooding. Based on the permeability changes after each step, they report that 83% of the total
41
42
permeability reduction was contributed by asphaltene plugging, but only 17% was caused by
43
44
45 asphaltene adsorption.
46
47 X-ray CT scanning experiments. CT scanning can be used to monitor changes in fluid
48
49 saturation in core samples. Tovar et al. performed fourteen total core-flooding huff-n-puff
50
51
52 experiments: nine injecting CO2 into shale cores, two injecting nitrogen into shale cores and three
53
54 injecting CO2 into Berea sandstone cores (Table 2, entry 20).78 CT-scanning during core-flooding
55
56
57
58
59
60 ACS Paragon Plus Environment
47
Energy & Fuels Page 48 of 170

1
2
3 was used to study the mechanism of oil recovery. Initially, the core composition was uniform
4
5
6 throughout and contained no CO2. As time proceeded, the CO2 diffused throughout the entire core.
7
8 Further, it was observed that lower molecular weight hydrocarbons were produced preferentially,
9
10 which lead the investigators to propose a peripheral, slow-kinetics vaporizing gas drive as the main
11
12
13
mechanism of oil recovery.
14
15 Fogden et al. used CT to visually confirm where diffusion occurs in shales and developed
16
17 a technique to determine diffusion rates.159 Sub-resolution porosity was determined on samples of
18
19
several mm in diameter by performing image registration and differentiation of samples scanned
20
21
22 dry and with CH2I2. Samples were then placed in a pressure vessel with toluene and CT scanned
23
24 sequentially to determine where the toluene infiltrated the shale matrix. Fractures (both filled and
25
26 open) in shale and higher porosity zones associated with organics demonstrated a higher rate of
27
28
29 fluid infiltration. This methodology has the potential to evaluate sub-micron features within the
30
31 matrix without the need for synchrotron techniques.
32
33 Zhang et al. performed an experimental and numerical study on the use of CO2 for EOR in
34
35
36
ULRs (Table 2, entry 16).120 First they quantified the recovery potential of CO2 through core flood
37
38 experiments. They used CT scanning to get more insight about the multiphase flow and fluid
39
40 distribution. They found that CO2 penetrates from fractures into matrix through a diffusive
41
42
mechanism to achieve miscibility. The recovery from core flood experiments was more than 70%
43
44
45 for Bakken samples with an average porosity of 7.5% and permeability of 1.8 μD.
46
47 Reactions between CO2 and shale. Fracturing and EOR processes can have a significant
48
49 effect on the petrophysical properties of shale.160 In the presence of water, CO2 forms a carbonic
50
51
52 acid solution which can lead to include mineral dissolution and precipitation, volume changes in
53
54 clays due to pH and ion concentration changes, dehydration.161 Chemical reactions between CO2,
55
56
57
58
59
60 ACS Paragon Plus Environment
48
Page 49 of 170 Energy & Fuels

1
2
3 water and shale have the potential to both increase or decrease shale permeability. Due to the
4
5
6 mineralogical complexity of shale, the causes of permeability changes can be difficult to identify.
7
8 The reader is referred to a recent review on CO2-shale interactions for further information.161
9
10 Mineral dissolution caused by acidic carbonic acid solutions has often been shown to
11
12
13
increase shale permeability. For example, Zhang et al. studied solubility, brine composition,
14
15 morphology, mineralogy and wettability changes resulting from interactions between CO2 and
16
17 tight sandstone.119 Dissolution of feldspar and cement was observed, which increased the pore
18
19
connectivity of the tight rocks.119 Wang et al. observed changes in mesopore volume as a result of
20
21
22 chemical reactions between CO2 and shale minerals.162 Zhu et al. found that exposure of Chinese
23
24 Wufeng Shale to supercritical CO2 in both the presence and absence of water caused dissolution
25
26 of calcite, dolomite, and pyrite minerals.163 Similarly, Lu et al. examined the interaction of CO2
27
28
29 saturated brine with carbonate (especially calcite) present in Middle Bakken core samples.164 They
30
31 observed an improvement in the residual conductivity of the fracture. Sanguinito and Goodman
32
33 found that dry CO2 is capable of reacting with water in the clay layers and can cause carbonate
34
35
36
dissolution.112 These reactions are magnified when water is available.165 Figure 15 combines
37
38 scanning electron microscopy, infrared spectroscopy, and pore size analysis to show pores opening
39
40 at the micro-scale and closing at the nano scale. These pore size changes may lead to changes in
41
42
the overall flow properties of both the matrix and fractures.165-166
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
49
Energy & Fuels Page 50 of 170

1
2
3
4
5
6
7
8
9
10
11 40 m 40 m
12
13
Figure 15. Scanning electron microscopy images of Utica Shale exposed to CO2 (first image) and
14 CO2 and water exposed Utica Shale (fourth image). These images are of the same location before
15 and after exposure to CO2 and water. Etching and pitting of this matrix occurs under CO2 exposure
16 (first image) but is amplified by the presence of CO2 and water (fourth image). Attenuated total
17 reflectance-Fourier transform infrared spectroscopy (ATR-FT-IR) spectra of the interface of the
18 CO2 fluid and adsorbed water film on the Utica Shale surface recorded as a function of time from
19
0 minutes to 35 days after the initial exposure to CO2 at 0.5 MPa (second image). Comparison of
20
21 unexposed, CO2-exposed, and CO2+H2O-exposed Utica shale pore size distribution from nitrogen
22 adsorption isotherms analysis (third image). Reprinted from Goodman, A.; Sanguinito, S.; Tkach,
23 M.; Natesakhawat, S.; Kutchko, B.; Fazio, J.; Cvetic, P., Investigating the role of water on CO2-
24 Utica Shale interactions for carbon storage and shale gas extraction activities – Evidence for pore
25 scale alterations. Fuel 2019, 242, 744-755 with permission from Elsevier.
26
27
28
29 Reactions of CO2 and shale in the absence of water can also cause permeability changes.
30
31 Dissolution of organic content by CO2 can increase permeability. Reynolds showed that in cases
32
33 where the core samples contained high levels of organic matter, permeability increased after
34
35
36
soaking in pressurized CO2.167 Adsorption of CO2 in shale can decrease permeability. Al Ismail et
37
38 al. studied the effect of CO2 adsorption on Utica and Permian Shale samples. They found that the
39
40 amount of CO2 adsorbed on pore walls of shale was dependent on the amount of clay in the
41
42
sample,104 and that adsorbed CO2 reduced Eagle Ford shale permeability by an order of magnitude
43
44
45 perpendicular to the bedding plane but by only 10% parallel to the bedding plane.168 In situations
46
47 where CO2 storage is possible, reduction in vertical permeability may help prevent CO2 leakage.
48
49 168
50
51
52 Combining CO2 with aqueous solutions. Most hydraulic fracturing involves the use of
53
54 aqueous fracturing fluids. Data is sparse, however, for laboratory-scale EOR experiments
55
56
57
58
59
60 ACS Paragon Plus Environment
50
Page 51 of 170 Energy & Fuels

1
2
3 involving ultralow-permeability cores that have been immersed first in brine and then in oil during
4
5
6 core preparation. A report on such a treatment of a core showed that the presence of imbibed water
7
8 reduced oil production by more than 40% compared to cores not exposed to brine.136 Han et al.
9
10 studied water-alternating-gas (WAG) injection on shale cores with permeabilities in the range of
11
12
13
0.016-0.132 mD (Table 2, entry 15).118 Here, the incorporation of water in the extraction process
14
15 increased oil recovery. For example, CO2 alone gave a total recovery factor of 63%, while WAG
16
17 with a 0.125 pore-volume (PV) water slug recovered 93% of oil. The authors postulated that WAG
18
19
injection controls CO2 mobility and delays CO2 breakthrough.
20
21
22 CO2-rock-brine contact angle studies. Several research groups have demonstrated that
23
24 exposure to CO2 changes the wettability of shale and tight sandstone from oil wet toward water
25
26 wet. For example, the interactions of reservoir rocks, brine and supercritical CO2 was investigated
27
28
29 for a tight sandstone reservoir in Lucaogou Formation (China).8 Upon exposure to CO2, the
30
31 wettability of the tight sandstone sample became more water-wet. This observation was attributed
32
33 to mineral dissolution, kaolinite formation, and surface corrosion. In another study, the contact
34
35
36
angle of oil on a Three Forks shale core was measured, and the shale surface was oil wet (Figure
37
38 16A).76 The core was then immersed in CO2 at 17.2 MPa for two days. The contact angle of the
39
40 surface was measured in water again and the oil droplet beaded up, indicating that the exposure to
41
42
CO2 had made the core surface more water-wet (Figure 16B). Finally, in a contact angle study
43
44
45 using carbonate rock samples showed that contact angles decreased dramatically from 134° in the
46
47 presence of seawater, to 36° in the presence of a mixture of seawater and CO2.169 This change was
48
49 attributed to H+ adsorption on the interface of oil/brine and brine/carbonate due to CO2 dissolution.
50
51
52 The ability of CO2 to change the surface of shale from oil wet to water wet could be a key
53
54 mechanism in increasing oil recovery by CO2 injection.
55
56
57
58
59
60 ACS Paragon Plus Environment
51
Energy & Fuels Page 52 of 170

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 16. Apparent increase in water-wet character of a core sample after 2 days of exposure to
17 high pressure CO2. (Data from Alharthy et al. 2015). 76
18
19
20 CO2 huff-n-puff experiments using microfluidics. Microfluidics offer the ability to measure
21
22
23 fluid properties with direct visualization of relevant mechanisms at the pore scale. Nguyen et al.
24
25 at Los Alamos National Laboratory studied CO2, nitrogen, and water EOR using microfluidics
26
27 experiments.170 Experiments were performed using glass micromodels containing both dead-end
28
29
30
and connected fracture networks, and decane was used as a model oil. The microfluidics chips
31
32 were pressurized to 10 MPa at 50 °C, the system allowed to equilibrate, and then depressurized,
33
34 to simulate a huff-n-puff injection. The process was observed by taking fluorescence microscope
35
36
images throughout the experiments. During depressurization (i.e., the “puff” phase), oil production
37
38
39 was driven by gas exsolution from the oil liquid phase, which involved bubble nucleation, growth
40
41 and coalescence. CO2 showed the highest oil recovery—with 90% of oil in the connected fracture
42
43 network and 60% of oil in the dead-end fracture network being recovered. Nitrogen recovered
44
45
46 40% of oil in the connected fracture network and 25% in the dead-end fracture network. Water did
47
48 not recover any oil. The authors concluded that higher gas solubility in oil enables more bubble
49
50 nucleation and expansion during depressurization, which promote recovery.
51
52
53
The role of nanoconfinement was explored by Zhong et al. at the University of Toronto
54
55 using nanofluidics experiments.171 Nanofluidic devices with mean pore sizes (60 nm), permeability
56
57
58
59
60 ACS Paragon Plus Environment
52
Page 53 of 170 Energy & Fuels

1
2
3 (10 μD) and porosity (14%) were fabricated to mimick the properties of ULRs. Huff-n-puff
4
5
6 injections were performed under both miscible conditions using CO2 and immiscible conditions
7
8 using nitrogen. Devices were saturated with light Texas crude oil, excess oil cleaned out with air,
9
10 and subsequently injected with gas (CO2 or nitrogen). Tests were run at 323 K at a range of
11
12
13
pressures from 5-11 MPa. The high capillary pressure in the nanoconfined system decreased gas
14
15 solubility, reducing the formation of bubbles during depressurization which drive oil recovery.
16
17 The authors concluded that during huff-n-puff injections it is important to allow sufficient time for
18
19
CO2 diffusion during injection. During production, however, fast pressure release allows more gas
20
21
22 breakout and facilitates oil recovery.
23
24 Nano- and microfluidics experiments offer several advantages for studying EOR in ULRs.
25
26 First, in glass microfluidics devices, the mechanisms of oil recovery can be directly observed using
27
28
29 microscopes. Gas dissolution and bubble formation was observed in microfluidics experiments by
30
31 both Nguyen and Zhong.170, 171 Another advangage is the short duration of experiments. The small
32
33 size of the devices means that oil saturation and production can be achieved in a short amount of
34
35
36
time. For example, during huff-n-puff nanofluidics experiments conducted by Zhong, the soak
37
38 time lasted only one hour.171 Oil saturation and recovery from a shale core with similar lithology
39
40 (60 nm pores, 10 μD permeability) could take months. Although the short duration of experiments
41
42
is useful in making mechanistic observations, the high oil recovery observed in these experiments
43
44
45 is certainly optimistic when compared to recoveries obtained in the field.
46
47 CO2-in-brine foams for conformance control. Conformance control is a major challenge in
48
49 fractured heterogeneous shale formations. Therefore, many researchers have investigated CO2
50
51
52 foams as a strategy to improve conformance control. Zhang et al. from China University of
53
54 Petroleum studied the use of an anionic surfactant to stabilize CO2 foams in the presence of
55
56
57
58
59
60 ACS Paragon Plus Environment
53
Energy & Fuels Page 54 of 170

1
2
3 ethanol.172 The solubility of sodium bis(2-ethylhexyl) sulfosuccinate surfactant in supercritical
4
5
6 CO2 was studied. Addition of ethanol was required to achieve a meaningful solubility of the
7
8 surfactant in CO2. Dissolution of surfactant in CO2 with ethanol led to a three-fold increase in CO2
9
10 viscosity; and dissolution of the surfactant in CO2—rather than water—significantly reduced the
11
12
13
adsorption of the surfactant on the rock. The injectivity of the CO2-ethanol-surfactant solutions
14
15 was tested in cores ranging in permeability from 0.62-0.80 mD. The injectivity of the CO2-ethanol-
16
17 surfactant solution was much higher than that of the water-surfactant solution.
18
19
Tamayo et al. reviewed various practices of using CO2 foams to mitigate clay swelling,
20
21
22 capillarity effects, near-wellbore restrictions, and formation complexity and heterogeneities.54-55
23
24 They recommended practices based on the experience gained after performing 192 fracturing jobs
25
26 in Cadomin Formation, Wild River field in Canada. Based on their review, some researchers had
27
28
29 combined traditionally incompatible fluids to create effective fracturing fluids. Others combined
30
31 borate-crosslinked fluids with CO2 foam as the fracturing fluid to maximize gas recovery while
32
33 maintaining optimal distribution of proppant. Others used surfactants to prepare chemical
34
35
36
packages that resolve water phase trapping issues in the tight gas formation.
37
38 CO2 diffusion coefficients. The rate at which CO2 diffuses in oil is an important parameter
39
40 for CO2 EOR which is often uncertain. Some researchers state that values for this coefficient can
41
42
vary from 1 × 10-6 cm2/s to 1× 10-3 cm2/s at reservoir conditions.173 Precise values for high-pressure
43
44
45 diffusion coefficients are challenging to obtain experimentally. Diffusion coefficient values can
46
47 be obtained using several techniques, including monitoring the weight of methane added to a
48
49 hydrocarbon liquid phase as a function of the square root of time, 174 and a dynamic pendant drop
50
51
52 analysis.100 Studies of diffusion coefficients over a broad range of pressures are somewhat rare.
53
54 Diffusion coefficients were recently presented for CO2 in eight hydrocarbons at temperatures
55
56
57
58
59
60 ACS Paragon Plus Environment
54
Page 55 of 170 Energy & Fuels

1
2
3 between 298 K and 423 K and pressures to 69 MPa.175 These researchers employed a Taylor
4
5
6 dispersion experimental technique that exploits the effects of axial dispersion due to the laminar
7
8 flow velocity profile and radial dispersion due to diffusion on CO2 in laminar flow. In general,
9
10 diffusion coefficients fell in the 5× 10-6 cm2/s to 2 × 10-4 cm2/s range, with the diffusion coefficient
11
12
13
increasing with increasing pressure and decreasing with increasing pressure. Further, this team
14
15 correlated the diffusion coefficient results with a novel rough hard sphere-based theory.
16
17 The carbon storage potential associated with CO2 EOR in ULRs. The design, planning,
18
19
operation and economic assessment of a CO2 EOR pilot projects may not take into account an
20
21
22 important environmental aspect of this technology—CO2 EOR in ULRs will provide a means of
23
24 permanently disposing of CO2. If the CO2 purchased for the EOR project is sourced from power
25
26 plants, gas processing facilities or chemical manufacturing facilities, then CO2 EOR will provide
27
28
29 a mode of carbon sequestration of anthropogenic CO2. This occurs because CO2 EOR in
30
31 unconventional formations will not result in the production of all of the CO2 that was injected.
32
33 Therefore, when the EOR project is complete and the well is shut in, the CO2 remaining in the
34
35
36
formation will be permanently stored in the fractured formation. When one considers the
37
38 economics of field-scale, long term CO2 EOR in ULRs, the revamped 45Q tax credit provides a
39
40 10-year ramp up to $35/ton for CO2 stored geologically through EOR.
41
42
The U.S. Department of Energy provided a methodology to estimate CO2 storage in
43
44
45 depleted shale formations.176-177 The methodology consists of a statistical tool and is primarily
46
47 directed for gas shale formations but can be applied to oil shales with modifications. Three criteria
48
49 are specific for CO2 storage in shale formations: (1) hydrocarbons are commercially produced
50
51
52 through artificially stimulated fracture networks, (2) the formation is deep enough to store CO2 in
53
54 a liquid or supercritical state, and (3) there is an appropriate seal in place to prevent migration of
55
56
57
58
59
60 ACS Paragon Plus Environment
55
Energy & Fuels Page 56 of 170

1
2
3 the CO2. The volume of high-pressure CO2 stored at reservoir conditions would be roughly of the
4
5
6 same order of magnitude as the combined volumes of incremental brine and incremental oil
7
8 produced as a result of CO2 injection during EOR in the ULR. A few other studies have attempted
9
10 to accurately quantify CO2 storage potential in shale during EOR. 29, 105, 110, 178-180 Sorensen et al.
11
12
13
applied the U.S. Department of Energy Methodology for Estimating CO2 Storage Capacity
14
15 demonstrated in the Carbon Sequestration Atlas of the United States and Canada, 2007, and
16
17 concluded that, depending on the scenario employed in the analysis, the CO2 storage potential
18
19
could range from 0.12 to 3.16 Gtons of CO2.145
20
21
22 3.4 Laboratory experiments related to methane and natural gas EOR in ULRs. The
23
24 mechanisms for oil recovery using natural gas are analogous to those observed with CO2. The
25
26 contributions of four different recovery mechanisms were recently evaluated. Oil swelling plays
27
28
29 a more prominent role in oils with a low GOR, while vaporization is a more important mechanism
30
31 for recovering oil from gas condensate formations with extremely high GOR.81 When natural gas
32
33 is studied as an injection fluid for EOR in ULR, laboratory investigations of the efficacy of natural
34
35
36
gas for EOR typically provide a composition of the natural gas (Table 3). The natural gas may be
37
38 a model fluid, pure methane, or as a more realistic mixture of methane and one or more light
39
40 alkanes (ethane, propane, butane, pentane).91, 181-182 Typically, the composition of the natural gas
41
42
is classified as either lean (e.g. CH4 + ~10% C2+) or rich (CH4 + ~35% C2+) in ethane and heavier
43
44
45 gases (C2+, natural gas liquids, NGL).
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
56
Page 57 of 170 Energy & Fuels

1
2
3 Table 3. Immersed core huff-n-puff experiments with methane.
4
5
Very Tight, 0.01-0.001 mD Extremely tight, 0.001-0.0001 mD
6
7
8 entry Core Reservoir core permeability core core Injection temp recovery reference
9 prep (mD) length diameter pressure (°C) factor
a
(cm) (cm) (MPa) (% OOIP)
10 1 C Middle Bakken 0.002-0.04 3-4 1.1 34.5 110 92 Alharthy et al.76
11 2b C Middle Bakken 0.002-0.04 3-4 1.1 34.5 110 95 Alharthy et al.76
12 3 B Wolfcamp n.r. 5.1 3.8-10.2 13.8 35 9-39 Li et al.183
13 4 B Wolfcamp n.r. 3.8-5.1 2.5-10.2 13.8 35 9-50 Li et al.184
14 5 B Wolfcamp 0.0003-0.0005 5.1 3.8 13.8 40 10-40 Li et al.137
6 B Eagle Ford n.r. 5.1 3.8 10.3-15.2 r.t. 46-53 Meng et al.185
15
7 B Wolfcamp n.r. 7.6 3.8 10.1-15.2 r.t. 70 Meng et al.186
16 8 B Eagle Ford 0.0001 10.2 3.8 10.1-13.1 20 19-25 Meng et al.187
17 9 C Upper, Lower 0.00075, 0.00525 3-4 1.1 34.5 110 18, 24 Jin et al.143
18 Bakken
19 10 b C Lower Bakken 0.00525 3-4 1.1 34.5 110 27 Jin et al.143
a b
20 Core preparation procedure, as described in Figure 10. CH4 (85%), ethane (15%)
21
22
23 Effect of huff-n-puff parameters on natural gas EOR. Li et al. conducted an experimental
24
25
26
natural gas huff-n-puff study (Table 3, entry 3).183 The results were then compared to a
27
28 composition model to upscale methane huff-n-puff from laboratory to field. Wolfcamp plugs
29
30 saturated with Wolfcamp crude oil were used in a laboratory study to optimize the duration of the
31
32
gas injection rate, production time and production rate, soaking time, and amount of gas injected.
33
34
35 As was observed in CO2 extraction experiments, the amount of oil recovered by individual cycles
36
37 was highest during the first cycle and decreased with subsequent cycles. They also investigated
38
39 how the core size effects oil recovery (Table 3, entry 4).184 Using an injection pressure of 13.8
40
41
42 MPa, oil recoveries diminished as the size of the core increased—likely due to the lower surface
43
44 area-to-volume ratio in larger cores. The same researchers compared the oil extraction
45
46 performance of methane that of CO2 and nitrogen, and methane extracted the lowest amount of oil
47
48
49
(Figure 11D, Section 3.2, Table 3, entry 5).137 Meng et al. studied methane EOR using Eagle Ford
50
51 cores (Table 3, entry 6).185 Recovery was measured by NMR. Condensate recovery increased with
52
53 increasing number of cycles, and soak time did not have a large effect on recovery. In a subsequent
54
55
experiment using Wolfcamp cores, recovery by natural gas huff-n-puff was increased to 70%
56
57
58
59
60 ACS Paragon Plus Environment
57
Energy & Fuels Page 58 of 170

1
2
3 (Table 3, entry 7).186 Laboratory experiments using huff-n-puff resulted in condensate recovery of
4
5
6 25% on an Eagle Ford outcrop core. In these experiments, recovery was measured by CT scan.
7
8 Additionally, the effect of injection mode was investigated (Table 3, entry 8).187 Oil recoveries
9
10 were measured from an both immersed core huff-n-puff and confined-core multiple-well cyclic
11
12
13
injections. The huff-n-puff injection afforded higher oil recovery (25%) than the multiple-well
14
15 cyclic injection (19%).
16
17 Several experiments performed at the EERC compared the oil recovery performance of
18
19
various gases, including methane and natural gas mixtures (Table 3, entries 1, 2, 9 and 10). These
20
21
22 experiments are discussed in Section 3.2. In general, natural gas mixtures (15% ethane/ 85%
23
24 methane) gave higher recoveries than methane alone. Compared to the large number of CO2 EOR
25
26 experiments in the literature, relatively few core flooding experiments to study natural gas EOR in
27
28
29 ULRs have been reported. This difference is likely due—at least in part—to the safety concerns
30
31 associated with using high-pressure flammable gases in a laboratory setting. The smaller number
32
33 of natural gas experiments should not discourage the use of natural gas for EOR in ULRs in the
34
35
36
field. The same mechanisms and trends regarding soak times, injection pressures, and core surface
37
38 area observed during CO2 EOR experiments are applicable in natural gas EOR in ULRs.
39
40 Natural gas foams for conformance control during EOR in ULR. A hydrocarbon in brine
41
42
foam was developed as a means to improve conformance control during gas injection.188 Foam-
43
44
45 stabilizing surfactants were screened for their ability to generate foams, their static adsorption on
46
47 reservoir rock and emulsion tendency. In the field, a surfactant solution in brine was injected for
48
49 five weeks in a fractured tight reservoir along with hydrocarbon gas to form foam in situ. Foam
50
51
52 generation and improved mobility control were observed. Increased oil production continued for
53
54 six weeks after surfactant injection, with more than 2000 bbl of incremental oil recovered. 188
55
56
57
58
59
60 ACS Paragon Plus Environment
58
Page 59 of 170 Energy & Fuels

1
2
3 3.5 Laboratory experiments related to ethane and propane EOR in ULRs. Natural gas
4
5
6 liquids (NGLs), most notably ethane and propane, are excellent low molecular weight, low
7
8 viscosity solvents for crude oil found in conventional and unconventional formations. Ethane’s
9
10 properties related to oil recovery are clearly superior to those of methane and nitrogen and are
11
12
13
generally considered to be slightly better than those of CO2 (Table 1, Section 2.0). Further, the
14
15 diffusion coefficients and swelling factors for propane and ethane are high, which is beneficial for
16
17 EOR in conventional formations and ULRs.100
18
19
Most studies involving ethane or propane have been conducted during the analysis of
20
21
22 various gases for EOR in ULRs. A significant concentration of ethane and other C2+ components
23
24 in natural gas will significantly enhance its EOR performance. Therefore, studies of ethane and
25
26 propane have been conducted to demonstrate why these fluids have such a favorable effect on
27
28
29 natural gas solvent strength. For example, in a comparison of ethane, CO2, methane, ethane and
30
31 nitrogen, Jin et al. at University of North Dakota observed that ethane gave the best results—
32
33 recovering ~100% of the oil in about 6 h (Figure 11B, Section 3.2).143 It was also recently
34
35
36
demonstrated that pure propane exhibits a much lower MMP for a typical Bakken crude oil (3.8
37
38 MPa) than pure ethane (9.26 MPa), with both of thee NGLs being substantially more miscible with
39
40 Bakken crude than methane (MMP 31.1).141 Therefore natural gases that have high concentrations
41
42
of NGLs become increasingly effective solvents, and one can enrich natural gas streams with
43
44
45 ethane to improve oil.116 To the best of our knowledge, there have not been reports of studies
46
47 utilizing pure butane, but it would be expected that butane would have an even lower MMP than
48
49 propane and its presence in natural gas would be beneficial for EOR.
50
51
52 Ethane foams for conformance control during EOR in ULR A high-pressure ethane-in-
53
54 water foams have been assessed on the laboratory scale for improved gas conformance and high
55
56
57
58
59
60 ACS Paragon Plus Environment
59
Energy & Fuels Page 60 of 170

1
2
3 oil recovery in low-permeability (<15 mD), harsh environments (74 °C and 136,000 ppm TDS
4
5
6 brine).115 EOR tests were conducted in cores using gravity-unstable miscible ethane foam, gravity-
7
8 stable miscible ethane (no foam), and gravity-unstable miscible ethane. Oil recovery was 98%,
9
10 62%, and 43% OOIP, respectively. These promising results indicated that miscible ethane injection
11
12
13
can yield high oil recovery even for a gravity-unstable conditions, with the addition of foam further
14
15 enhancing overall recovery.
16
17 3.6 Laboratory experiments related to nitrogen EOR in ULRs. Nitrogen has been
18
19
employed for EOR in conventional formations due to its low cost and its benefit of being non-
20
21
22 corrosive. Several studies exploring the feasibility of nitrogen EOR in ULRs have been conducted.
23
24 In 2013, Gamadi et al. determined oil recovery by nitrogen flooding using Marcos, Barnett and
25
26 Eagle Ford shale cores (Table 4, entry 1).189 Mineral oil was used to saturate cores and recoveries
27
28
29 were measured at pressures of 6.9, 13.8, 20.7 and 24.1 MPa, with 1-day shut in periods. Highest
30
31 oil production was observed in the first cycle and plateaued at the seventh cycle. Increasing the
32
33 pressure improved oil recovery more than increasing the shut-in periods (from 1 to 3 days).
34
35
36
Maximum recoveries were between 33 and 72%.
37
38
39
40 Table 4. Immersed core huff-n-puff experiments with nitrogen.
41
42 Extremely tight, 0.001-0.0001 mD
43
44 entry Reservoir Core permeability (mD) core core Injection temp recovery reference
45 prep length diameter pressure (°C) factor (%
a
(cm) (cm) (MPa) OOP)
46 1 Barnett, Mancos, B n.r. 5.1 3.8 6.9-24.1 35 33-72 Gamadi et al.189
47 Eagle Ford
48 2 Eagle Ford B n.r. 5.1 3.8 6.9 40 24-51 Yu et al. 190
49 3 Eagle Ford B 0.0003-0.0005 10.2 3.8 6.9, 34.5 40 12-26 Yu et al.191
50 4 Eagle Ford B 0.00008 5.1 3.8 6.9-34.5 22 8-40 Yu et al. 192
5 Eagle Ford B 0.00032-0.00039 5.1 3.8 6.9-34.5 22 8-25 Yu et al.193
51
6 Wolfcamp B 0.0003-0.0005 5.1 3.8 13.8 40 33, 48 Li et al. 137
52 7 Bakken C 0.00525-0.1035 4 1.1 34.5 110 8-26 Jin et al.143
53 8 Wolfcamp B n.r. n.r. 2.5 21.4-34.5 74 0 Tovar et al.78
54 a
Core preparation procedure, as described in Figure 10.
55
56
57
58
59
60 ACS Paragon Plus Environment
60
Page 61 of 170 Energy & Fuels

1
2
3 Yu et al. at Texas Technical University investigated the use of nitrogen injection for EOR
4
5
6 in ULRs. Shale cores with average porosity of 8.5% were cleaned and saturated with mineral oil
7
8 at 6.9 MPa, and the unconfined cores were subjected to nitrogen huff-n-puff for ten cycles (Table
9
10 4, entry 2).190 After each soak period, the pressure was allowed to drop at a set rate, and the cores
11
12
13
were removed from the holders and weighed to determine oil recovery. Recovery rates were
14
15 improved by faster pressure depletion rates and longer soaking times (but not more than 72 hours).
16
17 Subsequent experiments by Yu et al. employed Eagle Ford outcrop cores and shale oil instead of
18
19
mineral oil (Table 4, entry 3).191, 194 Cores with 9.7% porosity and between 300 and 500 nD
20
21
22 permeabilities were soaked in oil, then ten nitrogen huff-n-puff experiments were performed with
23
24 variable soaking time, injection pressure and pressure depletion time. Longer soaking times up to
25
26 12 hours improved recovery, but beyond 12 hours, no improvement in recovery was observed.
27
28
29 Increasing the number of huff-n-puff cycles and pressure depletion rate improved recovery. In
30
31 conjunction with this experimental study, numerical simulations were also performed and history-
32
33 matched to the results. Next, Yu et al. performed immiscible nitrogen flooding experiments on
34
35
36
core samples with permeabilities of 80 nD at three different pressures: 6.9, 20.7 and 34.5 MPa
37
38 (Table 4, entry 4).192 With an initial injection pressure of 34.5 MPa, a maximum recovery factor
39
40 of 40% was observed after 2 days. The results of 32 nitrogen flooding experiments were analyzed
41
42
extrapolated to a 300-day time point, where oil recovery of 80.5 to 85% were calculated. In 2017,
43
44
45 Yu et al. compared brine huff-n-puff with nitrogen huff-n-puff extraction (Table 4, entry 5).193
46
47 Nitrogen performed better than water under the same operating conditions. After 12 cycles with
48
49 various soak times, nitrogen afforded about 20% oil recovery while water afforded only about
50
51
52 10%.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
61
Energy & Fuels Page 62 of 170

1
2
3 Several research groups have conducted studies to compare the oil recovery abilities of
4
5
6 different gases, including nitrogen. In 2017, Li et al. compared the oil recovery abilities of
7
8 methane, CO2 and nitrogen using cores with permeabilities between 300 and 500 nD (Figure 11D,
9
10 Section 3.2, Table 4, entry 6).137 In this study, nitrogen performed better than methane but worse
11
12
13
than CO2. Jin et al. found that nitrogen injection provided lower recoveries than other gases—8%
14
15 in a Lower Bakken core and 26% in a Middle Bakken core (Figure 11B and C, Section 3.2, Table
16
17 4, entry 7).143 Tovar et al. (2018) at Texas A&M University flooded shale core with nitrogen, using
18
19
a huff-n-puff injection with a soak time of 21 hours at 27.6 MPa for 21 hours (Table 4, entry 8).78
20
21
22 Although the experimental pressure was 6.9 MPa above the oil-nitrogen MMP, no oil was
23
24 extracted. Even when the injection pressure was increased to 34.5 MPa, no oil was extracted.
25
26 In summary, several studies of nitrogen extraction from shale cores have been reported
27
28
29 with promising results.137, 189 190-193
When nitrogen is compared to methane and CO2 under the
30
31 same extraction conditions however, nitrogen performs the worst.78, 143 Based on the poor oil
32
33 recovery results commonly observed in experimental tests and the nitrogen contamination issues
34
35
36
associated with nitrogen injection, this fluid is a less attractive choice for EOR in ULRs compared
37
38 to CO2 and natural gas.
39
40 3.7 Conclusions from high-pressure gas laboratory experiments. In huff-n-puff
41
42
experiments involving shale cores, oil recovery increases as the miscibility of the oil and the gas
43
44
45 increases. Roughly speaking, CO2 and ethane are comparable solvents for recovering oil, with
46
47 ethane typically exhibiting greater miscibility with oil than CO2 at the same conditions. CO2 is a
48
49 much stronger solvent than either methane or nitrogen. Relatively few studies are conducted with
50
51
52 ethane as the EOR solvent because in most locations, CO2 and natural gas are more readily
53
54 available.115 However, natural gas with a significant C2+ content is a much better EOR fluid than
55
56
57
58
59
60 ACS Paragon Plus Environment
62
Page 63 of 170 Energy & Fuels

1
2
3 pure methane and can exhibit oil recovery and MMP values comparable to those of CO2.
4
5
6 Recoveries during huff-n-puff laboratory experiments are improved by longer soak times and
7
8 increased number of cycles—although increasing the number of cycles appears to be the more
9
10 important parameter. Higher pressures also improve recoveries, even beyond the oil-gas MMP.
11
12
13
Although recoveries in laboratory-scale experiments involving extremely low-permeability cores
14
15 can be very high (approaching 100% oil recovery), these experiments are likely overly optimistic
16
17 due to the high surface area exposed to CO2 (or other fluid), relative to the volume of the core. As
18
19
will be discussed in Section 6, these high recoveries in the laboratory do not translate to the field.
20
21
22 4.0 Simulation studies related to gas-based EOR in ULRs. Many modeling studies have
23
24 focused on miscible gas EOR in ULRs. These models account for both shale and tight sandstone
25
26 or siltstone formations, various injection strategies (continuous injection, huff-n-puff, or WAG),
27
28
29 and injection fluids (CO2, natural gas, or nitrogen). The reader should keep in mind that each
30
31 simulation model has its own assumptions, grid dimensions, fluid and formation properties and
32
33 production profiles. This review cannot provide accurate, generalized conclusions based upon this
34
35
36
large body of modeling studies because that would require history matching of past production
37
38 profiles and/or other forms of fair and appropriate comparison. Therefore, the following summary
39
40 is primarily intended to inform the reader of ongoing simulation efforts without rendering
41
42
definitive conclusions concerning lessons learned or comparisons/rankings of these models.
43
44
45 In general, the transport expressions for fluid flow in the high-permeability, hydraulic
46
47 fractures are coupled with equations governing diffusion-based mass transport mechanisms within
48
49 the natural fractures and shale matrix to accurately account for as many of the mechanisms of oil
50
51
52 recovery as possible. Some researchers employed equations of state within compositional transport
53
54 models to compensate for the limited phase behavior capabilities of commercial simulation
55
56
57
58
59
60 ACS Paragon Plus Environment
63
Energy & Fuels Page 64 of 170

1
2
3 software developed for unconventional reservoirs. Others settled with the built-in governing flow
4
5
6 equations of those simulators and hoped to tune various parameters to get a promising match with
7
8 the historical data. Oftentimes, these models have been tuned to match laboratory-scale and pilot-
9
10 scale EOR. Sensitivity analyses have also been performed on a wide variety of operating
11
12
13
parameters, such as fracture-matrix characteristics and parameters associated with huff-n-puff
14
15 (duration of injection, pressure, soaking period, number and length of cycles). Table 5 lists the
16
17 simulation studies of EOR in ULRs grouped by organization. The table also indicates whether the
18
19
model was applied to laboratory-scale or field-scale scenarios. Modeling studies can be performed
20
21
22 using either single-porosity or dual-porosity models. A single-porosity model assumes that natural
23
24 fractures are closed and flow is assumed only originates from the matrix to the wellbore. A dual-
25
26 porosity model is more complex, and considers both matrix and fracture to be contributing to the
27
28
29 flow. The porosity model used in each simulation (single, dual or both) is included in Table 5. The
30
31 injection gas and mode of injection are provided. Hydraulic fractures were modelled in these
32
33 studies, and the table also provides an indication as to whether natural fractures were also modeled.
34
35
36
The table also indicates whether diffusion-based mechanisms were incorporated, and the specific
37
38 simulation tool that was employed.”
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
64
Page 65 of 170 Energy & Fuels

1
2
3 Table 5. Simulation studies related to high-pressure gas EOR in ULRs.
4
5 Entry Organization Lab/ Reservoir Fluid EOR Method Natu- Single Diffus Simul- Reference
6 Field a
ral /Dual -ion? ation Tool
7 Frac- poro-
8 tures? sity
1 EERC Field Bakken CO2 Cont. No - No CMG Liu et al.195
9 2 Field Bakken CO2 Cont. Yes Dual Yes CMG, Sorensen et al.145
10 SCS
11 3 Both Bakken CO2, HnP-Imm Yes Both Yes CMG Alharthy et al.76
12 NGL, N2 (L), HnP (F)
4 Both Bakken CO2, HnP-Imm Yes Both Yes CMG Alharthy et al.18
13 NGL, N2 (L)., HnP (F)
14 5 Both Bakken CO2 HnP-Imm Yes Both Yes CMG Torres et al.196
15 (L), HnP (F)
6 Texas Tech Field Eagle Nat gas, Cont. No - No CMG Sheng et al. 12
16 University Ford H2O
17 7 Field Eagle Nat gas HnP No - No ECLIPSE Gamadi et al.197
18 Ford
19 8 Field Eagle Nat gas HnP No - No CMG Wan et al. 198
Ford
20 9 Field Eagle Nat gas HnP Yes No CMG Wan et al.199
21 Ford
-
22 10 Field Eagle Nat gas, HnP Yes Dual Yes CMG Wan et al. 198
23 Ford CO2
11 Field Eagle CO2 Cont. Yes Dual Yes CMG Wan et al.200
24 Ford
25 12 Lab Wolfcamp CH4 HnP-Cnf No - No CMG Li et al. 183
26 13 Both Wolfcamp CH4 HnP No - No CMG Li et al.140
27 14 Both Wolfcamp CO2, HnP Yes Dual No CMG Li et al.137
NGL, N2
28 15 Lab Eagle CO2 HnP-Cnf No - No CMG Shen et al.135
29 Ford
30 16 University of Field Bakken CO2 HnP No - Yes UT- Chen et al.61
Texas at COMP
31 17 Austin Field Bakken CO2 HnP Yes Single Yes CMG Sanchez-Rivera et
32 al.201
33 18 Field Bakken CO2, nat Cont. No - Yes CMG Zhu et al.202
34 gas
19 Field Bakken CO2 Cont, HnP Yes Single Yes CMG Zuloaga-Molero et
35 al.203
36 20 Field Bakken CO2 HnP Yes Single Yes CMG Zhang et al.96
37 21 Field Bakken CO2 HnP No - No CMG Sanaei et al.204
38 22 Field Bakken CO2 HnP Yes Dual Yes CMG Wang et al.173
23 Texas A&M Field Bakken CO2 Cont. No - Yes N/A Fai-Yengo et al.205
39 24 University Field N/A CO2 HnP Yes Dual Yes CMG Sun et al.206
40 25 Both Bakken CO2 HnP-Cnf (L), Yes Single Yes CMG Sun et al.207
41 HnP (F)
42 26 Field Eagle CO2 Cont, HnP, Yes Dual No CMG Phi et al.208
Ford WAG
43
44 27 Field Eagle CO2, nat Cont, HnP Yes Dual Yes CMG Iino et al.209
45 Ford gas
28 Field Eagle CO2 HnP Yes Single Yes CMG Yu et al.210
46 Ford
47 29 Field Bakken CO2 HnP No - Yes CMG Yu et al.211
48 30 Colorado Field Bakken CO2, nat Cont Yes Single No CMG Hoffman et al.60
School of gas
49
31 Mines Field Bakken CO2 Cont Yes Single No CMG Xu et al.212
50
b
51 32 Field Bakken, CO2, Cont No - No N/A Teklu et al.213
Monterey, C2H6,
52 Eagle CH4, N2
53 Ford,
54 Niobrara
33 Field Bakken CO2, nat Cont, WAG No - No CMG Pu et al.214
55 gas, CH4
56 34 Field Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.14
57
58
59
60 ACS Paragon Plus Environment
65
Energy & Fuels Page 66 of 170

1
2
3 35 Missouri Field Bakken CO2, HnP Yes Dual Yes CMG Alfarge et al.215
4 University of lean gas,
Science and rich gas
5
36 Technology Field Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.216
6 37 Field Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.217
7 38 Field Bakken CO2 Cont. No - No CMG Alfarge et al.218
8 39 Field Bakken CO2 HnP No - Yes CMG Alfarge et al. 219
9 40 Field Bakken CO2, nat HnP, Cont. Yes Both Yes CMG Alfarge et al. 220
gas
10 41 Field Bakken CO2, HnP Yes Both Yes CMG Alfarge et al.221
11 lean gas,
12 rich gas
42 Field Bakken CO2, nat HnP, Cont. No - Yes CMG Alfarge et al.222
13 gas
14 43 Both Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.223
15 44 Field Bakken CO2, HnP Yes Dual Yes CMG Alfarge et al.224
16 lean gas,
rich gas
17 45 Both Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.224
18 46 Field Bakken CO2 HnP Yes Dual Yes CMG Alfarge et al.225
19 47 China Field Bakken CO2 Cont. No - Yes ECLIPSE Dong et al.226
20 48 University of Field Eagle CO2 HnP No - No CMG Wan et al. 227
Petroleum Ford
21 49 Both Jiangsu CO2 HnP-cnf (L), No - No CMG Wang et al.228
22 HnP (F)
23 50 Field N/A CO2, HnP No - No CMG Cao et al.229
H2O,
24
nano-
25 fluid,
26 WAG
27 51 University of Field Bakken CO2 Cont Yes Single No CMG Kalra et al.179
52 Oklahoma Field Wolfcamp CO2 HnP Yes Dual No CMG Phan et al.230
28 53 Field Eagle CO2 HnP, Cont No Single Yes N/A Akita et al.231
29 Ford
30 54 University of Field Cardium CO2 HnP No - Yes CMG Kong et al.84
31 55 Calgary Field Bakken CO2 HnP No - Yes CMG Kanfar et al.232
56 Field Western CO2 HnP No - Yes CMG Haghshenaset al.233
32 Canada
33 57 Field Bakken CO2 HnP Yes Dual No CMG Hejazi et al.178
34 58 University of Field Forest CO2 HnP No - No N/Ab Mohammed-Singh
35 Regina Reserve et al.53
59 Lab Bakken CO2 HnP-Cnf No - No CMG Song et al.132
36 60 Both Bakken CO2 HnP No - No CMG Song et al.128
37 61 University of Field Chatanoo CO2 HnP Yes Dual Yes CMG Cudjoe et al.234
38 Kansas ga
39 62 Field Bakken CO2 HnP Yes Dual Yes CMG Jia et al.235
63 Stanford Lab Monterey CO2 HnP-Cnf No - Yes CMG Vega et al.58
40 64 University Lab Bakken CO2 Cont. Yes Single No CMG Zhang et al.120
41 65 InPetro Tech. Field Bakken CO2 Cont. No - No CMG Pu et al.236
42 66 Field Bakken CO2 HnP No - No CMG Pu et al.237
43 67 Occidental Field N/A CO2 HnP Yes Single Yes CMG Sahni et al.238
68 Virginia Tech Field Bakken CO2 HnP No - Yes CMG Chen et al.239
44
69 Montana Tech Field Bakken CO2 Cont. Yes Single No CMG Shoaib et al.57
45
46 70 Schlumberger Field Eagle CO2 HnP Yes Single Yes ECLIPSE Pankaj et al.240
47 Ford
71 Nalco- Field Bakken, Surfac- HnP Yes Single No CMG Kazempour et al. 241
48 Champion Niobrara tant
49
50 a
51
Cont. = continuous, HnP-cnf = huff-n-puff, confined core, HnP-imm = huff-n-puff, immersed
52 core., (L) =laboratory, (F) = Field. b Simulation tool was not reported. c This is a simulation of a
53 chemical EOR surfactant injection.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
66
Page 67 of 170 Energy & Fuels

1
2
3 4.1 Simulations performed at the EERC. Researchers at the EERC modeled both field-
4
5
6 and laboratory-scale EOR experiments in the Bakken Formation.145, 195, 18, 76, 196 An early single-
7
8 porosity-single-permeability study by Liu—which did not account for the effect of natural
9
10 fractures and diffusivity—concluded that continuous CO2 injection in the field would lead to 43-
11
12
13
58% oil recovery from the matrix and hydraulic fractures (Table 5, entry 1).195 Dual-porosity-
14
15 dual-permeability simulations by Sorensen indicated that diffusion is a key mechanism for oil
16
17 recovery, and prolonged CO2 injection time leads to a substantial increase in oil production (Table
18
19
5, entry 2).145 In these simulations, oil recovery was increased by placing the producing well
20
21
22 between two injectors. Alharthy et al. reported that both CO2 and rich natural gas huff-n-puff
23
24 injections produced almost same amount of oil. Molecular diffusion and mass transfer across the
25
26 matrix-fracture interface were the main oil recovery mechanisms (Table 5, entries 3 and 4).18 They
27
28
29 also suggested that shorter fracture spacing would improve oil recovery by increasing the ratio of
30
31 formation surface area exposed to the gas in fractures to the volume of the formation. Higher values
32
33 of surface area/volume were more important than prolonging soaking times. Torres et al. found
34
35
36
the presence of natural fractures in CO2 huff-n-puff to be significant in increasing fluid flow (Table
37
38 5, entry 5).196
39
40 4.2 Simulations performed at Texas Technical University. Researchers at Texas
41
42
Technical University modeled the Wolfcamp and Eagle Ford Formations on laboratory-scale 135,
43
44 140, 183
45 and field-scale.12, 181-182, 197-198, 200 After 10 years of primary production and 60 years of
46
47 continuous gas injection from separate vertical injection and production wells, 15% oil recovery
48
49 resulted from the low-permeability homogeneous medium (Table 5, entry 6).9 These results were
50
51
52 superior to primary production alone (6.5% recovery) or primary production followed by
53
54 waterflooding (12%). A study related to natural gas huff-n-puff injection into a horizontal well in
55
56
57
58
59
60 ACS Paragon Plus Environment
67
Energy & Fuels Page 68 of 170

1
2
3 the Eagle Ford Formation revealed that oil recovery increased as pressure was elevated above the
4
5
6 MMP (Table 5, entry 7).197 Over the same total duration, greater oil recovery was obtained using
7
8 many short cycles as opposed to fewer long cycles. In one scenario where gross and net gas
9
10 utilization was considered for a lean natural gas near the MMP, the optimal soak time was about
11
12
13
two months.
14
15 Wan et al. modeled natural gas huff-n-puff in the Eagle Ford Formation (100 nD and 6%
16
17 porosity) (Table 5, entries 8 to 11).182 They modeled 20 years of primary production, followed by
18
19
1000 days of gas injection, followed by 60 cycles of 100 days injection and 100 days of production
20
21
22 with no soak time. This scenario resulted in 21% incremental oil recovery. Prolonged cycle time
23
24 of nearly three years yielded only 10% oil recovery. Closer fracture spacing and higher miscibility
25
26 increased recovery.182,199 Having the hydraulic fractures connected to the natural fracture network
27
28
29 yielded a 20% increase in oil recovery compared to simulations without natural fractures. In a
30
31 simulation of CO2 EOR, Wan et al. modeled the dispersive-convective flux through nanopores in
32
33 shale formations using a dual-permeability model coupled with diffusion (Table 5, entry 11).200
34
35
36
They confirmed that diffusion is critical for oil recovery, especially in the absence of convective
37
38 displacement. Fast injection rates were not beneficial because injection at a slow rate provided the
39
40 necessary time for diffusion in the shale matrix. Although conductive natural fractures enhanced
41
42
oil diffusion, they reduced conformance control. Wan et al. also modeled huff-n-puff processes
43
44
45 using CO2, lean natural gas and rich natural gas (Table 5, entry 10).198 Very good oil recovery
46
47 was attained with each gas when the formation was pressurized to the MMP for the specific gas
48
49 composition. C3+ components in natural gas enhanced miscibility.
50
51
52 Li et al. modeled laboratory-scale methane huff-n-puff laboratory experiments and scaled
53
54 results for methane huff-n-puff field cases (Table 5, entries 12 to 14).140, 183 They found that
55
56
57
58
59
60 ACS Paragon Plus Environment
68
Page 69 of 170 Energy & Fuels

1
2
3 increasing the injection time improved oil recovery because it increased the average pressure and
4
5
6 miscibility. Further, they recommended reducing the pressure as much as possible during the puff
7
8 cycle. Although oil recovery increased with soaking time during a cycle, more cycles within a
9
10 fixed duration with shorter times increased recovery more (Table 5, entry 12).183 Li et al. included
11
12
13
a pressure term that reflected the time-integrated average pressure difference between the huff and
14
15 the puff cycles—providing the user with an easy way to evaluate numerous parameters
16
17 simultaneously (Table 5, entry 13).140 Li et al. also demonstrated that most effective solvents for
18
19
oil recovery from strongest to weakest at similar conditions are ethane, CO2, methane and nitrogen.
20
21
22 However, they then demonstrated that natural gas mixtures enriched in ethane, propane and butane
23
24 exhibited superior solvent strength to CO2 and CO2-rich mixtures (Table 5, entry 14).137 Shen
25
26 modeled the experimental CO2 huff-n-puff with an oil that contained asphaltenes (Table 5, entry
27
28
29 15).135 Permeability reduction caused by asphaltene precipitation near the core surface was most
30
31 significant during the first cycle. Overall, a 49% loss of permeability was observed after six cycles.
32
33 4.3 Simulations performed at the University of Texas at Austin. At the University of
34
35
36
Texas at Austin, researchers employed UT-COMP, a non-isothermal, three-dimensional, equation-
37
38 of-state based compositional reservoir simulator, and found that reservoir heterogeneity reduced
39
40 production rate and increased soak times did not improve production (Table 5, entry 16).61
41
42
Production could be increased by increasing the number of hydraulic fractures to promote more
43
44
45 uniform CO2 transport into the matrix. Other researchers at UT Austin noted that—although CO2
46
47 injection suffers from poor economics—re-injecting mixtures of CO2 and produced hydrocarbon
48
49 gas could make huff-n-puff a profitable operation (Table 5, entry 17).201 Further, they noted that
50
51
52 CO2 huff-n-puff should not be initiated too soon during primary production because it could lower
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
69
Energy & Fuels Page 70 of 170

1
2
3 oil recovery; waiting too long would not diminish oil recovery but the delay would diminish the
4
5
6 net present value of the associated cash flow.
7
8 A novel injection scheme was modeled by Zhu (Table 5, entry 18).202 In this simulation,
9
10 CO2 (or natural gas) was continuously injected into a hydraulic fracture of a horizontal well and
11
12
13
produced from the adjacent fracture of the same well. In general, oil recovery from this fracture-
14
15 to-fracture technique was promising. Oil recovery increased with smaller fracture spacing,
16
17 injection pressures near the MMP, and more homogenous reservoirs. Rich natural gas produced
18
19
slightly more oil than CO2. This difference was attributed to the lower viscosity of natural gas
20
21
22 relative to CO2, which enhanced injectivity.202 For reservoirs with permeabilities less than 0.03
23
24 mD, CO2 huff-n-puff injection was more effective for oil recovery, whereas for higher permeability
25
26 reservoirs, continuous CO2 injection was favored (Table 5, entry 19). 203
The most important
27
28
29 parameters were matrix permeability, well pattern, and the interaction between fracture half-length
30
31 and the number of wells. Zhang et al. developed a model that accounted for nanopore confinement
32
33 effects on phase behavior (Table 5, entry 20).96 This model illuminated the roles of CO2 molecular
34
35
36
diffusion and capillary pressure. MMP was more accurately predicted when the effects of nano-
37
38 confinement were included.242 Sanaei et al. compared four gases (CO2, N2, methane, and an 85%
39
40 methane-15% ethane mixture) in a study of huff-n-puff in the Middle Bakken. Re-pressurization
41
42
and oil swelling (not diffusion) were identified as the dominant mechanisms of oil recovery (Table
43
44
45 5, entry 21).204 Because the injection pressure of 48 MPa was well above the MMP of all of these
46
47 fluids, they yielded nearly identical oil recovery results except for CO2, which recovered slightly
48
49 more oil. Given the high utilization ratio of CO2, and lower cost of injecting produced gas, it was
50
51
52 suggested that natural gas injection may yield more favorable economics. Further, they noted that
53
54 the number of cycles should be increased to optimize oil recovery. Gas injection rate should be
55
56
57
58
59
60 ACS Paragon Plus Environment
70
Page 71 of 170 Energy & Fuels

1
2
3 optimized to reach the MMP, while not being so high that significant operating costs are incurred
4
5
6 while inducing small increases in oil recovery. In 2019, Wang and Yu implemented diffusion,
7
8 natural fractures, huff-n-puff timing in a comprehensive simulation study (Table 5, entry 22).173
9
10 They found the role of natural fractures significantly important, and diffusion to play a relatively
11
12
13
small beneficial role.
14
15 4.4 Simulations performed at Texas A&M University. Simulation studies of EOR in
16
17 ULR have also been conducted at Texas A&M University. In a simple, single block matrix/fracture
18
19
model of continuous CO2 injection in the Bakken Formation, the combined mechanisms of
20
21
22 diffusion, vaporizing and condensing accounted for oil recovery (Table 5, entry 23).205 Rich gas
23
24 (50% CO, 25% C1, 25% C2) was more effective than CO2 in recovering oil.
25
26 Sun et al. scaled laboratory CO2 huff-n-puff results to the field-scale (Table 5, entries 24
27
28
29 and 25).206-207 Oil recovery was not sensitive to capillary pressure, diffusion coefficient, or soaking
30
31 time. Other factors, however, such as injection above the MMP during the “huff” period, oil
32
33 swelling, viscosity reduction, and gas expansion during the “puff” period, were vital to success at
34
35
36
the field-scale. The highest oil recovery was attained by starting huff-n-puff injections after
37
38 primary production at low pressures, and then later using high pressures during huff periods and
39
40 low pressures during puff periods.
41
42
Phi and Schechter simulated EOR in the Eagle Ford Formation (Table 5, entry 26).208 They
43
44
45 noted the importance of natural fractures that are perpendicular to the horizontal wellbore. CO2
46
47 huff-n-puff injection performed better than continuous CO2 injection and CO2-WAG. They
48
49 suggested a shut-in time of 1000 days to pressurize more of the formation before opening the
50
51
52 production well. Although there would be no production for a while, the subsequent production
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
71
Energy & Fuels Page 72 of 170

1
2
3 from higher reservoir pressure would be at such a fast rate and that greater cumulative recovery
4
5
6 would occur less than a year later.
7
8 In another study, Fast-Marching Method (FMM) flow simulation was used to optimize CO2
9
10 huff-n-puff injection (Table 5, entry 27).209 The FMM-based model gave good agreement with a
11
12
13
commercial flow dynamics simulator in a synthetic dual-porosity reservoir with multi-stage
14
15 hydraulic fractures while using less computing time. Oil recovery was optimized with 6-7 million
16
17 scf/day injection for 33-40 days, followed by a relatively short soak time of 8-12 days, followed
18
19
by production for 110-150 days.
20
21
22 In another study, CO2 molecular diffusion and oil swelling were identified as major EOR
23
24 mechanisms in the Bakken Formation (Table 5, entry 28).210 The most important parameters for
25
26 improved oil recovery were high injection rates, increased cycle times, greater number of cycles
27
28
29 and increased CO2 diffusivity.211 Molecular diffusion and nanopore confinement were important
30
31 mechanisms in predicting the performance of CO2 huff-n-puff in Eagle Ford (Table 5, entry 29).
32
33 211
For CO2 diffusion coefficients of 0.00001 – 0.0001 cm2/s, CO2 huff-n-puff would underperform
34
35
36
continued primary production. Comparable results would occur for diffusion coefficients of
37
38 0.0001–0.001 cm2/s, and better oil recovery would be obtained with CO2 huff-n-puff for diffusion
39
40 coefficients greater than 0.001 cm2/s.
41
42
4.5 Simulations performed at the Colorado School of Mines. At the Colorado School of
43
44
45 Mines, Hoffman modeled CO2 injection in the tight Middle Bakken (Table 5, entry 30).60 Due to
46
47 both pressure maintenance and solvent strength, the miscible injection scenarios of CO2 and
48
49 hydrocarbon gas yielded more oil recovery than immiscible scenarios. An equation-of-state
50
51
52 compositional model was used to test CO2, natural gas, and lean gas (pure methane) using
53
54 continuous and WAG injection schemes (Table 5, entry 33).214 CO2 WAG utilized less CO2 but
55
56
57
58
59
60 ACS Paragon Plus Environment
72
Page 73 of 170 Energy & Fuels

1
2
3 yielded much less incremental oil than continuous CO2 injection. A rich natural gas composed of
4
5
6 55% methane and 45% C2-C7 led to slightly better simulated oil recovery than CO2. The lean gas
7
8 yielded the lowest recovery due to its inability to attain miscibility at reservoir conditions.
9
10 A comparison of fracture orientations showed that slightly higher oil recovery and
11
12
13
production rates occurred with transverse fractures (i.e. fracture planes perpendicular to the
14
15 horizontal well) (Table 5, entry 31).206 The authors calculated the miscibility of a wide array of
16
17 gases, including CO2, N2, hydrogen sulfide, methane, ethane, propane, gas mixtures, LPG, and
18
19
NGL with the oil from Bakken, Monterey, Eagle Ford, and Niobrara shale reservoirs. CO2, ethane,
20
21
22 LPG, and NGL were effective EOR fluids due to their solvent strength and oil swelling capability
23
24 (Table 5, entry 32).213 In all cases, pore confinement enhanced the miscibility of gases with the
25
26 oil.
27
28
29 4.6 Simulations performed at Missouri University of Science and Technology.
30
31 Researchers at Missouri University of Science and Technology reported that the biggest
32
33 hinderances for CO2 huff-n-puff in the Bakken Formation were low total porosity and low intensity
34
35
36
of natural fractures (Table 5, entry 36).243, 216 They postulated that CO2 diffusion into shale pores
37
38 does not even occur in the field because oil production occurs faster than CO2 diffusion (Table 5,
39
40 entries 34, 37, 39, 40 and 45).14, 217, 219, 224, 220 The modest oil recovery results in the field were
41
42
best simulated with low CO2 diffusivity. Therefore, increased contact area and exposure time
43
44
45 would be required for a more successful project. Further, they suggested slowing oil production
46
47 during the puff portion of the cycle (Table 5, entry 44).224 Both the contact area and exposure time
48
49 of CO2 to the matrix are higher in the laboratory than in the field.223 For natural gas EOR, higher
50
51
52 porosity, natural fracture intensity, and diffusion rates were also beneficial (Table 5, entry 41).215,
53
221
54 However, the oil recovery performance of the natural gas was more sensitive to molecular
55
56
57
58
59
60 ACS Paragon Plus Environment
73
Energy & Fuels Page 74 of 170

1
2
3 diffusion than CO2 and increased more dramatically with increases in diffusion rate. These results
4
5
6 imply that natural gas is the preferred injection fluid when the permeability of the formation is less
7
8 than 1 mD because CO2 has higher molecular weight and needs larger pore throats to invade the
9
10 matrix.82 In simulations focused on well spacing for continuous CO2 injection into horizontal
11
12
13
wells, shorter producer-injector spacing improved oil recovery but diminished carbon storage
14
15 because more CO2 was produced relatively quickly (Table 5, entry 38).218
16
17 4.7 Simulations performed at China University of Petroleum. Researchers from China
18
19
University of Petroleum estimated that continuous CO2 injection in the Bakken Formation would
20
21
22 increase oil production by 5-21% (Table 5, entry 47).226 Additional horizontal injection wells
23
24 would benefit oil recovery. An interconnected fracture network must be present, and the fracture-
25
26 network spacing is more critical than the conductivity of the fractures (Table 5, entry 48).227 In a
27
28
29 comparison of various EOR techniques for ULRs, 3% incremental oil recovery was attained using
30
31 CO2 huff-n-puff, which was superior to water huff-n-puff, nanofluids huff-n-puff, and water-
32
33 alternating-gas (Table 5, entry 50).229 Low bottom-hole pressure during production and a higher
34
35
36
number of cycles were more beneficial than longer soaking time, injection and production time,
37
38 and injection rate.
39
40 4.8 Simulations performed at the University of Oklahoma. At the University of
41
42
Oklahoma, researchers developed a model for the injection of CO2 into fractured Middle Bakken
43
44
45 wells to produce oil and to store CO2 (Table 5, entry 51).179 Natural fractures were deemed critical,
46
47 along with a reduced distance between injector and producer. To improve the incremental oil
48
49 recovery by 1%, injection of about 3% hydrocarbon pore volume of CO2 would be required.
50
51
52 Another study focused on CO2 huff-n-puff in the Wolfcamp Formation (Table 5, entry 52).230 An
53
54 integrated modeling framework using complex fracture geometry and heterogeneous fracture
55
56
57
58
59
60 ACS Paragon Plus Environment
74
Page 75 of 170 Energy & Fuels

1
2
3 conductivity was used. They recommended 30 days of injection and 90 days of soaking for optimal
4
5
6 oil recovery. Numerous short cycles over a specified duration were more beneficial than a single
7
8 cycle. Akita et al. presented an equation to upscale experimental EOR results to field-scale to
9
10 account for all the unknown parameters in the field (Table 5, entry 53).231 They suggest that the
11
12
13
high laboratory oil recoveries are not observed in the field because of the low area of the matrix
14
15 exposed to EOR fluid in the field. The authors noted that oil recovery is dependent upon the Biot
16
17 number, which is the product of the matrix volume and fracture permeability divided by the
18
19
product of the fracture surface area and the total gas diffusion coefficient in the matrix. Biot
20
21
22 number represents the ratio of flow resistance outside the matrix to that inside the matrix, which
23
24 is very small in the laboratory but large in the field. They also found that, while pressure changes
25
26 occur quickly in the laboratory, it may take on the order of one year for pressure waves to propagate
27
28
29 to the field-scale system boundary.
30
31 4.9 Simulations performed at the University of Calgary. Two studies from the
32
33 University of Calgary focused CO2 huff-n-puff in ULRs suggested that the duration of injection
34
35
36
and production steps of each successive cycle should be prolonged (Table 5, entries 54 and 55).84,
37
232
38 Successive cycles of increasing duration resulted in more oil recovery using less CO2 than
39
40 cycles of fixed duration because the longer duration of the later huffs allowed the CO2 to access
41
42
oil that was further from the injector. Further, they suggested that a multi-well asynchronous CO2
43
44
45 huff-n-puff (i.e. a central well that is huffing while two parallel side wells are puffing) can increase
46
47 oil recovery. 84
48
49 The profitability of CO2 huff-n-puff in the Bakken was optimized using an algorithm to
50
51
52 maximize net present value (NPV) (Table 5, entry 56).232 Primary production should continue until
53
54 production rates decline and boundary-dominated flow is attained. For example, the highest NPV
55
56
57
58
59
60 ACS Paragon Plus Environment
75
Energy & Fuels Page 76 of 170

1
2
3 involved 2100 days of primary production, 80 cycles of 10-day huff, 125-day soak, and 220-day
4
5
6 puff. In general, their solutions favored shortened injection period, longer soaking periods to
7
8 promote molecular diffusion, and even longer production periods. Haghshenas et al. used a
9
10 compositional model to match the data from a producing multi-fractured horizontal well (Table 5,
11
12
13
entry 56).233 They stressed the importance of including adsorption and diffusion effects in CO2
14
15 huff-n-puff models. These parameters account for the variability of reservoir fluid component
16
17 adsorption and reservoir fluid property variability. In their 1000-day example for a single huff-n-
18
19
puff cycle, they recommended 55 days of injection time and 20 days of soaking time followed by
20
21
22 production.
23
24 4.10 Simulations performed at the University of Regina. At the University of Regina,
25
26 history-matching simulations were performed for CO2 huff-n-puff experiments for a confined
27
28
29 Bakken core (0.27 – 0.83 mD) (Table 5, entry 59).132 The optimum injection pressure of was at
30
31 the MMP and further increases in pressure did not result in substantial oil recovery increases.
32
33 Longer soaking time (up to about 15 days) resulted in higher recovery. The results confirmed that
34
35
36
multiple mechanisms, including oil swelling, viscosity reduction, IFT reduction, solution gas
37
38 drive, and light oil component extraction by CO2, were responsible for the oil recovery. A later
39
40 study, which included simulation of a five-layer reservoir in southern Saskatchewan, showed that
41
42
either a higher injection pressure or a lower wellhead pressure increased oil recovery during CO2
43
44
45 huff-n-puff (Table 5, entry 60).128
46
47 4.11 Simulations performed at the University of Kansas. After characterizing
48
49 Chattanooga shale by measuring geological, geophysical, geomechanical, and geochemical
50
51
52 parameters, Cudjoe et al. from the University of Kansas demonstrated that CO2 huff-n-puff
53
54 injection could increase the oil recovery from 10% to 53% (Table 5, entry 61).234 Jia and colleagues
55
56
57
58
59
60 ACS Paragon Plus Environment
76
Page 77 of 170 Energy & Fuels

1
2
3 developed a field-scale dual permeability compositional model for CO2 huff-n-puff in the Bakken
4
5
6 Formation in which the permeability from natural fractures was established with a series of
7
8 Dykstra-Parsons (DP) coefficients and spatial correlation lengths (Table 5, entry 62).235 Their
9
10 objective was to assess CO2 injectivity and the role of diffusion in heterogeneous reservoirs. Oil
11
12
13
production during primary recovery was decreased by higher degrees of heterogeneity and long
14
15 fracture lengths. Based on their study, discounting the molecular diffusion in the field would result
16
17 in up to 39% underestimation of oil recovery after 500 days of production in the puff period.
18
19
A review of CO2-EOR in shale oil reservoirs by Jia et al. stated the importance of
20
21
22 appropriately modeling the complex fracture network in unconventional reservoirs. The dual-
23
24 continuum (dual-porosity, dual-permeability) assumption refers to matrix/fracture characteristics
25
26 and is a classical way to account for natural fractures of conventional reservoirs. In unconventional
27
28
29 reservoirs, however, dual-continuum cannot fully represent the irregular shape and orientation of
30
31 the fractures. Further, in the dual-continuum model, the natural fracture properties are different
32
33 around the stimulated reservoir volume from those outside the region, where natural fractures are
34
35
36
mostly unstimulated. Instead of using the dual-continuum as the implicit way of dealing with
37
38 complex fracture network, the explicit approach would be the discrete fracture model (DFM), in
39
40 which geometries of the grids are dealt with in a structured or unstructured manner. Simulations
41
42
have improved greatly, but confidence in a model must be tempered against the complexity of the
43
44
45 geology and the accurate modeling of fractures, adsorption, nano-confinement, and multiple oil
46
47 recovery mechanisms.16
48
49 4.12 Simulations performed at Stanford University. In an early study from Stanford
50
51
52 University, simulations of continuous and huff-n-puff CO2 injection in a confined siliceous
53
54 Monterey core were able to match the experimental oil recovery (Table 5, entry 63). The results
55
56
57
58
59
60 ACS Paragon Plus Environment
77
Energy & Fuels Page 78 of 170

1
2
3 revealed the importance of countercurrent diffusion of CO2 into the matrix.58 Later, a model was
4
5
6 developed to reproduce the experimental results of CO2 continuous injection in the Bakken (Table
7
8 5, entry 64).120 The ability of CO2 to extract light hydrocarbons into the oil was an important
9
10 mechanism, supplemented by re-pressurization, oil swelling, viscosity reduction, and IFT
11
12
13
reduction. Challenges included low permeability and reservoir heterogeneities.
14
15 4.13 Simulations performed at other institutions: InPetro Tech, Occidental, Virginia
16
17 Tech, Montana Tech, Schlumberger, and Nalco Champion. Investigators from InPetro
18
19
included capillarity, pore size distribution, and adsorption in their simulations of CO2 huff-n-puff
20
21
22 in the Bakken (Table 5, entries 65 and 66). 236, 237 Results indicated that incremental recovery could
23
24 be underestimated if these effects are not modeled. Further, it was suggested that an insufficient
25
26 volume of CO2 may have been used in North Dakota Bakken CO2 pilot tests. Similar trends were
27
28
29 observed when the same investigators considered continuous CO2 injection, with the inclusion of
30
31 a distribution of more small pores yielding a several %OOIP increase in recovery.
32
33 An investigation of CO2 huff-n-puff at Occidental suggested that oil recovery was
34
35
36
proportional to the mass of injected CO2 (Table 5, entry 67).238 According to this model, molecular
37
38 diffusion and increased soaking time had no major beneficial effects on incremental recovery.
39
40 They found that the best time to switch from the puff production back to huff was when the oil
41
42
recovery rate returns to the pre-treatment base decline rate.
43
44
45 Researchers at Virginia Tech modeled CO2 huff-n-puff injection in the Bakken Formation.
46
47 The duration of primary production, injection period and puff production length were the most
48
49 important EOR parameters (Table 5, entry 68). 224 They did not implement a soaking period and
50
51
52 neglected the effect of CO2 diffusion on recovery—they assumed that pressure-driven viscous flow
53
54 dominated. Short injection and production periods with a higher number of cycles were favorable.
55
56
57
58
59
60 ACS Paragon Plus Environment
78
Page 79 of 170 Energy & Fuels

1
2
3 They also suggested selecting an optimal point, at which primary depletion should cease and huff-
4
5
6 n-puff commence.
7
8 In an early feasibility study, Shoaib and Hoffman from Montana Tech University (Table 5,
9
10 entry 69) showed that continuous injection of CO2 into Bakken using horizontal injectors
11
12
13
performed better than vertical ones.57 Further, they tested different scenarios with horizontal and
14
15 vertical injectors and found that oil recovery from single-well cyclic CO2 injection was
16
17 significantly less than that of continuous injection using both vertical and horizontal injectors.
18
19
By accounting for the existence of natural fractures, and diffusion coefficients and oil
20
21
22 properties, researchers at Schlumberger evaluated the performance of CO2 huff-n-puff in the Eagle
23
24 Ford Formation (Table 5, entry 70).240 They concluded that 9% OOIP additional oil recovery
25
26 would be achieved over hydraulic fracturing alone. They also suggested increasing hydraulic
27
28
29 fractures as much as possible to improve the well productivity by increasing the surface area
30
31 exposed to CO2.
32
33 The last entry in Table 5 corresponds to a simulation of an aqueous surfactant solution
34
35
36
(rather than high pressure gas) used for EOR in an unconventional formation. Kazempour et al.
37
38 from Nalco-Champion history-matched the permeability and capillary pressure curves from a field
39
40 trial of surfactant injection in the Middle Bakken (Table 5, entry 71).241 Even though they used a
41
42
single-porosity model, they included both micro fractures and main fractures in the stimulated
43
44
45 region and showed that increased oil recovery was possible through the injection of surfactant after
46
47 a shut-in period.
48
49 4.14 Conclusions from simulations of high-pressure gas EOR in ULRs. Simulations of
50
51
52 EOR in ULRs can provide valuable insights which are not observable in the laboratory. For
53
54 example, several reports suggest that CO2 and other EOR fluids are more miscible with oil when
55
56
57
58
59
60 ACS Paragon Plus Environment
79
Energy & Fuels Page 80 of 170

1
2
3 confined in nanopores. 96, 211, 207 Insights can also be extended to the field scale using simulations.
4
5
6 For example, many simulation studies agree that EOR should not be started until primary
7
8 production is complete,201, 232, 238-239 closer fracture spacing can improve recovery, 18, 182, 199, 202, 227
9
10 and numerous short cycles can recover more oil than one long cycle. 197, 182, 183, 61, 204, 210, 229, 230
11
12
13
Often, conclusions from simulation studies are supported by observations in laboratory-scale
14
15 experiments. For example, several simulation reports state that injection pressures at or above the
16
17 MMP for a given gas can improve oil recovery,132, 60, 206-207, 198, 197 that a rich natural gas can
18
19
improve oil recovery over methane alone,214, 205, 202, 137, 198, 18 and that increasing the surface area
20
21
22 of the rock exposed to EOR fluid can increase recovery.18, 240, 220, 223 In many cases, however,
23
24 simulations studies reach conflicting conclusions. For example, recommendations of huff-n-puff
25
26 soak times vary from to 8-20 days, 209, 132, 232 to two months,197 to 125 days (4 months),232 to 1000
27
28
29 days (2.7 years),208 to five years.9 Some reports recommend a longer injection period, 140, 183, 200,
30
145
31 while others recommended a shorter injection period.224, 232, 211
Some studies recognize
32
33 molecular diffusion as the main mechanism of improving oil recovery, 18, 195, 200, 205, 210, 235 while
34
35
36
others suggested that recovery was not sensitive to diffusion. 173, 204, 192, 193, 243, 216, 224, 207, 238
37
38 Despite so many efforts dedicated to predicting oil recovery performance of EOR in ULRs,
39
40 improvement is needed. While it is understandable that every simulation case takes on a particular
41
42
approach based on the parametric assumptions, geological features, and availability of resources,
43
44
45 there is no standard model available for unconventional reservoirs. The existence of natural
46
47 fractures in the matrix and their contribution to diffusion process and the oil recovery is a
48
49 significant parameter than has been neglected by many simulation studies. Commercial simulators
50
51
52 are designed to model conventional reservoirs have limited capabilities to model multi-scale
53
54 fractured unconventional reservoirs. So, unless the governing flow equations are self-developed
55
56
57
58
59
60 ACS Paragon Plus Environment
80
Page 81 of 170 Energy & Fuels

1
2
3 and implemented into models for each individual case, like some studies,240, 244 the true flow
4
5
6 behavior in the nanopores of a tight reservoir is not represented. Additionally, lack of available
7
8 field data for history-matching leads to building models based on assumptions, such as using
9
10 experimental diffusion coefficients to model field-scale diffusion process. Therefore, the need for
11
12
13
a more thorough understanding and more comprehensive assessment of the flow transport
14
15 mechanisms to simulate the EOR process in the ULRs is needed.
16
17 5.0 Chemical EOR using water in ULRs. Water remains the predominant liquid of choice
18
19
for hydraulic fracturing due to its low cost, ease of handling, and ability to dissolve performance-
20
21
22 enhancing additives. EOR in ULRs may also be conducted with water-based solutions, and this is
23
24 commonly referred to as “chemical EOR”. This section covers water-based fluids for EOR in
25
26 ULRs, including the use of brine, surfactants, water-based emulsions of organic liquids, and
27
28
29 nanoparticle dispersions in water. While we will refer to water-based EOR as “chemical EOR”,
30
31 note that chemicals can also be added to fluids other than water. For example, nonionic surfactants
32
33 have been added to CO2 to generate mobility control foams in situ during EOR in conventional
34
35
36
formations.245 However, there have yet to be reports of the addition of chemicals dissolved in CO2,
37
38 natural gas or nitrogen for EOR in unconventional formations.
39
40 5.1 Salt solutions for EOR in ULRs. Low salinity water (LSW) is a solution of dissolved
41
42
salts in water at a concentration significantly less than that of produced water or brine within the
43
44
45 formation. LSW can be formulated by adding salt to fresh water, but it is often more practical to
46
47 blend fresh water with produced brine. If available, seawater or blends of seawater and fresh water
48
49 can also be used. LSW is appealing because it inexpensive and can be conveniently prepared in
50
51
52 the field. Oil recovery using LSW in conventional formations can exceed that obtained using
53
54 produced high-salinity brines.246 Laboratory-scale oil recoveries from ULR cores indicate that
55
56
57
58
59
60 ACS Paragon Plus Environment
81
Energy & Fuels Page 82 of 170

1
2
3 several mechanisms may be responsible for the oil recovery observed during the immersion of oil-
4
5
6 rich shale cores in water or LSW.247-248, 249 Possible mechanisms of oil recovery by LSW include
7
8 wettability alteration,248, 250, 251, 252
detachment of clays,253 multicomponent ion exchange,254
9
10 desorption of polar components from rock surfaces,255, 256-257, 258 formation of cracks in shale due
11
12
13
to LSW imbibition,259-262, 263 and osmosis.264
14
15 Laboratory experiments involving low salinity brine for EOR. Oil recovery by LSW
16
17 soaking is typically low and slow. An early look into the imbibition of brine into thin sections
18
19
(0.65 – 5mm) of Pierre Shale indicated an increase in permeability and oil recovery of up to 40%
20
21
22 OOIP.259 This increase in recovery is likely due to mineral dissolution and cracking due to clay
23
24 swelling. This experiment alleviated concerns that LSW causes permeability reduction due to clay
25
26 swelling. A study of brine imbibition in the Spraberry Formation included both low-salinity brine
27
28
29 and a 1,000 ppm NaCl solution.265 Oil recovery ranged from 19-35% for ~1,000 ppm brine to 15-
30
31 23% for 138,000 ppm brine. The higher recovery obtained using low salinity brine was possibly
32
33 due to the dilution of the surface concentration of divalent ions.
34
35
36
A direct comparison of LSW (0.5 wt% TDS, total dissolved solids, fracturing fluid) and
37
38 high salinity brine (28.5 wt% TDS formation brine from the Bakken Formation) illustrated that oil
39
40 recovery was much higher for the low salinity brine.266 The authors concluded that the shale could
41
42
be viewed as a semi-permeable membrane that initially promotes the flux of water into the shale
43
44
45 at a greater rate than the high salinity formation brine, leading to increased oil recovery due to
46
47 osmosis.266-267
48
49 The spontaneous imbibition of 4 wt% KI (potassium iodide), into Permian Basin cores
50
51
52 resulted in the recovery of 11 % OOIP after about 150 hours.247, 268 Similar experiments with cores
53
54 from the Wolfcamp Formation resulted in the recovery of ~8% OOIP after about 100 hours, 269 or
55
56
57
58
59
60 ACS Paragon Plus Environment
82
Page 83 of 170 Energy & Fuels

1
2
3 ~2% OOIP after 65 hours,270 and 2% OOIP from an Eagle Ford core after about 50 hours.269 In a
4
5
6 thorough study of the characteristics of several ULRs, imbibition of water into short, 2.5-cm-
7
8 diameter Bakken, Wolfcamp, or Eagle Ford cores was slow and required 100-200 hours of
9
10 immersion in order to reach an ultimate oil recovery of only 3-8% OOIP.69
11
12
13
Other tests conducted with Bakken cores noted a higher recovery of oil with LSW than
14
15 high salinity brine during the long imbibition process, however the cores had very different
16
17 permeability values; 58% of the oil was recovered using 2 wt% KCl brine in 14 days from a 0.94
18
19
mD Bakken core, while it took 23 days to recover 43% of the oil using 24 wt% KCl brine from a
20
21
22 0.001 mD Bakken core.248 Another investigation with Bakken cores indicated that while the
23
24 amount of oil recoverable with LSW was not large, it was significantly greater than that obtained
25
26 with high salinity brine.271 During a study of capillary pressure and wettability, the same team
27
28
29 demonstrated that when LSW was used, oil production from Middle Bakken cores via spontaneous
30
31 imbibition was comparable to that achieved with forced imbibition, which indicated that molecular
32
33 diffusion and capillary osmosis were more dominant than wettability effects regarding oil
34
35
36
recovery.272
37
38 5.2 Alkaline solutions for EOR in ULRs. Alkaline floods in conventional formations
39
40 generate surfactants in situ as the injected caustic solution (e.g. a solution of sodium hydroxide,
41
42
NaOH, in water) reacts with carboxylic acids in the crude oil. There have been no reports of
43
44
45 alkaline floods for EOR in ULRs. We concur with a recent review that states that, given the low
46
47 cost and small molecular size of the alkaline additives (e.g. NaOH), aqueous caustic solution
48
49 injection may be a viable candidate for EOR in ULR that contain acidic crudes.13 The rate of
50
51
52 surfactant generation, however, would have to be fast and the surfactants would need to have a
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
83
Energy & Fuels Page 84 of 170

1
2
3 favorable composition and concentration to achieve the desired change in wettability and/or IFT
4
5
6 reduction.
7
8 Alkaline-surfactant-polymer floods in conventional formations also include viscosity-
9
10 enhancing polymers for mobility control. The addition of polymer thickeners would be undesirable
11
12
13
for EOR in ULRs. The polymer would have difficulty being transported within the very low
14
15 permeability matrix. We agree with others who have concluded that the implementation of
16
17 alkaline-surfactant-polymer floods for EOR in ULR are unlikely to be successful.13
18
19
5.3 Surfactant solutions for EOR in ULRs. The use of surfactants—likely in conjunction
20
21
22 with adjustments in salinity and pH—appears to be the most promising route for chemical EOR in
23
24 ULRs. Surfactant solutions can yield significantly higher recoveries than the addition of salts
25
26 alone. The ability of surfactants to reduce IFT and alter the wettability of a rock surface from oil-
27
28
29 wet to water-wet has been exploited for EOR in conventional reservoirs, hydraulic fracturing in
30
31 ULRs, and EOR in ULRs.
32
33 Although surfactants can be used to both reduce IFT and alter rock wettability toward
34
35
36
water-wet, it may not be beneficial to do both during EOR in ULRs.67 There are three phases of
37
38 hydraulic fracturing which each have unique pressure conditions. In Phase 1, a high-pressure
39
40 aqueous fracturing fluid is injected into a wellbore to create hydraulic fractures. Here, the
41
42
fracturing fluid pressure is greater than the formation pressure, resulting in a large pressure
43
44
45 differential (ΔP). During Phase 2, the well is shut-in for a period of time, and the pressure
46
47 differential (ΔP) between the formation and the injected hydraulic fracturing fluid diminishes.
48
49 Finally, during Phase 3, the well is opened for the flowback of fluids as the pressure of the
50
51
52 fracturing fluid declines to values below the formation pressure. Conceptually, during Phases 1
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
84
Page 85 of 170 Energy & Fuels

1
2
3 and 3, viscous forces driven by the large pressure difference between the formation and fracturing
4
5
6 fluid dominate. In Phase 2, the pressure differential is close to zero, and capillary forces dominate.
7
8 The relationship between viscous forces and capillary forces is often expressed using the
9
10 capillary number (𝑁𝐶𝑎 ) (eq. 1),
11
12
13
14
15 𝜇𝑢
𝑁𝐶𝑎 = (1)
16 ϒ𝑐𝑜𝑠𝜃
17
18
19
20
21
where  is the displacing fluid viscosity, u is the Darcy velocity of the displacing fluid,  is the
22
23 water-oil IFT, and  is the contact angle (measured within a water droplet). A high capillary
24
25 number, NCa means that viscous forces dominate over capillary forces, and oil flows from the rock.
26
27 273, 67, 274-275
28 Because most early studies dealt with strongly water-wet formations, the contact angle
29
30 approached 0° and cos approached unity. Therefore, common forms of the capillary number are:
31
32
33
34 𝜇𝑢
35 𝑁𝐶𝑎 = (2)
36 ϒ
37
38
39 and, in combination with Darcy’s Law,
40
41
42 𝑘 𝛥𝑃
43 𝑁𝐶𝑎 = (3)
44 ϒ𝐿
45
46
47
48 where k is the permeability of a brine saturated medium, and ΔP/L is the pressure drop gradient.
49
50
51
When NCa < 10-5 oil is trapped in the porous medium, as NCa increases from 10-5 to ~ 10-2 oil is
52
53 mobilized, and at NCa > 10-2 oil recovery is essentially complete.276-277
54
55
56
57
58
59
60 ACS Paragon Plus Environment
85
Energy & Fuels Page 86 of 170

1
2
3 Now, consider the addition of surfactants to the aqueous fluid.41 One would desire the
4
5
6 surfactants to reduce IFT values () to increase the capillary number and increase the relative
7
8 permeability of the oil phase, promoting the flow of oil (see Eq. 1). Therefore, during injection
9
10
(Phase 1) and flowback (Phase 3), an ultra-low IFT () between water and oil promotes oil
11
12
13 recovery. However, ultra-low IFT can be problematic during the soaking period (Phase 2) when
14
15 the capillary number (NCa) is very small because the pressure difference between the injected fluid
16
17
18
and the formation pressure approaches zero (△P, Eq. 3). During Phase 2, oil recovery is not
19
20 attained by viscous forces, rather, oil is displaced by spontaneous imbibition of the aqueous
21
22
solution into the shale that displaces oil from the pores. Spontaneous imbibition is determined by
23
24
25 capillary pressure, Pc, in the pores, where r is the radius of the capillary:
26
27
28
29 2 ϒ 𝑐𝑜𝑠𝜃
30 𝑃𝑐 = (4)
31 𝑟
32
33
34 A large and positive capillary pressure, Pc, promotes flow of water into pores and oil out
35
36 of pores. Here, a higher IFT (ϒ) increases capillary pressure (Pc), which is advantageous for oil
37
38
39 recovery (Eq. 4). A water-wet pore with a contact angle between 0 and 90° (where 0 < cosθ < 1)
40
41 will give a positive capillary pressure, Pc, which promotes oil recovery by spontaneous imbibition
42
43 (Eq. 4, illustrated in Figure 17). Therefore, during the soaking period (Period 2)—where the
44
45
46
pressure differential (△P) approaches zero (Eq. 3)—a surfactant that can alter wettability without
47
48 decreasing IFT is desirable. Because the soaking period is the most important phase during EOR
49
50
in ULRs, the surfactants used in ULRs should change wettability to water-wet without large
51
52
53 reductions in IFT. Equations 1-4 are also applicable to alkaline flooding and CO2 flooding.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
86
Page 87 of 170 Energy & Fuels

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 17. Desired wettability change from oil-wet to water-wet during the soak period. If the
20
21
rock surface is oil-wet, then  > 90o and cos is negative. It is desirable for Pc to be positive and
22 large; therefore, it is advantageous for the surfactant to make the surface as water-wet as possible
23 ( → 0; cos  → +1 ) while reducing the IFT  by as little as possible.
24
25
26 Mechanisms of wettability alteration by surfactants. The ability to alter the wettability of
27
28
29 oil-wet portions of shale to more water-wet is generally recognized as effective strategy for
30
31 improving oil recovery in ULRs. 278-283 Although mineral surfaces themselves are water-wet, four
32
33 mechanisms have been identified that are likely sources of the inherent oil-wet nature of shale
34
35
36
surfaces: (1) adsorption of polar crude oil constituents,284 283-284 (2) the deposition of asphaltenes
37
38 from crude oil onto the rock surfaces,283-284 (3) acid-base interactions between polar functional
39
40 groups in the oil and the thin layer of brine on the mineral surface,284 and (4) formation of ion
41
42
bridges between polar components of oil and charged surface sites on the mineral surface.284
43
44
45 Three mechanisms have been proposed for changing an oil-wet surface to a more water-
46
47 wet surface with surfactants: ion pair formation,285 surfactant adsorption,70 and micellar
48
49 solubilization.70 Ion pair formation involves an aqueous cationic surfactant solution entering an
50
51
52 oil-wet pore that is coated with adsorbed natural anionic compounds present in crude oil (Scheme
53
54 18A). The cationic surfactant in the water “pairs” with the oppositely charged, natural, adsorbed
55
56
57
58
59
60 ACS Paragon Plus Environment
87
Energy & Fuels Page 88 of 170

1
2
3 anionic compounds.285-286 This pairing weakens the adsorption of the natural surfactant to the
4
5
6 mineral surface.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26 Figure 18. Mechanisms of wettability alteration by surfactants.
27
28
29
30
31 Wettability can also be altered via formation of a surfactant bilayer (Figure 18B). 285, 287 If
32
33 anionic surfactant is injected, ion pairs will not form because the adsorbed natural surfactants and
34
35 the injected surfactants both have negative charges. Instead, the hydrocarbon tails of the anionic
36
37
38
surfactants can interact via van der Waals forces with the hydrocarbon portion of the adsorbed
39
40 natural surfactants. The anionic head group of the injected surfactant extends outward from this
41
42 hydrocarbon layer, rendering the surface water-wet. This type of wettability alteration is not
43
44
irreversible and may lead to re-imbibition of oil.
45
46
47 Anionic natural surfactants can also be “cleaned” off a mineral surface, altering wettability
48
49 via a micellar solubilization (Figure 18C). Natural anionic surfactants adhered to the mineral
50
51 surface can be slowly solubilized by injected surfactants. The hydrocarbon tails interact by van der
52
53
54 Waals forces, exposing the charged surfactant heads to the aqueous phase. The micelle can then
55
56 leave the mineral surface and be dissolved in the aqueous phase, leaving behind a water-wet
57
58
59
60 ACS Paragon Plus Environment
88
Page 89 of 170 Energy & Fuels

1
2
3 mineral surface.288,289 Although Figure 18 illustrates the three mechanisms using adsorbed anionic
4
5
6 oil components and charged injected surfactants, these mechanisms can also take place with
7
8 cationic oil components and nonionic injected surfactants. 289
9
10 Experiments related to surfactant EOR in ULRs. Surfactant formulations should be
11
12
13
designed for the specific formations under consideration.290-291, 292 There are many examples of
14
15 surfactant solutions for hydraulic fracturing, 293, 294, 295, 296, 297, 249, 298, 268, 269, 270, 250, 299, 300, 301, 302, 283,

16
17 284, 303, 304, 305, 306
and for EOR in conventional formations.307, 16, 344, 345, 346, 245, 308, 287, 309, 286, 310, 311,
18
19 38, 351, 312
Here, we will focus on publications focused on surfactant solutions for EOR in ULRs.
20
21
22 Many papers state that their results related to EOR in ULRs could also benefit the fracturing
23
24 process.
25
26 Ionic surfactants are quite commonly considered for EOR in ULRs because they are low-
27
28
29 cost, tolerant to high salinity, and do not form viscous microemulsions. Nonionic surfactants may
30
31 exhibit enough adsorption and oil solubilization to favorably alter wettability while inducing
32
33 relatively small reductions in IFT. (Excessive adsorption is not desirable because this would lead
34
35
36
to wettability changes only in the near wellbore region rather than deeper into the formation. 358)
37
38 Therefore nonionic surfactants are often considered alongside ionic candidates as viable candidates
39
40 for oil recovery.289, 313, 314 Oil recovery by aqueous surfactant solutions is typically measured using
41
42
spontaneous imbibition experiments, wherein a core is soaked in a solution of surfactant in water
43
44
45 or brine. Spontaneous imbibition experiments can be performed either at atmospheric pressure, or
46
47 at elevated pressure. Results of experiments performed on lower permeability cores are more
48
49 applicable to shale reservoirs. Table 6 lists spontaneous imbibition experiments by permeability
50
51
52 classification (low-permeability, tight, very tight or extremely tight). The reservoirs being studied,
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
89
Energy & Fuels Page 90 of 170

1
2
3 any information provided about the surfactants used, the salinity of the brine, and the range of
4
5
6 recovery factors reported are also included.
7
8
9
10 Table 6. Spontaneous imbibition experiments involving surfactants using cores from ULRs.
11
12
13
Low, 1-0.1 mD Tight, 0.1-0.01 mD Very Tight, 0.01-0.001 mD Extremely tight, 0.001-0.0001 mD
14
15
entry core reservoir surfactant Brine recovery reference
16 permeability (TDS, factor
17 (mD) %) (%OOIP)
18 1 0.20-1.61 Ordos TRS10, alkyl sulfonate (anionic) 8 5.4 Chen et al. 133
19 2 0.10-0.28 Changqing Carbonated water, betaine (zwitterionic) 5.0 2.1-9.4 Yu et al.315
3 0.06-0.01 Middle amphoteric dimethyl amine oxide, nonionic 15-30 16-25 Wang et al.316
20
Bakken ethoxylated alcohol, an anionic internal olefin
21 sulfonate, anionic linear α-olefin sulfonate (0.05-
22 0.1%)
23 4 0.035-0.15 Permian sulfates, internal olefin sulfonates (anionic) (0.1%) 14 40-68 Mohan et
24 Basin al.265
5 0.03-0.24 clay-rich alkyl ether sulfates, internal olefin sulfonates 13 46-58 Kathel et al.
25 sandstone (anionic) ethoxylated alcohols (nonionic) (0.1%) 317-318
26 6 n.r.a Eagle Ford, anionic, cationic, cationic-nonionic blend (0.1%) 7, 16 18-27 Alamdari et
27 Bakken al.319
28 7 0.006 Bakken internal olefin sulfonate (anionic), ethoxylated 0-30 30-60 Zhang et al.
252
29 alcohol (nonionic), ethoxylated tallow amine
(cationic), dimethyl amine oxide (amphoteric)
30 (0.05-2%)
31 8 0.0015- Middle Nalco surfactants (P300 Kit A, P300 Kit, R-50634) 23 1.7-5.2 Wang et al.320
32 0.096 Bakken and Liberty SF-72
33 9 0.001-2.74 Bakken anionic 2, 24 2.7-3.8 Teklu et al.248
34 10 0.005-0.08 Middle cationic, nonionic, amphoteric, anionic, tallow 2, 27 48-55 Nguyen et
Bakken, amine, anionic/nonionic blend al.321
35 Eagle Ford
36 11 0.001-0.02 5 reservoirs b amphoteric and anionic 4, 22 10-90 Nguyen et
37 al.322
38 12 0.002-0.007 Tavel, alkyl ethoxylate (nonionic), “low IFT formulation” 1.6-4.0 20-47 Chevalier et
39 Moliere, al.323
Lavoux
40 13 0.001-0.020 Bakken nonionic “production enhancer” 4.0 30-40 Dawson et
41 al.324-325
42 14 0.0002- Permian Branched alcohol oxyalkylate (nonionic), 4% KI 4-17 Alvarez et al.
43 0.001 Basin ethoxylated isodecylalcohol (nonionic), quaternary 70

44 ammonium (cationic), sulfonate (anionic), alkyl


ethoxylate and sulfonate (nonionic-anionic blend)
45 15 0.0002- Permian same as entry 13 4% KI 7-34 Alvarez et
46 0.001 Basin al.247
47 16 0.000098- Dubernay nonionic n.r. e 0-25 Begum et al.
48 0.000142 67, 326-327

49
50 a
Permeability not reported. Porosities between 12-14% b Bakken, Niobrara, Eagle Ford,
51 Wolfcamp, Spraberry. c Brine was used, but the composition was not reported.
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
90
Page 91 of 170 Energy & Fuels

1
2
3 Spontaneous imbibition experiments using low-permeability and tight cores (1-0.01 mD).
4
5
6 An online NMR study of oil recovery in tight formations confirmed that surfactant solutions are
7
8 more effective than water in displacing oil as it flows through porous media (as opposed to
9
10 imbibition), however neither proved to be as effective as high pressure nitrogen or CO2 (Table 6,
11
12
13
entry 1).133 Yu et al. showed oil recovery from low-permeability cores using carbonated water and
14
15 “active” carbonated water, which contained surfactant (Table 6, entry 2). They found that “active”
16
17 carbonated water injection gave higher oil recovery than carbonated water alone. In another
18
19
experiment by Wang et al., oil was recovered from Bakken cores at reservoir conditions via the
20
21
22 imbibition of surfactant solutions (Table 6, entry 3).316 The cores were 25 or 38 mm in diameter
23
24 with lengths varying from 2-52 mm. The surfactants included an amphoteric dimethyl amine
25
26 oxide, a nonionic ethoxylated alcohol, an anionic internal olefin sulfonate, and an anionic linear
27
28
29 α-olefin sulfonate. An additional incremental 7-10% OOIP was recovered after brine imbibition
30
31 in three tests, while 15-25% OOIP was realized in ten other EOR surfactant imbibition tests. None
32
33 of the surfactants were obviously superior to the others. Each was shown to alter wettability and
34
35
36
each yielded promising EOR results. The effect of increasing alkalinity was unclear—in some
37
38 cases it enhanced oil recovery and in others it reduced oil recovery. 316
39
40 A study of surfactant solution imbibition in the tight (<0.1 mD), weakly water-wet
41
42
Spraberry Formation of the Permian Basin was restricted to anionic and non-ionic surfactants
43
44
45 (Table 6, entry 4).265 The strong interaction of cationic surfactant head groups with negatively
46
47 charged clay particle surfaces was thought to render the rock more oil-wet. The 2.5 cm-diameter,
48
49 5.1-cm long cores were cleaned, dried, evacuated, saturated with brine, flooded with oil and aged.
50
51
52 Oil recovery of 40 – 63% OOIP was obtained with 0.1 wt% anionic surfactant dissolved in
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
91
Energy & Fuels Page 92 of 170

1
2
3 formation brine, with recovery as high as 68% OOIP after 60 days. Although some degree of IFT
4
5
6 reduction was attained, the authors attributed the oil recovery to capillary and buoyancy forces. 265
7
8 Kathel et al. demonstrated that non-ionic ethoxylated alcohols with more than 20 ethylene
9
10 oxide (EO) groups increased the water-wet nature of sandstone (Table 6, entry 5).317-318 The control
11
12
13
experiment, which used formation brine only, resulted in the recovery of 15% OOIP after about
14
15 one month. Although the nonionic surfactants increased water-wetness, none could alter the
16
17 wettability to water-wet (contact angle less than 80°), therefore only anionic surfactants were used
18
19
for testing of oil recovery via imbibition. Oil recovery ranged from 54-68% OOIP for an anionic
20
21
22 C13-EO27 sulfate. An alkyl-propoxylated-ethoxylated sulfate, C20-PO7-EO30-sulfate (PO represents
23
24 propylene oxide), yielded 46-52% OOIP, while C24-PO25-EO46-sulfate yielded 46-52% OOIP.317-
25
26 318
27
28
29 Spontaneous imbibition experiments using very tight cores (0.01-0.001 mD). A study was
30
31 conducted that compared fresh water, brine with 30% TDS, and aqueous solutions of anionic,
32
33 cationic and nonionic surfactants at 0.05 - 2 wt% in brine, or formation brine at 30 wt% TDS with
34
35
36
either no alkaline content or a dilute alkaline concentration (0.2 – 0.25%) (Table 6, entry 7).252
37
38 Water and brine results were poor. Surfactant solutions were promising, with both IFT reduction
39
40 and wettability alteration deemed to be important for EOR from Bakken cores. Wettability
41
42
alteration was particularly relevant for the nonionic surfactant formulation. Oil recoveries—which
43
44
45 occurred quickly during the first 25 hours and leveled off after about 100-200 hours—ranged from
46
47 about 30-60% OOIP, with the highest recoveries occurring for the tests conducted at the optimal
48
49 salinity. Combined with their previous work that indicated the best pH range for nonionic
50
51
52 surfactants at 1000 ppm concentration was 7-8,259, 328 this team thoroughly considered thermal
53
54 stability and imbibition as a function of salinity, pH, and surfactant concentration. They
55
56
57
58
59
60 ACS Paragon Plus Environment
92
Page 93 of 170 Energy & Fuels

1
2
3 recommended a wettability-altering formulation of 0.1% surfactant, 4% TDS, and 1000 ppm of an
4
5
6 alkaline buffer NaBO2.4H2O at a pH of 7-8.329
7
8 A series of brine and surfactant solution imbibition tests was conducted with Middle
9
10 Bakken cores with permeability ranging from 0.0015 to 0.096 mD, and synthetic Bakken brine
11
12
13
(Table 6, entry 8).320 The surfactants included three Nalco surfactants (P300 Kit A, P300 Kit, R-
14
15 50634) and Liberty SF-72. In some of the oil recovery tests, the cores had no initial water
16
17 saturation, while in others, the initial water saturation was 53%. Oil recoveries were higher from
18
19
the cores with no initial water, ranging from 35-58% OOIP after 10 days. Somewhat surprisingly,
20
21
22 although the oil recovery with brine was slower than that attained with surfactant solutions, the
23
24 recovery was ultimately comparable. In an attempt to scale the rates of imbibition to the field
25
26 scale, the investigators used Ma’s scaling group for laboratory-scale tests.330
27
28
29 Teklu et al. studied oil recovery via imbibition using 2.5-cm diameter Bakken cores,
30
31 Bakken crude oil, high salinity brine (240,000 ppm) and low salinity formation brine (20,000 ppm)
32
33 (Table 6, entry 9).248 The anionic surfactant solutions were more effective than low salinity brine,
34
35
36
which in turn was more effective than high salinity brine. This observation prompted the team to
37
38 suggest that high salinity brines be avoided during fracturing and that EOR in ULRs be conducted
39
40 first with low salinity brine, and then—in subsequent cycles—with surfactant solutions.
41
42
Nguyen et al. assessed nonionic, cationic, anionic and amphoteric surfactants at
43
44
45 concentrations of 0.05 – 0.2% in brine for the recovery of oil from ULR (Table 6, entry 10).321
46
47 Middle Bakken and Eagle Ford cores were soaked in Bakken crude oil and Eagle Ford condensate,
48
49 respectively, various brines up to 30% TDS, and an alkaline additive. The total acid number and
50
51
52 resin content of the Bakken crude oil was higher than that of the Eagle Ford condensate, which
53
54 caused the Bakken cores to be more oil-wet than the Eagle Ford cores. In experiments involving
55
56
57
58
59
60 ACS Paragon Plus Environment
93
Energy & Fuels Page 94 of 170

1
2
3 Bakken cores, the best results were obtained using a nonionic surfactant. In experiments involving
4
5
6 Eagle Ford cores, higher recoveries of 48% were realized with the anionic surfactant. Alkaline
7
8 addition (intended to suppress adsorption) diminished oil recovery for cationic and anionic
9
10 surfactants, but increased recovery for nonionic surfactants. No clear correlation between IFT and
11
12
13
oil recovery was observed, and the authors suggest that wettability alteration is the dominant
14
15 mechanism of recovery. In a study related to oil adhesion on Bakken samples, it was shown that
16
17 an anionic surfactant solution was salt tolerant (30% TDS brine) and thermally stable at 115 °C
18
19
(Table 6, entry 11).322 LSW (4% TDS) was better than high salinity brine (22% TDS) in oil
20
21
22 recovery via imbibition, and that the addition of 1000 ppm of anionic surfactant increased oil
23
24 recovery for both brines. Similar (but less dramatic) improvements were attained with an
25
26 amphoteric surfactant.
27
28
29 In a paper and a related commentary, formulations of wettability-altering nonionic
30
31 surfactant formulations which also included corrosion inhibitors, scale inhibitors, and biocides
32
33 were considered for EOR the Bakken Formation (Table 6, entry 13).324-325 The “production
34
35
36
enhancers” used in this study exhibited IFT reductions (to 1-2 mN/m), low critical micelle
37
38 concentration (5-20 ppm), thermal stability and a strong emulsion-breaking tendency. These
39
40 formulations yielded 30-40% OOIP recovery in Bakken cores, initial spontaneous imbibition rates
41
42
of more than 10 cm/day in 1-20 µD cores, indicating that surfactant flooding EOR could be viable
43
44
45 even in extremely tight formations.
46
47 Spontaneous imbibition experiments using extremely tight cores (0.001-0.0001 mD).
48
49 Alvarez and colleagues compared the performance of two unspecified nonionic and anionic
50
51
52 surfactants at 1,000 ppm concentrations.331 The anionic surfactants proved superior to nonionic
53
54 surfactants with regard to oil displacement during spontaneous imbibition tests. Anionic
55
56
57
58
59
60 ACS Paragon Plus Environment
94
Page 95 of 170 Energy & Fuels

1
2
3 surfactants are generally considered to be better for IFT reduction, but at a fixed concentration of
4
5
6 1000 ppm, the nonionic and anionic IFT reductions were comparable. CT imaging showed that
7
8 both types of surfactants demonstrated comparable penetration into cores. In spontaneous
9
10 imbibition experiments, however, anionic surfactants proved superior to nonionic surfactants in
11
12
13
producing oil. The higher oil recovery using anionic surfactants was attributed to the greater degree
14
15 of wettability alteration by anionic surfactants, which was observed with contact angle experiments
16
17 that showed anionic surfactants imparting more hydrophilicity. Alvarez and Schechter reached
18
19
similar conclusions when they assessed several nonionic, anionic and cationic surfactants (Table
20
21
22 6, entries 14 and 15).70,247 The best oil recovery of oil from 0.2 – 1 μD cores via imbibition was
23
24 attained using anionic surfactants. The authors attributed this to lower IFT and more effective
25
26 alterations of wettability to water-wet. However, they emphasized that the reduction of IFT should
27
28
29 be sufficient only to allow water imbibition into the pores and the subsequent oil counterflow; they
30
31 asserted that ultralow IFT is not desirable for imbibition because it can favor oil re-deposition on
32
33 surfaces and water movement outwards the matrix. The same researchers continued their
34
35
36
exploration of surfactant solutions, noting that contact angle determination, the NMR method, and
37
38 zeta potential measurements are the most important tools for assessing wettability, and oil recovery
39
40 via imbibition that is enhanced by dissolved surfactants is best assessed with a suite of tests,
41
42
including solubility of surfactants in brine, changes in contact angle, IFT measurement, surfactant
43
44
45 adsorption, spontaneous imbibition tests and forced imbibition tests.332
46
47 A recent study of imbibition oil recovery from the oil-wet Duvernay Shale Formation
48
49 provided a comparison of distilled water, produced brine, a nonionic surfactant, and clay stabilizer
50
51 67, 326-327
52 (Table 6, entry 16). The greatest oil recovery was attained with the surfactant solution,
53
54 which was attributed to surfactant solution being capable of displacing oil from hydrophilic pores
55
56
57
58
59
60 ACS Paragon Plus Environment
95
Energy & Fuels Page 96 of 170

1
2
3 and a portion of the hydrophobic pores. These researchers concluded that water cannot displace
4
5
6 oil from hydrophobic organic pores.
7
8 Surfactant imbibition experiments favoring IFT reduction over wettability-alteration.
9
10 Although most reports focused on surfactants for EOR in ULRs agree that wettability-alteration is
11
12
13
the more important oil recovery mechanism during spontaneous imbibition of tight rocks, a few
14
15 reports support the use of IFT-reducing surfactants. A recent study has shown that oil recovery
16
17 with cationic IFT-reducing surfactant can give higher oil recovery than wettability-altering
18
19
surfactants in mixed wettability media (Table 6, entry 6).319 In experiments using Eagle Ford
20
21
22 cores, low salinity brine (7,000 ppm TDS) yielded greater oil recovery than any of the 0.1wt%
23
24 wettability-altering surfactant formulations after 50 hours of immersion. The LSW was even more
25
26 effective than an anionic surfactant solution that reduced the water-shale contact angle from 170°
27
28
29 to 44° at a concentration of 0.1 wt%. Rather than continuing to explore wettability changes, the
30
31 researchers turned their attention to IFT-reducing formulations based on anionic or cationic
32
33 surfactants that were able to attain IFT as low as 0.01 dyne/cm. Oil recovery via imbibition
34
35
36
increased dramatically, with recovery of OOIP as high as 68% being attained when the cores were
37
38 immersed in the surfactant solutions. The cationic IFT-reducing surfactant also exhibited stability
39
40 at formation temperature and absence of emulsion formation.319
41
42
Another group used NMR monitoring of fluid saturations during imbibition tests as a rapid
43
44
45 way to compare the performance of surfactant formulations (Table 6, entry 12).323 An ultralow IFT
46
47 surfactant (which also behaved as a wettability-altering surfactant) was compared to a benchmark
48
49 wettability-altering surfactant (a nonionic alkyl ethoxylate). The undisclosed formulation
50
51
52 “LowIFT_3” afforded the highest oil recovery in imbibition experiments. The authors concluded
53
54 that these results support the importance of lowering IFT for oil recovery by surfactants.
55
56
57
58
59
60 ACS Paragon Plus Environment
96
Page 97 of 170 Energy & Fuels

1
2
3 Alternative wettability-altering water-soluble chemicals for ULR. There are chemicals
4
5
6 other than traditional surfactants that may alter wettability from mixed-wet and oil-wet toward
7
8 water-wet. For example, it was recently shown that 3-pentanone, a symmetric ketone, is about 1.1
9
10 wt% soluble in water and will also partition into oil (in which it is completely miscible). 333
11
12
13
Although 3-pentanone has no impact on IFT, it does induce significant reductions in water-oil-
14
15 rock surface contact angle from 123° in reservoir brine to 26° with a 1.1 wt% solution of 3-
16
17 pentanone in reservoir brine.
18
19
A novel dimethyl ether (DME)−brine enhanced oil recovery (EOR) method has been
20
21
22 developed for the North Sea tight chalk oil reservoirs.334 Both secondary and tertiary DME−brine
23
24 injection presented a potential to improve oil recovery in tight chalk oil reservoirs. The results of
25
26 core flooding experiments using conventional cores revealed that oil swelling caused by
27
28
29 segregation of DME into the oil phase is the dominant oil recovery mechanism.
30
31 A combination of surfactant injection and CO2 huff-n-puff is also being considered. Using
32
33 Eagle Ford core plugs and crude oil, Zhang et al. demonstrated that after several CO2 huff-n-puff
34
35
36
cycles, surfactant injection accounted for an additional ~5-10% OOIP.335 This increase was
37
38 especially noteworthy for the CO2 huff-n-puff test conducted at a pressure below the MMP of the
39
40 CO2-oil mixture. The combination of CO2 and a “chemical blend”, composed of an aqueous
41
42
solution of an anionic surfactant and a persulfate compound in brine, was proposed for elevating
43
44
45 the oil recovery relative to CO2 alone. 336 The aqueous chemical blend was used first, followed by
46
47 three cycles of CO2 huff-n-puff. These combined surfactant-gas injection strategies could also be
48
49 employed using natural gas or nitrogen rather than CO2.
50
51
52 5.4 Aqueous nanofluids and nanoparticle dispersions. Complex nanofluid (CNF)
53
54 emulsions of solvents in water. Surfactant-stabilized aqueous microemulsions of solvent droplets
55
56
57
58
59
60 ACS Paragon Plus Environment
97
Energy & Fuels Page 98 of 170

1
2
3 suspended in water have been proposed for use during hydraulic fracturing. The intended purpose
4
5
6 of these “complex nanofluids” (CNF) is to combine the IFT-reducing and wettability-altering
7
8 characteristics of the surfactant with the oil viscosity-reducing properties of the solvent. 337-338,337,
9
10 339
Molecular dynamics simulations have shown that nonionic alkyl ethoxylate surfactants can
11
12
13
stabilize limonene droplets that—due to their elasticity—can pass through pores smaller than the
14
15 droplet diameter. This ability allows them to transport solvents and surfactants throughout the
16
17 shale matrix to increase oil recovery.338, 340 In a study of hydraulic fracturing fluids for the Bakken
18
19
Formation, CNFs composed of water, citrus terpenes and isopropanol reduced water-oil IFT and
20
21
22 induced favorable wettability changes.298 Relative to surfactant solutions, the CNFs yielded the
23
24 second-best oil recovery in siliceous Bakken cores and the best oil recovery Bakken cores with
25
26 high carbonate content.
27
28
29 Aqueous nanofluids composed of nanoparticle dispersions. Water can also serve as a
30
31 carrier fluid for suspensions of nanoparticles. Nanoparticles are defined as any solid particle with
32
33 a diameter less than 100 nm. Nanoparticles are typically composed of silicon oxide particles (i.e.
34
35
36
fumed silica),341-345, 346 or metal oxides of aluminum, titanium, zirconium, and zinc. 347-349, 346
In
37
38 petroleum engineering applications, nanoparticles are dispersed in a liquid (water, brine or a
39
40 surfactant solution) and must remain dispersed (not aggregate and settle) in order to be useful.
41
42
Most of the potential oil recovery applications of nanoparticles are related to EOR processes in
43
44
45 conventional formations.21-22, 350-351 A comprehensive overview of nanoparticle usage in hydraulic
46
47 fracturing of ULRs summarizes of all of the uses of nanoparticles for hydraulic fracturing,
48
49 including their use as gel crosslinkers, viscoelastic surfactant solution viscosity enhancers, micelle
50
51
52 viscosity enhancers, foam stabilizers, nanosensors, nanoproppants, fluid loss control agents, gel
53
54 breakers, and constituents of biopolymer nanocomposites.352 Most nanoparticle research has been
55
56
57
58
59
60 ACS Paragon Plus Environment
98
Page 99 of 170 Energy & Fuels

1
2
3 performed using nanoparticles dispersed at concentrations of 0.001-10 wt%.353 Silica nanoparticles
4
5
6 successfully increased oil recovery when added to hydraulic fracturing fluid in the Permian Basin
7
8 wells.354 Nanoparticles can be used in place of surfactants to stabilize CO2-in-brine mobility
9
10 control foams, especially in high-temperature, high-salinity formations where surfactants are prone
11
12
13
to precipitation, adsorption, or decomposition.245 Nanoparticle-induced wettability changes of the
14
15 rock was covered to some extent in the last section of the Yekeen et al review.352
16
17 While surfactants act via a “chemical” mechanism of wettability alteration (shown in
18
19
Figure 18), nanoparticles act by a “mechanical” mechanism of wettability alteration. Due to their
20
21
22 high surface area to volume ratio, nanoparticles tend to accumulate at phase interfaces. Between
23
24 oil-water interfaces, nanoparticle accumulation results in reduction of IFT. At oil-water-rock
25
26 interfaces, nanoparticle accumulation results in wettability alteration. Nanoparticles form a wedge
27
28
29 between the rock surface and the oil, and a structural disjoining pressure is exerted on the oil
30
31 (Figure 19). The interface of oil/brine with nanoparticles moves forward, decreasing the contact
32
33 angle between the oil and the rock surface.355-358 Structural disjoining pressure is the sum of several
34
35
36
interactions: dispersion forces, electrostatic forces between charged surfaces, interactions due to
37
38 layers of neutral molecules adsorbed on the two surfaces, and the structural effects of the solvent.
39
40 The structural disjoining pressure increases with the volume fraction of nanoparticles.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56 Figure 19. Mechanism of wettability alteration by nanoparticles.
57
58
59
60 ACS Paragon Plus Environment
99
Energy & Fuels Page 100 of 170

1
2
3 Many investigations involving nanoparticles have focused on interfacial phenomenon,
4
5
6 such as surface energy and wettability,329, 353, 347, 359, 360-362, 346, 347, 363, 354 which is related to reservoir
7
8 surface mineralogy,364 fluid saturation, and relative permeability of rock.365 Naik suggested adding
9
10 nanoparticles into drilling or fracturing fluids to improve gas production by increasing contact
11
12
13
angle to reduce water retention, since they altered rock surface to neutral or gas-wet.366 361, 367

14
15 Kuang et al. (2018) found that electrostatic attractive/repulsive forces of nanoparticles, i.e.,
16
17 negatively charged carboxyl groups or hydroxyl groups on hydrophilic nanoparticles, could be
18
19
essential to co-adsorption of nanoparticles and surfactant molecules on rock surface.360 Kondiparty
20
21
22 et al. showed that smaller nanoparticles exhibit more favorable surface spreading, which is
23
24 advantageous for wettability alteration.357
25
26 Nanoparticle-Surfactant Dispersions. Both nanoparticles and surfactants were combined
27
28
29 in a study of a nanoparticle surfactant (NPS) dispersion.368 The NPS dispersion was intended to
30
31 carry low-salinity hydraulic fracturing fluid deeper in the matrix and to produce oil by
32
33 nanoparticle-induced IFT reduction and wettability alteration. Therefore, to encourage
34
35
36
spontaneous imbibition, the nanoparticle-surfactant dispersion was not produced rapidly after
37
38 fracturing. Instead the NPS dispersion was allowed to soak into the fractured rock. The promising
39
40 laboratory results were confirmed by an improvement in oil recovery using NPS in field trials in
41
42
the Montney Formation.
43
44
45 5.5 Conclusions from reports of chemical EOR technologies for EOR in ULRs.
46
47 Looking forward, the implementation of surfactant solutions and nanoparticle dispersions hold
48
49 promise for EOR in ULRs. These chemical EOR techniques have resulted in favorable laboratory-
50
51
52 scale EOR in ULR results. The injected fluids can be formulated for specific formations via the
53
54 changes in the number, type and concentration of surfactants, selection of brine salinity, and the
55
56
57
58
59
60 ACS Paragon Plus Environment
100
Page 101 of 170 Energy & Fuels

1
2
3 introduction of other additives.369 Anionic surfactants appear to be most effective, 317-318, 321, 331, 70,
4
5 247
6 but some researchers recommend the use of nonionic,321 or cationic 319 surfactants. Due to the
7
8 predominance of capillary effects during long soak times, wettability-alteration is likely more
9
10 important than IFT reduction for surfactants used in ULRs. 316, 265, 317-318, 329, 331, 321, 70, 247 Still, some
11
12
13
researchers maintain that achieving ultralow IFT reduction is more important than wettability
14
15 alteration oil recovery.319, 323 Excessive adsorption remains a common concern for surfactants
16
17 intended to alter wettability.324-325 Operators can further enhance the recovery of oil with
18
19 370, 245, 309, 248, 322
surfactant solutions with optimal—usually low (~4%)—salinity. Despite the
20
21
22 preponderance of these research efforts, there appears to be a consensus that the extremely low
23
24 permeability of the most significant ULR targets renders aqueous solutions less likely to yield
25
26 economically viable rates of oil production that high-pressure gas-based EOR strategies such as
27
28
29 natural gas or CO2 huff-n-puff. This consensus is also reflected in the field trials, which are
30
31 currently dominated by high pressure gas-based EOR efforts.
32
33 6.0 EOR pilot tests in ULRs. In this section, we summarize field tests of EOR in ULRs
34
35
36
reported in publications about individual projects and previous literature reviews of field tests.13,
37
24-26, 371
38 Many miscible gases and aqueous chemical EOR fluids have been assessed in the
39
40 laboratory and in computational models. However, only three types of fluids (fresh or produced
41
42
water, natural gas, and CO2) have been tested for EOR in ULRs.13-14, 24-26 Table 7 lists all EOR in
43
44
45 ULR pilot tests reported in the literature. The pilots are organized by formation and fluid tested.
46
47 The year of the pilot test, whether the test was to enhance oil recovery (EOR) or simply test the
48
49 fluid injectivity (inj), and the mode (huff-n-puff, HnP, or continuous, cont.) are also included.
50
51
52 Reported results of each test are included, along with any information available regarding the rock
53
54 of fluid properties of the wells.
55
56
57
58
59
60 ACS Paragon Plus Environment
101
Energy & Fuels Page 102 of 170

1
2
3 Table 7. Summary of EOR pilot tests in ULRs.
4
5
entry Location Fluid Year EOR Mode b Results Reference
6 or Inj.a
7 1 CO2 2008 EOR HnP No significant increase in oil production. Hoffman et al.26
8 2 CO2 2009 EOR HnP No significant increase in oil production. Hoffman et al.26
9 3 CO2 2014 Inj Ver. Early breakthrough. Hoffman et al.26
10 4 CO2 2014 EOR Cont. Results unavailable. Sorensen et al.25
5 CO2 2017 EOR Cont. Unfractured reservoir, injectivity was good and oil was produced. Reservoir pressure 8,668 Sorensen et al.372
11 psi (60 MPa). Lighter hydrocarbons were produced preferentially.
12 6 Nat gas 2014 EOR Cont. Modest oil increase in offset wells after a few months. Hoffman et al.26
13 U.S.
7 Water/ Field gas 2014 EOR Cont. Production increase. Sorensen et al.25
Bakken
14 8 Water 1994 EOR Cont. Production increase. Sorensen et al.25
15 9 Produced Water 2012 EOR HnP Oil production enhanced due to frac-hits. Hoffman et al.26
10 Water 2012 EOR Cont. No significant increase in oil production. Hoffman et al.26
16
11 Water 2014 EOR Cont. Oil production enhanced due to frac-hits. Hoffman et al.26
17 12 Surfactant 2018 EOR HnP Oil recovery improved after 4-month shut in. Salinity and pH of formation brine Kazempour et
18 equilibrated, indicating good mixing between injected and formation brine. al.241
19 13 Ethane Foam 2019 EOR HnP Mobility control achieved by foam generation. Oil production doubled. Porosity 13%, Katiyar et al.188
20 permeability 0.1-15 mD, 39˚ API oil,
21 14 Nat gas 2011 EOR Cont. Oil production increase from 135 bbl/day to 295 bbl/day during 1-year period. Porosity 9- Schmidt et al.373
Canadian 12%, permeability < 1 mD, 42˚ API oil. Water saturation 55-59%, reservoir pressure 2,320
22 Bakken psi (16 MPa). oil viscosity 2-3 cP, OOIP 8,000 MSTB (Pilot area), pore throats 0.1-0.2 µm.
23 15 Water 2011 EOR Cont. Oil production increased from 75 bbl/day to 550 bbl/day. Wood et al.271
24 16 Nat Gas 2012 EOR HnP 15 wells, 30-70% increase in oil production. Wang et al.15
25 17 Nat Gas 2012- EOR HnP Pilot A, Preliminary results only. Production rate increase for each cycle. Hoffman et al.24,
374
26 2013
18 Eagle Nat gas 2015 EOR HnP Pilot B, 17% increase in incremental oil production. Hoffman et al.374
27 19 Nat gas 2015 EOR HnP Pilot C, 20% increase in incremental oil production. Hoffman et al.374
Ford
28 20 Nat gas 2015 EOR HnP Pilot D, 30% increase in incremental oil production. Hoffman et al.374
29 21 Nat gas 2015 EOR HnP Pilot E, Possible increase, production obscured by 3 other wells in lease. Hoffman et al.374
30 22 Nat gas 2015 EOR HnP Pilot F, Inconclusive results because 60 other wells in lease production report. Hoffman et al.374
31 23 Nat gas 2016 EOR HnP Pilot G, Promising initial incremental oil results. Hoffman et al.374
32 24 Argentina Water 2017 EOR HnP Oil production reported only for 2-day flowback. Tuero et al.375
25 CO2 2003 Inj. Ver. Low recovery, project considered unsuccessful. Permeability 0.79 mD, oil viscosity 6.6 cP. Jiang et al.376
33 26 CO2 2008 Inj. Ver. High injectivity, 21% increase in oil recovery. Permeability 0.96 mD, oil viscosity 3.6 cP. Wang et al.377
34 27 Water 2008 EOR Pulsed Sharp production after shut-in, no significant reduction of water-cut. Permeability 0.17 mD, Wang et al.377
35 oil viscosity 1.01 cP.
36 28 China Water 2008 EOR Asynch- Increased oil production from 26 bbl/day to 40 bbl/day. Permeability 0.17 mD, oil viscosity Wang et al.377
37 ronous 1.01 cP.
29 Water 2014 EOR HnP Increased oil production in neighboring wells Lin et al.378
38
30 Surfactant 2015 EOR HnP Production rate increased after 7-day shut in. Porosity 8%, permeability 0.54 mD. Oil Wei et al.379
39 viscosity at 50 ˚C, 4.67 cP. Reservoir pressure 4.25 MPa, bubble point pressure 1.12 MPa.
40 31 Water 2014 EOR HnP Production rate increased from 48 bbl/day to 269 bbl/day. Li et al.380
41 a
Inj. = Injectivity only, HnP = Huff-n-puff, Cont. = continuous flood, Ver. = vertical injection.
b

42
43
44
45 ACS Paragon Plus Environment
46 102
47
Page 103 of 170 Energy & Fuels

1
2
3 6.1 EOR pilot tests in the U.S. Bakken Formation. Twelve pilot tests performed in the
4
5
6 U.S. Bakken have been documented in the literature. Five tests involve CO2 injection, two involve
7
8 natural gas or field gas injection, and five involve water or surfactant injection (Table 7, entries 1-
9
10 12).26, 25, 241, 372, 241 In general, the injectivity of these fluids was high, but the oil recovery results
11
12
13
were disappointing. Oil recoveries from pilot tests in the Bakken Formation have yet to meet the
14
15 favorable expectations based on promising laboratory-scale experiments and numerical
16
17 simulations.
18
19
CO2 EOR pilot tests in the U.S. Bakken Formation. In two of the huff-n-puff projects, high-
20
21
22 pressure CO2 was injected (Table 7, entries 1 and 2). Even though the CO2 huff-n-puff pilot tests
23
24 were performed by two different operators at two different locations in the Bakken in 2008 and
25
26 2009, the results were similar. The planned injection rate of 1,000-2,000 Mscf/day was met without
27
28
29 going beyond the fracture pressure limit. In the first project, CO2 was injected at 1,000 Mscf/day
30
31 for about 30 days. After less than two weeks, an increase in the CO2 concentration was observed
32
33 5,000 feet from the wellhead (Table 7, entry 1). In the second project, CO2 was injected into a
34
35
36
vertical well in the Elm Coulee Field at a rate of 1,500-2,000 Mscf/day for 45 days (Table 7, entry
37
38 2). A history-matching simulation was performed based on this field test, which supported
39
40 diffusion as the recovery mechanism.224 No significant increase in the oil production using CO2
41
42
huff-n-puff was reported for either project.26, 381
43
44
45 Another CO2 huff-n-puff test in the Middle Bakken was planned for 2014, but its scope
46
47 was limited to injectivity (Table 7, entry 3). The injectivity was 20 times higher than other huff-n-
48
49 puff pilot tests. Early CO2 breakthrough occurred at a well about 900 feet away in less than 24
50
51
52 hours. Therefore, the huff-n-puff oil production study was not performed. The Whiting company
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
103
Energy & Fuels Page 104 of 170

1
2
3 conducted another pilot test in 2014, to see whether CO2 could be injected in the Middle Bakken
4
5
6 Formation (Table 7, entry 4).25 The results were unavailable.
7
8 In 2017, a pilot test was conducted in a virgin reservoir by injecting CO2 into a vertical
9
10 well (with no hydraulic fractures) (Table 7, entry 5).372 CO2 (726 bbl) of was injected for 4 days
11
12
13
into the Middle Bakken at pressures below the fracturing pressure, soaked for 15 days, and then
14
15 fluid samples were collected, monitored, and analyzed during the flowback. Even though the
16
17 injectivity was low, they found that CO2 could penetrate the Middle Bakken and mobilize oil from
18
19
the matrix. Analysis of the produced oil showed that lighter hydrocarbons were produced
20
21
22 preferentially,372 which supports diffusion as the recovery mechanism.80
23
24 Natural Gas EOR pilot tests in the U.S. Bakken Formation. In 2014, the injection well of
25
26 a previous CO2 EOR pilot which suffered from early breakthrough (Table 7, entry 3) was instead
27
28
29 used to test the continuous injection of an enriched natural gas (Table 7, entry 6).26 The gas was
30
31 composed of 55% methane, 10% nitrogen, and 35% C2+, and was injected at a rate of 1,600
32
33 Mscf/day. All offset wells showed oil production enhancement after a few months. However, the
34
35
36
interpretation of the favorable results was complicated by frac-hits and activities in nearby wells
37
38 that may have caused the increase in the oil production. The gas flood test was shut down after 55
39
40 days. This pilot demonstrated the possibility of using rich natural gas as the injection fluid in the
41
42
Bakken Formation.
43
44
45 In 2014, EOG achieved a production increase in another well after injecting water followed
46
47 by field gas into the Middle Bakken (Table 7, entry 7).25 Their goal was to evaluate the feasibility
48
49 and production performance of injecting produced gas into the Bakken. The cumulative injected
50
51
52 field gas volume was 89,000 Mcf. No post-test production data was available.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
104
Page 105 of 170 Energy & Fuels

1
2
3 Chemical EOR pilot tests in the U.S. Bakken Formation. In 1994, Meridian Oil Company
4
5
6 conducted a fresh water injection in the Upper Bakken Shale in North Dakota (Table 7, entry 8).25
7
8 The injection lasted for 50 days. No increase in oil production was observed and the test was
9
10 considered to be unsuccessful.
11
12
13
A CO2 EOR project started by EOG in 2008 was switched to a huff-n-puff injection using
14
15 produced water in 2012 (Table 7, entry 9).381, 26 The project took place in the Parshall Field in
16
17 North Dakota. Water was injected at a rate of about 1,200 bbl/day. The operators used two injection
18
19
cycles of almost one-month, soaking periods of about two weeks, and production intervals of 3-4
20
21
22 months. No significant amount of additional oil was produced after each cycle. An increase in
23
24 production was observed about one year after both injection cycles, however this was attributed to
25
26 frac-hits or well-bashing caused by the hydraulic fracturing of another nearby well.
27
28
29 In 2012, a continuous waterflooding pilot was conducted for eight months in North Dakota,
30
31 where water was injected at a rate of about 1,350 bbl/day (Table 7, entry 10).26 The bottomhole
32
33 pressure was elevated to 6000 psi (41.4 MPa), but no enhancement in the oil production was
34
35
36
observed at any offset wells. The water partially remained in the reservoir and provided a pressure
37
38 support with no improvement in the oil sweep.
39
40 In 2014, a continuous waterflood pilot test was carried out in Montana at an injection rate
41
42
of about 1,700 bbl/day for three months and then 900 bbl/day for four months (Table 7, entry 11).26
43
44
45 Water and oil production rates were enhanced at almost all the surrounding wells but was attributed
46
47 to the frac-hits. One of the offset wells was shut in for a couple of months in early 2015 and its oil
48
49 production rate increased after opening, which was attributed to the waterflood, since no
50
51
52 surrounding well was being fractured at the time.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
105
Energy & Fuels Page 106 of 170

1
2
3 A field trial of a “Production Enhancer” (PE), which appears to be a non-ionic surfactant-
4
5
6 based aqueous solution, was reported in 2018 (Table 7, entry 12).241 Solutions of “Production
7
8 Enhancer A” and “Production Enhancer B” were designed for the Middle Bakken and Niobrara
9
10 Formations, which are characterized by high temperature (> 100 °C), high salinity (> 220,000 ppm
11
12
13
TDS) and hardness (> 15,000 ppm). The huff-n-puff field trial was conducted in a 10,000 ft long
14
15 hydraulically fractured horizontal well the Middle Bakken Formation after 2.5 years of primary
16
17 production. A 1,500-ppm solution was prepared in low salinity Bakken brine (obtained by dilution
18
19
of the Bakken brine with fresh water) and allowed to soak during a four-month long shut-in period.
20
21
22 A distinct increase in oil production (relative to the extrapolated decline curve from primary
23
24 production) occurred over the following 1.5 years. It was estimated that the additional cumulative
25
26 oil production increased by 25% compared to the cumulative oil recovery that would have resulted
27
28
29 solely from continued primary production.
30
31 A pilot test of an ethane-water-surfactant co-injection was reported in 2019 (Table 7, entry
32
33 13).188 The mixture generated a foam which improved mobility control. Oil production rates were
34
35
36
doubled, and the gas utilization ratio and volumetric sweep were improved. Over 2000 bbl of
37
38 incremental oil was produced over 11 weeks. This pilot test represents a combined EOR strategy
39
40 (surfactant and gas injection) which could prove beneficial in other formations.
41
42
6.2 EOR pilot tests in the Canadian Bakken Formation. A pilot test in the Canadian
43
44
45 Bakken was reported in which dry natural gas was injected at a rate of 350-1,000 mcf/day (Table
46
47 7, entry 14).373 The well had an average permeability below 1 mD, and 42˚ API oil. Oil production
48
49 increased from 135 bbl/day to 295 bbl/day during a 1-year period.
50
51
52 Eight Canadian Bakken wells with Toe-Heel injection patterns were subjected to
53
54 waterflooding (Table 7, entry 15).14, 273 The spacings between the injection and production wells
55
56
57
58
59
60 ACS Paragon Plus Environment
106
Page 107 of 170 Energy & Fuels

1
2
3 were only 200 feet, which is much shorter than those of the U.S. Bakken. The oil production rate
4
5
6 was increased from 75 bbl/day to 550 bbl/day by waterflooding. The short spacing and
7
8 comparatively higher permeability of these formations were suggested to be the reason behind the
9
10 success of some of these pilot tests.
11
12
13
6.3 EOR pilot tests in the Eagle Ford Formation. EOG has conducted many pilot tests
14
15 in Eagle Ford in the last decade. One brief summary of Eagle Ford EOR activities noted that four
16
17 successful pilots injecting produced gas into 15 horizontal wells consistently resulted in significant
18
19
oil production increase of about 1.3-1.7 times of the primary recovery (Table 7, entry 16).15 This
20
21
22 success was followed by another 32-well program that produced 80 bbl/day for an average finding
23
24 cost of $6 per barrel.10 EOG has never released details of their technique and results.
25
26 Seven huff-n-puff pilot tests were performed in the Eagle Ford Formation using natural gas
27
28
29 with different compositions.374 These pilot tests were reported by Hoffman, and are referred to as
30
31 Pilot tests A-G. Pilot test A started in late 2012 with an isolated single well. Lean natural gas (90-
32
33 95% methane) was injected in three cycles at a rate around 2,000-3,000 Mscf/day with 6000 psi
34
35
36
(41.4 MPa) injection pressure (Table 7, entry 17).24 Each cycle led to increases in oil production,
37
38 but the interpretation of the data was complicated by more lease wells coming on line.
39
40 Pilot tests B and C were multi-well pilots started in early 2015 with lease oil-production
41
42
rate of 370 bbl/day and 1,065 bbl/day for Pilots B and C, respectively (Table 7, entries 18 and
43
44
45 19).374 Pilot test B resulted in a 17% increase in cumulative production with 1.5 years of injection
46
47 and then producing for another year without injection. Pilot test C consisted of eight horizontal
48
49 wells with other wells running perpendicular at the toe and heel of the original eight wells. Pilot C
50
51
52 resulted in 20% increase in cumulative production after huff-n-puff injection for 2.5 years. Pilot D
53
54 started in 2015 with four huff-n-puff wells into which natural gas was injected (Table 7, entry
55
56
57
58
59
60 ACS Paragon Plus Environment
107
Energy & Fuels Page 108 of 170

1
2
3 20).374 This pilot produced 300,000 bbl in three years. The huff-n-puff injection resulted in a 30%
4
5
6 increase in incremental oil recovery.
7
8 Pilot tests E and F were both single well huff-n-puff pilots that were performed in 2015 by
9
10 a different operator using a richer gas (70% methane, 30% C2+) (Table 7, entries 21 and 22).374
11
12
13
The injection with a rate of 2,000-2,500 Mscf/day was planned for one month, followed by a short
14
15 soaking period and about one month of production. The pilot included two injection periods, but
16
17 the amount of incremental production could not be determined from the reported results because
18
19
oil production was reported for the entire lease, which also included three other wells in lease E
20
21
22 and 60 wells in lease F. Pilot G began as an extension of Pilot test C, which was extended to
23
24 include two offset leases (Table 7, entry 23). Early reports of incremental oil recovery were
25
26 promising.
27
28
29 6.4 EOR pilot test in Argentina. A water huff-n-puff field trial in a three-stage vertical
30
31 well in the lower portion of the Vaca Muerta Formation in Argentina involved the injection of
32
33 ~9,400 bbl water, followed by a 50-day soak (Table 7, entry 24).375 Oil was detected at the
34
35
36
wellhead 12 days after the well was shut-in, and oil percolation into the wellbore was indirectly
37
38 monitored by measuring the hydrostatic head between the wellhead and bottom hole that decreased
39
40 as water was imbibed into the formation while lower density oil moved into the wellbore. At the
41
42
time of the report, the well was still shut-in, and the only oil production was associated with a brief
43
44
45 two day-long flowback test that had an oil cut more than 50%.375, 382
46
47 6.5 EOR pilot tests in China. Several CO2 EOR pilot tests were conducted in China. In
48
49 the Song-Fang-Dun field in China, which has an average porosity of 12% and permeability of 0.79
50
51
52 mD, CO2 EOR pilot tests were performed with one injector and five producers in 2003 (Table 7,
53
54 entry 25).371, 376 Later, the project was expanded to 14 injectors and 26 producers.376 By 2015, 1.5
55
56
57
58
59
60 ACS Paragon Plus Environment
108
Page 109 of 170 Energy & Fuels

1
2
3 million bbl of CO2 were injected, and 66,000 bbl of oil were produced. Even though some
4
5
6 additional oil was recovered, production was low given the large amount of CO2 injected. The low
7
8 oil recovery was attributed to the heterogeneity of the reservoir.
9
10 In the Yu-Shu-Lin field, which has an average permeability of 0.96 mD, a waterflooding
11
12
13
was performed, followed by CO2 injection using 7 injectors and 17 producers (Table 7, entry
14
15 26).377 A total of 810,000 bbl CO2 were injected, and 403,000 bbl of oil were produced. The CO2
16
17 injectivity was found to be more than 4 times the water injectivity. No fracturing occurred during
18
19
CO2 injection. Oil recovery following CO2 injection was estimated to be 21%, compared to 12%
20
21
22 from waterflooding. Following the success of this pilot, the test was later expanded to have 70
23
24 injectors and 140 producers.
25
26 Other EOR activities in China involved water injection using different injection modes.
27
28
29 For example, pulsed injection, which involves changing the injection rate at some cycles of time
30
31 to generate elastic energy was implemented (Table 7, entry 27).371, 377 This strategy resulted in a
32
33 sharp increase in production after the shut-in.377 Another test involved asynchronous injection—
34
35
36
keeping the injectors open and producers shut-in and vice versa—which increased the oil rate from
37
38 26 bbl/day 40 bbl/day in a field with only 0.17 mD permeability (Table 7, entry 28).377 In another
39
40 test, a huff-n-puff injection resulted in an increase in the oil recovery, especially in the neighboring
41
42
wells (Table 7, entry 29).378 A test of a surfactant flood was performed in a field with 0.54 mD
43
44
45 permeability and 8% porosity. Oil production rate increased after 7 days of soaking (Table 7, entry
46
47 30).379 Details regarding the surfactant formulation were not included in the report. Finally, a
48
49 waterflooding test led to an increase in the oil production rate from 48 bbl/day to 269 bbl/day after
50
51
52 7 days of injection (Table 7, entry 31).380
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
109
Energy & Fuels Page 110 of 170

1
2
3 6.6 Conclusions from EOR pilot tests. Unfortunately, reported field tests typically do not
4
5
6 reflect the high oil recoveries predicted by simulations. Even though injectivity was not an issue
7
8 with any of the pilot tests, in most cases, the injected fluid did not produce significant improvement
9
10 in oil recovery. High injectivity was attributed to flow through hydraulic fractures, which led to
11
12
13
conformance problems and early breakthrough. Another primary difficulty associated with CO2
14
15 EOR in shale is the high heterogeneity in the rock lithology. Further, in most of the wells, the
16
17 interpretation of results was muddled by fracturing activities in nearby wells.
18
19
CO2 injection—primarily in the Bakken Formation—did not result in an improvement in
20
21
22 production. Natural gas injection—primarily in the Eagle Ford Formation—on the other hand, did
23
24 give moderate increases in oil production. It is not clear at this time whether the more promising
25
26 EOR results obtained with natural gas in the Eagle Ford indicate that natural gas is a better fluid
27
28
29 for EOR than CO2. Rather, the differences may be attributable to differences between the Bakken
30
31 and Eagle Ford Formations and the design of the EOR pilots test themselves (e.g. the well patterns,
32
33 the ability to isolate the wells, conformance control problems). Both natural gas and CO2 should
34
35
36
both continue to be considered for EOR in ULRs.
37
38 The improvement of CO2 or natural gas EOR in ULR will undoubtedly require
39
40 confirmation from more field trials than those found in Table 7. In addition to improving the
41
42
timing, location, pattern layout, fluid type, and economics for an effective pilot test, it is necessary
43
44
45 to find a solution for conformance control problem.26, 24 Well patterns, soak times, and injection
46
47 and production modes must be optimized. In the future, the application of combined strategies—
48
49 such as the use of CO2 and aqueous surfactant solutions—could be economic.371
50
51
52 7.0 Fluid selection, key findings and significant recommendations.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
110
Page 111 of 170 Energy & Fuels

1
2
3 Fluid selection. Given the numerous strategies available for EOR in ULRs, one must match
4
5
6 the appropriate method to each particular formation. Alfarge et al. provided guidelines for
7
8 consideration of thermal, chemical, microbial, gas injection and water-based EOR methods.13
9
10 Thermal methods (e.g. steam injection) commonly used to reduce viscosity of heavy oils are
11
12
13
inappropriate for light shale oils. Techniques involving high molecular weight polymers (e.g.
14
15 biopolymeric microbial solutions or synthetic polymer solutions) are also poorly suited for EOR
16
17 in ULRs due to the small pore sizes of shale matrices. Several other techniques—such as alkaline
18
19
chemical solutions and bio-surfactants—were judged to be premature for pilot-scale consideration
20
21
22 because they had not even been studied thoroughly in the laboratory.13 Ling et al. noted that
23
24 aqueous surfactant injection is preferable for matrix permeabilities greater than 10 mD—and we
25
26 agree with this recommendation.383
27
28
29 In selecting which fluid to use in low-permeability ULRs (less than 10 mD), one can choose
30
31 between propane, ethane, CO2, natural gas, methane, and nitrogen. Propane and ethane are
32
33 excellent solvents for oil, and before dismissing these fluids based on cost, one should determine
34
35
36
if either (especially ethane) may actually present the most favorable logistics and economics.115
37
38 Because CO2 and rich natural gas generally yield comparable, very good performance at the
39
40 laboratory-scale, the selection of one gas over the other is almost exclusively governed by their
41
42
availability at the location of interest. Lean natural gas is a weaker solvent for EOR, yet it may be
43
44
45 the only fluid available for consideration. Nitrogen is also a poor solvent for oil, but there are
46
47 experiments showing that its oil extraction ability is comparable to that of methane.137 If used
48
49 above their MMP values, nitrogen and methane could be effective. If none of these gas-based
50
51
52 strategies are viable at a particular location, then one can consider LSW or aqueous surfactant
53
54 solutions, especially if the matrix permeability is at the high end of permeabilities associated with
55
56
57
58
59
60 ACS Paragon Plus Environment
111
Energy & Fuels Page 112 of 170

1
2
3 unconventional formations. In addition to selecting the appropriate fluid for the EOR project, one
4
5
6 may need to consider conformance control strategies (e.g. co-injection of gas and water, gas-in-
7
8 water foams, gels).383 A recent industry-first pilot test of a hydrocarbon gas (ethane)-in-water
9
10 conformance control foam yielded better conformance control than the injection of water and
11
12
13
gas.188
14
15 Key Findings:
16
17 • In miscible gas EOR in ULRs, oil is extracted from pores via a slow diffusion process
18
19
during the “huff” period, followed by expansion during the “puff” period. Miscible gas EOR in
20
21
22 unconventional liquid reservoirs (ULRs) is attained through a combination of possible
23
24 mechanisms, including re-pressurization, injection-induced fracturing, diffusion of CO2 into
25
26 matrix pores, CO2-oil interfacial tension (IFT) reduction, oil swelling, oil viscosity reduction,
27
28
29 wettability alteration toward increasing water wetness, vaporization of lighter oil components,
30
31 relative permeability hysteresis, and solution gas drive. In tight reservoirs where gas cannot flow
32
33 through the porous rock matrix, diffusion is the dominant mechanism by which gas moves into
34
35
36
pores and oil moves out of pores. Therefore, the following strategies can improve recovery for
37
38 CO2 huff-n-puff after primary recovery has reached very low rates: (1) exploit and increase the
39
40 extent of hydraulic fractures and natural fracture networks to increase the ratio of the matrix
41
42
surface area exposed to fractures to the volume of matrix bounded by fractures, (2) increase
43
44
45 injection pressures until the MMP is attained or exceeded, and (3) optimize the schedule and
46
47 conditions for multiple cycles of CO2 injection, soak time (typically on the order of a few weeks
48
49 to a few months), and production.
50
51
52 • CO2 and natural gas (rich natural gas) are the most promising high-pressure gaseous
53
54 solvents for EOR in ULR. Highest oil recovery is more likely to be attained with fluids that exhibit
55
56
57
58
59
60 ACS Paragon Plus Environment
112
Page 113 of 170 Energy & Fuels

1
2
3 a high degree of miscibility with oil. The best solvents for oil are (from strongest to weakest)
4
5
6 propane, ethane, CO2, methane and nitrogen. Despite their excellent solvent strength for oil,
7
8 propane and ethane are rarely used in laboratory studies because in most situations the economics,
9
10 constraints and logistics of using produced natural gas or CO2 in the field would be more favorable.
11
12
13
CO2 has demonstrated high (90-100%) oil recoveries in huff-n-puff experiments. Methane, the
14
15 main constituent of natural gas, is a far weaker solvent for oil than CO2, therefore laboratory-scale
16
17 comparisons of CO2 vs. methane for EOR usually indicate that CO2 is a superior solvent for oil
18
19
recovery from shale. However, natural gas—which contains roughly 15-25+% of oil-miscible C2+
20
21
22 hydrocarbons—performs comparably to CO2 in many laboratory studies. Nitrogen exhibits much
23
24 lower miscibility with oil; however, if extremely high-pressure conditions are encountered, then
25
26 gases with an MMP lower than the formation pressure, including nitrogen, may yield reasonable
27
28
29 oil recovery performance. Laboratory-scale EOR experiments for each of these fluids have shown
30
31 promising results.
32
33 CO2 and natural gas have a variety of advantages for EOR in ULRs. Advantages of CO2
34
35
36
include its high solvent strength, pressure-adjustable density and solvent strength, non-
37
38 flammability, and availability as a cold, saturated liquid in refrigerated trucks that can be delivered
39
40 to remote locations. Natural gas also typically has an even lower viscosity than CO2, and the
41
42
logistics of using produced natural gas from nearby wells for EOR are generally more favorable
43
44
45 than the truck delivery of CO2 or other gases. Therefore, because neither fluid is clearly superior
46
47 to the other, both CO2 and a produced natural gas should be considered for both laboratory-scale
48
49 and field-scale evaluation.
50
51
52 • In laboratory experiments, miscible gas injection tends to recover more oil than chemical
53
54 EOR injections. However, interest will persist in the lower cost, water-based EOR in ULR,
55
56
57
58
59
60 ACS Paragon Plus Environment
113
Energy & Fuels Page 114 of 170

1
2
3 especially chemical EOR in which adding salt and surfactants to the water to make the shale more
4
5
6 water-wet and/or reduce water-oil IFT. Although miscible gases (CO2, natural gas) give higher
7
8 EOR recoveries in the laboratory than aqueous fluids, a very large body of work exists and
9
10 continues for water-based fluids due to their low cost, logistical ease, and the ability to design
11
12
13
chemical EOR formulations without the high-pressure equipment required to confine natural gas
14
15 or CO2 in a dense state. Further, water remains the predominant fluid of choice for hydraulic
16
17 fracturing, and many of the same additives used to improve the performance of water during the
18
19
hydraulic fracturing process are applicable to the use of water for EOR. The logistics and cost of
20
21
22 using water for EOR are also more favorable than the logistics and costs of high-pressure gases,
23
24 most notably CO2 and nitrogen. Water is a very strong solvent for many ionic and polar chemicals,
25
26 and the addition of salt and wettability-altering surfactants (mixed- or oil-wet altered to become
27
28
29 more water-wet) and/or oil-water IFT-reducing surfactants are the most commonly touted water-
30
31 soluble additives that can be used to improve EOR. The design of these aqueous solutions is not
32
33 straightforward; there is evidence that oil recovery via imbibition is improved by inducing shifts
34
35
36
in wettability while minimizing IFT reduction. However, laboratory-scale studies have also shown
37
38 that IFT-reducing surfactants can also result in better oil recovery is some situations. Therefore,
39
40 it is not uncommon for numerous aqueous formulations with wide ranges of salt type, salt
41
42
concentration, surfactant type, and surfactant concentrations to be assessed in laboratory-scale
43
44
45 studies of EOR in ULR.
46
47 • Pilot-scale EOR results are inherently less successful than laboratory-scale results
48
49 primarily because of the small ratio of rock surface area to rock volume in the field, and the serious
50
51
52 conformance control issues that can occur in the field. Small rock samples or cores have a large
53
54 ratio of surface area to volume. This allows the injected gas (CO2 or natural gas) to access a
55
56
57
58
59
60 ACS Paragon Plus Environment
114
Page 115 of 170 Energy & Fuels

1
2
3 relatively large amount of surface area while reducing the distance that the gas must diffuse to
4
5
6 penetrate within the relatively small rock. As a result, oil recovery can be very high in the
7
8 laboratory, even for short tests. For example, 24-h CO2 huff-n-puff at pressures near or above the
9
10 MMP using small, unconfined 1 cm diameter, 4 cm long Bakken cores can result in oil recovery
11
12
13
over 90%. However, larger unconfined 4 cm diameter, 5 cm long Bakken cores have oil recoveries
14
15 of only 65 - 72%. When these larger Bakken cores are confined such that only one flat end of the
16
17 core is exposed to CO2, the area to volume ratio attains an even lower value and oil recovery drops
18
19
to about 25%. The ratio of shale surface area to volume in the field is governed by the
20
21
22 characteristics and spacing of hydraulic fractures and natural fractures. This area to volume ratio
23
24 can be orders of magnitude smaller in the field than in the laboratory. Another problem observed
25
26 in EOR in ULR field tests relates to conformance control. Injected fluid quickly finds fast paths
27
28
29 through to nearby wells through high-permeability fractures rather than soaking through the shale
30
31 matrix. Conformance control issues can even prevent the operator from attaining the desired MMP
32
33 during EOR. As a result, CO2 injectivity may be high, but oil recovery can be low because the CO2
34
35
36
is unable to attain its desired pressure and/or the CO2 may be flowing toward and into another well
37
38 rather than flowing or diffusing into the shale matrix. The ability to increase the surface area to
39
40 volume ratio and improve conformance control could make EOR in ULRs much more successful.
41
42
• Initial EOR field tests in the US Bakken have not yet performed well using either CO2 or
43
44
45 natural gas. However, EOR in the Eagle Ford utilizing produced natural gas has enjoyed recent
46
47 success (CO2 pilots in the Eagle Ford Formation have not yet been reported). Results from pilot
48
49 tests of natural gas EOR in the Eagle Ford Formation have been more promising than the results
50
51
52 of pilot tests of CO2 EOR and natural gas EOR in the Bakken Formation. Because of the
53
54 availability of produced natural gas in the Eagle Ford, no pilots using CO2 have yet been reported
55
56
57
58
59
60 ACS Paragon Plus Environment
115
Energy & Fuels Page 116 of 170

1
2
3 in that field. Although some may have inferred that these field tests indicate natural gas may be a
4
5
6 better solvent for EOR in ULRs, it is more likely that these field tests indicate that the Eagle Ford
7
8 Formation is more amenable to gas-based EOR than the Bakken Formation. The differences
9
10 between these fields may be related to permeability, reservoir heterogeneity, fracture density, the
11
12
13
design of well patterns, conformance control issues, and the ability of activities at neighboring
14
15 wells on the oil production from the pilot well. We anticipate that CO2 EOR pilots in the Eagle
16
17 Ford will soon be reported in the future, and the results of these trials will provide an important
18
19
metric for the comparison of CO2 and natural gas. A final point is that most EOR attempts in
20
21
22 unconventional formations have been a modification from EOR for conventional reservoirs. New
23
24 approaches can be developed in the future as our understanding of unconventional formations
25
26 improve.
27
28
29 • CO2 EOR in ULRs reduce the carbon intensity of the oil produced by associated CO2
30
31 storage. A significant benefit of using CO2 for EOR in ULRs is that the suppliers providing CO2
32
33 for the project can obtain a large portion of the CO2 from anthropogenic sources. Because a
34
35
36
significant portion of the injected CO2 will remain in the formation after enhanced oil production
37
38 is stopped, CO2 EOR will also serve to store anthropogenic CO2. The CO2 retained by the shale or
39
40 tight formation will be stored as either an adsorbed gas in the organic material, a free gas phase,
41
42
gas dissolved in unrecovered oil, gas dissolved in the brine, or carbonates that slowly form as CO2
43
44
45 and minerals react. Although volumes of CO2 stored may not be great enough to create net zero
46
47 emission oil, storage should decrease the carbon intensity associated with the shale oil production,
48
49 as determined by a life cycle analysis.
50
51
52 Significant Recommendations:
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
116
Page 117 of 170 Energy & Fuels

1
2
3 • The acquisition of field cores “from depth” and reservoir crude oil (rather than outcrop
4
5
6 cores) would improve the reliability and of laboratory-scale results. Laboratory experiments
7
8 would be improved by better access to relatively large “at depth” cores (~2.5 cm in diameter, ~5.0
9
10 cm long) from oil-producing shales and reservoir fluids (e.g. crude oil) obtained from the same
11
12
13
formation (rather than using outcrop cores that have been aged and saturated with high pressure
14
15 crude oil). Although larger reservoir cores could be preferable for EOR studies, reservoir core
16
17 plugs are commonly cut from the butt end of a 4" diameter core, which limits the length of the
18
19
core. Also, the standard diameter of most service labs is 1”. Therefore, reservoir cores that are 1”
20
21
22 in diameter and 2.0-3.0 inches long are typical.
23
24 • Huff-n-puff experiments should not only be conducted with small chips immersed in CO2
25
26 or small or relatively large core samples surrounded by CO2 in high permeability flow paths, but
27
28
29 also with relatively large confined cores that have a relatively small amount of surface (e.g. one
30
31 side) exposed to CO2; in this manner a wide range of surface area to volume ratio can be explored.
32
33 Oil recovery experiments conducted with very small prices of porous media surrounded by CO2
34
35
36
are informative. However, if possible, these experiments should be complemented with laboratory-
37
38 scale huff-n-puff with relatively large cores that are retained in a core holder such that both or only
39
40 one circular end of the core is exposed to the CO2. In this manner, the effect of a wide range of
41
42
surface area/volume values on oil recovery can be determined. This is important because the low
43
44
45 ratio of surface area to volume in the field has been cited as a possible reason for disappointing
46
47 levels of oil recovery in pilot tests.
48
49 • Laboratory-studies and pilot tests using both CO2 and natural gas should continue; there
50
51
52 is no reason to favor one over the other based on laboratory-scale or pilot-scale results to date.
53
54 In general, the CO2 pilots in the Bakken were disappointing in that very little incremental oil
55
56
57
58
59
60 ACS Paragon Plus Environment
117
Energy & Fuels Page 118 of 170

1
2
3 production was reported. In general, the recently reported results from the natural gas EOR in the
4
5
6 Eagle Ford Formation are much more promising. However, there are no reports of CO2 huff-n-
7
8 puff in the Eagle Ford for comparison, probably because of the current widespread availability of
9
10 produced natural gas and lack of CO2 source in this region. The Bakken Formation appears to
11
12
13
challenge gas-based EOR projects. The success of natural gas huff-n-puff in the Eagle Ford does
14
15 not indicate that natural gas is better than CO2 for EOR in that field. Only future tests of CO2 in
16
17 the Eagle Ford would provide a clear comparison of these two fluids in a field that appears to be a
18
19
good target for EOR. We therefore suggest that it would be wise to continue exploring both CO2
20
21
22 and natural gas for EOR in ULR given their very similar performance in nearly all laboratory-scale
23
24 experiments and most field-scale simulation studies.
25
26 • Due to heterogeneous geology, numerous oil recovery mechanisms, complex fracture
27
28
29 networks, uncertainty in physical parameter values, and a small number of field results for history
30
31 matching, numerical simulations of CO2 or natural gas EOR should be used only as qualitative
32
33 guides for the design of pilot-tests. The simulation of EOR in ULR is extremely challenging,
34
35
36
especially in comparison to the relatively mature state-of-the-art for modelling EOR in
37
38 conventional formations. There are very few detailed reports on pilot-scale field tests in
39
40 unconventional formations using high pressure gases, and oftentimes the production data from
41
42
these pilots is influenced by activities at nearby wells, which complicates interpretation. Further,
43
44
45 there are multiple mechanisms at play in a complex geologic setting. The characteristics of the
46
47 hydraulic and natural fractures, which are critically important to the oil recovery, are often not
48
49 clearly understood. Further the presence of heterogeneities and high-permeability flow paths can
50
51
52 lead to conformance control issues. Some of the important parameters that exert a tremendous
53
54 influence on simulation results, such as the diffusion coefficient of gases, the permeability of the
55
56
57
58
59
60 ACS Paragon Plus Environment
118
Page 119 of 170 Energy & Fuels

1
2
3 shale matrix, and the characteristics of the natural fracture network, are often not well specified.
4
5
6 For these reasons, many simulation studies invoke sensitivity studies such that a wide range or
7
8 parameter values can be considered and useful recommendations for future field-scale EOR
9
10 projects can be made based on the general trends that are observed. Despite these challenges,
11
12
13
simulations studies should continue because it is not possible to use laboratory-scale and pilot scale
14
15 experimental results alone to accurately understand or predict the oil recovery process at the pilot-
16
17 scale and full field scale.
18
19
20
21
22 AUTHOR INFORMATION
23
24
25
Corresponding Author
26
27 *Angela Goodman, angela.goodman@netl.doe.gov, *Robert M. Enick, rme@pitt.edu
28
29
30
Disclaimer
31
32
33 This project was funded by the Department of Energy, National Energy Technology Laboratory,
34
35 an agency of the United States Government, through a support contract with Leidos Research
36
37 Support Team. Neither the United States Government nor any agency thereof, nor any of their
38
39
40 employees, nor Leidos Research Support Team, nor any of their employees, makes any warranty,
41
42 expressed or implied, or assumes any legal liability or responsibility for the accuracy,
43
44 completeness, or usefulness of any information, apparatus, product, or process disclosed, or
45
46
47
represents that its use would not infringe privately owned rights. Reference herein to any specific
48
49 commercial product, process, or service by trade name, trademark, manufacturer, or otherwise,
50
51 does not necessarily constitute or imply its endorsement, recommendation, or favoring by the
52
53
United States Government or any agency thereof. The views and opinions of authors expressed
54
55
56
57
58
59
60 ACS Paragon Plus Environment
119
Energy & Fuels Page 120 of 170

1
2
3 herein do not necessarily state or reflect those of the United States Government or any agency
4
5
6 thereof.
7
8
Author Contributions
9
10
11 The manuscript was written through contributions of all authors. All authors have given approval
12
13 to the final version of the manuscript.
14
15
16
17 Funding Sources
18
19 This technical effort was performed in support of the National Energy Technology Laboratory’s
20
21 ongoing research under Award Number 1610260 for Environmentally Prudent Development
22
23
24
25
26 ABBREVIATIONS
27
28 Bbl, barrel; C2H6, Ethane; CNF, complex nanofluid; CO2, carbon dioxide; CH4, methane; CT,
29
30
31 computerized tomography; EERC, Energy and Environmental Research Center; EO, ethylene
32
33 oxide; EOR, enhanced oil recovery; GOR, gas:oil ratio; IFT, interfacial tension; LSW, low salinity
34
35 water; MMP, minimum miscibility pressure; N2, nitrogen; NGL, natural gas liquids; NMR, nuclear
36
37
38
magnetic resonance; NPS, nanoparticle surfactant; OOIP, original oil in place; ppm, parts per
39
40 million; psi, pounds per square inch; scf, standard cubic foot; TDS, total dissolved solids; ULRs,
41
42 unconventional liquid reservoirs; WAG, water alternating gas.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
120
Page 121 of 170 Energy & Fuels

1
2
3 TOC Graphic
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 REFERENCES
39
40
41
1. Soeder, D. J., The Successful Development of Gas and Oil Resources from Shales in North
42
43 America. J. Pet. Sci. Eng. 2018, 163, 399-420.
44
45 2. Harju, J. In Enhanced Oil Recovery in North Dakota: Opportunities and Challenges, North
46
47
Dakota Interim Taxation Committee Meeting, Bismark, ND, Center, E. a. E. R., Ed. Bismark, ND,
48
49
50 2016.
51
52 3. Dlouhy, J. Enhanced Oil Recovery Techniques Limited in Shale. Houston Chronicle
53
54 [Online], 2014.
55
56
57
58
59
60 ACS Paragon Plus Environment
121
Energy & Fuels Page 122 of 170

1
2
3 4. von Flatern, R. Tackling the Tight Oil Recovery Challenge Upstream Technology [Online],
4
5
6 2017.
7
8 5. COGA. Oil Shale Vs. Shale Oil The Basics [Online], 2013.
9
10 http://www.coga.org/?cpt_energy=/oil-shale-vs-shale-oil-basics/ (accessed June 18, 2013).
11
12
13
6. Castro-Alvarez, F.; Marsters, P.; Ponce de León Barido, D.; Kammen, D. M., Sustainability
14
15 Lessons from Shale Development in the United States for Mexico and Other Emerging
16
17 Unconventional Oil and Gas Developers. Renewable Sustainable Energy Rev. 2018, 82, 1320-
18
19
1332.
20
21
22 7. Rassenfoss, S., Carbon Dioxide May Offer an Unconventional EOR Option. J. Pet.
23
24 Technol. 2014, 66, 52-56.
25
26 8. Joshi, S., EOR: Next Frontier for Unconventional Oil. J. Pet. Technol. 2014, 66, 20-22.
27
28
29 9. Babadagli, T., Philosophy of EOR. J. Pet. Sci. Eng. 2020, 188, 106930.
30
31 10. Rassenfoss, S., Shale EOR Works, but Will It Make a Difference? J. Pet. Technol. 2017,
32
33 69, 34-40.
34
35
36
11. Balasubramanian, S.; Chen, P.; Bose, S.; Alzahabi, A.; Thakur, G. C., Recent Advances in
37
38 Enhanced Oil Recovery Technologies for Unconventional Oil Reservoirs. Offshore Technology
39
40 Conference 2018, 7, 10.4043/28973-MS.
41
42
12. Sheng, J. J.; Chen, K., Evaluation of the EOR Potential of Gas and Water Injection in Shale
43
44
45 Oil Reservoirs. Journal of Unconventional Oil and Gas Resources 2014, 5, 1-9.
46
47 13. Alfarge, D.; Wei, M.; Bai, B., Ior Methods in Unconventional Reservoirs of North
48
49 America: Comprehensive Review. SPE Western Regional Meeting 2017, 24, 10.2118/185640-MS.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
122
Page 123 of 170 Energy & Fuels

1
2
3 14. Alfarge, D.; Wei, M.; Bai, B., Feasibility of CO2-EOR in Shale-Oil Reservoirs: Numerical
4
5
6 Simulation Study and Pilot Tests. Carbon Management Technology Conference 2017, 28,
7
8 10.7122/485111-MS.
9
10 15. Wang, L.; Tian, Y.; Yu, X.; Wang, C.; Yao, B.; Wang, S.; Winterfeld, P. H.; Wang, X.;
11
12
13
Yang, Z.; Wang, Y.; Cui, J.; Wu, Y.-S., Advances in Improved/Enhanced Oil Recovery
14
15 Technologies for Tight and Shale Reservoirs. Fuel 2017, 210, 425-445.
16
17 16. Jia, B.; Tsau, J.-S.; Barati, R., A Review of the Current Progress of CO2 Injection EOR and
18
19
Carbon Storage in Shale Oil Reservoirs. Fuel 2019, 236, 404-427.
20
21
22 17. Du, F.; Nojabaei, B., A Review of Gas Injection in Shale Reservoirs: Enhanced Oil/Gas
23
24 Recovery Approaches and Greenhouse Gas Control. Energies 2019, 12, 1-33.
25
26 18. Alharthy, N.; Teklu, T. W.; Kazemi, H.; Graves, R. M.; Hawthorne, S. B.; Braunberger, J.;
27
28
29 Kurtoglu, B., Enhanced Oil Recovery in Liquid-Rich Shale Reservoirs: Laboratory to Field. SPE
30
31 Reservoir Eval. Eng. 2018, 21, 137-159.
32
33 19. Alfarge, D.; Wei, M.; Bai, B.; Alsaba, M., Selection Criteria for Miscible-Gases to Enhance
34
35
36
Oil Recovery in Unconventional Reservoirs of North America. SPE Kuwait Oil & Gas Show and
37
38 Conference 2017, 17, 10.2118/187576-MS.
39
40 20. Carpenter, C., A Review of Improved-Oil-Recovery Methods in North American
41
42
Unconventional Reservoirs. J. Pet. Technol. 2018, 70, 42-44.
43
44
45 21. Lau, H. C.; Yu, M.; Nguyen, Q. P., Nanotechnology for Oilfield Applications: Challenges
46
47 and Impact. Abu Dhabi International Petroleum Exhibition & Conference 2016, 17,
48
49 10.2118/183301-MS.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
123
Energy & Fuels Page 124 of 170

1
2
3 22. Idogun, A. K.; Iyagba, E. T.; Ukwotije-Ikwut, R. P.; Aseminaso, A., A Review Study of
4
5
6 Oil Displacement Mechanisms and Challenges of Nanoparticle Enhanced Oil Recovery. SPE
7
8 Nigeria Annual International Conference and Exhibition 2016, 11, 10.2118/184352-MS.
9
10 23. Dang, C. T. Q.; Nguyen, N. T. B.; Chen, Z.; Nguyen, H. X.; Bae, W.; Phung, T. H., A
11
12
13
Comprehensive Evaluation of the Performances of Alkaline/Surfactant/Polymer Flooding in
14
15 Conventional and Unconventional Reservoirs. SPE Asia Pacific Oil and Gas Conference and
16
17 Exhibition 2012, 11, 10.2118/160444-MS.
18
19
24. Hoffman, T. B.; Evans, J., Unconventional Enhanced Oil Recovery Pilot Projects in the
20
21
22 Bakken Formation. Meeting, A. R. M. S. A., Ed. 2017,
23
24 25. Sorensen, J.; Hamling, J., Enhanced Oil Recovery (EOR) in Tight Oil: Lessons Learned
25
26 from Pilot Tests in the Bakken Tight Oil Optimization Workshop, Center, E. a. E. R., Ed. 2015,
27
28
29 26. Hoffman, T. B.; Evans, J. G., Improved Oil Recovery Ior Pilot Projects in the Bakken
30
31 Formation. SPE Low Perm Symposium 2016, 22, 10.2118/180270-MS.
32
33 27. Eshkalak, M. O.; Al-shalabi, E. W.; Sanaei, A.; Aybar, U.; Sepehrnoori, K., Enhanced Gas
34
35
36
Recovery by CO2 Sequestration Versus Re-Fracturing Treatment in Unconventional Shale Gas
37
38 Reservoirs. Abu Dhabi International Petroleum Exhibition and Conference 2014, 18,
39
40 10.2118/172083-MS.
41
42
28. Hasan, M.; Eliebid, M.; Mahmoud, M.; Elkatatny, S.; Shawabkeh, R., Enhanced Gas
43
44
45 Recovery (EGR) Methods and Production Enhancement Techniques for Shale & Tight Gas
46
47 Reservoirs. SPE Kingdom of Saudi Arabia Annual Technical Symposium and Exhibition 2017, 9,
48
49 10.2118/188090-MS.
50
51
52 29. Zhan, J.; Soo, E.; Fogwill, A.; Cheng, S.; Cai, H.; Zhang, K.; Chen, Z., A Systematic
53
54 Reservoir Simulation Study on Assessing the Feasibility of CO2 Sequestration in Liquid-Rich
55
56
57
58
59
60 ACS Paragon Plus Environment
124
Page 125 of 170 Energy & Fuels

1
2
3 Shale Gas Reservoirs with Potential Enhanced Gas Recovery. Offshore Technology Conference
4
5
6 Asia 2018, 14, 10.4043/28395-MS.
7
8 30. Jiang, J.; Shao, Y.; Younis, R. M., Development of a Multi-Continuum Multi-Component
9
10 Model for Enhanced Gas Recovery and CO2 Storage in Fractured Shale Gas Reservoirs. SPE
11
12
13
Improved Oil Recovery Symposium 2014, 17, 10.2118/169114-MS.
14
15 31. Yu, W.; Al-Shalabi, E. W.; Sepehrnoori, K., A Sensitivity Study of Potential CO2 Injection
16
17 for Enhanced Gas Recovery in Barnett Shale Reservoirs. SPE Unconventional Resources
18
19
Conference 2014, 16, 10.2118/169012-MS.
20
21
22 32. Eliebid, M.; Mahmoud, M.; Elkatatny, S.; Abouelresh, M.; Shawabkeh, R., Adsorption
23
24 Role in Shale Gas Recovery and the Feasibility of CO2 in Shale Enhanced Gas Recovery: A Study
25
26 on Shale Gas from Saudi Arabia. SPE Kuwait Oil & Gas Show and Conference 2017, 12,
27
28
29 10.2118/187667-MS.
30
31 33. Xiaoqi, W.; Xu, J.; Jianming, L.; Liang, S.; Xiaodan, L.; Zengqiang, Z., New
32
33 Considerations of Shale Gas CO2-EGR from Molecular Simulation. SPE/IATMI Asia Pacific Oil
34
35
36
& Gas Conference and Exhibition 2017, 9, 10.2118/186406-MS.
37
38 34. Kang, S. M.; Fathi, E.; Ambrose, R. J.; Akkutlu, I. Y.; Sigal, R. F., Carbon Dioxide Storage
39
40 Capacity of Organic-Rich Shales. SPE J. 2011, 16, 842-855.
41
42
35. Bodini, S. A.; Forni, L. P.; Tuero, F.; Crotti, M. A.; Labayen, I. L., Unconventional EOR:
43
44
45 Field Tests Results in Vaca Muerta Shale Play: A Capillary Based Improved Oil Recovery Case
46
47 Study for Shale/Tight Oil Scenarios. SPE Argentina Exploration and Production of
48
49 Unconventional Resources Symposium 2018, 12, 10.2118/191877-MS.
50
51
52 36. Soliman, M. Y.; Boonen, P., Review of Fractured Horizontal Wells Technology. Abu
53
54 Dhabi International Petroleum Exhibition and Conference 1996, 17, 10.2118/36289-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
125
Energy & Fuels Page 126 of 170

1
2
3 37. King, G. E., Hydraulic Fracturing 101: What Every Representative, Environmentalist,
4
5
6 Regulator, Reporter, Investor, University Researcher, Neighbor and Engineer Should Know About
7
8 Estimating Frac Risk and Improving Frac Performance in Unconventional Gas and Oil Wells. SPE
9
10 Hydraulic Fracturing Technology Conference 2012, 80, 10.2118/152596-MS.
11
12
13
38. Bunger, A.; Lecampion, B.; Feng, X.-T., Four Critical Issues for Successful Hydraulic
14
15 Fracturing Applications. In Rock Mechanics and Engineering, Feng, X.-T., Ed. CRC Press: 2017.
16
17 39. Gandossi, L.; Von Estorff, U. An Overview of Hydraulic Fracturing and Other Formation
18
19
Stimulation Technologies for Shale Gas Production: Update 2015; 9789279538940; EUR 26347:
20
21
22 2015.
23
24 40. Alfarge, D.; Wei, M.; Bai, B., Evaluating the Performance of Hydraulic-Fractures in
25
26 Unconventional Reservoirs Using Production Data: Comprehensive Review. J. Nat. Gas Sci. Eng.
27
28
29 2019, 61, 133-141.
30
31 41. Roussel, N. P.; Sharma, M. M., Optimizing Fracture Spacing and Sequencing in
32
33 Horizontal-Well Fracturing. SPE Prod. Oper. 2011, 26, 173-184.
34
35
36
42. Al-Arfaj, M. K.; Amanullah, M.; Sultan, A. S.; Hossain, M. E.; Abdulraheem, A., Chemical
37
38 and Mechanical Aspects of Wellbore Stability in Shale Formations: A Literature Review. Abu
39
40 Dhabi International Petroleum Exhibition and Conference 2014, 11, 10.2118/171682-MS.
41
42
43. Wutherich, K.; Walker, K. J., Designing Completions in Horizontal Shale Gas Wells:
43
44
45 Perforation Strategies. SPE Americas Unconventional Resources Conference 2012, 10,
46
47 10.2118/155485-MS.
48
49 44. McLennan, J. D.; Green, S. J.; Bai, M., Proppant Placement During Tight Gas Shale
50
51
52 Stimulation: Literature Review and Speculation. The 42nd U.S. Rock Mechanics Symposium
53
54 (USRMS) 2008, 14.
55
56
57
58
59
60 ACS Paragon Plus Environment
126
Page 127 of 170 Energy & Fuels

1
2
3 45. Lecampion, B.; Bunger, A.; Zhang, X., Numerical Methods for Hydraulic Fracture
4
5
6 Propagation: A Review of Recent Trends. J. Nat. Gas Sci. Eng. 2018, 49, 66-83.
7
8 46. Karantinos, E.; Sharma, M. M.; Ayoub, J. A.; Parlar, M.; Chanpura, R. A., Choke
9
10 Management Strategies for Hydraulically Fractured Wells and Frac–Pack Completions in Vertical
11
12
13
Wells. SPE International Conference and Exhibition on Formation Damage Control 2016, 20,
14
15 10.2118/178973-MS.
16
17 47. Potapenko, D. I.; Williams, R. D.; Desroches, J.; Enkababian, P.; Theuveny, B.; Willberg,
18
19
D. M.; Moncada, K.; Deslandes, P.; Wilson, N.; Neaton, R.; Mikovich, M.; Han, Y.; Conort, G.,
20
21
22 Securing Long-Term Well Productivity of Horizontal Wells through Optimization of
23
24 Postfracturing Operations. SPE Annual Technical Conference and Exhibition 2017, 23,
25
26 10.2118/187104-MS.
27
28
29 48. Al-Muntasheri, G. A.; Li, L.; Liang, F.; Gomaa, A. M., Concepts in Cleanup of Fracturing
30
31 Fluids Used in Conventional Reservoirs: A Literature Review. SPE Prod. Oper. 2018, 33, 196-
32
33 213.
34
35
36
49. Jacobs, T., EOR-for-Shale Ideas to Boost Output Gain Traction. J. Pet. Technol. 2016, 68,
37
38 28-31.
39
40 50. Malpani, R.; Sinha, S.; Charry, L.; Sinosic, B.; Clark, B.; Gakhar, K., Improving
41
42
Hydrocarbon Recovery of Horizontal Shale Wells through Refracturing. SPE/CSUR
43
44
45 Unconventional Resources Conference 2015, 28, 10.2118/175920-MS.
46
47 51. Clarkson, C. R., Production Data Analysis of Unconventional Gas Wells: Review of
48
49 Theory and Best Practices. Int. J. Coal Geol. 2013, 109-110, 101-146.
50
51
52 52. Yousefzadeh, A.; Li, Q.; Aguilera, R., Microseismic 101: Monitoring and Evaluating
53
54 Hydraulic Fracturing to Improve the Efficiency of Oil and Gas Recovery from Unconventional
55
56
57
58
59
60 ACS Paragon Plus Environment
127
Energy & Fuels Page 128 of 170

1
2
3 Reservoirs. SPE Latin American and Caribbean Petroleum Engineering Conference 2015, 63,
4
5
6 10.2118/177277-MS.
7
8 53. Mohammed-Singh, L. J.; Singhal, A. K.; Sim, S. S.-K., Screening Criteria for CO2 Huff
9
10 'N' Puff Operations. SPE/DOE Symposium on Improved Oil Recovery 2006, 10, 10.2118/100044-
11
12
13
MS.
14
15 54. Tamayo, H. C.; Lee, K. J.; Taylor, R. S., Enhanced Aqueous Fracturing Fluid Recovery
16
17 from Tight Gas Formations: Foamed Co Pre-Pad Fracturing Fluid and More Effective Surfactant
18
19
Systems. Canadian International Petroleum Conference 2007, 10, 10.2118/2007-112.
20
21
22 55. Tamayo, H. C.; Lee, K. J.; Taylor, R. S., Enhanced Aqueous Fracturing Fluid Recovery
23
24 from Tight Gas Formations: Foamed CO2 Pre-Pad Fracturing Fluid and More Effective Surfactant
25
26 Systems. J. Can. Pet. Technol. 2008, 47, 6.
27
28
29 56. Kovscek, A. R.; Tang, G.-Q.; Vega, B., Experimental Investigation of Oil Recovery from
30
31 Siliceous Shale by CO2 Injection. SPE Annual Technical Conference and Exhibition 2008, 17,
32
33 10.2118/115679-MS.
34
35
36
57. Shoaib, S.; Hoffman, B. T., CO2 Flooding the Elm Coulee Field. SPE Rocky Mountain
37
38 Petroleum Technology Conference 2009, 11, 10.2118/123176-MS.
39
40 58. Vega, B.; O'Brien, W. J.; Kovscek, A. R., Experimental Investigation of Oil Recovery from
41
42
Siliceous Shale by Miscible CO2 Injection. SPE Annual Technical Conference and Exhibition
43
44
45 2010, 21, 10.2118/135627-MS.
46
47 59. Hoteit, H., Proper Modeling of Diffusion in Fractured Reservoirs. SPE Reservoir
48
49 Simulation Symposium 2011, 22, 10.2118/141937-MS.
50
51
52 60. Hoffman, B. T., Comparison of Various Gases for Enhanced Recovery from Shale Oil
53
54 Reservoirs. SPE Improved Oil Recovery Symposium 2012, 8, 10.2118/154329-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
128
Page 129 of 170 Energy & Fuels

1
2
3 61. Chen, C.; Balhoff, M. T.; Mohanty, K. K., Effect of Reservoir Heterogeneity on Primary
4
5
6 Recovery and CO2 Huff 'N' Puff Recovery in Shale-Oil Reservoirs. SPE Reservoir Eval. Eng.
7
8 2014, 17, 404-413.
9
10 62. Mohanty, K. K.; Chen, C.; Balhoff, M. T., Effect of Reservoir Heterogeneity on Improved
11
12
13
Shale Oil Recovery by CO2 Huff-N-Puff. SPE Unconventional Resources Conference-USA 2013,
14
15 16, 10.2118/164553-MS.
16
17 63. Loucks, R. G.; Reed, R. M.; Ruppel, S. C.; Hammes, U., Spectrum of Pore Types and
18
19
Networks in Mudrocks and a Descriptive Classification for Matrix-Related Mudrock
20
21
22 Poresspectrum of Pore Types and Networks in Mudrocks. AAPG Bull. 2012, 96, 1071-1098.
23
24 64. Liu, K.; Ostadhassan, M.; Zhou, J.; Gentzis, T.; Rezaee, R., Nanoscale Pore Structure
25
26 Characterization of the Bakken Shale in the USA. Fuel 2017, 209, 567-578.
27
28
29 65. Nelson, P. H., Pore-Throat Sizes in Sandstones, Tight Sandstones, and Shales. Am. Assoc.
30
31 Pet. Geol. Bull. 2009, 93, 329-340.
32
33 66. Zhang, Y., Barber, T., Hu, Q., Kibria, M.G., , Quantifying Hydrophilic and Hydrophobic
34
35
36
Pore Networks of the Bakken Shale, Search and Discovery AAPG Annual Convention and
37
38 Exhibition 2013.
39
40 67. Begum, M.; Yassin, M. R.; Dehghanpour, H., An Experimental Study of Imbibition Oil
41
42
Recovery from Oil-Wet Shales. SPE Western Regional Meeting 2018, 23, 10.2118/190044-MS.
43
44
45 68. Morsy, S.; Gomaa, A.; Sheng, J. J., Improvement of Eagle Ford Shale Formations Water
46
47 Imbibition by Mineral Dissolution and Wettability Alteration. SPE Unconventional Resources
48
49 Conference 2014, 9, 10.2118/168985-MS.
50
51
52 69. Alvarez, J. O.; Schechter, D. S., Wettability, Oil and Rock Characterization of the Most
53
54 Important Unconventional Liquid Reservoirs in the United States and the Impact on Oil Recovery.
55
56
57
58
59
60 ACS Paragon Plus Environment
129
Energy & Fuels Page 130 of 170

1
2
3 SPE/AAPG/SEG Unconventional Resources Technology Conference 2016, 21,
4
5
6 10.15530/2461651.
7
8 70. Alvarez, J. O.; Schechter, D. S., Wettability Alteration and Spontaneous Imbibition in
9
10 Unconventional Liquid Reservoirs by Surfactant Additives. SPE Latin American and Caribbean
11
12
13
Petroleum Engineering Conference 2015, 16, 10.2118/177057-MS.
14
15 71. Hawthorne, S. B.; Gorecki, C. D.; Sorensen, J. A.; Miller, D. J.; Harju, J. A.; Melzer, L. S.,
16
17 Hydrocarbon Mobilization Mechanisms Using CO2 in an Unconventional Oil Play. Energy
18
19
Procedia 2014, 63, 7717-7723.
20
21
22 72. Zhang, N.; Wei, M.; Bai, B., Statistical and Analytical Review of Worldwide CO2
23
24 Immiscible Field Applications. Fuel 2018, 220, 89-100.
25
26 73. Hawthorne, S. B.; Miller, D. J.; Grabanski, C. B.; Sorensen, J. A.; Pekot, L. J.; Kurz, B.
27
28
29 A.; Gorecki, C. D.; Steadman, E. N.; Harju, J. A.; Melzer, S., Measured Crude Oil Mmps with
30
31 Pure and Mixed CO2, Methane, and Ethane, and Their Relevance to Enhanced Oil Recovery from
32
33 Middle Bakken and Bakken Shales. SPE Unconventional Resources Conference 2017, 8,
34
35
36
10.2118/185072-MS.
37
38 74. Hawthorne, S. B.; Gorecki, C. D.; Sorensen, J. A.; Steadman, E. N.; Harju, J. A.; Melzer,
39
40 S., Hydrocarbon Mobilization Mechanisms from Upper, Middle, and Lower Bakken Reservoir
41
42
Rocks Exposed to CO2. SPE Unconventional Resources Conference Canada 2013, 9,
43
44
45 10.2118/167200-MS.
46
47 75. Hawthorne, S. B.; Jin, L.; Kurz, B. A.; Miller, D. J.; Grabanski, C. B.; Sorensen, J. A.;
48
49 Pekot, L. J.; Bosshart, N. W.; Smith, S. A.; Burton-Kelly, M. E.; Heebink, L. V.; Gorecki, C. D.;
50
51
52 Steadman, E. N.; Harju, J. A. In Integrating Petrographic and Petrophysical Analyses with CO2
53
54 Permeation and Oil Extraction and Recovery in the Bakken Tight Oil Formation, SPE
55
56
57
58
59
60 ACS Paragon Plus Environment
130
Page 131 of 170 Energy & Fuels

1
2
3 Unconventional Resources Conference, Calgary, Alberta, Canada, 2017/2/15/; Society of
4
5
6 Petroleum Engineers: Calgary, Alberta, Canada, 2017; 11, 10.2118/185081-MS.
7
8 76. Alharthy, N.; Teklu, T.; Kazemi, H.; Graves, R.; Hawthorne, S.; Braunberger, J.; Kurtoglu,
9
10 B., Enhanced Oil Recovery in Liquid-Rich Shale Reservoirs: Laboratory to Field. SPE Annual
11
12
13
Technical Conference and Exhibition 2015, 29, 10.2118/175034-MS.
14
15 77. Adel, I. A.; Tovar, F. D.; Zhang, F.; Schechter, D. S., The Impact of Mmp on Recovery
16
17 Factor During CO2 – EOR in Unconventional Liquid Reservoirs. SPE Annual Technical
18
19
Conference and Exhibition 2018, 17, 10.2118/191752-MS.
20
21
22 78. Tovar, F. D.; Barrufet, M. A.; Schechter, D. S., Gas Injection for EOR in Organic Rich
23
24 Shales. Part Ii: Mechanisms of Recovery. SPE/AAPG/SEG Unconventional Resources Technology
25
26 Conference 2018, 21, 2903026-MS.
27
28
29 79. Hawthorne, S. B.; Miller, D. J.; Grabanski, C. B.; Azzolina, N.; Kurz, B. A.; Ardakani, O.
30
31 H.; Smith, S. A.; Sanei, H.; Sorensen, J. A., Hydrocarbon Recovery from Williston Basin Shale
32
33 and Mudrock Cores with Supercritical CO2: Part 1. Method Validation and Recoveries from Cores
34
35
36
Collected across the Basin. Energy Fuels 2019, 33, 6857-6866.
37
38 80. Hawthorne, S. B.; Grabanski, C. B.; Miller, D. J.; Kurz, B. A.; Sorensen, J. A.,
39
40 Hydrocarbon Recovery from Williston Basin Shale and Mudrock Cores with Supercritical CO2: 2.
41
42
Mechanisms That Control Oil Recovery Rates and CO2 Permeation. Energy Fuels 2019, 33, 6867-
43
44
45 6877.
46
47 81. Hoffman, B. T.; Reichhardt, D., Quantitative Evaluation of Recovery Mechanisms for
48
49 Huff-N-Puff Gas Injection in Unconventional Reservoirs. SPE/AAPG/SEG Unconventional
50
51
52 Resources Technology Conference 2019, 15, 10.105530/147-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
131
Energy & Fuels Page 132 of 170

1
2
3 82. Alfarge, D.; Wei, M.; Bai, B., A Parametric Study on the Applicability of Miscible Gases
4
5
6 Based EOR Techniques in Unconventional Liquids Rich Reservoirs. SPE Canada Unconventional
7
8 Resources Conference 2018, 23, 10.2118/189785-MS.
9
10 83. Haines, H. K.; Monger, T. G., A Laboratory Study of Natural Gas Huff `N' Puff. CIM/SPE
11
12
13
International Technical Meeting 1990, 16, 10.2118/21576-MS.
14
15 84. Kong, B.; Wang, S.; Chen, S., Simulation and Optimization of CO2 Huff-and-Puff
16
17 Processes in Tight Oil Reservoirs. SPE Improved Oil Recovery Conference 2016, 14,
18
19
10.2118/179668-MS.
20
21
22 85. Mehio, N.; Dai, S.; Jiang, D.-e., Quantum Mechanical Basis for Kinetic Diameters of Small
23
24 Gaseous Molecules. J. Phys. Chem. A 2014, 118, 1150-1154.
25
26 86. Geier, S. J.; Mason, J. A.; Bloch, E. D.; Queen, W. L.; Hudson, M. R.; Brown, C. M.; Long,
27
28
29 J. R., Selective Adsorption of Ethylene over Ethane and Propylene over Propane in the Metal–
30
31 Organic Frameworks M2(Dobdc) (M = Mg, Mn, Fe, Co, Ni, Zn). Chem. Sci. 2013, 4, 2054-2061.
32
33 87. Jin, L.; Hawthorne, S.; Sorensen, J.; Pekot, L.; Bosshart, N.; Gorecki, C.; Steadman, E.;
34
35
36
Harju, J., Utilization of Produced Gas for Improved Oil Recovery and Reduced Emissions from
37
38 the Bakken Formation. SPE Health, Safety, Security, Environment, & Social Responsibility
39
40 Conference - North America 2017, 12, 10.2118/184414-MS.
41
42
88. Saraf, M. K. Polymerization of Vinylidene Fluoride in Supercritical CO2: Molecular
43
44
45 Weight Distribution. North Carolina State University, Raleigh, NC, 2001.
46
47 89. Williams, L. L.; Rubin, J. B.; Edwards, H. W., Calculation of Hansen Solubility Parameter
48
49 Values for a Range of Pressure and Temperature Conditions, Including the Supercritical Fluid
50
51
52 Region. Ind. Eng. Chem. Res. 2004, 43, 4967-4972.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
132
Page 133 of 170 Energy & Fuels

1
2
3 90. Adekunle, O.; Hoffman, B. T., Experimental and Analytical Methods to Determine
4
5
6 Minimum Miscibility Pressure (Mmp) for Bakken Formation Crude Oil. J. Pet. Sci. Eng. 2016,
7
8 146, 170-182.
9
10 91. Adekunle, O. O.; Hoffman, B. T., Minimum Miscibility Pressure Studies in the Bakken.
11
12
13
SPE Improved Oil Recovery Symposium 2014, 16, 10.2118/169077-MS.
14
15 92. Li, S.; Luo, P., Experimental and Simulation Determination of Minimum Miscibility
16
17 Pressure for a Bakken Tight Oil and Different Injection gases. Petroleum 2017, 3, 79-86.
18
19
93. Ghorbani, M.; Momeni, A.; Safavi, S.; Gandomkar, A., Modified Vanishing Interfacial
20
21
22 Tension (Vit) Test for CO2–Oil Minimum Miscibility Pressure (Mmp) Measurement. J. Nat. Gas
23
24 Sci. Eng. 2014, 20, 92-98.
25
26 94. Orr, F. M.; Jessen, K., An Analysis of the Vanishing Interfacial Tension Technique for
27
28
29 Determination of Minimum Miscibility Pressure. Fluid Phase Equilib. 2007, 255, 99-109.
30
31 95. Hawthorne, S. B.; Miller, D. J.; Jin, L.; Gorecki, C. D., Rapid and Simple Capillary-
32
33 Rise/Vanishing Interfacial Tension Method to Determine Crude Oil Minimum Miscibility
34
35
36
Pressure: Pure and Mixed CO2, Methane, and Ethane. Energy Fuels 2016, 30, 6365-6372.
37
38 96. Zhang, K.; Jia, N.; Zeng, F.; Luo, P., A New Diminishing Interface Method for
39
40 Determining the Minimum Miscibility Pressures of Light Oil–CO2 Systems in Bulk Phase and
41
42
Nanopores. Energy Fuels 2017, 31, 12021-12034.
43
44
45 97. Macleod, D. B., On a Relation between Surface Tension and Density. Transactions of the
46
47 Faraday Society 1923, 19, 38-41.
48
49 98. O'Bryan, P. L.; Bourgoyne, A. T., Jr., Swelling of Oil-Based Drilling Fluids Resulting from
50
51
52 Dissolved Gas. SPE Drilling Eng. 1990, 5, 149-155.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
133
Energy & Fuels Page 134 of 170

1
2
3 99. O'Bryan, P. L.; Bourgoyne, A. T., Jr.; Monger, T. G.; Kopcso, D. P., An Experimental
4
5
6 Study of Gas Solubility in Oil-Based Drilling Fluids. SPE Drilling Eng. 1988, 3, 33-42.
7
8 100. Yang, C.; Gu, Y., Diffusion Coefficients and Oil Swelling Factors of Carbon Dioxide,
9
10 Methane, Ethane, Propane, and Their Mixtures in Heavy Oil. Fluid Phase Equilib. 2006, 243, 64-
11
12
13
73.
14
15 101. Pereira, L. M. C.; Chapoy, A.; Burgass, R.; Tohidi, B., Measurement and Modelling of
16
17 High Pressure Density and Interfacial Tension of (Gas+N-Alkane) Binary Mixtures. J. Chem.
18
19
Thermodyn. 2016, 97, 55-69.
20
21
22 102. Zhu, C.; Li, Y.; Gong, H.; Sang, Q.; Li, Z.; Dong, M., Adsorption and Dissolution
23
24 Behaviors of Carbon Dioxide and N-Dodecane Mixtures in Shale. Energy Fuels 2018, 32, 1374-
25
26 1386.
27
28
29 103. Ho, T. A.; Wang, Y.; Xiong, Y.; Criscenti, L. J., Differential Retention and Release of CO2
30
31 and Ch4 in Kerogen Nanopores: Implications for Gas Extraction and Carbon Sequestration. Fuel
32
33 2018, 220, 1-7.
34
35
36
104. Al Ismail, M.; Zoback, M., CO2-Based Technologies in Unconventional Resources: Impact
37
38 of Rock Mineralogy on Adsorption. SPE Kingdom of Saudi Arabia Annual Technical Symposium
39
40 and Exhibition 2017, 12, 10.2118/188168-MS.
41
42
105. Kim, T. H.; Park, S. S.; Lee, K. S., Modeling of CO2 Injection Considering Multi-
43
44
45 Component Transport and Geomechanical Effect in Shale Gas Reservoirs. SPE/IATMI Asia
46
47 Pacific Oil & Gas Conference and Exhibition 2015, 17, 10.2118/176174-MS.
48
49 106. Wang, R.; Peng, F.; Song, K.; Feng, G.; Guo, Z., Molecular Dynamics Study of Interfacial
50
51
52 Properties in CO2 Enhanced Oil Recovery. Fluid Phase Equilib. 2018, 467, 25-32.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
134
Page 135 of 170 Energy & Fuels

1
2
3 107. Yang, S.; Wei, Y.; Chen, Z.; Wu, W.; Xu, J., Effects of Multicomponent Adsorption on
4
5
6 Enhanced Shale Reservoir Recovery by CO2 Injection Coupled with Reservoir Geomechanics.
7
8 SPE Low Perm Symposium 2016, 9, 10.2118/180208-MS.
9
10 108. Zhang, K.; Liu, Q.; Wang, M.; Kong, B.; Lv, J.; Wu, K.; Chen, S.; Chen, Z., Investigation
11
12
13
of CO2 Enhanced Gas Recovery in Shale Plays. SPE Europec featured at 78th EAGE Conference
14
15 and Exhibition 2016, 10, 10.2118/180174-MS.
16
17 109. Lan, Y.; Yang, Z.; Wang, P.; Yan, Y.; Zhang, L.; Ran, J., A Review of Microscopic
18
19
Seepage Mechanism for Shale Gas Extracted by Supercritical CO2 Flooding. Fuel 2019, 238, 412-
20
21
22 424.
23
24 110. Tang, L.; Sun, S.; Mi, L.; Wang, L.; Qi, H.; Yu, R.; Bai, F.; Chaoyang, J., Analysis of
25
26 Enhancement Shale Gas Recovery by CO2 Storage. SPE Argentina Exploration and Production of
27
28
29 Unconventional Resources Symposium 2016, 9, 10.2118/180983-MS.
30
31 111. Jin, L.; Hawthorne, S.; Sorensen, J.; Pekot, L.; Kurz, B.; Smith, S.; Heebink, L.; Herdegen,
32
33 V.; Bosshart, N.; Torres, J.; Dalkhaa, C.; Peterson, K.; Gorecki, C.; Steadman, E.; Harju, J.,
34
35
36
Advancing CO2 Enhanced Oil Recovery and Storage in Unconventional Oil Play—Experimental
37
38 Studies on Bakken Shales. Appl. Energy 2017, 208, 171-183.
39
40 112. Sanguinito, S.; Goodman, A.; Tkach, M.; Kutchko, B.; Culp, J.; Natesakhawat, S.; Fazio,
41
42
J.; Fukai, I.; Crandall, D., Quantifying Dry Supercritical CO2-Induced Changes of the Utica Shale.
43
44
45 Fuel 2018, 226, 54-64.
46
47 113. Zhao, H.; Lai, Z.; Firoozabadi, A., Sorption Hysteresis of Light Hydrocarbons and Carbon
48
49 Dioxide in Shale and Kerogen. Sci. Rep. 2017, 7, 16209.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
135
Energy & Fuels Page 136 of 170

1
2
3 114. Vetter, O. J.; Farone, W. A.; Veith, E.; Lankford, S., Calcium Carbonate Scale
4
5
6 Considerations: A Practical Approach. SPE Production Technology Symposium 1987, 14,
7
8 10.2118/17009-MS.
9
10 115. Salman, M.; Kostarelos, K.; Sharma, P.; Lee, J. H., Application of Miscible Ethane Foam
11
12
13
for Gas EOR Conformance in Low-Permeability Heterogeneous Harsh Environments.
14
15 SPE/AAPG/SEG Unconventional Resources Technology Conference 2019, 18, 10.105530/urtec-
16
17 2019-1009.
18
19
116. Ning, Y.; Kazemi, H., Ethane-Enriched Gas Injection EOR in Niobrara and Codell: A
20
21
22 Dual-Porosity Compositional Model. SPE Improved Oil Recovery Conference 2018, 18,
23
24 10.2118/190226-MS.
25
26 117. McGuire, P. L.; Okuno, R.; Gould, T. L.; Lake, L. W., Ethane-Based Enhanced Oil
27
28
29 Recovery: An Innovative and Profitable Enhanced-Oil-Recovery Opportunity for a Low-Price
30
31 Environment. SPE Reservoir Eval. Eng. 2017, 20, 42-58.
32
33 118. Han, L.; Gu, Y., Optimization of Miscible CO2 Water-Alternating-Gas Injection in the
34
35
36
Bakken Formation. Energy Fuels 2014, 28, 6811-6819.
37
38 119. Zhang, X.; Wei, B.; Shang, J.; Gao, K.; Pu, W.; Xu, X.; Wood, C.; Sun, L., Alterations of
39
40 Geochemical Properties of a Tight Sandstone Reservoir Caused by Supercritical CO2-Brine-Rock
41
42
Interactions in CO2-EOR and Geosequestration. J. CO2 Util. 2018, 28, 408-418.
43
44
45 120. Zhang, K., Experimental and Numerical Investigation of Oil Recovery from Bakken
46
47 Formation by Miscible CO2 Injection. SPE Annual Technical Conference and Exhibition 2016, 23,
48
49 10.2118/184486-STU.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
136
Page 137 of 170 Energy & Fuels

1
2
3 121. Tovar, F. D.; Barrufet, M. A.; Schechter, D. S., Gas Injection for EOR in Organic Rich
4
5
6 Shale. Part I: Operational Philosophy. SPE Improved Oil Recovery Conference 2018, 25,
7
8 10.2118/190323-MS.
9
10 122. Wei, B.; Lu, L.; Pu, W.; Wu, R.; Zhang, X.; Li, Y.; Jin, F., Production Dynamics of CO2
11
12
13
Cyclic Injection and CO2 Sequestration in Tight Porous Media of Lucaogou Formation in Jimsar
14
15 Sag. J. Pet. Sci. Eng. 2017, 157, 1084-1094.
16
17 123. Zhou, X.; Yuan, Q.; Zhang, Y.; Wang, H.; Zeng, F.; Zhang, L., Performance Evaluation of
18
19
CO2 Flooding Process in Tight Oil Reservoir Via Experimental and Numerical Simulation Studies.
20
21
22 Fuel 2019, 236, 730-746.
23
24 124. Habibi, A.; Yassin, M. R.; Dehghanpour, H.; Bryan, D., Experimental Investigation of Co
25
26 2 -Oil Interactions in Tight Rocks: A Montney Case Study. Fuel 2017, 203, 853-867.
27
28
29 125. Habibi, A.; Yassin, M. R.; Dehghanpour, H.; Bryan, D., CO2-Oil Interactions in Tight
30
31 Rocks: An Experimental Study. SPE Unconventional Resources Conference 2017, 26,
32
33 10.2118/185047-MS.
34
35
36
126. Tovar, F. D.; Eide, O.; Graue, A.; Schechter, D. S., Experimental Investigation of
37
38 Enhanced Recovery in Unconventional Liquid Reservoirs Using CO2: A Look Ahead to the Future
39
40 of Unconventional EOR. SPE Unconventional Resources Conference 2014, 9, 10.2118/169022-
41
42
MS.
43
44
45 127. Carpenter, C., Enhanced Recovery in Unconventional Liquid Reservoirs by Use of CO2. J.
46
47 Pet. Technol. 2014, 66, 138-141.
48
49 128. Song, C.; Yang, D., Experimental and Numerical Evaluation of CO2 Huff-N-Puff
50
51
52 Processes in Bakken Formation. Fuel 2017, 190, 145-162.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
137
Energy & Fuels Page 138 of 170

1
2
3 129. Qian, K.; Yang, S.; Dou, H.; Wang, Q.; Wang, L.; Huang, Y., Experimental Investigation
4
5
6 on Microscopic Residual Oil Distribution During CO2 Huff-and-Puff Process in Tight Oil
7
8 Reservoirs. Energies 2018, 11, 1-16.
9
10 130. Cao, M.; Gu, Y., Oil Recovery Mechanisms and Asphaltene Precipitation Phenomenon in
11
12
13
Immiscible and Miscible CO2 Flooding Processes. Fuel 2013, 109, 157-166.
14
15 131. Li, Z.; Gu, Y., Optimum Timing for Miscible CO2-EOR after Waterflooding in a Tight
16
17 Sandstone Formation. Energy Fuels 2014, 28, 488-499.
18
19
132. Song, C.; Yang, D., Performance Evaluation of CO2 Huff-N-Puff Processes in Tight Oil
20
21
22 Formations. SPE Unconventional Resources Conference Canada 2013, 17, 10.2118/167217-MS.
23
24 133. Chen, T.; Yang, Z.; Luo, Y.; Lin, W.; Xu, J.; Ding, Y.; Niu, J., Evaluation of Displacement
25
26 Effects of Different Injection Media in Tight Oil Sandstone by Online Nuclear Magnetic
27
28
29 Resonance. Energies 2018, 11, 1-16.
30
31 134. Gamadi, T. D.; Sheng, J. J.; Soliman, M. Y.; Menouar, H.; Watson, M. C.; Emadibaladehi,
32
33 H., An Experimental Study of Cyclic CO2 Injection to Improve Shale Oil Recovery. SPE Improved
34
35
36
Oil Recovery Symposium 2014, 9, 10.2118/169142-MS.
37
38 135. Shen, Z.; Sheng, J. J., Experimental and Numerical Study of Permeability Reduction
39
40 Caused by Asphaltene Precipitation and Deposition During CO2 Huff and Puff Injection in Eagle
41
42
Ford Shale. Fuel 2018, 211, 432-445.
43
44
45 136. Li, L.; Sheng, J. J.; Su, Y.; Zhan, S., Further Investigation of Effects of Injection Pressure
46
47 and Imbibition Water on CO2 Huff-N-Puff Performance in Liquid-Rich Shale Reservoirs. Energy
48
49 Fuels 2018, 32, 5789-5798.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
138
Page 139 of 170 Energy & Fuels

1
2
3 137. Li, L.; Sheng, J. J.; Xu, J., Gas Selection for Huff-N-Puff EOR in Shale Oil Reservoirs
4
5
6 Based Upon Experimental and Numerical Study. SPE Unconventional Resources Conference
7
8 2017, 15, 10.2118/185066-MS.
9
10 138. Jin, L.; Hawthorne, S.; Sorensen, J.; Pekot, L.; Kurz, B.; Smith, S.; Heebink, L.; Bosshart,
11
12
13
N.; Torres, J.; Dalkhaa, C.; Gorecki, C.; Steadman, E.; Harju, J., Extraction of Oil from the Bakken
14
15 Shales with Supercritical CO2. SPE/AAPG/SEG Unconventional Resources Technology
16
17 Conference 2017, 17, 10.15530/2671596.
18
19
139. Jin, L.; Sorensen, J. A.; Hawthorne, S. B.; Smith, S. A.; Pekot, L. J.; Bosshart, N. W.;
20
21
22 Burton-Kelly, M. E.; Miller, D. J.; Grabanski, C. B.; Gorecki, C. D.; Steadman, E. N.; Harju, J.
23
24 A., Improving Oil Recovery by Use of Carbon Dioxide in the Bakken Unconventional System: A
25
26 Laboratory Investigation. SPE Reservoir Eval. Eng. 2017, 20, 602-612.
27
28
29 140. Li, L.; Sheng, J. J., Upscale Methodology for Gas Huff-N-Puff Process in Shale Oil
30
31 Reservoirs. J. Pet. Sci. Eng. 2017, 153, 36-46.
32
33 141. Hawthorne, S.; Sorensen, J.; Miller, D. J.; Gorecki, C.; Harju, J.; Pospisil, G., Laboratory
34
35
36
Studies of Rich Gas Interactions with Bakken Crude Oil to Support Enhanced Oil Recovery.
37
38 SPE/AAPG/SEG Unconventional Resources Technology Conference 2019, 8, 10.105530/961-MS.
39
40 142. Li, L.; Zhang, Y.; Sheng, J. J., Effect of the Injection Pressure on Enhancing Oil Recovery
41
42
in Shale Cores During the CO2 Huff-N-Puff Process When It Is above and Below the Minimum
43
44
45 Miscibility Pressure. Energy Fuels 2017, 31, 3856-3867.
46
47 143. Jin, L.; Hawthorne, S.; Sorensen, J.; Kurz, B.; Pekot, L.; Smith, S.; Bosshart, N.; Azenkeng,
48
49 A.; Gorecki, C.; Harju, J., A Systematic Investigation of Gas-Based Improved Oil Recovery
50
51
52 Technologies for the Bakken Tight Oil Formation. SPE/AAPG/SEG Unconventional Resources
53
54 Technology Conference 2016, 15, 10.15530/2433692.
55
56
57
58
59
60 ACS Paragon Plus Environment
139
Energy & Fuels Page 140 of 170

1
2
3 144. Jin, L.; Sorensen, J. A.; Hawthorne, S. B.; Smith, S. A.; Bosshart, N. W.; Burton-Kelly, M.
4
5
6 E.; Miller, D. J.; Grabanski, C. B.; Harju, J. A., Improving Oil Transportability Using CO2 in the
7
8 Bakken System – a Laboratory Investigation. SPE International Conference and Exhibition on
9
10 Formation Damage Control 2016, 17, 10.2118/178948-MS.
11
12
13
145. Sorensen, J. A.; Braunberger, J. R.; Liu, G.; Smith, S. A.; Hawthorne, S. A.; Steadman, E.
14
15 N.; Harju, J. A., Characterization and Evaluation of the Bakken Petroleum System for Co
16
17 Enhanced Oil Recovery. Unconventional Resources Technology Conference 2015, 21,
18
19
10.15530/2169871.
20
21
22 146. Fakher, S.; Imqam, A., Asphaltene Precipitation and Deposition During CO2 Injection in
23
24 Nano Shale Pore Structure and Its Impact on Oil Recovery. Fuel 2019, 237, 1029-1039.
25
26 147. Mohammed, S.; Gadikota, G., The Influence of CO2 on the Structure of Confined
27
28
29 Asphaltenes in Calcite Nanopores. Fuel 2019, 236, 769-777.
30
31 148. Qian, K.; Yang, S.; Dou, H.-e.; Pang, J.; Huang, Y., Formation Damage Due to Asphaltene
32
33 Precipitation During CO2 Flooding Processes with NMR Technique. Oil Gas Sci. Technol. – Rev.
34
35
36
IFP Energies nouvelles 2019, 74, 1-11.
37
38 149. Shen, Z.; Sheng, J. J., Experimental Study of Asphaltene Aggregation During CO2 and Ch4
39
40 Injection in Shale Oil Reservoirs. SPE Improved Oil Recovery Conference 2016, 11,
41
42
10.2118/179675-MS.
43
44
45 150. Shen, Z.; Sheng, J. J., Investigation of Asphaltene Deposition Mechanisms During CO2
46
47 Huff-N-Puff Injection in Eagle Ford Shale. Pet. Sci. Technol. 2017, 35, 1960-1966.
48
49 151. Ding, M.; Gao, M.; Wang, Y.; Qu, Z.; Chen, X., Experimental Study on CO2-EOR in
50
51
52 Fractured Reservoirs: Influence of Fracture Density, Miscibility and Production Scheme. J. Pet.
53
54 Sci. Eng. 2019, 174, 476-485.
55
56
57
58
59
60 ACS Paragon Plus Environment
140
Page 141 of 170 Energy & Fuels

1
2
3 152. Ma, J.; Wang, X.; Gao, R.; Zeng, F.; Huang, C.; Tontiwachwuthikul, P.; Liang, Z.,
4
5
6 Enhanced Light Oil Recovery from Tight Formations through CO2 Huff ‘N’ Puff Processes. Fuel
7
8 2015, 154, 35-44.
9
10 153. Wang, H.; Lun, Z.; Lv, C.; Lang, D.; Luo, M.; Zhao, Q.; Zhao, C., Nuclear Magnetic
11
12
13
Resonance Study on Oil Mobilization in the Shale Exposed to CO2. SPE Improved Oil Recovery
14
15 Conference 2018, 8, 10.2118/190185-MS.
16
17 154. Liu, S.; Sahni, V.; Tan, J.; Beckett, D.; Vo, T., Laboratory Investigation of EOR
18
19
Techniques for Organic Rich Shales in the Permian Basin. SPE/AAPG/SEG Unconventional
20
21
22 Resources Technology Conference 2018, 9, 2890074-MS.
23
24 155. Meng, X.; Meng, Z.; Ma, J.; Wang, T., Performance Evaluation of CO2 Huff-N-Puff Gas
25
26 Injection in Shale Gas Condensate Reservoirs. Energies 2018, 12, 42.
27
28
29 156. Li, L.; Su, Y.; Lv, Y.; Tu, J., Asphaltene Deposition and Permeability Impairment in Shale
30
31 Reservoirs During CO2 Huff-N-Puff EOR Process. Pet. Sci. Technol. 2020, 38, 384-390.
32
33 157. Li, L.; Su, Y.; Hao, Y.; Zhan, S.; Lv, Y.; Zhao, Q.; Wang, H., A Comparative Study of
34
35
36
CO2 and N2 Huff-N-Puff EOR Performance in Shale Oil Production. J. Pet. Sci. Eng. 2019, 181,
37
38 106174.
39
40 158. Shen, Z.; Sheng, J. J., Experimental Study of Permeability Reduction and Pore Size
41
42
Distribution Change Due to Asphaltene Deposition During CO2 Huff and Puff Injection in Eagle
43
44
45 Ford Shale. Asia-Pac. J. Chem. Eng. 2017, 12, 381-390.
46
47 159. Fogden, A.; Cheng, Q.; Middleton, J.; Kingston, A.; Turner, M.; Sheppard, A.; Olson*, T.;
48
49 Armstrong, R., Dynamic Micro-Ct Imaging of Diffusion in Unconventionals. In Unconventional
50
51
52 Resources Technology Conference, San Antonio, Texas, 20-22 July 2015, Society of Exploration
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
141
Energy & Fuels Page 142 of 170

1
2
3 Geophysicists, American Association of Petroleum Geologists, Society of Petroleum Engineers:
4
5
6 2015; 1244-1259.
7
8 160. Liu, G.; Yin, H.; Lan, Y.; Fei, S.; Yang, D., Experimental Determination of Dynamic Pore-
9
10 Throat Structure Characteristics in a Tight Gas Sandstone Formation with Consideration of
11
12
13
Effective Stress. Mar. Pet. Geol. 2020, 113, 104170.
14
15 161. Bhuiyan, H. M.; Agofack, N.; Gawel, M. K.; Cerasi, R. P., Micro- and Macroscale
16
17 Consequences of Interactions between CO2 and Shale Rocks. Energies 2020, 13, 1-30.
18
19
162. Wang, S.; Liu, K.; Han, J.; Ling, K.; Wang, H.; Jia, B., Investigation of Properties
20
21
22 Alternation During Super-Critical CO2 Injection in Shale. Appl. Sci. 2019, 9, 1686.
23
24 163. Zhu, H.-Y.; Tao, L.; Liu, Q.-Y.; Lei, Z.-D., Multiscale Change Characteristics of Chinese
25
26 Wufeng Shale under Supercritical Carbon Dioxide. SPE Annual Technical Conference and
27
28
29 Exhibition 2016, 15, 10.2118/181369-MS.
30
31 164. Lu, J.; Nicot, J.-P.; Mickler, P. J.; Ribeiro, L. H.; Darvari, R., Alteration of Bakken
32
33 Reservoir Rock During CO2-Based Fracturing—an Autoclave Reaction Experiment. Journal of
34
35
36
Unconventional Oil and Gas Resources 2016, 14, 72-85.
37
38 165. Goodman, A.; Sanguinito, S.; Tkach, M.; Natesakhawat, S.; Kutchko, B.; Fazio, J.; Cvetic,
39
40 P., Investigating the Role of Water on CO2-Utica Shale Interactions for Carbon Storage and Shale
41
42
Gas Extraction Activities – Evidence for Pore Scale Alterations. Fuel 2019, 242, 744-755.
43
44
45 166. Goodman, A.; Sanguinito, S.; Kutchko, B.; Natesakhawat, S.; Cvetic, P.; Allen, A. J., Shale
46
47 Pore Alteration: Potential Implications for Hydrocarbon Extraction and CO2 Storage. Fuel 2020,
48
49 265, 116930.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
142
Page 143 of 170 Energy & Fuels

1
2
3 167. Reynolds, M.; Britten, D.; Cui, X.; Twemlow, C.; Cook, M., A Laboratory Study of CO2
4
5
6 Interactions within Shale and Tight Sand Cores - Duvernay, Montney and Wolfcamp Formations.
7
8 SPE Canada Unconventional Resources Conference 2018, 18, 10.2118/189800-MS.
9
10 168. Al Ismail, M. I.; Hol, S.; Reece, J. S.; Zoback, M. D., The Effect of CO2 Adsorption on
11
12
13
Permeability Anisotropy in the Eagle Ford Shale. Unconventional Resources Technology
14
15 Conference 2014, 9, 10.15530/1921520.
16
17 169. Xie, Q.; Chen, Y.; Sari, A.; Pu, W.; Saeedi, A.; Liao, X., A Ph-Resolved Wettability
18
19
Alteration: Implications for CO2-Assisted EOR in Carbonate Reservoirs. Energy Fuels 2017, 31,
20
21
22 13593-13599.
23
24 170. Nguyen, P.; Carey, J. W.; Viswanathan, H. S.; Porter, M., Effectiveness of Supercritical-
25
26 CO2 and N2 Huff-and-Puff Methods of Enhanced Oil Recovery in Shale Fracture Networks Using
27
28
29 Microfluidic Experiments. Appl. Energy 2018, 230, 160-174.
30
31 171. Zhong, J.; Abedini, A.; Xu, L.; Xu, Y.; Qi, Z.; Mostowfi, F.; Sinton, D., Nanomodel
32
33 Visualization of Fluid Injections in Tight Formations. Nanoscale 2018, 10, 21994-22002.
34
35
36
172. Zhang, C.; Li, Z.; Li, S.; Lv, Q.; Wang, P.; Liu, J.; Liu, J., Enhancing Sodium Bis(2-
37
38 Ethylhexyl) Sulfosuccinate Injectivity for CO2 Foam Formation in Low-Permeability Cores:
39
40 Dissolving in CO2 with Ethanol. Energy Fuels 2018, 32, 5846-5856.
41
42
173. Wang, L.; Yu, W., Mechanistic Simulation Study of Gas Puff and Huff Process for Bakken
43
44
45 Tight Oil Fractured Reservoir. Fuel 2019, 239, 1179-1193.
46
47 174. Reamer, H. H.; Sage, B. H., Diffusion Coefficients in Hydrocarbon Systems: Methane in
48
49 the Liquid Phase of the Methane–N-Heptane System. AIChE J. 1957, 3, 449-453.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
143
Energy & Fuels Page 144 of 170

1
2
3 175. Cadogan, S. P.; Mistry, B.; Wong, Y.; Maitland, G. C.; Trusler, J. P. M., Diffusion
4
5
6 Coefficients of Carbon Dioxide in Eight Hydrocarbon Liquids at Temperatures between (298.15
7
8 and 423.15) K at Pressures up to 69 Mpa. J. Can. Pet. Technol. 2016, 61, 3922-3932.
9
10 176. Levine, J. S.; Fukai, I.; Soeder, D. J.; Bromhal, G.; Dilmore, R. M.; Guthrie, G. D.;
11
12
13
Rodosta, T.; Sanguinito, S.; Frailey, S.; Gorecki, C.; Peck, W.; Goodman, A. L., U.S. Doe Netl
14
15 Methodology for Estimating the Prospective CO2 Storage Resource of Shales at the National and
16
17 Regional Scale. Int. J. Greenhouse Gas Control 2016, 51, 81-94.
18
19
177. Myshakin, E. M.; Singh, H.; Sanguinito, S.; Bromhal, G.; Goodman, A. L., Numerical
20
21
22 Estimations of Storage Efficiency for the Prospective CO2 Storage Resource of Shales. Int. J.
23
24 Greenhouse Gas Control 2018, 76, 24-31.
25
26 178. Hejazi, S. H.; Assef, Y.; Tavallali, M.; Popli, A., Cyclic Co 2 -EOR in the Bakken
27
28
29 Formation: Variable Cycle Sizes and Coupled Reservoir Response Effects. Fuel 2017, 210, 758-
30
31 767.
32
33 179. Kalra, S.; Tian, W.; Wu, X., A Numerical Simulation Study of CO2 Injection for Enhancing
34
35
36
Hydrocarbon Recovery and Sequestration in Liquid-Rich Shales. Pet. Sci. 2017, 15, 103-115.
37
38 180. Pranesh, V., Subsurface CO2 Storage Estimation in Bakken Tight Oil and Eagle Ford Shale
39
40 Gas Condensate Reservoirs by Retention Mechanism. Fuel 2018, 215, 580-591.
41
42
181. Wan, T.; Sheng, J. J.; Soliman, M. Y., Evaluate EOR Potential in Fractured Shale Oil
43
44
45 Reservoirs by Cyclic Gas Injection. Unconventional Resources Technology Conference 2013, 10,
46
47 1611383-MS.
48
49 182. Wan, T.; Sheng, J. J.; Soliman, M. Y., Evaluation of the EOR Potential in Shale Oil
50
51
52 Reservoirs by Cyclic Gas Injection. SPWLA 54th Annual Logging Symposium 2013, 16.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
144
Page 145 of 170 Energy & Fuels

1
2
3 183. Li, L.; Sheng, J. J.; Sheng, J., Optimization of Huff-N-Puff Gas Injection to Enhance Oil
4
5
6 Recovery in Shale Reservoirs. SPE Low Perm Symposium 2016, 18, 10.2118/180219-MS.
7
8 184. Li, L.; Sheng, J. J., Experimental Study of Core Size Effect on Ch4 Huff-N-Puff Enhanced
9
10 Oil Recovery in Liquid-Rich Shale Reservoirs. J. Nat. Gas Sci. Eng. 2016, 34, 1392-1402.
11
12
13
185. Meng, X.; Yu, Y.; Sheng, J. J.; Watson, M.; Mody, F., An Experimental Study on Huff-N-
14
15 Puff Gas Injection to Enhance Condensate Recovery in Shale Gas Reservoirs. Unconventional
16
17 Resources Technology Conference 2015, 11, 10.15530/URTEC-2015-2153322.
18
19
186. Meng, X.; Sheng, J. J.; Yu, Y., Evaluation of Enhanced Condensate Recovery Potential in
20
21
22 Shale Plays by Huff-N-Puff Gas Injection. SPE Eastern Regional Meeting 2015, 12,
23
24 10.2118/177283-MS.
25
26 187. Meng, X.; Sheng, J. J.; Yu, Y., Experimental and Numerical Study of Enhanced
27
28
29 Condensate Recovery by Gas Injection in Shale Gas-Condensate Reservoirs. SPE Reservoir Eval.
30
31 Eng. 2017, 20, 471-477.
32
33 188. Katiyar, A.; Patil, P.; Rohilla, N.; Rozowski, P.; Evans, J.; Bozeman, T.; Nguyen, Q.,
34
35
36
Industry-First Hydrocarbon-Foam EOR Pilot in an Unconventional Reservoir: Design,
37
38 Implementation, and Performance Analysis. SPE/AAPG/SEG Unconventional Resources
39
40 Technology Conference 2019, 23, 10.105530/103-MS.
41
42
189. Gamadi, T. D.; Sheng, J. J.; Soliman, M. Y., An Experimental Study of Cyclic Gas
43
44
45 Injection to Improve Shale Oil Recovery. SPE Annual Technical Conference and Exhibition 2013,
46
47 9, 10.2118/166334-MS.
48
49 190. Yu, Y.; Sheng, J. J., An Experimental Investigation of the Effect of Pressure Depletion
50
51
52 Rate on Oil Recovery from Shale Cores by Cyclic N2 Injection. Unconventional Resources
53
54 Technology Conference 2015, 12, 10.15530/URTEC-2015-2144010.
55
56
57
58
59
60 ACS Paragon Plus Environment
145
Energy & Fuels Page 146 of 170

1
2
3 191. Yu, Y.; Li, L.; Sheng, J. J., Further Discuss the Roles of Soaking Time and Pressure
4
5
6 Depletion Rate in Gas Huff-N-Puff Process in Fractured Liquid-Rich Shale Reservoirs. SPE
7
8 Annual Technical Conference and Exhibition 2016, 17, 10.2118/181471-MS.
9
10 192. Yu, Y.; Meng, X.; Sheng, J. J., Experimental and Numerical Evaluation of the Potential of
11
12
13
Improving Oil Recovery from Shale Plugs by Nitrogen Gas Flooding. Journal of Unconventional
14
15 Oil and Gas Resources 2016, 15, 56-65.
16
17 193. Yu, Y.; Sheng, J. J., A Comparative Experimental Study of Ior Potential in Fractured Shale
18
19
Reservoirs by Cyclic Water and Nitrogen Gas Injection. J. Pet. Sci. Eng. 2017, 149, 844-850.
20
21
22 194. Yu, Y.; Sheng, J. J., Experimental Evaluation of Shale Oil Recovery from Eagle Ford Core
23
24 Samples by Nitrogen Gas Flooding. SPE Improved Oil Recovery Conference 2016, 8,
25
26 10.2118/179547-MS.
27
28
29 195. Liu, G.; Sorensen, J. A.; Braunberger, J. R.; Klenner, R.; Ge, J.; Gorecki, C. D.; Steadman,
30
31 E. N.; Harju, J. A., CO2-Based Enhanced Oil Recovery from Unconventional Reservoirs: A Case
32
33 Study of the Bakken Formation. SPE Unconventional Resources Conference 2014, 7,
34
35
36
10.2118/168979-MS.
37
38 196. Torres, J. A.; Jin, L.; Bosshart, N. W.; Pekot, L. J.; Sorensen, J. A.; Peterson, K.; Anderson,
39
40 P. W.; Hawthorne, S. B., Multiscale Modeling to Evaluate the Mechanisms Controlling CO2-
41
42
Based Enhanced Oil Recovery and CO2 Storage in the Bakken Formation. SPE/AAPG/SEG
43
44
45 Unconventional Resources Technology Conference 2018, 20, 2902837-MS.
46
47 197. Gamadi, T.; Elldakli, F.; Sheng, J. J., Compositional Simulation Evaluation of EOR
48
49 Potential in Shale Oil Reservoirs by Cyclic Natural Gas Injection. In Unconventional Resources
50
51
52 Technology Conference, Denver, Colorado, 25-27 August 2014, 2014; 1625-1639.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
146
Page 147 of 170 Energy & Fuels

1
2
3 198. Wan, T.; Meng, X.; Sheng, J. J.; Watson, M., Compositional Modeling of EOR Process in
4
5
6 Stimulated Shale Oil Reservoirs by Cyclic Gas Injection. SPE Improved Oil Recovery Symposium
7
8 2014, 19, 10.2118/169069-MS.
9
10 199. Wan, T.; Sheng, J.; Soliman, M. Y.; Zhang, Y., Effect of Fracture Characteristics on
11
12
13
Behavior of Fractured Shale-Oil Reservoirs by Cyclic Gas Injection. SPE Reservoir Eval. Eng.
14
15 2016, 19, 350-355.
16
17 200. Wan, T.; Sheng, J., Compositional Modelling of the Diffusion Effect on EOR Process in
18
19
Fractured Shale-Oil Reservoirs by Gasflooding. J. Can. Pet. Technol. 2015, 54, 107-115.
20
21
22 201. Sanchez-Rivera, D.; Mohanty, K.; Balhoff, M., Reservoir Simulation and Optimization of
23
24 Huff-and-Puff Operations in the Bakken Shale. Fuel 2015, 147, 82-94.
25
26 202. Zhu, P.; Balhoff, M. T.; Mohanty, K. K., Simulation of Fracture-to-Fracture Gas Injection
27
28
29 in an Oil-Rich Shale. SPE Annual Technical Conference and Exhibition 2015, 18,
30
31 10.2118/175131-MS.
32
33 203. Zuloaga, P.; Yu, W.; Miao, J.; Sepehrnoori, K., Performance Evaluation of CO2 Huff-N-
34
35
36
Puff and Continuous CO2 Injection in Tight Oil Reservoirs. Energy 2017, 134, 181-192.
37
38 204. Sanaei, A.; Abouie, A.; Tagavifar, M.; Sepehrnoori, K., Comprehensive Study of Gas
39
40 Cycling in the Bakken Shale. SPE/AAPG/SEG Unconventional Resources Technology Conference
41
42
2018, 15, 2902940-MS.
43
44
45 205. Fai-Yengo, V. A.; Rahnema, H.; Alfi, M., Impact of Light Component Stripping During
46
47 CO2 Injection in Bakken Formation. SPE Unconventional Resources Conference 2014,
48
49 10.15530/urtec-2014-1922932.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
147
Energy & Fuels Page 148 of 170

1
2
3 206. Sun, J.; Zou, A.; Sotelo, E.; Schechter, D., Numerical Simulation of CO2 Huff-N-Puff in
4
5
6 Complex Fracture Networks of Unconventional Liquid Reservoirs. J. Nat. Gas Sci. Eng. 2016, 31,
7
8 481-492.
9
10 207. Sun, J.; Zou, A.; Schechter, D., Experimental and Numerical Studies of CO2 EOR in
11
12
13
Unconventional Liquid Reservoirs with Complex Fracture Networks. SPE Improved Oil Recovery
14
15 Conference 2016, 28, 10.2118/179634-MS.
16
17 208. Phi, T.; Schechter, D., CO2 EOR Simulation in Unconventional Liquid Reservoirs: An
18
19
Eagle Ford Case Study. SPE Unconventional Resources Conference 2017, 16, 10.2118/185034-
20
21
22 MS.
23
24 209. Iino, A.; Datta-Gupta, A., Optimizing CO2 and Field Gas Injection EOR in Unconventional
25
26 Reservoirs Using the Fast Marching Method. SPE Improved Oil Recovery Conference 2018, 21,
27
28
29 10.2118/190304-MS.
30
31 210. Yu, W.; Zhang, Y.; Varavei, A.; Sepehrnoori, K.; Zhang, T.; Wu, K.; Miao, J.,
32
33 Compositional Simulation of CO2 Huff-N-Puff in Eagle Ford Tight Oil Reservoirs with CO2
34
35
36
Molecular Diffusion, Nanopore Confinement and Complex Natural Fractures. SPE Improved Oil
37
38 Recovery Conference 2018, 26, 10.2118/190325-MS.
39
40 211. Yu, W.; Lashgari, H. R.; Wu, K.; Sepehrnoori, K., CO2 Injection for Enhanced Oil
41
42
Recovery in Bakken Tight Oil Reservoirs. Fuel 2015, 159, 354-363.
43
44
45 212. Xu, T.; Hoffman, T., Hydraulic Fracture Orientation for Miscible Gas Injection EOR in
46
47 Unconventional Oil Reservoirs. Unconventional Resources Technology Conference 2013, 11,
48
49 1580226-MS.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
148
Page 149 of 170 Energy & Fuels

1
2
3 213. Teklu, T. W.; Alharthy, N.; Kazemi, H.; Yin, X.; Graves, R. M., Hydrocarbon and Non-
4
5
6 Hydrocarbon Gas Miscibility with Light Oil in Shale Reservoirs. SPE Improved Oil Recovery
7
8 Symposium 2014, 9, 10.2118/169123-MS.
9
10 214. Pu, W.; Hoffman, B. T., Eos Modeling and Reservoir Simulation Study of Bakken Gas
11
12
13
Injection Improved Oil Recovery in the Elm Coulee Field, Montana. Unconventional Resources
14
15 Technology Conference 2014, 12, 10.15530/1922538.
16
17 215. Alfarge, D.; Wei, M.; Bai, B., Comparative Study for CO2-EOR and Natural Gases
18
19
Injection-Techniques for Improving Oil Recovery in Unconventional Oil Reservoirs. Carbon
20
21
22 Management Technology Conference 2017, 21, 10.7122/485175-MS.
23
24 216. Alfarge, D.; Wei, M.; Bai, B., Applicability of CO2-EOR in Shale-Oil Reservoirs Using
25
26 Diagnostic Plots. Carbon Management Technology Conference 2017, 18, 10.7122/502143-MS.
27
28
29 217. Alfarge, D.; Wei, M.; Bai, B., Factors Affecting CO2-EOR in Shale-Oil Reservoirs:
30
31 Numerical Simulation Study and Pilot Tests. Energy Fuels 2017, 31, 8462-8480.
32
33 218. Alfarge, D.; Wei, M.; Bai, B.; Almansour, A., Optimizing Injector-Producer Spacing for
34
35
36
CO2 Injection in Unconventional Reservoirs of North America. SPE Kingdom of Saudi Arabia
37
38 Annual Technical Symposium and Exhibition 2017, 16, 10.2118/188002-MS.
39
40 219. Alfarge, D.; Wei, M.; Bai, B.; Almansour, A., Effect of Molecular-Diffusion Mechanisim
41
42
on CO2 Huff-N-Puff Process in Shale-Oil Reservoirs. SPE Kingdom of Saudi Arabia Annual
43
44
45 Technical Symposium and Exhibition 2017, 17, 10.2118/188003-MS.
46
47 220. Alfarge, D.; Wei, M.; Bai, B.; Alsaba, M., Analysis of Ior Pilots in Bakken Formation by
48
49 Using Numerical Simulation. Abu Dhabi International Petroleum Exhibition & Conference 2017,
50
51
52 26, 10.2118/188633-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
149
Energy & Fuels Page 150 of 170

1
2
3 221. Alfarge, D.; Wei, M.; Bai, B., Numerical Simulation Study on Miscible EOR Techniques
4
5
6 for Improving Oil Recovery in Shale Oil Reservoirs. J. Pet. Explor. Prod. Technol. 2018, 8, 901-
7
8 916.
9
10 222. Alfarge, D.; Wei, M.; Bai, B.; Alsaba, M., Lessons Learned from Ior Pilots in Bakken
11
12
13
Formation by Using Numerical Simulation. J. Pet. Sci. Eng. 2018, 171, 1-15.
14
15 223. Alfarge, D.; Wei, M.; Bai, B., Co 2 -EOR Mechanisms in Huff-N-Puff Operations in Shale
16
17 Oil Reservoirs Based on History Matching Results. Fuel 2018, 226, 112-120.
18
19
224. Alfarge, D.; Wei, M.; Bai, B., Integrated Investigation of CO2-EOR Mechanisms in Huff-
20
21
22 N-Puff Operations Based on History Matching Results. SPE Improved Oil Recovery Conference
23
24 2018, 19, 10.2118/190234-MS.
25
26 225. Alfarge, D.; Wei, M.; Bai, B., Mechanistic Study for the Applicability of CO2-EOR in
27
28
29 Unconventional Liquids Rich Reservoirs. SPE Improved Oil Recovery Conference 2018, 29,
30
31 10.2118/190277-MS.
32
33 226. Dong, C.; Hoffman, B. T., Gas Injection to Increase Recovery from the Bakken Formation.
34
35
36
Int. J. Oil, Gas Coal Technol. 2015, 9, 148-168.
37
38 227. Wan, T.; Liu, H.-X., Exploitation of Fractured Shale Oil Resources by Cyclic CO2
39
40 Injection. Pet. Sci. 2018, 15, 552-563.
41
42
228. Wang, Y.; Hou, J.; Song, Z.; Yuan, D.; Zhang, J.; Zhao, T., A Case Study on Simulation
43
44
45 of in-Situ CO2 Huff-‘N’-Puff Process. SPE Reservoir Eval. Eng. 2018, 21, 109-121.
46
47 229. Cao, M.; Wu, X.; An, Y.; Zuo, Y.; Wang, R.; Li, P., Numerical Simulation and
48
49 Performance Evaluation of CO2 Huff-N-Puff Processes in Unconventional Oil Reservoirs. Carbon
50
51
52 Management Technology Conference 2017, 9, 10.7122/486556-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
150
Page 151 of 170 Energy & Fuels

1
2
3 230. Phan, T. N.; Cronk, B. R.; Almasoodi, M. M.; Reza, Z. A., Lithologic and Geomechanical
4
5
6 Control on CO2 Huff-N-Puff Enhanced Oil Recovery Processes Using Integrated Modeling
7
8 Framework in Wolfcamp. SPE/AAPG/SEG Unconventional Resources Technology Conference
9
10 2018, 20, 2901346-MS.
11
12
13
231. Akita, E.; Moghanloo, R. G.; Davudov, D.; Tinni, A., A Systematic Approach for
14
15 Upscaling of the EOR Results from Lab-Scale to Well-Scale in Liquid-Rich Shale Plays. SPE
16
17 Improved Oil Recovery Conference 2018, 15, 10.2118/190188-MS.
18
19
232. Kanfar, M. S.; Ghaderi, S. M.; Clarkson, C. R.; Reynolds, M. M.; Hetherington, C., A
20
21
22 Modeling Study of EOR Potential for CO2 Huff-N-Puff in Tight Oil Reservoirs - Example from
23
24 the Bakken Formation. SPE Unconventional Resources Conference 2017, 12, 10.2118/185026-
25
26 MS.
27
28
29 233. Haghshenas, B.; Qanbari, F.; Clarkson, C. R., Simulation of Enhanced Recovery Using
30
31 CO2 in a Liquid-Rich Western Canadian Unconventional Reservoir: Accounting for Reservoir
32
33 Fluid Adsorption and Compositional Heterogeneity. SPE Unconventional Resources Conference
34
35
36
2017, 19, 10.2118/185069-MS.
37
38 234. Cudjoe, S.; Vinassa, M.; Henrique Bessa Gomes, J.; Barati, R. G., A Comprehensive
39
40 Approach to Sweet-Spot Mapping for Hydraulic Fracturing and Co 2 Huff-N-Puff Injection in
41
42
Chattanooga Shale Formation. J. Nat. Gas Sci. Eng. 2016, 33, 1201-1218.
43
44
45 235. Jia, B.; Tsau, J.-S.; Barati, R., Role of Molecular Diffusion in Heterogeneous Shale
46
47 Reservoirs During CO2 Huff-N-Puff. SPE Europec featured at 79th EAGE Conference and
48
49 Exhibition 2017, 29, 10.2118/185797-MS.
50
51
52 236. Pu, H.; Li, Y., Novel Capillarity Quantification Method in Ior Process in Bakken Shale Oil
53
54 Reservoirs. SPE Improved Oil Recovery Conference 2016, 13, 10.2118/179533-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
151
Energy & Fuels Page 152 of 170

1
2
3 237. Pu, H.; Li, Y., CO2 EOR Mechanisms in Bakken Shale Oil Reservoirs. Carbon
4
5
6 Management Technology Conference 2015, 19, 10.7122/439769-MS.
7
8 238. Sahni, V.; Liu, S., Miscible EOR Process Assessment for Unconventional Reservoirs:
9
10 Understanding Key Mechanisms for Optimal Field Test Design. SPE/AAPG/SEG Unconventional
11
12
13
Resources Technology Conference 2018, 12, 2870010-MS.
14
15 239. Chen, C.; Gu, M., Investigation of Cyclic Co 2 Huff-and-Puff Recovery in Shale Oil
16
17 Reservoirs Using Reservoir Simulation and Sensitivity Analysis. Fuel 2017, 188, 102-111.
18
19
240. Pankaj, P.; Mukisa, H.; Solovyeva, I.; Xue, H., Boosting Oil Recovery in Naturally
20
21
22 Fractured Shale Using CO2 Huff-N-Puff. SPE Argentina Exploration and Production of
23
24 Unconventional Resources Symposium 2018, 15, 10.2118/191823-MS.
25
26 241. Kazempour, M.; Kiani, M.; Nguyen, D.; Salehi, M.; Bidhendi, M. M.; Lantz, M., Boosting
27
28
29 Oil Recovery in Unconventional Resources Utilizing Wettability Altering Agents: Successful
30
31 Translation from Laboratory to Field. SPE Improved Oil Recovery Conference 2018, 19,
32
33 10.2118/190172-MS.
34
35
36
242. Zhang, Y.; Di, Y.; Yu, W.; Sepehrnoori, K., A Comprehensive Model for Investigation of
37
38 Carbon Dioxide Enhanced Oil Recovery with Nanopore Confinement in the Bakken Tight Oil
39
40 Reservoir. SPE Reservoir Eval. Eng. 2018, Preprint, 15.
41
42
243. Alfarge, D.; Wei, M.; Bai, B., Data Analysis for CO2-EOR in Shale-Oil Reservoirs Based
43
44
45 on a Laboratory Database. J. Pet. Sci. Eng. 2018, 162, 697-711.
46
47 244. Jiang, J.; Younis, R. M., Compositional Modeling of Enhanced Hydrocarbons Recovery
48
49 for Fractured Shale Gas-Condensate Reservoirs with the Effects of Capillary Pressure and
50
51
52 Multicomponent Mechanisms. SPE Improved Oil Recovery Conference 2016, 27,
53
54 10.2118/179704-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
152
Page 153 of 170 Energy & Fuels

1
2
3 245. Enick, R. M.; Olsen, D. K.; Ammer, J. R.; Schuller, W., Mobility and Conformance Control
4
5
6 for CO2 EOR Via Thickeners, Foams, and Gels -- a Literature Review of 40 Years of Research
7
8 and Pilot Tests. SPE Improved Oil Recovery Symposium 2012, 12, 10.2118/154122-MS.
9
10 246. Zhang, P.; Tweheyo, M. T.; Austad, T., Wettability Alteration and Improved Oil Recovery
11
12
13
by Spontaneous Imbibition of Seawater into Chalk: Impact of the Potential Determining Ions
14
15 Ca2+, Mg2+, and So42−. Colloids and Surfaces A: Physicochemical and Engineering Aspects
16
17 2007, 301, 199-208.
18
19
247. Alvarez, J. O.; Schechter, D. S., Wettability Alteration and Spontaneous Imbibition in
20
21
22 Unconventional Liquid Reservoirs by Surfactant Additives. SPE Reservoir Eval. Eng. 2017, 20,
23
24 107-117.
25
26 248. Teklu, T. W.; Li, X.; Zhou, Z.; Alharthy, N.; Wang, L.; Abass, H., Low-Salinity Water and
27
28
29 Surfactants for Hydraulic Fracturing and EOR of Shales. J. Pet. Sci. Eng. 2018, 162, 367-377.
30
31 249. Nguyen, C.; Kothamasu, R.; He, K.; Xu, L., Low-Salinity Brine Enhances Oil Production
32
33 in Liquids-Rich Shale Formations. SPE Western Regional Meeting 2015, 8, 10.2118/174041-MS.
34
35
36
250. He, K.; Xu, L., Unique Mixtures of Anionic/Cationic Surfactants: A New Approach to
37
38 Enhance Surfactant Performance in Liquids-Rich Shale Reservoirs. SPE International Conference
39
40 on Oilfield Chemistry 2017, 10, 10.2118/184515-MS.
41
42
251. Valluri, M. K.; Alvarez, J. O.; Schechter, D. S., Study of the Rock/Fluid Interactions of
43
44
45 Sodium and Calcium Brines with Ultra-Tight Rock Surfaces and Their Impact on Improving Oil
46
47 Recovery by Spontaneous Imbibition. SPE Low Perm Symposium 2016, 17, 10.2118/180274-MS.
48
49 252. Zhang, J.; Wang, D.; Butler, R., Optimal Salinity Study to Support Surfactant Imbibition
50
51
52 into the Bakken Shale. SPE Unconventional Resources Conference Canada 2013, 17,
53
54 10.2118/167142-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
153
Energy & Fuels Page 154 of 170

1
2
3 253. Tang, G.-Q.; Morrow, N. R., Influence of Brine Composition and Fines Migration on
4
5
6 Crude Oil/Brine/Rock Interactions and Oil Recovery. J. Pet. Sci. Eng. 1999, 24, 99-111.
7
8 254. Lager, A.; Webb, K. J.; Black, C. J. J.; Singleton, M.; Sorbie, K. S. In Low Salinity Oil
9
10 Recovery - an Experimental Investigation, International Symposium of the Society of Core
11
12
13
Analysts Trondheim, Norway, 12-16 September, 2006; Society of Core Analysts: Trondheim,
14
15 Norway, 2006.
16
17 255. Lager, A.; Webb, K. J.; Collins, I. R.; Richmond, D. M., Losal Enhanced Oil Recovery:
18
19
Evidence of Enhanced Oil Recovery at the Reservoir Scale. SPE Symposium on Improved Oil
20
21
22 Recovery 2008, 12, 10.2118/113976-MS.
23
24 256. Ligthelm, D. J.; Gronsveld, J.; Hofman, J.; Brussee, N.; Marcelis, F.; van der Linde, H.,
25
26 Novel Waterflooding Strategy by Manipulation of Injection Brine Composition.
27
28
29 EUROPEC/EAGE Conference and Exhibition 2009, 22, 10.2118/119835-MS.
30
31 257. Alameri, W.; Teklu, T. W.; Graves, R. M.; Kazemi, H.; AlSumaiti, A. M., Wettability
32
33 Alteration During Low-Salinity Waterflooding in Carbonate Reservoir Cores. SPE Asia Pacific
34
35
36
Oil & Gas Conference and Exhibition 2014, 18, 10.2118/171529-MS.
37
38 258. Mahani, H.; Keya, A. L.; Berg, S.; Bartels, W.-B.; Nasralla, R.; Rossen, W. R., Insights
39
40 into the Mechanism of Wettability Alteration by Low-Salinity Flooding (LSF) in Carbonates.
41
42
Energy Fuels 2015, 29, 1352-1367.
43
44
45 259. Wang, D.; Butler, R.; Liu, H.; Ahmed, S., Flow-Rate Behavior and Imbibition in Shale.
46
47 SPE Reservoir Eval. Eng. 2011, 14, 485-492.
48
49 260. Morsy, S.; Sheng, J. J.; Soliman, M. Y., Waterflooding in the Eagle Ford Shaleformation:
50
51
52 Experimental and Simulation Study. SPE Unconventional Resources Conference and Exhibition-
53
54 Asia Pacific 2013, 11, 10.2118/167056-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
154
Page 155 of 170 Energy & Fuels

1
2
3 261. Wang, D.; Zhang, J.; Butler, R.; Koskella, D.; Rabun, R.; Clark, A., Flow Rate Behavior
4
5
6 and Imbibition Comparison between Bakken and Niobrara Formationsrch. Unconventional
7
8 Resources Technology Conference 2014, 13, 10.15530/URTEC-2014-1920887.
9
10 262. Morsy, S.; Sheng, J., Effect of Water Salinity on Shale Reservoir Productivity. Adv. Pet.
11
12
13
Explor. Dev. 2014, 8, 9-14.
14
15 263. Yu, Y.; Sheng, J. J., Experimental Investigation of Light Oil Recovery from Fractured
16
17 Shale Reservoirs by Cyclic Water Injection. SPE Western Regional Meeting 2016, 10,
18
19
10.2118/180378-MS.
20
21
22 264. Li, X.; Teklu, T. W.; Abass, H.; Cui, Q., The Impact of Water Salinity/Surfactant on
23
24 Spontaneous Imbibition through Capillarity and Osmosis for Unconventional Ior. SPE/AAPG/SEG
25
26 Unconventional Resources Technology Conference 2016, 19, 10.15530/2461736.
27
28
29 265. Mohan, K.; Leonard, P. A., Evaluating EOR Techniques in the Spraberry. SPE/AAPG/SEG
30
31 Unconventional Resources Technology Conference 2013, 8, 10.15530/URTEC-1579528-MS.
32
33 266. Kurtoglu, B.; Kazemi, H.; Rosen, R.; Mickelson, W.; Kosanke, T., A Rock and Fluid Study
34
35
36
of Middle Bakken Formation: Key to Enhanced Oil Recovery. SPE/CSUR Unconventional
37
38 Resources Conference – Canada 2014, 19, 10.2118/171668-MS.
39
40 267. Fakcharoenphol, P.; Kurtoglu, B.; Kazemi, H.; Charoenwongsa, S.; Wu, Y.-S., The Effect
41
42
of Osmotic Pressure on Improve Oil Recovery from Fractured Shale Formations. SPE
43
44
45 Unconventional Resources Conference 2014, 12, 10.2118/168998-MS.
46
47 268. Alvarez, J. O.; Schechter, D. S., Improving Oil Recovery in the Wolfcamp Unconventional
48
49 Liquid Reservoir Using Surfactants in Completion Fluids. J. Pet. Sci. Eng. 2017, 157, 806-815.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
155
Energy & Fuels Page 156 of 170

1
2
3 269. Alvarez, J. O.; Saputra, I. W. R.; Schechter, D. S., The Impact of Surfactant Imbibition and
4
5
6 Adsorption for Improving Oil Recovery in the Wolfcamp and Eagle Ford Reservoirs. SPE Annual
7
8 Technical Conference and Exhibition 2017, 25, 10.2118/187176-MS.
9
10 270. Alvarez, J. O.; Tovar, F. D.; Schechter, D. S., Improving Oil Recovery in Unconventional
11
12
13
Liquid Reservoirs by Soaking-Flowback Production Schedule with Surfactant Additives. SPE
14
15 Liquids-Rich Basins Conference - North America 2017, 18, 10.2118/187483-MS.
16
17 271. Karimi, S.; Kazemi, H., Characterizing Pores and Pore-Scale Flow Properties in Middle
18
19
Bakken Cores. SPE J. 2018, 23, 1343-1358.
20
21
22 272. Karimi, S.; Kazemi, H.; Simpson, G., Capillary Pressure and Wettability Indications of
23
24 Middle Bakken Core Plugs for Improved Oil Recovery. SPE Reservoir Eval. Eng. 2019, 22, 310-
25
26 325.
27
28
29 273. Wood, T. Waterflood Potential Could Unlock Billions of Barrels; 2011; 1-12.
30
31 274. Lake, L. W., Enhanced Oil Recovery. Prentice Hall: 1989, 550.
32
33 275. Delshad, M.; Bhuyan, D.; Pope, G. A.; Lake, L. W., Effect of Capillary Number on the
34
35
36
Residual Saturation of a Three-Phase Micellar Solution. SPE Enhanced Oil Recovery Symposium
37
38 1986, 12, 10.2118/14911-MS.
39
40 276. Terry, R., Enhanced Oil Recovery. Encyclopedia of Physical Science and Technology, 3rd
41
42
Edition ed.; Meyers, R. A., Ed. Academic Press Inc: 2001, 503-518.
43
44
45 277. Khishvand, M.; Akbarabadi, M.; Piri, M., Micro-Scale Experimental Investigation of the
46
47 Effect of Flow Rate on Trapping in Sandstone and Carbonate Rock Samples. Adv. Water Resour.
48
49 2016, 94, 379-399.
50
51
52 278. Jadhunandan, P. P.; Morrow, N. R., Effect of Wettability on Waterflood Recovery for
53
54 Crude-Oil/Brine/Rock Systems. SPE Reservoir Eng. 1995, 10, 40-46.
55
56
57
58
59
60 ACS Paragon Plus Environment
156
Page 157 of 170 Energy & Fuels

1
2
3 279. Morrow, N. R.; Mason, G., Recovery of Oil by Spontaneous Imbibition. Curr. Opin.
4
5
6 Colloid Interface Sci. 2001, 6, 321-337.
7
8 280. Hirasaki, G.; Zhang, D. L., Surface Chemistry of Oil Recovery from Fractured, Oil-Wet,
9
10 Carbonate Formations. SPE J. 2004, 9, 151-162.
11
12
13
281. Ayirala, S. C.; Yousef, A. A., Injection Water Chemistry Requirement Guidelines for
14
15 Ior/EOR. SPE Improved Oil Recovery Symposium 2014, 24, 10.2118/169048-MS.
16
17 282. Ayirala, S. C.; Xu, W.; Rao, D. N., Interfacial Behaviour of Complex Hydrocarbon Fluids
18
19
at Elevated Pressures and Temperatures. The Canadian Journal of Chemical Engineering 2006,
20
21
22 84, 22-32.
23
24 283. Wang, Y.; Xu, H.; Yu, W.; Bai, B.; Song, X.; Zhang, J., Surfactant Induced Reservoir
25
26 Wettability Alteration: Recent Theoretical and Experimental Advances in Enhanced Oil Recovery.
27
28
29 Pet. Sci. 2011, 8, 463-476.
30
31 284. Buckley, J. S.; Liu, Y.; Monsterleet, S., Mechanisms of Wetting Alteration by Crude Oils.
32
33 SPE J. 1998, 3, 54-61.
34
35
36
285. Standnes, D. C.; Austad, T., Wettability Alteration in Chalk: 2. Mechanism for Wettability
37
38 Alteration from Oil-Wet to Water-Wet Using Surfactants. J. Pet. Sci. Eng. 2000, 28, 123-143.
39
40 286. Standnes, D. C.; Nogaret, L. A. D.; Chen, H.-L.; Austad, T., An Evaluation of Spontaneous
41
42
Imbibition of Water into Oil-Wet Carbonate Reservoir Cores Using a Nonionic and a Cationic
43
44
45 Surfactant. Energy Fuels 2002, 16, 1557-1564.
46
47 287. Chen, H. L.; Lucas, L. R.; Nogaret, L. A. D.; Yang, H. D.; Kenyon, D. E., Laboratory
48
49 Monitoring of Surfactant Imbibition Using Computerized Tomography. SPE International
50
51
52 Petroleum Conference and Exhibition in Mexico 2000, 14, 10.2118/59006-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
157
Energy & Fuels Page 158 of 170

1
2
3 288. Kumar, K.; Dao, E. K.; Mohanty, K. K., Atomic Force Microscopy Study of Wettability
4
5
6 Alteration. SPE International Symposium on Oilfield Chemistry 2005, 11, 10.2118/93009-MS.
7
8 289. Patil, P. D.; Rohilla, N.; Katiyar, A.; Yu, W.; Nelson, C.; Falcone, S.; Rozowski, P.,
9
10 Surfactant Based EOR for Tight Oil Unconventional Reservoirs through Wettability Alteration:
11
12
13
Novel Surfactant Formulations and Their Efficacy to Induce Spontaneous Imbibition.
14
15 SPE/AAPG/SEG Unconventional Resources Technology Conference 2018, 10, 2896289-MS.
16
17 290. Rickman, R. D.; Jaripatke, O. A., Optimizing Micro-Emulsion / Surfactant Packages for
18
19
Shale and Tight Gas Reservoirs. SPE Deep Gas Conference and Exhibition 2010, 7,
20
21
22 10.2118/131107-MS.
23
24 291. Shuler, P. J.; Tang, H.; Lu, Z.; Tang, Y., Chemical Process for Improved Oil Recovery
25
26 from Bakken Shale. Canadian Unconventional Resources Conference 2011, 8, 10.2118/147531-
27
28
29 MS.
30
31 292. Rane, J. P.; Xu, L., New Dynamic-Surface-Tension Analysis Yields Improved Residual
32
33 Surfactant Measurements in Flowback and Produced Waters. SPE Prod. Oper. 2015, 30, 223-228.
34
35
36
293. Kim, J.; Zhang, H.; Sun, H.; Li, B.; Carman, P., Choosing Surfactants for the Eagle Ford
37
38 Shale Formation: Guidelines for Maximizing Flowback and Initial Oil Recovery. SPE Low Perm
39
40 Symposium 2016, 12, 10.2118/180227-MS.
41
42
294. He, K.; Nguyen, C.; Kothamasu, R.; Xu, L., Insights into Whether Low Salinity Brine
43
44
45 Enhances Oil Production in Liquids-Rich Shale Formations. EUROPEC 2015 2015, 10,
46
47 10.2118/174362-MS.
48
49 295. He, K.; Xu, L.; Kenzhekhanov, S.; Yin, X.; Neeves, K. B., A Rock-on-a-Chip Approach
50
51
52 to Study Fluid Invasion and Flowback in Liquids-Rich Shale Formations. SPE Oklahoma City Oil
53
54 and Gas Symposium 2017, 8, 10.2118/185088-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
158
Page 159 of 170 Energy & Fuels

1
2
3 296. Morsy, S.; Zhou, J.; Lant, K.; Cutler, J.; Sun, H.; Qu, Q.; Shuler, P., Optimizing Surfactant
4
5
6 Additives for Enhanced Well Stimulation in Bakken Formation. SPE International Symposium
7
8 and Exhibition on Formation Damage Control 2014, 11, 10.2118/168180-MS.
9
10 297. Xu, L.; Fu, Q., Ensuring Better Well Stimulation in Unconventional Oil and Gas
11
12
13
Formations by Optimizing Surfactant Additives. SPE Western Regional Meeting 2012, 7,
14
15 10.2118/154242-MS.
16
17 298. Alvarez, J. O.; Schechter, D. S., Altering Wettability in Bakken Shale by Surfactant
18
19
Additives and Potential of Improving Oil Recovery During Injection of Completion Fluids. SPE
20
21
22 Improved Oil Recovery Conference 2016, 20, 10.2118/179688-MS.
23
24 299. He, K.; Xu, L., Unique Mixtures of Anionic/Cationic Surfactants: A New Approach to
25
26 Enhance Surfactant Performance in Liquids-Rich Shale Reservoirs. SPE Prod. Oper. 2018, 33,
27
28
29 363-370.
30
31 300. Zeng, T.; S. Miller, C.; Mohanty, K., Application of Surfactants in Shale Chemical EOR
32
33 at High Temperatures. SPE Improved Oil Recovery Conference 2018, 14, 10.2118/190318-MS.
34
35
36
301. Mohanty, K. K.; Tong, S.; Miller, C.; Honarpour, M. M.; Turek, E.; Peck, D. D., Improved
37
38 Hydrocarbon Recovery Using Mixtures of Energizing Chemicals in Unconventional Reservoirs.
39
40 SPE Annual Technical Conference and Exhibition 2017, 21, 10.2118/187240-MS.
41
42
302. He, K.; Yue, Z.; Fan, C.; Xu, L., Minimizing Surfactant Adsorption Using Polyelectrolyte
43
44
45 Based Sacrificial Agent: A Way to Optimize Surfactant Performance in Unconventional
46
47 Formations. SPE International Symposium on Oilfield Chemistry 2015, 12, 10.2118/173750-MS.
48
49 303. Yarveicy, H.; Habibi, A.; Pegov, S.; Zolfaghari, A.; Dehghanpour, H., Enhancing Oil
50
51
52 Recovery by Adding Surfactants in Fracturing Water: A Montney Case Study. SPE Canada
53
54 Unconventional Resources Conference 2018, 13, 10.2118/189829-MS.
55
56
57
58
59
60 ACS Paragon Plus Environment
159
Energy & Fuels Page 160 of 170

1
2
3 304. Longoria, R. A.; Liang, T.; Huynh, U. T.; Nguyen, Q. P.; DiCarlo, D. A., Water Blocks in
4
5
6 Tight Formations: The Role of Matrix/Fracture Interaction in Hydrocarbon-Permeability
7
8 Reduction and Its Implications in the Use of Enhanced Oil Recovery Techniques. SPE J. 2017, 22,
9
10 1393-1401.
11
12
13
305. Liang, T.; Longoria, R. A.; Lu, J.; Nguyen, Q. P.; DiCarlo, D. A.; Huynh, U. T., The
14
15 Applicability of Surfactants on Enhancing the Productivity in Tight Formations. Unconventional
16
17 Resources Technology Conference 2015, 12, 10.15530/URTEC-2015-2154284.
18
19
306. Saputra, I. W. R.; Schechter, D. S., Comprehensive Workflow for Laboratory to Field-
20
21
22 Scale Numerical Simulation to Improve Oil Recovery in the Eagle Ford Shale by Selective Testing
23
24 and Modelling of Surfactants for Wettability Alteration. SPE/AAPG/SEG Unconventional
25
26 Resources Technology Conference 2018, 22, 2884598-MS.
27
28
29 307. Heyob, K. M.; Blotevogel, J.; Brooker, M.; Evans, M. V.; Lenhart, J. J.; Wright, J.;
30
31 Lamendella, R.; Borch, T.; Mouser, P. J., Natural Attenuation of Nonionic Surfactants Used in
32
33 Hydraulic Fracturing Fluids: Degradation Rates, Pathways, and Mechanisms. Environ. Sci.
34
35
36
Technol. 2017, 51, 13985-13994.
37
38 308. Chen, Y.; Elhag, A. S.; Poon, B. M.; Cui, L.; Ma, K.; Liao, S. Y.; Reddy, P. P.; Worthen,
39
40 A. J.; Hirasaki, G. J.; Nguyen, Q. P.; Biswal, S. L.; Johnston, K. P., Switchable Nonionic to
41
42
Cationic Ethoxylated Amine Surfactants for CO2 Enhanced Oil Recovery in High-Temperature,
43
44
45 High-Salinity Carbonate Reservoirs. SPE J. 2014, 19, 249-259.
46
47 309. Austad, T.; Milter, J., Spontaneous Imbibition of Water into Low Permeable Chalk at
48
49 Different Wettabilities Using Surfactants. International Symposium on Oilfield Chemistry 1997,
50
51
52 10, 10.2118/37236-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
160
Page 161 of 170 Energy & Fuels

1
2
3 310. Vijapurapu, C. S.; Rao, D. N., Compositional Effects of Fluids on Spreading, Adhesion
4
5
6 and Wettability in Porous Media. Colloids and Surfaces A: Physicochemical and Engineering
7
8 Aspects 2004, 241, 335-342.
9
10 311. Xie, X.; Weiss, W. W.; Tong, Z. J.; Morrow, N. R., Improved Oil Recovery from Carbonate
11
12
13
Reservoirs by Chemical Stimulation. SPE J. 2005, 10, 276-285.
14
15 312. Zhang, K.; Li, Y.; Hong, A.; Wu, K.; Jing, G.; Torsæter, O.; Chen, S.; Chen, Z., Nanofluid
16
17 Alternating Gas for Tight Oil Exploitation. SPE/IATMI Asia Pacific Oil & Gas Conference and
18
19
Exhibition 2015, 17, 10.2118/176241-MS.
20
21
22 313. Mirchi, V.; Saraji, S.; Goual, L.; Piri, M., Experimental Investigation of Surfactant
23
24 Flooding in Shale Oil Reservoirs: Dynamic Interfacial Tension, Adsorption, and Wettability.
25
26 Unconventional Resources Technology Conference 2014, 6, 10.15530/URTEC-2014-1913287.
27
28
29 314. Mirchi, V.; Saraji, S.; Goual, L.; Piri, M., Dynamic Interfacial Tensions and Contact
30
31 Angles of Surfactant-in-Brine/Oil/Shale Systems: Implications to Enhanced Oil Recovery in Shale
32
33 Oil Reservoirs. SPE Improved Oil Recovery Symposium 2014, 15, 10.2118/169171-MS.
34
35
36
315. Yu, H.; Rui, Z.; Chen, Z.; Lu, X.; Yang, Z.; Liu, J.; Qu, X.; Patil, S.; Ling, K.; Lu, J.,
37
38 Feasibility Study of Improved Unconventional Reservoir Performance with Carbonated Water and
39
40 Surfactant. Energy 2019, 182, 135-147.
41
42
316. Wang, D.; Butler, R.; Zhang, J.; Seright, R., Wettability Survey in Bakken Shale with
43
44
45 Surfactant-Formulation Imbibition. SPE Reservoir Eval. Eng. 2012, 15, 695-705.
46
47 317. Kathel, P.; Mohanty, K. K., EOR in Tight Oil Reservoirs through Wettability Alteration.
48
49 SPE Annual Technical Conference and Exhibition 2013, 15, 10.2118/166281-MS.
50
51
52 318. Kathel, P.; Mohanty, K. K., Wettability Alteration in a Tight Oil Reservoir. Energy Fuels
53
54 2013, 27, 6460-6468.
55
56
57
58
59
60 ACS Paragon Plus Environment
161
Energy & Fuels Page 162 of 170

1
2
3 319. Alamdari, B.; Hsu, T.-P.; Nguyen, D.; Kiani, M.; Salehi, M., Understanding the Oil
4
5
6 Recovery Mechanism in Mixed-Wet Unconventional Reservoirs: Uniqueness and Challenges of
7
8 Developing Chemical Formulations. SPE Improved Oil Recovery Conference 2018, 12,
9
10 10.2118/190201-MS.
11
12
13
320. Wang, D.; Zhang, J.; Butler, R.; Olatunji, K., Scaling Laboratory-Data Surfactant-
14
15 Imbibition Rates to the Field in Fractured-Shale Formations. SPE Reservoir Eval. Eng. 2016, 19,
16
17 440-449.
18
19
321. Nguyen, D.; Wang, D.; Oladapo, A.; Zhang, J.; Sickorez, J.; Butler, R.; Mueller, B.,
20
21
22 Evaluation of Surfactants for Oil Recovery Potential in Shale Reservoirs. SPE Improved Oil
23
24 Recovery Symposium 2014, 12, 10.2118/169085-MS.
25
26 322. Nguyen, D.; Phan, T.; Phan, J.; Hsu, T. P., Investigation of Oil Adhesion to Shale Rocks
27
28
29 for EOR. SPE International Conference on Oilfield Chemistry 2017, 14, 10.2118/184550-MS.
30
31 323. Chevalier, T.; Labaume, J.; Gautier, S.; Chevallier, E.; Chabert, M., A Novel Experimental
32
33 Approach for Accurate Evaluation of Chemical EOR Processes in Tight Reservoir Rocks. SPE
34
35
36
Improved Oil Recovery Conference 2018, 12, 10.2118/190171-MS.
37
38 324. Dawson, M.; Nguyen, D.; Champion, N.; Li, H., Designing an Optimized Surfactant Flood
39
40 in the Bakken. SPE/CSUR Unconventional Resources Conference 2015, 18, 10.2118/175937-MS.
41
42
325. Carpenter, C., Designing an Optimized Surfactant Flood in the Bakken. J. Pet. Technol.
43
44
45 2016, 68, 58-92.
46
47 326. Begum, M.; Reza Yassin, M.; Dehghanpour, H.; Dunn, L., Rock-Fluid Interactions in the
48
49 Duvernay Formation: Measurement of Wettability and Imbibition Oil Recovery. SPE
50
51
52 Unconventional Resources Conference 2017, 27, 10.2118/185065-MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
162
Page 163 of 170 Energy & Fuels

1
2
3 327. Yassin, M. R.; Dehghanpour, H.; Begum, M.; Dunn, L., Evaluation of Imbibition Oil
4
5
6 Recovery in the Duvernay Formation. SPE Reservoir Eval. Eng. 2018, 21, 257-272.
7
8 328. Wang, D.; Butler, R.; Liu, H.; Ahmed, S., Surfactant Formulation Study for Bakken Shale
9
10 Imbibition. SPE Annual Technical Conference and Exhibition 2011, 14, 10.2118/145510-MS.
11
12
13
329. Wang, D.; Dawson, M.; Butler, R.; Li, H.; Zhang, J.; Olatunji, K., Optimizing Water
14
15 Chemistry to Improve Oil Recovery from the Middle Bakken Formation. SPE Improved Oil
16
17 Recovery Conference 2016, 17, 10.2118/179541-MS.
18
19
330. Ma, S.; Morrow, N. R.; Zhang, X., Generalized Scaling of Spontaneous Imbibition Data
20
21
22 for Strongly Water-Wet Systems. J. Pet. Sci. Eng. 1997, 18, 165-178.
23
24 331. Alvarez, J. O.; Neog, A.; Jais, A.; Schechter, D. S., Impact of Surfactants for Wettability
25
26 Alteration in Stimulation Fluids and the Potential for Surfactant EOR in Unconventional Liquid
27
28
29 Reservoirs. SPE Unconventional Resources Conference 2014, 18, 10.2118/169001-MS.
30
31 332. Alvarez, J. O.; Schechter, D. S., Application of Wettability Alteration in the Exploitation
32
33 of Unconventional Liquid Resources. Adv. Pet. Explor. Dev. 2016, 43, 832-840.
34
35
36
333. Wang, M.; Abeykoon, G. A.; Vivas, F. A.; Okuno, R., Novel Wettability Modifiers for
37
38 Improved Oil Recovery in Tight Oil Reservoirs. SPE/AAPG/SEG Unconventional Resources
39
40 Technology Conference 2019, 20, 10.105530/1069.
41
42
334. Javanmard, H.; Seyyedi, M.; Nielsen, S. M., On Oil Recovery Mechanisms and Potential
43
44
45 of Dme–Brine Injection in the North Sea Chalk Oil Reservoirs. Ind. Eng. Chem. Res. 2018, 57,
46
47 15898-15908.
48
49 335. Zhang, F.; Adel, I. A.; Park, K. H.; Saputra, I. W. R.; Schechter, D. S., Enhanced Oil
50
51
52 Recovery in Unconventional Liquid Reservoir Using a Combination of CO2 Huff-N-Puff and
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
163
Energy & Fuels Page 164 of 170

1
2
3 Surfactant-Assisted Spontaneous Imbibition. SPE Annual Technical Conference and Exhibition
4
5
6 2018, 20, 10.2118/191502-MS.
7
8 336. Mohanty, K.; Zeng, T.; Miller, C.; Mohanty, K., Chemical Blend-CO2 Huff-N-Puff for
9
10 Enhanced Oil Recovery in Shales. SPE/AAPG/SEG Unconventional Resources Technology
11
12
13
Conference 2019, 10, 10.105530/urtec-2019-362.
14
15 337. Bui, K.; Akkutlu, I.; Zelenev, A.; Saboowala, H.; Gillis, J.; Silas, J., Insights into
16
17 Mobilization of Shale Oil Using Microemulsion. In Unconventional Resources Technology
18
19
Conference, San Antonio, Texas, 20-22 July 2015, Society of Exploration Geophysicists,
20
21
22 American Association of Petroleum Geologists, Society of Petroleum Engineers: 2015; 837-848.
23
24 338. Bui, K.; Akkutlu, I. Y.; Zelenev, A. S.; Silas, J., Understanding Penetration Behavior of
25
26 Microemulsions into Shale Nanopores. SPE Europec featured at 79th EAGE Conference and
27
28
29 Exhibition 2017, 12, 10.2118/185787-MS.
30
31 339. Bui, B. T. A Multi-Physics Model for Enhanced Oil Recovery in Liquid-Rich
32
33 Unconventional Reservoirs. Colorado School of Mines, Golden, Colorado, 2016.
34
35
36
340. Bui, K.; Akkutlu, I. Y.; Zelenev, A.; Hill, W. A., Kerogen Maturation Effects on Pore
37
38 Morphology and Enhanced Shale Oil Recovery. SPE Europec featured at 80th EAGE Conference
39
40 and Exhibition 2018, 13, 10.2118/190818-MS.
41
42
341. Sheshdeh, M. J., A Review Study of Wettability Alteration Methods with Regard to Nano-
43
44
45 Materials Application. SPE Bergen One Day Seminar 2015, 12, 10.2118/173884-MS.
46
47 342. Onyekonwu, M. O.; Ogolo, N. A., Investigating the Use of Nanoparticles in Enhancing Oil
48
49 Recovery. Nigeria Annual International Conference and Exhibition 2010, 14, 10.2118/140744-
50
51
52 MS.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
164
Page 165 of 170 Energy & Fuels

1
2
3 343. Hendraningrat, L.; Li, S.; Torsæter, O., A Coreflood Investigation of Nanofluid Enhanced
4
5
6 Oil Recovery. J. Pet. Sci. Eng. 2013, 111, 128-138.
7
8 344. Hendraningrat, L.; Li, S.; Torsaeter, O., Enhancing Oil Recovery of Low-Permeability
9
10 Berea Sandstone through Optimised Nanofluids Concentration. SPE Enhanced Oil Recovery
11
12
13
Conference 2013, 10, 10.2118/165283-MS.
14
15 345. Hendraningrat, L.; Torsæter, O., Metal Oxide-Based Nanoparticles: Revealing Their
16
17 Potential to Enhance Oil Recovery in Different Wettability Systems. Appl. Nanosci. 2015, 5, 181-
18
19
199.
20
21
22 346. Ogolo, N. C.; Olafuyi, O. A.; Onyekonwu, M., Effect of Nanoparticles on Migrating Fines
23
24 in Formations. SPE International Oilfield Nanotechnology Conference and Exhibition 2012, 12,
25
26 10.2118/155213-MS.
27
28
29 347. Karimi, A.; Fakhroueian, Z.; Bahramian, A.; Pour Khiabani, N.; Darabad, J. B.; Azin, R.;
30
31 Arya, S., Wettability Alteration in Carbonates Using Zirconium Oxide Nanofluids: EOR
32
33 Implications. Energy Fuels 2012, 26, 1028-1036.
34
35
36
348. Giraldo, J.; Benjumea, P.; Lopera, S.; Cortés, F. B.; Ruiz, M. A., Wettability Alteration of
37
38 Sandstone Cores by Alumina-Based Nanofluids. Energy Fuels 2013, 27, 3659-3665.
39
40 349. Mohammadi, M. S.; Moghadasi, J.; Naseri, S., An Experimental Investigation of
41
42
Wettability Alteration in Carbonate Reservoir Using Γ-Al2o3 Nanoparticles. Iranian J. Oil & Gas
43
44
45 Sci. and Tech. 2014, 3, 18-26.
46
47 350. Al-Muntasheri, G. A.; Liang, F.; Hull, K. L., Nanoparticle-Enhanced Hydraulic-Fracturing
48
49 Fluids: A Review. SPE Prod. Oper. 2017, 32, 186-195.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
165
Energy & Fuels Page 166 of 170

1
2
3 351. Mohammadi, M.; Hemmati-Sarapardeh, A.; Sedighi, M., Application of Nanofluids in
4
5
6 Enhanced Oil Recovery: A Systematic Literature Review and Organizing Framework. 2019; 433-
7
8 451.
9
10 352. Yekeen, N.; Padmanabhan, E.; Idris, A. K.; Chauhan, P. S., Nanoparticles Applications for
11
12
13
Hydraulic Fracturing of Unconventional Reservoirs: A Comprehensive Review of Recent
14
15 Advances and Prospects. J. Pet. Sci. Eng. 2019, 178, 41-73.
16
17 353. Cheraghian, G.; Hendraningrat, L., A Review on Applications of Nanotechnology in the
18
19
Enhanced Oil Recovery Part A: Effects of Nanoparticles on Interfacial Tension. Int. Nano Lett.
20
21
22 2016, 6, 129-138.
23
24 354. Syfan, F. E.; Holcomb, D. L.; Lowrey, T. A.; Nickerson, R. L.; Sam, A. B.; Ahmad, Y.,
25
26 Enhancing Delaware Basin Stimulation Results Using Nanoparticle Dispersion Technology. SPE
27
28
29 Hydraulic Fracturing Technology Conference and Exhibition 2018, 24, 10.2118/189876-MS.
30
31 355. Wasan, D. T.; McNamara, J. J.; Shah, S. M.; Sampath, K.; Aderangi, N., The Role of
32
33 Coalescence Phenomena and Interfacial Rheological Properties in Enhanced Oil Recovery: An
34
35
36
Overview. J. Rheol. 1979, 23, 181-207.
37
38 356. Wasan, D. T.; Nikolov, A. D., Spreading of Nanofluids on Solids. Nature 2003, 423, 156.
39
40 357. Kondiparty, K.; Nikolov, A.; Wu, S.; Wasan, D., Wetting and Spreading of Nanofluids on
41
42
Solid Surfaces Driven by the Structural Disjoining Pressure: Statics Analysis and Experiments.
43
44
45 Langmuir 2011, 27, 3324-3335.
46
47 358. Zhang, H.; Nikolov, A.; Wasan, D., Enhanced Oil Recovery (EOR) Using Nanoparticle
48
49 Dispersions: Underlying Mechanism and Imbibition Experiments. Energy Fuels 2014, 28, 3002-
50
51
52 3009.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
166
Page 167 of 170 Energy & Fuels

1
2
3 359. Cheraghian, G., Effects of Nanoparticles on Wettability: A Review on Applications of
4
5
6 Nanotechnology in the Enhanced Oil Recovery. Int. J. Nano Dimens. 2015, 6, 443-452.
7
8 360. Kuang, W.; Saraji, S.; Piri, M., A Systematic Experimental Investigation on the Synergistic
9
10 Effects of Aqueous Nanofluids on Interfacial Properties and Their Implications for Enhanced Oil
11
12
13
Recovery. Fuel 2018, 220, 849-870.
14
15 361. Wang, T.; Zhang, Y.; Li, L.; Yang, Z.; Liu, Y.; Fang, J.; Dai, C.; You, Q., Experimental
16
17 Study on Pressure-Decreasing Performance and Mechanism of Nanoparticles in Low Permeability
18
19
Reservoir. J. Pet. Sci. Eng. 2018, 166, 693-703.
20
21
22 362. Dai, C.; Wang, S.; Li, Y.; Gao, M.; Liu, Y.; Sun, Y.; Zhao, M., The First Study of Surface
23
24 Modified Silica Nanoparticles in Pressure-Decreasing Application. RSC Adv. 2015, 5, 61838-
25
26 61845.
27
28
29 363. Franco-Aguirre, M.; Zabala, R. D.; Lopera, S. H.; Franco, C. A.; Cortés, F. B., Interaction
30
31 of Anionic Surfactant-Nanoparticles for Gas - Wettability Alteration of Sandstone in Tight Gas-
32
33 Condensate Reservoirs. J. Nat. Gas Sci. Eng. 2018, 51, 53-64.
34
35
36
364. Willhite, G. P., Waterflooding. Society of Petroleum Engineers: Richardson, TX, 1986,
37
38 Vol. 3.
39
40 365. Green, D. W.; Willhite, G. P., Enhanced Oil Recovery. Society of Petroleum Engineers:
41
42
Richardson, Texas, 1998, Vol. 6.
43
44
45 366. Naik, S.; You, Z.; Bedrikovetsky, P., Effect of Wettability Alteration on Productivity
46
47 Enhancement in Unconventional Gas Reservoirs: Application of Nanotechnology. SPE Asia
48
49 Pacific Unconventional Resources Conference and Exhibition 2015, 17, 10.2118/177021-MS.
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
167
Energy & Fuels Page 168 of 170

1
2
3 367. Li, J.; Li, X.; Wu, K.; Feng, D.; Zhang, T.; Zhang, Y., Thickness and Stability of Water
4
5
6 Film Confined inside Nanoslits and Nanocapillaries of Shale and Clay. Int. J. Coal Geol. 2017,
7
8 179, 253-268.
9
10 368. Quintero, H.; Mattucci, M.; Hawkes, R.; Zhang, K.; O'Neil, B., Nano-Particle Surfactant
11
12
13
in Hydraulic Fracturing Fluids for Enhanced Post Frac Oil Recovery. SPE Canada Unconventional
14
15 Resources Conference 2018, 25, 10.2118/189780-MS.
16
17 369. Mahmoudkhani, A.; O'Neil, B.; Wylde, J. J.; Kakadjian, S.; Bauer, M., Microemulsions as
18
19
Flowback Aids for Enhanced Oil and Gas Recovery after Fracturing, Myth or Reality: A Turnkey
20
21
22 Study to Determine the Features and Benefits. SPE International Symposium on Oilfield Chemistry
23
24 2015, 21, 10.2118/173729-MS.
25
26 370. Li, J.; Yu, W.; Wu, K., Analyzing the Impact of Fracture Complexity on Well Performance
27
28
29 and Wettability Alteration in Eagle Ford Shale. SPE/AAPG/SEG Unconventional Resources
30
31 Technology Conference 2018, 15, 2899349-MS.
32
33 371. Sheng, J. J., Critical Review of Field EOR Projects in Shale and Tight Reservoirs. J. Pet.
34
35
36
Sci. Eng. 2017, 159, 654-665.
37
38 372. Sorensen, J. A.; Pekot, L. J.; Torres, J. A.; Jin, L.; Hawthorne, S. B.; Smith, S. A.; Jacobson,
39
40 L. L.; Doll, T. E., Field Test of CO2 Injection in a Vertical Middle Bakken Well to Evaluate the
41
42
Potential for Enhanced Oil Recovery and CO2 Storage. SPE/AAPG/SEG Unconventional
43
44
45 Resources Technology Conference 2018, 18, 2902813-MS.
46
47 373. Schmidt, M.; Sekar, B. K., Innovative Unconventional2 EOR-a Light EOR an
48
49 Unconventional Tertiary Recovery Approach to an Unconventional Bakken Reservoir in Southeast
50
51
52 Saskatchewan. 21st World Petroleum Congress 2014, 12.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
168
Page 169 of 170 Energy & Fuels

1
2
3 374. Hoffman, B. T., Huff-N-Puff Gas Injection Pilot Projects in the Eagle Ford. SPE Canada
4
5
6 Unconventional Resources Conference 2018, 18, 10.2118/189816-MS.
7
8 375. Tuero, F. R.; Crotti, M.; Labayen, I., Unconventional EOR: A Capillary Based Improved
9
10 Oil Recovery Case Study for Shale Oil Scenarios in the Vaca Muerta Resource Play.
11
12
13
SPE/AAPG/SEG Unconventional Resources Technology Conference 2017, 15,
14
15 10.15530/2659910.
16
17 376. Jiang, H.; Lei, Y.; Xiong, X.; Yan, L.; Pi, X.; Li, X.; Yu, C., A CO2 Immiscible
18
19
Displacement Experiment Study Aiming at Fuyang Extra-Low Permeability Layer at Peripheral
20
21
22 of Daqing Placanticline. Geoscience 2008, 22, 659-663.
23
24 377. Wang, Y.-Y., CO2 Flooding Test in Fayang Reservoir in Daqing Yushulin Oilfield.
25
26 Petroleum Geol. Oilfield Dev. Daqing 2015, 34, 136–139.
27
28
29 378. Lin, Y. Y.; Wang, P. P.; Li, Q. D.; Han, B. M.; Li, H. L.; Wu, Y. P.; He, W. S., Performance
30
31 Analysis of Different Modes of Huff-N-Puff Water Injection in Horizontal Wells in an-83 Chang-
32
33 7 Tight Oil Reservoir. Petrochem. Ind. Appl. 2016, 35, 94-97.
34
35
36
379. Wei, F. R., Discussion of High-Volume Water Injection and Soaking Effect in an Ultra-
37
38 Low Permeability Tight Formation. Explor. Dev. 2016, 3, 155-157.
39
40 380. Li, Z.; Qu, X.; Liu, W.; Lei, Q.; Sun, H.; He, Y., Development Modes of Triassic Yanchang
41
42
Formation Chang 7 Member Tight Oil in Ordos Basin, Nw China. Adv. Pet. Explor. Dev. 2015,
43
44
45 42, 241-246.
46
47 381. Sorensen, J. A.; Hamling, J. A. Historical Bakken Test Data Provide Critical Insights on
48
49 EOR in Tight Oil Plays The American Oil and Gas Reporter [Online], 2016.
50
51
52 https://www.aogr.com/magazine/cover-story/historical-bakken-test-data-provide-critical-
53
54 insights-on-eor-in-tight-oil-p.
55
56
57
58
59
60 ACS Paragon Plus Environment
169
Energy & Fuels Page 170 of 170

1
2
3 382. Tuero, F.; Crotti M, M.; Labayen, I., Water Imbibition EOR Proposal for Shale Oil
4
5
6 Scenarios. SPE Latin America and Caribbean Petroleum Engineering Conference 2017, 18,
7
8 10.2118/185560-MS.
9
10 383. Ling, K.; Shen, Z.; Han, G.; He, J.; Peng, P., A Review of Enhanced Oil Recovery Methods
11
12
13
Applied in Williston Basin. Unconventional Resources Technology Conference 2014, 16,
14
15 10.15530/1891560.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
170

You might also like