You are on page 1of 29

Energy Reports 6 (2020) 3150–3178

Contents lists available at ScienceDirect

Energy Reports
journal homepage: www.elsevier.com/locate/egyr

Review article

The use of surfactants in enhanced oil recovery: A review of recent


advances

Osama Massarweh, Ahmad S. Abushaikha
Division of Sustainable Development, College of Science and Engineering, Hamad Bin Khalifa University, Education City, Qatar Foundation, P.O. Box
34110, Doha, Qatar

article info a b s t r a c t

Article history: Currently, there is a widespread interest in the different methods of chemical enhanced oil recovery
Received 2 September 2020 (EOR) as a result of the continuous decline in the conventional oil reserves and the accelerated increase
Received in revised form 27 October 2020 in the global energy demand. Surfactant flooding is a well-established method of chemical EOR. This
Accepted 1 November 2020
method has proven successful as it increases oil recovery through a combination of mechanisms. These
Available online xxxx
include interfacial tension (IFT) reduction, wettability alteration, foam generation and emulsification.
Keywords: Despite its popularity, surfactant flooding is still challenged by issues including instability under harsh
Enhanced oil recovery (or normal) reservoir conditions and excessive adsorption. These issues affect the expected oil recovery
Surfactant flooding and thereby reduce the economic returns of EOR projects. Nevertheless, surfactants can be properly
Interfacial tension (IFT) selected according to reservoir conditions and rock type. This is usually carried out using surfactant
Wettability alteration
screening methods, which impose limits related to the IFT, surfactant adsorption and other factors
under given temperature and salinity conditions. This paper reviews surfactant characterization and
phase behavior, the role of surfactants in oil recovery, surfactant adsorption onto reservoir rock,
and the application of surfactants in EOR on both laboratory and field scales. Finally, the review
presents current research trends and future prospects based on recently published studies in the area
of surfactant flooding.
© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).

Contents

1. Introduction..................................................................................................................................................................................................................... 3151
2. Background on surfactants ............................................................................................................................................................................................ 3152
2.1. Definition of surfactants ................................................................................................................................................................................... 3152
2.2. Characterization of surfactants ........................................................................................................................................................................ 3152
2.2.1. Critical micelle concentration ........................................................................................................................................................... 3153
2.2.2. Krafft temperature and Krafft point ................................................................................................................................................ 3153
2.2.3. Cloud point ......................................................................................................................................................................................... 3153
2.2.4. Hydrophile–lipophile balance ........................................................................................................................................................... 3154
2.2.5. Solubilization parameters.................................................................................................................................................................. 3154
2.2.6. Microemulsion systems ..................................................................................................................................................................... 3155
2.2.7. Winsor’s R-ratio.................................................................................................................................................................................. 3155
2.2.8. Hydrophilic–lipophilic deviation (HLD) ........................................................................................................................................... 3155
2.2.9. Molecular packing parameter ........................................................................................................................................................... 3156
3. The role of surfactants in oil recovery ........................................................................................................................................................................ 3156
3.1. Interfacial tension reduction ............................................................................................................................................................................ 3156
3.2. Wettability alteration ........................................................................................................................................................................................ 3157
3.3. Foam generation ................................................................................................................................................................................................ 3158
3.4. Emulsification ..................................................................................................................................................................................................... 3158
4. Surfactant adsorption onto reservoir rock .................................................................................................................................................................. 3159
4.1. Influence of surfactant adsorption on EOR..................................................................................................................................................... 3159
4.2. Surfactant adsorption isotherms ...................................................................................................................................................................... 3159

∗ Corresponding author.
E-mail address: aabushaikha@hbku.edu.qa (A.S. Abushaikha).

https://doi.org/10.1016/j.egyr.2020.11.009
2352-4847/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

4.3. Factors affecting surfactant adsorption ........................................................................................................................................................... 3160


5. Application of surfactants in chemical EOR ................................................................................................................................................................ 3160
5.1. Anionic surfactants ............................................................................................................................................................................................ 3160
5.1.1. Sodium dodecyl sulfate ..................................................................................................................................................................... 3160
5.1.2. Internal olefin sulfonates .................................................................................................................................................................. 3161
5.1.3. Alpha olefin sulfonates ...................................................................................................................................................................... 3162
5.1.4. Alkyl aryl sulfonates .......................................................................................................................................................................... 3163
5.2. Cationic surfactants ........................................................................................................................................................................................... 3163
5.2.1. Cetyltrimethylammonium bromide (CTAB)..................................................................................................................................... 3164
5.2.2. Dodecyltrimethylammonium bromide (DTAB) ............................................................................................................................... 3164
5.2.3. Alkyltrimethylammonium chlorides ................................................................................................................................................ 3164
5.3. Nonionic surfactants.......................................................................................................................................................................................... 3164
5.4. Zwitterionic surfactants .................................................................................................................................................................................... 3165
5.5. Surfactant screening criteria............................................................................................................................................................................. 3166
5.6. Field-scale application of surfactant-assisted chemical EOR ........................................................................................................................ 3166
5.7. Application of nanoparticles in surfactant flooding ...................................................................................................................................... 3168
6. Summary and future prospects .................................................................................................................................................................................... 3170
Declaration of competing interest................................................................................................................................................................................ 3170
Acknowledgments .......................................................................................................................................................................................................... 3170
References ....................................................................................................................................................................................................................... 3170

Once primary oil recovery is no more feasible, secondary oil


1. Introduction recovery is implemented to drive additional crude oil out of the
reservoir using pressure maintenance techniques such as water-
Worldwide energy demand is estimated to increase by 30% flooding and gas injection (Haq et al., 2020b,a). In waterflooding,
in 2040 compared to 2010 (Karatayev et al., 2019). It is also water is injected into a reservoir using a number of injection
estimated that by 2040, oil consumption will reach 111.1 mil- wells to maintain the pressure. The displaced oil is collected
lion barrels per day (Zhang et al., 2020). With depleting oil and produced using a number of production wells. However,
reserves and growing energy demand, due to population growth water injection does not extract all the oil from the reservoir
and rapid industrial development, it has become more important for two main reasons. Firstly, due to reservoir heterogeneities,
to improve oil recovery from the declining oil reservoirs (Joshi water may flow in highly permeable pathways that exist between
et al., 2015). The process of oil recovery is divided into three injection wells and production wells. This leaves several regions
main phases including primary, secondary and tertiary phases. of the reservoir unswept by the waterflood. Secondly, oil ganglia
In the petroleum industry, the primary and secondary phases surrounded by water are trapped within the small interstices of
are referred to as the conventional methods of oil extraction. On the rock matrix due to oil–water surface tension preventing oil
the other hand, the tertiary phase is referred to as enhanced oil from flowing (Blunt et al., 1993).
recovery (EOR) (Pogaku et al., 2018). As indicated in Table 1, significant quantities of crude oil
The primary recovery produces less than 30% of original-oil- remain unrecovered following the primary and secondary phases
in-place (OOIP) through natural flow and artificial lift (Pogaku of oil recovery. For this reason, a tertiary recovery phase, also
et al., 2018; Saravanan and Keerthana, 2017). Initially, crude oil known as EOR, has been introduced to boost the recovery from oil
naturally flows out of a reservoir due to its own pressure (Grassia reservoirs. EOR involves a variety of operations such as chemical
et al., 2016; Miller and Sorrell, 2014). Several mechanisms are flooding (Brantson et al., 2020; Jin et al., 2020), gas injection (Song
involved in this process including solution-gas drive (important et al., 2020; Yu et al., 2020), microbial recovery (Nikolova and
in heavy-oil reservoirs) (He et al., 2019), gas-cap drive (Rezaei Gutierrez, 2020; Niu et al., 2020) and thermal recovery (Dorane-
et al., 2020b), water drive (Tayari et al., 2018), rock and liquid hgard and Siavashi, 2018; Mokheimer et al., 2018). Fig. 1 presents
expansion (Agi et al., 2018; Gyan et al., 2019) and gravity drainage the main methods of EOR.
(Aljuboori et al., 2019). In many cases, the primary recovery is The different EOR operations ultimately aim to improve the
supported by a combination of these driving mechanisms (Gyan overall oil displacement efficiency, which depends on the mi-
et al., 2019). croscopic and macroscopic displacement efficiencies (Kumar and
Continuous oil extraction causes the pressure gradient to drop Mandal, 2017b; Riahinezhad et al., 2017). Such improvement is
in the reservoir and hence oil production rates decrease, in accor- achieved by influencing one or more of the following: oil viscosity
dance with Darcy’s law (Höök et al., 2014). To achieve higher rates (Zhu et al., 2018), rock wettability (El-hoshoudy et al., 2017),
of oil production, the drawdown pressure is usually increased by interfacial tension (IFT) (AfzaliTabar et al., 2020) and capillary
lowering the bottom-hole pressure (BHP) in the production well. forces (You et al., 2018) in addition to altering the mobility ratio
This is accomplished by applying artificial lift, which compensates between the displacing and displaced fluids to more favorable
for the reduction in energy supplied by natural drive mechanisms values (Haruna et al., 2020).
after years of oil production (Davarpanah and Mirshekari, 2018). A wide variety of chemical agents are employed in chemical
There are several artificial lift systems that have been used world- EOR including surfactants (Sheng, 2015), ionically modified water
wide. Examples of these systems include hydraulic jet pumping, (Rezaeian et al., 2020), alkalis (Fakher et al., 2019), polymers
gas lift, plunger lift, beam pumping, hydraulic piston pumping (Agi et al., 2018) and nanoparticles (Olayiwola and Dejam, 2019).
and others. The selection of an artificial lift system depends on Surfactants are widely used in the petroleum industry because
determining factors including downhole pressure/temperature, they have the ability to affect the water/oil interface and the
fluid properties, completion type and hole characteristics. The properties of the rock surface (Zulkifli et al., 2020).
factors also include well location, operating and service per- Surfactant formulations should be optimized under reservoir
sonnel, surface climate, available power sources, economics and conditions before being employed in oil recovery applications.
others (Brown, 1982). Usually, this is carried out using surfactant screening, which is
3151
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 1. Main methods of EOR (Brantson et al., 2020; Gbadamosi et al., 2019b, 2018; Mokheimer et al., 2018; Shafiai and Gohari, 2020).

Table 1 Foroozesh and Kumar, 2020), which are discussed in other parts
Recovery factors following the primary and secondary phases of oil recovery. of this review.
Reference Oil recovery factor In this paper, we review the literature on the use of surfac-
Maurya et al. (2017) 20%–30% tants in chemical EOR. The review covers the following topics:
Sengupta et al. (2012) 30%–40% background on surfactants (Section 2), the role of surfactants
Riazi and Golkari (2016) 30%–50%
Kok (2012) 30%–50%
in oil recovery (Section 3), surfactant adsorption onto reservoir
Desnelli et al. (2015) 40%–50% rock (Section 4) and application of surfactants in chemical EOR
Kumar and Mandal (2017b) ∼30% (Section 5). Finally, in Section 6, we summarize the review and
Ali et al. (2020) ∼50% present future prospects based on recently published studies.
Ahmadi et al. (2016) <50%

2. Background on surfactants
still a challenging process that requires time and material re-
sources (Jin et al., 2017; Wang et al., 2019). There are two main Prior to discussing surfactants in relation to EOR, a brief back-
reasons that make surfactant screening a difficult task. First, most ground on surfactants is provided in this section. This background
types of surfactants are highly affected by the conditions present includes the definition of surfactants and the main parameters
in the oil reservoir in addition to the type of the reservoir rock examined during surfactant characterization.
(Ivanova et al., 2020; Riswati et al., 2019). Second, there is a wide
range of surfactants that have the potential to be used for EOR 2.1. Definition of surfactants
(Miller et al., 2020; Trickett et al., 2010); therefore, starting with
a certain group of surfactants to be screened is also challenging. Surfactants (surface active agents) are usually organic com-
Because of the significance of this issue, we present in this review pounds consist of two distinctive parts: a hydrophilic head and
the parameters and the methods that are usually considered a hydrophobic tail. There are four classifications of surfactants
during surfactant screening. We also present a classification of according to the charge existing on the hydrophilic head. These
the most commonly applied surfactants in chemical EOR. classifications include non-ionic (without any charge), anionic
Surfactant adsorption is an important parameter that affects (carrying negative charge), cationic (carrying positive charge),
the efficiency of surfactant flooding and hence the economic and zwitterionic (carrying both negative and positive charges)
viability of EOR projects (Amirmoshiri et al., 2020; ShamsiJazeyi (Kume et al., 2008), as shown in Fig. 2. Generally, surfactants
et al., 2014). Research studies showed that surfactant adsorption have a chain structure (Azam et al., 2013; Burnham et al., 2013),
is affected by the temperature, salinity, pH and other reservoir which is mainly attributed to the structure of the hydrophobic
parameters (Puntervold et al., 2018; Zhang et al., 2019a). Re- tail. According to Gbadamosi et al. (2019b), the tail group of
cently, significant research effort has focused on improving the a surfactant is often formed of a short polymer chain, a long
methods used to control surfactant adsorption, thereby reducing hydrocarbon chain, a siloxane chain or a fluorocarbon chain. The
its negative impacts on oil recovery. Common examples of these head group, on the other hand, is formed of moieties such as
methods include tuning the salinity and ionic composition of sulfates, sulfonates, polyoxyethylene chains, carboxylates, alco-
the injection water (Ahmadi and Shadizadeh, 2018; Hossein- hols or quaternary ammonium salts. The presence of these groups
zade Khanamiri et al., 2016a), alkali addition (Hazarika and Gogoi, dictates the amphiphilic nature of the surfactant.
2020; Phukan et al., 2020) and nanoparticle (NP) addition (Cher-
aghian, 2017; Cheraghian and Nezhad, 2016; Kumar and Mandal, 2.2. Characterization of surfactants
2020). Beside reducing surfactant adsorption, some types of NPs
tend to improve the rheological properties of surfactant solutions It is necessary characterize surfactants for technical and eco-
(Almahfood and Bai, 2018; Cheraghian, 2016a; Cheraghian et al., nomic reasons. Before using a surfactant in any application, it is
2017) and foams (Guo et al., 2017; Suleymani et al., 2020); there- important to assure its effectiveness and stability over its period
fore, oil recovery is improved. It is noteworthy that NPs provide of use. In this part of the review, the main parameters studied to
other benefits to chemical EOR processes (Cheraghian et al., 2020; characterize surfactants are presented.
3152
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

and the voltammetric method (Nesměrák and Němcová, 2006).


In terms of surfactant EOR applications, IFT measurement is a
commonly used method to determine the CMC (Haq et al., 2020b;
Liu et al., 2020a).

2.2.2. Krafft temperature and Krafft point


An important parameter related to the CMC is the Krafft tem-
perature (TK ), which is also known as the critical micelle temper-
ature (Nazrul Islam et al., 2015; Roy et al., 2014). Generally, the
Krafft temperature is regarded as the melting temperature of a
hydrated solid surfactant (Roy et al., 2014). Below the Krafft tem-
perature, surfactant micelles cannot be formed (Dicharry et al.,
2016; Dölle et al., 2012) and the surfactant tends to be insoluble
Fig. 2. Types of surfactants according to the headgroup charge: (a) non-ionic, (Makavipour et al., 2019; Moon et al., 2001). At the Krafft temper-
(b) cationic, (c) anionic, and (d) zwitterionic (amphoteric) surfactants. ature, however, the solubility of the surfactant increases sharply,
Source: Modified from Soleimani Zohr Shiri et al. (2019).
and further increase in temperature allows for micelle formation
upon reaching the CMC (Malik and Ali, 2016). One of the methods
used to determine the Krafft temperature involves measuring
2.2.1. Critical micelle concentration the electrical conductivity of a surfactant solution of a certain
Micelle is a term refers to an aggregate form of surfactant concentration at different values of temperature. In this case,
molecules dispersed in a liquid colloid (Bhosle et al., 2020; Naseri the Krafft temperature is identified by the temperature point at
et al., 2018), as shown in Fig. 3. The critical micelle concentration which a break is shown by the conductivity-vs-temperature plot
(CMC) is the surfactant concentration above which micelles can (Bales et al., 2002; Song et al., 2009), as depicted in Fig. 4. Another
be formed. At the CMC, several physicochemical properties of the related concept is the Kraft point that refers to the temperature
solution (e.g., thermal and electrical conductivities, viscosity and at which the surfactant solubility equals the CMC (Peris-Vicente
surface tension) vary abruptly (Sabahi et al., 2017). The CMC of a et al., 2016; Zapf et al., 2003). The Krafft point can be determined
particular surfactant depends on factors such surfactant molec- experimentally using solubility measurements. The value for the
ular structure (e.g., hydrophobic chain length) (Glennie et al., Krafft point corresponds to the intersection between the CMC
2006), pressure conditions (Thiruvengadam et al., 2020), solution curve and the solubility curve (Inada et al., 2017; Miyake and
salinity (Bratovcic and Nazdrajic, 2020), ionic composition, pH, Oyama, 2009; Shinoda and Hutchinson, 1962), as shown in Fig. 5.
temperature and others (Harutyunyan and Harutyunyan, 2019).
More than 30 methods have been developed for measuring the
CMC. These include the surface tension method, the dye adsorp- 2.2.3. Cloud point
tion method, the solubilization method and others (Suzuki, 1970). The cloud point is defined as the temperature at which aque-
According to Nesměrák and Němcová (2006), the methods that ous solutions of nonionic surfactants exhibit sudden cloudiness
are used for measuring the CMC are classified into direct and indi- or turbidity (Aktar et al., 2020; Khattab et al., 2020). The solu-
rect methods. In the direct methods, the change of some property bility of a broad range of nonionic surfactants, such as those of
of the surfactant solution is observed in response to the change the polyethoxylated type, depends on forming hydrogen bonds
in the surfactant concentration. Therefore, the CMC is determined between the headgroups of the surfactant and water molecules
based on discontinuity or slope change of the observed solution (Álvarez et al., 2019; Illous et al., 2020). The increase in so-
property such as viscosity, electrical conductivity, refractive index lution temperature, however, weakens these bonds and results
and osmotic pressure. On the other hand, the indirect methods in the dehydration of polyoxyethylene chains existing in the
are used to find the CMC by observing the change of a certain headgroups of the surfactant (Mukhopadhyay et al., 2019; Turan
property of a probe (a substance added to the surfactant solution) et al., 2020). Therefore, above the cloud point, surfactant solutions
in response to the change in the concentration of the surfactant. tend to undergo phase separation (i.e., they separate into two
Examples of such methods include the spectrometric method distinguished phases). The first phase is a surfactant-rich phase

Fig. 3. Micelle formation upon reaching the CMC.

3153
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 2
Relation between HLB values and the expected properties/applications of
surfactants (Sheng, 2010).
HLB value Property/application
0–3 Antifoaming agent
4–6 W/O (water in oil) emulsifier
7–9 Wetting agent
8–18 O/W emulsifier
13–15 Detergent
10–18 Hydrotrope or solubilizer

for low-salinity formations. Analogously, a high-HLB surfactant


should be selected for high-salinity formations (Sheng, 2010).
Researchers developed some equations to calculate the HLB
values of surfactants. Research reported by Griffin (1954, 1949)
and Davies (1957) was among the earliest research to provide
such equations. Royer et al. (2018) stated Griffin’s equation to
Fig. 4. Determination of the Krafft temperature (Tk ) for surfactants from the
calculate the HLB for nonionic ethoxylated surfactants as follows:
conductivity-vs-temperature plot (Bales et al., 2002; Song et al., 2009).
( )
1 MH
HLBGriffin = ∗ 100 (1)
5 MT
where MH denotes the molecular mass of the hydrophilic part
of the surfactant molecule and MT denotes the total molecular
mass of the surfactant molecule (Royer et al., 2018). Davies (1957)
found that HLB values for surfactants can be calculated from their
chemical formulae based on group numbers, as shown in the
following equation:

HLBDavies = 7 + (hydrophilic group numbers)
∑ (2)
− (lipophilic group numbers)
According to Davies (1957), given a surfactant containing a num-
ber n of –CH2 – groups, the HLB value is calculated as follows:


Fig. 5. Determination of the Krafft point from the phase diagram of surfactant HLBDavies = 7 + (hydrophilic group numbers)
solutions (Inada et al., 2017; Miyake and Oyama, 2009). (3)
− n(group number per CH2 group)
where the value of CH2 -group number is substituted as 0.475
that contains a high surfactant concentration. The second phase is and the hydrophilic group numbers are obtained from tables
a bulk aqueous phase that contains a low surfactant concentration provided by Davies (1957). Later, other researchers developed
compared with the first phase (Kachangoon et al., 2020; Pan experimental methods to determine the HLB values. Some of
et al., 2019). According to Puerto et al. (2012), the cloud-point these methods are based on the emulsion inversion point (EIP)
phenomenon of nonionic surfactants, such as AGES (alkoxylatedg- (Marszall, 1978, 1976), the phase-inversion temperature (PIT)
lycidyl ether sulfonate), adversely affects their EOR-performance (Kunieda and Ishikawa, 1985; Kunieda and Shinoda, 1985; Shin-
in reservoirs of high temperature. Solutions of AGES surfactant oda and Sagitani, 1978) and others.
separate into two phases upon reaching the cloud temperature,
which has lower values in the presence of salt (NaCl). Never- 2.2.5. Solubilization parameters
theless, this problem can be alleviated by combining AGES with Oil (or water) solubilization ratio is defined as the volume of
a surfactant such as IOS (internal olefin sulfonate), which is an oil (or water) solubilized per surfactant volume in a microemul-
anionic surfactant. This is because IOS exhibits higher solubility sion phase (Das et al., 2020; Sarmah et al., 2020). According
in aqueous solutions of NaCl as temperature increases. to Abalkhail et al. (2020), the solubilization ratio (SR) can be
expressed as follows:
2.2.4. Hydrophile–lipophile balance Volumeoil
An important criterion for surfactant characterization is the SRoil = (4)
Volumesurfactant
hydrophile–lipophile balance (HLB). This criterion measures the
Volumewater
extent to which a surfactant is lipophilic or hydrophilic (Kondo SRwater = (5)
et al., 2007; Reham et al., 2015). The HLB is a number on a Volumesurfactant
scale from 0 to 20 that indicates relatively the tendency of a Bera et al. (2011) and Hamidi et al. (2015) reported that the
surfactant to dissolve in oil or water. On that scale, a value of 0 optimum solubilization, which produces the ideal microemulsion
correlates with a completely hydrophobic (lipophilic) molecule, formulation needed for oil recovery, occurs when SRwater and
while a value of 20 correlates with a molecule composed entirely SRoil are equal. This is obtained by drawing oil-SR curve and
of hydrophilic components. Table 2 shows surfactant properties water-SR curve, where the intersection point between the two
predicted using the HLB values. To form proper microemulsions curves represents the optimum solubilization occurring at the
during oil recovery, a low-HLB surfactant needs to be selected optimum salinity, as shown in Fig. 6. Liyanage et al. (2015) and
3154
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

2.2.7. Winsor’s R-ratio


Winsor’s R-ratio is a useful tool to understand the phase
behavior of surfactant–oil–water systems, and it is defined as
follows:
ACO
R= (6)
ACW
where ACO and ACW denote the interactions of the surfactant
adsorbed at the interface with the oil molecules and the water
molecules, respectively (Nguyen et al., 2010; Salager et al., 2005;
Winsor, 1954). The value of R is interpreted as shown in Table 3,
where S, O and W indicate surfactant, oil and water phases,
respectively.
The interactions given in Table 3 change upon changing any
of the tuning parameters (e.g., temperature and salinity). For
Fig. 6. Determination of the optimal salinity for an oil–water–surfactant system instance, increasing the salinity will decrease the S–W interaction
using the solubilization-ratio curves (Das et al., 2020; Hamidi et al., 2015). (i.e., ACW decreases). As a result, the R-ratio will increase and
a transition from Type I to Type III microemulsion system may
happen. That is, increasing the salinity of water will increase the
Khaledialidusti et al. (2017) stated that studying the solubilization hydrophobicity of the surfactant system and intensify the S–O
parameters is important to optimize the oil recovery process. This interaction, which in turn raises the R-ratio. This implies that
is because the lowest oil–water IFT is normally observed at the surfactants with higher hydrophilicity require higher salinity to
optimum salinity. form proper microemulsions (Nguyen et al., 2010; Salager et al.,
2005).
2.2.6. Microemulsion systems
At equilibrium, microemulsion systems are classified, accord-
2.2.8. Hydrophilic–lipophilic deviation (HLD)
ing to Winsor (1948), into four main types. These systems are
The HLB concept and Winsor R ratio, presented earlier, have
shown in Fig. 7. Winsor’s Type I systems consist of an O/W (oil-
limitations when applied to calculate the optimum surfactant
in-water) microemulsion that coexists with an excess oil phase.
Conversely, in Type II systems, a W/O microemulsion coexists formulation for EOR and other surfactant-based processes. For
with an excess water phase. In Type III systems, three separate instance, the HLB concept does not consider the effects of vari-
phases coexist including microemulsion, water and oil. Finally, ables such as salinity, temperature, the nature of the oil and the
Type IV systems can be described as homogeneous, single-phase presence of co-surfactants or alcohols in the system (Bouton et al.,
systems formed entirely of one microemulsion phase (Nordiyana 2009; Nguyen et al., 2019; Witthayapanyanon et al., 2008). On the
et al., 2016; Salleh et al., 2019). Type IV microemulsions are other hand, the parameters related to Winsor R ratio are difficult
extension of Type III systems. However, they are formed by to estimate (Budhathoki et al., 2016b; Wang et al., 2019). Because
increasing the concentration of the surfactant until the whole of these limitations, researchers including Salager et al. (1979),
liquid transforms into a single phase (Salleh et al., 2019). For Bourrel et al. (1980) and Antón et al. (1997) proposed correlations
EOR applications, microemulsions of Type III are usually preferred that lead to more accurate prediction of microemulsion phase be-
as they provide the lowest IFT compared to the other systems havior for different types of surfactants. Unlike the HLB concept,
(Nordiyana et al., 2016). these correlations take into account several variables that have a

Fig. 7. Winsor’s classification of microemulsions.


Source: Modified from Winsor (1948).

3155
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 3
Interpretation of R ratio in relation to microemulsion systems (Nguyen et al., 2010; Salager et al., 2005).
R-ratio Interpretation Microemulsion system
R < 1 W–S interaction is stronger than O–S interaction Type I
R > 1 O–S interaction is stronger than W–S interaction Type II
R = 1 W–S and O–S interactions are balanced (optimum formulation) Type III

significant impact on the performance of surfactants used for oil


recovery.
Later, these proposed correlations led to the emergence of
the HLD concept (Salager et al., 2005, 2000, 2003). For ionic and
nonionic surfactants, HLD is numerically expressed according to
Eqs. (7) and (8), respectively:

HLD = ln (S) − kACN + f (A) + σ − aT (T − T0 ) (7)


HLD = α − EON + bS − kACN + f (A) + cT (T − T0 ) (8)
where S is the salinity (as wt.% NaCl in the aqueous phase), ACN
denotes the alkane carbon number for an alkane oil, b, k, aT and cT
are empirical constants that depend upon the type of the system,
f(A) is a function of alcohol type and concentration, σ and α
are characteristic parameters of the surfactant structure, T is the Fig. 8. Definition of the terms used in equation of the packing parameter
temperature (◦ C) and T0 has a value of 25 ◦ C (Antón et al., 1997; (Lombardo et al., 2015; Sheng, 2010).
Bourrel et al., 1980; Salager et al., 2005, 2003, 1979).
Based on Eqs. (7) and (8), the HLD basically measures the
departure from a reference state at which a surfactant–water– 3.1. Interfacial tension reduction
oil formulation represents a Winsor Type III emulsion. At this
reference state (i.e., HLD = 0), the oil–water IFT becomes ex- The interfacial tension is generally defined as the attraction
tremely low. However, when HLD is negative or positive, the force between the molecules existing at the interface of two
formulation represents an oil-in-water (O/W) emulsion or a W/O fluids (Cong et al., 2020; Kalantari Meybodi et al., 2015). The
emulsion, respectively (Salager et al., 2003). Recently, several interfacial tension between hydrocarbons and water molecules
research studies have investigated the use of the HLD concept to causes an increase in the capillary force that plays an essential
screen surfactants for EOR processes. These studies indicated that role in hydrocarbon trapping within porous media. Therefore,
the HLD concept is an effective tool that can be employed in the surfactant injection is used as an option to reduce the interfacial
design and optimization of surfactant–water–oil formulations for tension (Hosseininoosheri et al., 2016). Studies showed that many
specific temperature and salinity conditions (Budhathoki et al., types of surfactants can be used at low concentrations (∼0.05–
2016b; Cholpraves et al., 2017; Jin et al., 2015; Nguyen et al., 0.2%) to achieve low interfacial tension on the order of 10−2
2019; Wang et al., 2019). dynes/cm or less (Ge and Wang, 2015; Hosseinzade Khanamiri
et al., 2016b; Kumar and Mandal, 2017a; Rosen et al., 2005). In the
2.2.9. Molecular packing parameter absence of oil, the presence of a surfactant in water reduces the
The molecular packing parameter (PC ) is a concept that de- surface tension because the molecules of the surfactant replace
scribes the relation between the geometry of a surfactant part of the water molecules at the water surface. The attraction
molecule and its aggregate structure in aqueous solutions. Exam- forces between the water and surfactant molecules are less than
ples of aggregate structures include spherical micelles, rod-like the attraction forces between the water molecules themselves.
micelles, bilayer vesicles, etc. To calculate the Pc, the following Subsequently, the contraction force responsible for the surface
equation is used (Nagarajan, 2002): tension is reduced (Hu et al., 2016; Mohapatra et al., 2014).
vo In water–oil–surfactant systems, on the other hand, surfactant
PC = (9) molecules replace some of the oil and water molecules at the
lo a
original oil–water interface in a process called surfactant adsorp-
where vo is the volume of the surfactant tail, lo is the length of tion. This new arrangement of molecules involves an interaction
the surfactant tail, and a is the surface area of the hydrophilic between the hydrophobic components of the surfactant and oil
headgroup at the surface of the aggregate (Nagarajan, 2002). Fig. 8 on one side of the interface, and between the hydrophilic com-
depicts the terms used in the equation of the packing parameter. ponents of the surfactant and water on the other side. In fact,
the new interaction across the oil–water interface is significantly
According to the value of the packing parameter, the stronger than the original interaction between water and oil
molecules of the surfactant self-assemble into different micelle prior to surfactant addition. Therefore, the interfacial tension is
geometries, as shown in Fig. 9. These geometries affect the bulk reduced (Azodi and Nazar, 2013; Xu et al., 2013).
properties of the solution such as its viscoelastic properties (van The ability of surfactants to reduce the interfacial tension
Zanten, 2011) and its solubilization capacity (Nagarajan, 2002). between crude oil and brine is highly sensitive to the type and
concentration of ions existing in the brine. As stated earlier, there
3. The role of surfactants in oil recovery is an optimal salinity at which the interfacial tension between
crude oil and brine is reduced to ultra-low values. Usually, this
Generally, surfactants affect the processes of oil recovery optimal salinity is expressed in terms of NaCl content that is
through four main mechanisms: interfacial tension reduction, dissolved in the solution (Tichelkamp et al., 2014). However,
wettability alteration, foam generation and emulsification. In this research studies showed that introducing divalent/multivalent
part of the review, a brief discussion of these mechanisms is cations to a surfactant solution containing a given NaCl con-
presented. centration reduces the value of the optimal salinity. This means
3156
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 9. Effect of the critical packing parameter (PC ) on the aggregate structure of surfactants (Cornwell, 2018; Lombardo et al., 2015; van Zanten, 2011).

that the surfactant becomes less tolerant to NaCl salinity in the reduced (Jha et al., 2018; Maneedaeng and Flood, 2017; McGuire
presence of these divalent/multivalent cations, which adversely et al., 2005; Salager et al., 2013).
affects the interfacial tension between oil and brine. In a study
conducted by Bansal and Shah (1978), the effect of divalent 3.2. Wettability alteration
cations on the interfacial tension properties of a surfactant for-
mulation was investigated. The formulation contained petroleum Wettability is the ability (or tendency) of a fluid to adhere to,
sulfonate and ethoxylated sulfonate. It was demonstrated that or spread on, a solid surface in the presence of other immiscible
increasing the concentration of divalent cations (Ca2+ and Mg2+ ) fluids (Jing et al., 2019; Yang and Zhou, 2020). Basically, this
in surfactant–oil–brine systems drastically reduces the optimal type of adhesion occurs due to various forces including van der
salinity. In a similar study, Kumar et al. (1984) showed that Waals forces, structural forces and electrostatic forces, which
increasing the concentration of Ca2+ and Mg2+ in surfactant result in stable distribution of fluids in porous media (Wang
solutions containing petroleum sulfonate and lignosulfonate was et al., 2018). In the presence of two liquids, the one that has the
detrimental to their ability to reduce the oil–brine interfacial ten- strongest adhesion to the surface is referred to as the wetting
sion. Therefore, this effect should be considered during surfactant phase (Khaleel et al., 2019). Sandstones and carbonates have
screening because natural connotate water normally contains different wettability characteristics. Generally, sandstones exhibit
these divalent cations (Kumar et al., 1984). In fact, this phe- water-wet to intermediate-wet properties. On the other hand,
nomenon is mainly observed with anionic surfactants that tend carbonates frequently exhibit intermediate-wet to oil-wet prop-
to precipitate out of the bulk solution upon reacting with the erties (Schön, 2015). Several methods have been used to alter the
divalent/trivalent cations, and thus, oil recovery is expected to be rock wettability toward more water-wet states. These methods
3157
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

include surfactant injection (Nowrouzi et al., 2020), nanofluid surface of the carbonate rock, turning it into an oil-wet surface
injection (Zargar et al., 2020), low salinity waterflooding (Liu and (Standnes and Austad, 2003). Several studies also reported that
Wang, 2020), and thermal recovery processes (Doranehgard and carboxylates present in crude oil are strongly adsorbed on car-
Siavashi, 2018). bonate rocks causing them to be oil-wet (Awolayo et al., 2018; Bai
Rock wettability has a significant impact on oil recovery from et al., 2020; Karimi et al., 2015). Therefore, in order to effectively
naturally fractured carbonate reservoirs. Commonly, waterflood- change the wettability of carbonates using surfactant flooding,
ing is conducted to recover additional oil after the primary re- carboxylates usually need to be targeted in the process. In a
study carried out by Standnes and Austad (2000), fourteen types
covery period. The effectiveness of waterflooding in fractured
of anionic and cationic surfactants were examined in terms of
reservoirs depends on the spontaneous imbibition of water to
their ability to alter the wettability of chalk, which is a carbonate
displace the oil contained in the matrix to the system of fractures
rock. The results showed that cationic surfactants have signifi-
(Abushaikha and Gosselin, 2008). The process of spontaneous cantly larger potential to change the wettability of the rock. This
imbibition is mostly effective in water-wet reservoirs in which change is attributed to the mechanism of ion-pair formation. In
the capillary driving forces are strong. However, in fractured this mechanism, cationic surfactants have the ability to desorb
reservoirs of mixed- or oil-wet properties, the capillary driving the negatively charged surface-active materials (i.e., carboxylates)
forces for the spontaneous imbibition are weak, and hence the oil from the carbonate surface (Standnes and Austad, 2000). As a
recovery from waterflooding is low (Andersen et al., 2014; Salehi result, the surface returns to its original water-wet state, making
et al., 2008). spontaneous imbibition possible.
In many cases, capillarity can retain oil, and it may not be pos-
sible to apply an adequate pressure gradient across the fractured 3.3. Foam generation
oil-wet carbonate formations. Consequently, high oil saturation
remains in those formations. Therefore, oil recovery from car- Foam is a dispersion consists of two fluids: a gas and an
bonates (e.g., chalk and dolomite) relies largely on wettability aqueous surfactant solution. The gas is trapped in the dispersion
in the form of bubbles separated by thin films of liquid known
alteration, which can be achieved by injecting proper surfactants
as lamellae (Farajzadeh et al., 2020; Samimi et al., 2020). Several
(Leslie Zhang et al., 2006).
gases can be used in the process of foam generation including air,
Wettability alteration in carbonates by surfactant injection is
nitrogen, carbon dioxide, natural gas and others (Zhao and Torabi,
explained according to two main mechanisms. The first mecha- 2020). Typically, gas injection in subsurface porous media is chal-
nism is called the coating mechanism, and it is related to anionic lenged by issues such as formation heterogeneities and the low
surfactants. In this mechanism, the molecules of anionic surfac- gas density/viscosity compared to in-situ fluids. Therefore, foam
tants create a monolayer on the carbonate rock surface where has been introduced to EOR applications as it provides several
the hydrophobic surfactant tails interact with the adsorbed oil. advantages. First, foam can alleviate the viscosity issue and re-
This arrangement implies that the originally oil-wet rock surface duce the gas mobility. Second, foam has the ability to block some
is covered with hydrophilic head groups of the surfactant. These layers of the reservoir, especially those with high permeability,
groups modify the rock wettability to more water-wet conditions. in favor of injecting fluids into less permeable layers. Third,
The second mechanism is called the cleaning mechanism, and it foam can increase the lateral pressure gradient which relatively
is related to cationic surfactants. This mechanism is based on reduces the gravity effect and hence reduces the phenomenon
forming ion pairs between the surfactant cationic heads and the of gravity segregation. As a result, these advantages increase the
acidic portions of crude oil adsorbed onto the carbonate rock volume fraction of the reservoir that is swept by the injected
fluids (Farajzadeh et al., 2020). According to Pal et al. (2018), foam
surface. The formed ion pairs can strip the adsorbed oil layer off
injection is used as a method of mobility control in circumstances
the rock surface, which exposes the originally water-wet surface.
where polymer flooding, gas injection or water–alternating-gas
Generally, the cationic surfactants exhibit a better performance
(WAG) injection are not feasible. This mostly occurs in formations
of wettability alteration in carbonates compared with the anionic with unfavorable characteristics such as high heterogeneity or
surfactants. This is because the ion-pair interactions are stronger high salinity and temperature affecting the polymer stability.
than the hydrophobic interactions (Hammond and Unsal, 2011;
Salehi et al., 2008). In sandstones with oil-wet properties, wetta- 3.4. Emulsification
bility alteration toward water-wet conditions also occurs due to
ion-pair interactions or hydrophobic interactions depending on Emulsification is one of the mechanisms that can increase
the type of the injected surfactant (Hou et al., 2016). oil recovery in the tertiary phase. Emulsions mostly form dur-
As stated earlier, carbonates tend to exhibit more oil-wet ing alkaline and/or surfactant flooding under the conditions of
characteristics compared to sandstones. Because of that, a large low/ultra-low IFT and under shear resulting from fluid flow in
amount of research has been devoted to study the interactions rock pores (Bryan and Kantzas, 2009; Li et al., 2020d). Emulsi-
between surfactants, brine, oil and carbonate surfaces in relation fication usually improves oil recovery through two main mecha-
to rock wettability. According to Standnes and Austad (2003), nisms: (1) emulsification and entrainment; and (2) emulsification
and entrapment (Mehranfar and Ghazanfari, 2014; Shi et al.,
the brine in carbonate reservoirs usually has basic nature, with
2015; Türksoy and Bağci, 2000). In the case of emulsification
a pH approximately equal to 8. At this pH, the carboxylic groups
and entrainment, the IFT is significantly reduced allowing the
existing in the crude oil become negatively charged in contact
droplets of crude oil to be emulsified into the water phase.
with the formation brine, and they strongly adsorb onto the Then, the droplets are transported along with the water phase.
positively charged carbonate surface. Normally, carboxylic ma- This means that oil bulks are gradually produced in the form
terials are found in the heavy-end fractions of crude oil, such of fine particles. Usually, the emulsification-entrainment mech-
as asphaltene and resin. These carboxylic materials perform as anism occurs when the formed emulsion droplets have sizes that
‘‘anchors’’ for the polar molecules to adsorb onto the carbonate are equal to or smaller than the pore sizes. On the other hand,
surface, which occurs mainly due hydrophobic and dipole–dipole emulsification-entrapment mechanism involves the formation of
interactions. As a result, an organic layer will cover the entire emulsion droplets that plug pore throats and water channels. As
3158
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 4
List of the surfactants that appear in the review.
Surfactant name Abbreviation Type
Dodecyltrimethylammonium chloride DTAC Cationic
Tetradecyltrimethylammonium chloride TTAC Cationic
Octadecyltrimethylammonium chloride OTAC Cationic
Hexadecyl-N,N-dimethyl-2-ammonio-1-ethanecarbonate C16DmCB Zwitterionic
Triton X-100 TX-100 Nonionic
Cetyltrimethylammonium bromide CTAB Cationic
Sodium dodecyl sulfates SDS Anionic
Alkyl polyglycoside APG Nonionic
Internal olefin sulfonates IOS Anionic
Tridecyl alcohol propoxy sulfate TDA Anionic
Guerbet alkoxylate carboxylate GAC Anionic
Cocamidopropyl betaine CAPB Zwitterionic
Sodium laurilsulfate or sodium lauryl sulfate SLS Anionic
1-dodecyl-3-methylimidazolium chloride (surface-active ionic liquid) M12 Cationic
Alpha olefin sulfonates AOS Anionic
Alkyl aryl sulfonates AAS Anionic
Alkyl aryl disulfonate ADS Anionic
Naphthenic aryl sulfonate NAS Anionic
Alkyl aryl monosulfonate AMS Anionic
Dodecyltrimethylammonium bromide DTAB Cationic
Tergitol 15-s-7 Tg7 Nonionic
Tergitol 15-s-9 Tg9 Nonionic
Tergitol 15-S-12 Tg12 Nonionic
N, N-Dimethyl-N-[2-hydroxy-3-sulfopropyl]-N′ -phenyloctadecanoyl-1,3-diaminopropane SPODP Zwitterionic
Oleyl polyoxyethylene amidopropyl carboxybetaine OPAC Zwitterionic
Oleyl polyoxyethylene amidopropyl hydroxy sulfobetaine OPAH Zwitterionic
Oleyl polyoxyethylene amidopropyl sulfobetaine OPAS Zwitterionic
N-phenyloctade-canoicamidopropyl-N, N-dimethylcarboxylbetaine POAPMB Zwitterionic
Alcohol propoxy sulfate APS Anionic
Alkyl benzene sulfonate ABS Anionic
Cocamidopropyl hydroxysultaine CAHS Zwitterionic
Alkyl ether sulfate AES Anionic
Alkyldimethylamine oxides ADAO Zwitterionic
Lauryl alcohol polyoxyethylene ether LAPE Nonionic
Alkyl dimethyl ammonium chloride ADAC Cationic
Alkyl trimethyl ammonium chloride ATAC Cationic

a result, the water-to-oil mobility ratio is reduced and surfactant process (Belhaj et al., 2020). This in turn may result in significant
solution is diverted to upswept areas, which improves the areal economic drawbacks (Liu et al., 2020c; Ojo et al., 2020; Venancio
and vertical sweep efficiencies. This effect is important, particu- et al., 2020).
larly in the production of viscous oils where the sweep efficiency In reservoirs with intermediate to oil-wet nature, on the other
of waterflooding is poor. It is noteworthy that the entrapment hand, surfactant adsorption can play a beneficial role in wet-
in this mechanism occurs because the IFT is not low enough to tability alteration. This occurs when surfactant molecules are
reduce the average size of the emulsion droplets below that of adsorbed at the sites where crude oil was desorbed by the flood-
the pores in the reservoir rock (Bryan and Kantzas, 2009; Castor ing process. Subsequently, further interactions between crude oil
et al., 1981; Johnson, 1976). and the surface of the rock are prevented. Therefore, the rock
wettability is altered to water wet. However, excessive adsorption
4. Surfactant adsorption onto reservoir rock of surfactants reduces their concentration in the aqueous phase
and adversely impacts the efficiency of surfactant-EOR processes
In Section 3, we discussed the role of surfactants in the process (Saxena et al., 2019). For instance, the use of cationic surfactants
of oil recovery. In this section, we will discuss the adsorption in carbonates has a positive effect on the wettability state. How-
of surfactants during EOR processes. In addition, we will present ever, cationic surfactants exhibit strong adsorption in sandstones
the factors that affects the adsorption process based on recently and hence they cannot be used in sandstone reservoirs (Sheng,
published research that considered different types of surfactants. 2010).
Table 4 provides a list of abbreviations of surfactants as they
appear in the following parts of the review. 4.2. Surfactant adsorption isotherms

4.1. Influence of surfactant adsorption on EOR Adsorption of ionic surfactants on oppositely charged reservoir
rock can be described using surfactant adsorption isotherms, as
During the flooding process, surfactant retention in porous shown in Fig. 10. Typically, an adsorption isotherm is divided
media occurs according to three main mechanisms: precipita- into four regions. In region I, adsorption occurs over low concen-
tion, adsorption and phase trapping. Surfactant precipitation and trations of the surfactant as a result of electrostatic interactions
phase trapping can be prevented by selecting a surfactant that between the surfactant monomers and the solid surface (Bud-
tolerates temperature and salinity. However, the adsorption of hathoki et al., 2016a; Zhang and Somasundaran, 2006). The ad-
a surfactant onto reservoir rock can be only minimized. The sorption in this region increases linearly with the increase in the
adsorption process decreases the concentration of the surfactant concentration of the surfactant, following Henry’s law (Grządka
in the injected solution, which affects its ability to reduce the and Matusiak, 2020). In region II, surface aggregates of the sur-
oil–water IFT. Consequently, the effectiveness of surfactant-based factant start to form by reason of lateral interactions between
EOR to mobilize the trapped oil is influenced by this adsorption the hydrocarbon chains. Examples of surface aggregates include
3159
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 10. Surfactant adsorption isotherm (Budhathoki et al., 2016a; Zhang and Somasundaran, 2006).

surface colloids (solloids), hemimicelles (monolayer), admicelles 5.1. Anionic surfactants


(bilayers), etc. In addition to the lateral interactions, the adsorp-
tion process is still influenced by active electrostatic interactions. 5.1.1. Sodium dodecyl sulfate
Hence, a sharp increase in the slope of the isotherm occurs. In Sodium dodecyl sulfate (SDS) belongs to a group of sur-
region III, the solid surface becomes electrically neutralized by factants called alkyl sulfates, and it has the chemical formula
the adsorbed ions of the surfactant. Therefore, the electrostatic in- CH3 (CH2 )11 SO4 Na (Li et al., 2020a). SDS is also known as sodium
teractions become no longer active, which decreases the slope of laurilsulfate or sodium lauryl sulfate (SLS) (Bondi et al., 2015;
Wołowicz and Staszak, 2020). This surfactant has a linear molec-
the isotherm. At this stage, surfactant adsorption occurs only due
ular structure that consists of a 12-carbon hydrophobic chain
to the lateral interactions. Finally, the surfactant concentration
attached to a negatively charged sulfate group (Hou et al., 2020;
reaches the CMC as indicated by region IV. This means that any
Khan et al., 2019).
increase in the concentration contributes only to the formation of
Recent research investigated the use of SDS incorporated with
micelles in the solution without increasing the adsorption density nanoparticles (NPs) in EOR applications. For instance, Wu et al.
(Budhathoki et al., 2016a; Zhang and Somasundaran, 2006). (2017) studied SDS adsorption and its EOR performance in the
presence hydrophilic silica NPs. Static adsorption tests of SDS on
quartz sand showed that the use of 0.2 to 0.3 wt% silica NPs
4.3. Factors affecting surfactant adsorption
is optimal to effectively reduce surfactant adsorption. Core-flood
experiments showed that 0.2 wt% SDS could achieve incremental
There are several factors that affect surfactant adsorption onto oil recovery of 4.43% following waterflooding. This percentage
reservoir rock. Examples of these factors include surfactant type, was raised to 9.11% using 0.2 wt% SDS and 0.2 wt% silica NPs.
rock surface charge, salinity, pH and temperature. Therefore, ex- Rezk and Allam (2019) reported that the flooding of 0.2 wt%
tensive study is needed prior to selecting a surfactant for a SDS achieved about 16% increase in oil recovery in the tertiary
particular oil reservoir. Table 5 summarizes several research stud- phase. However, the use of 0.2 wt% SDS with 0.05 wt% ZnO NPs
ies highlighting the effect of salinity, temperature and pH on raised the value to 35%. This improvement was attributed to the
surfactant adsorption. The summarized studies considered the wettability alteration and significant reduction in the IFT between
type of the surfactant, such as cationic and anionic surfactants, in water and oil (n-dodecane) caused by ZnO NPs.
addition to the type of the mineral surface including carbonate, Researchers also studied the effect of SDS combined with
sandstone, shale, clay, Kaolinite, etc. other surface-active compounds on solution properties and EOR.
For example, Jia et al. (2019) investigated the synergistic effects
of SDS and 1-dodecyl-3-methylimidazolium chloride (M12) on
5. Application of surfactants in chemical EOR properties such as IFT, rock wettability and oil-recovery perfor-
mance. It was demonstrated that mixed-surfactant systems have
higher capability to reduce the IFT compared to SDS or M12
There is a broad range of surfactants that have been em-
systems at conditions of high salinity and temperature. It was
ployed in EOR applications. The use of such a wide range of
also demonstrated that all surfactant systems had the ability to
surfactants is attributed to technical, economic and environmen- alter rock wettability. However, mixed systems exhibited higher
tal considerations. This section of the review presents surfactants tendency to alter the wettability due to the pseudo two-tailed
frequently used in EOR. The surfactants are presented according co-surfactant formed upon mixing anionic (SDS) and cationic
to their type: anionic, cationic, nonionic and zwitterionic sur- (M12) surfactants. Lastly, Jia et al. (2019) demonstrated the ability
factants. In addition, this section presents field-scale application of SDS–M12 systems to improve oil recovery, at high temper-
of surfactant-based EOR. Finally, it discusses the role of NPs in ature and salinity, by more than 10%. However, the effect of
surfactant flooding, which is an emerging research topic. single-surfactant systems on oil recovery was not studied.
3160
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 5
Effect of salinity (S), temperature (T) and pH on surfactant adsorption.
Reference Surfactant name and Type of adsorbent Experimental Main findings
type or solid surface conditions
Effect of salinity
Chen et al. (2014) DTAC (cationic) Carbonate S = 13987.7, - Surfactant adsorption on limestone
TTAC (cationic) 27975.4, 55950.7, surfaces increased as the salinity
OTAC (cationic) 111901.4 and increased.
223802.8 ppm
Kumar and Mandal (2019) (C16DmCB) Carbonate S = 0, 15000, - At the CMC, higher adsorption
(zwitterionic) Sandstone 30000 and 45000 occurred due to salt addition which
ppm (NaCl) reduces the surfactant solubility and
T = 30 ◦ C reduces the repulsion between the
adsorbed surfactant molecules.
- Surfactant adsorption was higher in
carbonate rock compared to sandstone
rock.
Saxena et al. (2019) Soap-nut surfactant Sandstone S = 0, 10000, - UV spectroscopy can be used to study
(anionic) Carbonate 20000 and 30000 surfactant adsorption in the presence of
Bentonite ppm (NaCl) salts.
- Surfactant adsorption increased as the
salinity increased.
- Surfactant adsorption was highest in
bentonite clay and lowest in sandstone.
- Rock mineralogy and morphology
affected the adsorption process.
Yekeen et al. (2019) TX-100 (nonionic) Shale S = 0, 10000 and - The adsorption of CTAB was 2.57, 2.76
CTAB (cationic) 30000 ppm (NaCl) and 2.90 mg/g at 0, 10000 and 30000
ppm, respectively.
- The adsorption of TX-100 was 1.51,
1.82 and 1.99 mg/g at 0, 10000 and
30000 ppm, respectively.
- The high adsorption of the cationic
surfactant CTAB was related to the high
quartz content in the shale.
Zhong et al. (2019) BW (zwitterionic) Bakken rocks and S = 500 and - CS-50 exhibited a slight increase in the
CA (zwitterionic) Berea sandstone in 289820 ppm adsorption with the increase in salinity.
CS-50 (zwitterionic) addition to other T = 20 ◦ C BW and CA exhibited a reduction in the
minerals (calcite adsorption with the increase in salinity.
and clay) - Unlike CS-50, –COO− functional groups
in BW and CA have the potential to be
protonated in acidic solutions. Hence,
the cationic surfactant parts partially
overweigh the anionic parts, which
increases adsorption onto negatively
charged mineral surfaces.
Abbas et al. (2020) Aerosol-OT (anionic) Quartz sand S = 0, 20000, - In the presence of minerals, the value
Illite 35000 and 60000 of CMC increased as the salinity
Kaolinite ppm increased.
Montmorillonite T = 25 ◦ C - The minimum and maximum rates of
adsorption were observed in quartz sand
and kaolinite, respectively.
Effect of temperature
Azam et al. (2014) Synthesized Berea sandstone T = 25 and 70 ◦ C - The titration method was used to
surfactant consists determine the surfactant adsorption
mainly of C18 carbon based on the surfactant concentration.
chain and a sulfonate - Surfactant adsorption at 25 ◦ C was
headgroup (anionic) 0.9604 mg/g, and it decreased to 0.7491
mg/g at 70 ◦ C.
- Onset of higher translational kinetic
energy led to the reduction in surfactant
adsorption.

(continued on next page)

5.1.2. Internal olefin sulfonates These surfactants are twin-tailed, high resistant to temperature,
Internal olefin sulfonates (IOS) are a group of anionic surfac- of high solubilization activity and of low cost.
tants that are widely used in EOR due to their remarkable ability Rudyk et al. (2019) studied the effect of salinity on the foama-
to withstand harsh reservoir conditions of temperature and salin- bility of 0.5% IOS (C14 –C17 ). At a temperature of 22 ◦ C, the volume
ity (i.e., monovalent ions). However, IOS surfactants (i.e., C15−18 of foam generated in test tubes increased as the salinity increased
IOS, C19−13 IOS, and C24−26 IOS) tend to precipitate when used from 1% to 11% NaCl, with surfactant precipitation starting at
alone in the presence of divalent ions such as Ca2+ (Rattanaudom 6% NaCl. At 60 ◦ C, however, foam generation was observed at a
et al., 2019). As reported by (Nourafkan et al., 2018), IOS surfac- salinity of up to 7% NaCl and the onset of surfactant precipitation
tants have several advantageous properties when used in EOR. occurred at 8% NaCl salinity. To improve the effectiveness of
3161
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 5 (continued).
Reference Surfactant name and Type of adsorbent Experimental Main findings
type or solid surface conditions
Effect of salinity
Li et al. (2016) Nonylphenyl Sandstone S = 5002 ppm - At 76.5 ◦ C, the adsorption of Sa and
ethoxylate T = 76.5 ◦ C Sc was ∼0 and 0.55 mg/g.
carboxylate surfactant - A mixture of both surfactants (Sa/c )
(Sa ) (anionic) exhibits adsorption values between
Quaternary those of Sa and Sc .
ammonium salt (Sc ) - Sa is an anionic surfactant. Therefore,
(cationic) it exhibits low adsorption on the
Surfactant mixture negatively charged sandstone rock even
(Sa/c ) at high temperature.
Yekeen et al. (2017a) SDS (anionic) Kaolinite S = 0 ppm, 1000 - Surfactant adsorption was determined
ppm (CaCl2 ) and by measuring the surface tension.
10000 ppm (NaCl) - In the presence and absence of
T = 25.85, 49.85 electrolytes, surfactant adsorption
and 79.85 ◦ C decreased as the temperature increased.
Ahmadi et al. (2018) Saponin (nonionic) Shale–sandstone T = 28, 40, 60 - The conductivity method was used to
surface and 75 ◦ C determine the surfactant adsorption
based on the surfactant concentration.
- At various surfactant concentrations,
surfactant adsorption decreased as the
temperature increased.
Wei et al. (2020) APG (nonionic) Sandstone T = 20, 50, 80 - The adsorption of 0.6 wt% APG at
and 115 ◦ C 20 ◦ C was 24.2 mg/g, and it decreased
to 21.6 mg/g at 115 ◦ C.
- The change in adsorption was slight
due to the stable hydrophilicity of the
surfactant despite temperature change.
Abbas et al. (2020) Aerosol-OT (anionic) Quartz sand S = 35000 ppm − Surfactant adsorption decreased as
Illite T = 25, 45 and the temperature increased. This was
Kaolinite 85 ◦ C attributed to the exothermic nature of
Montmorillonite the adsorption process.
Effect of pH
Li and Ishiguro (2016) SDS (anionic) Porous SiO2 pH = 3 to 7 - SDS adsorption on SiO2 surface
powder gel decreased as the pH increased because
of the increased electrostatic repulsion.
- Further increase in electrostatic
repulsion prevented the detection of
SDS adsorption.
Southwick et al. (2016) IOS 2024 (anionic) Sandstone pH = 8.24 to 9.57 − At a concentration of 3000 ppm of
IOS 2024, increasing the pH from 8.24
to 9.57 decreased surfactant adsorption
from 0.760 to 0.161 meq/100 gram of
rock.
Tagavifar et al. (2018) IOS (anionic) Indiana limestone T = 24 and 78 ◦ C - The adsorption of two surfactant
TDA (anionic) pH = ∼7.6 to ∼12 mixtures was examined with the pH
GAC (anionic) adjusted using Na2 CO3 . The adsorption
decreased almost linearly as the pH
exceeded a value of 9.
- Experimental data did not clearly
exhibit the effect of temperature on
adsorption affinity unlike surface
complexation modeling.
Saxena et al. (2019) Soap-nut surfactant Sandstone Alkali addition = - Surfactant adsorption decreased as the
(anionic) Carbonate 0, 1, 2 and 3% Na2 CO3 concentration increased.
Bentonite (Na2 CO3 ) - Na2 CO3 dissociates in water to form
H2 CO3 and OH− ions. This increases the
pH and the negativity of the surface
leading to more electrostatic repulsion
of anionic surfactants.

(continued on next page)

EOR in liquid-rich shales, Teklu et al. (2018) investigated the 5.1.3. Alpha olefin sulfonates
effect of IOS and salinity on oil recovery using the process of Alpha olefin sulfonates (AOS) are a group of anionic surfactants
spontaneous imbibition in Bakken cores. It was found that low-
that exhibit high biodegradability and low toxicity. These sur-
salinity brine imbibition (20,000 ppm KCl) expels larger amount
factants are also distinguished by excellent wetting, dispersing,
of oil compared to high-salinity brine imbibition (240,000 ppm
KCl). Moreover, the addition of 1000 ppm IOS to the low-salinity foaming, emulsifying, and stabilizing abilities (Harutyunyan and
brine resulted in further increase in the amount of expelled oil. Harutyunyan, 2019). A general chemical formula of an AOS
3162
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 5 (continued).
Reference Surfactant name and Type of adsorbent Experimental Main findings
type or solid surface conditions
Effect of salinity
Liu et al. (2019) Alcohol alkoxy sulfate Clay surface T = 23 ± 0.1 ◦ C - The concentration of Na+ and Ca2+ in
(anionic) pH = 6 to 11 the surfactant solution was raised by
the addition of NaCl and CaCl2 .
- In the presence of Na+ ions,
adsorption of alcohol alkoxy sulfate
decreased as pH increased. However, the
opposite was observed in the presence
of Ca2+ ions.
Rezaei et al. (2020b) CAPB (zwitterionic) Dolomite pH = 7.3 to 11.2 − The addition of 0.1 wt% Na2 CO3
increased the solution pH from 7.3 to
11.2. It also reduced CAPB adsorption
from 0.795 to 0.320 mg/g-rock.

surfactant can be expressed as R–CH==CH–(CH2 )n –SO3 Na, where based on reacting propylene tetramer with benzene to form a
R is C10 –C20 (Negin et al., 2017). AOS surfactants are typically pre- group of alkyl aryl. Then, sulfonation of alkyl aryl was used
pared by reacting sulfur trioxide (SO3 ) with linear alpha-olefins. to produce AAS (Showell, 1997). Recently, AAS surfactants have
When the length of the hydrophobic chain of AOS surfactants been frequently used in EOR processes (Machale et al., 2019a;
ranges between C12 –C18 , they show some excellent character- Tackie-Otoo et al., 2020). Generally, these anionic surfactants
istics. These include stability over a broad range of pH, high are preferred to be used in sandstones and clays due to their
resistance to water hardness and other metal ions, solubility in low adsorption onto their surfaces. They are also stable at high
cold water and excellent foaming (Abed et al., 2004). temperatures and can be customized according to reservoir con-
In many cases, AOS surfactants were studied for EOR applica- ditions (Machale et al., 2019b). Furthermore, these surfactants are
tions in combination with other materials such gases (e.g., CO2 suitable to be used with polymers, such as polyacrylamide, during
and N2 ) and nanoparticles. For instance, Farajzadeh et al. (2010) surfactant–polymer (SP) flooding (Han et al., 2019; Yan and Ding,
investigated the performance of an AOS surfactant (with C12 2019).
hydrophobic chain) by conducting core-flood experiments in Luan et al. (2019) investigated the behavior of AAS surfac-
quartz-rich Bentheimer sandstone. In their experiments, CO2 was tants in terms of surface tension, micellization and adsorption
injected in the presence and absence of AOS under miscibility properties to screen them for EOR. The selected surfactants in-
conditions. They found that AOS injection prior to CO2 injection cluded alkyl aryl disulfonate (ADS), raw naphthenic aryl sulfonate
could improve the oil recovery by 40% of OOIP compared to 28% (NAS) and alkyl aryl monosulfonate (AMS). It was shown that
achieved by the injection of CO2 only. On a reservoir scale, they the surface tension was significantly reduced by each of the
indicated that surfactant flooding forms foam in regions of low surfactants in their aqueous solution. The lowest surface tension
oil saturation rather than in those of high oil saturation. This, in was attained by AMS at a value of 29.2 mN/m. The CMC varied
turn, reduces CO2 mobility in regions of low oil saturation and slightly between the surfactants and it was highest for AMS as
divert CO2 to regions of high oil saturation. it had the shortest hydrophobic chain. The highest adsorption
Chen et al. (2019a) studied the effect of clay NPs on foaming capacity of the surfactants at air–water interface was for AMS at
properties of AOS (C14 –C16 ). They found that clay NPs increased all surfactant concentrations. Luan et al. (2019) also determined
the bulk foam stability according to the concentration and hy- the micelle aggregation number (N), which indicates the num-
drophobicity of the NPs. It was also found that the addition of ber of the molecules aggregated in a formed micelle at a given
clay NPs to AOS foam enhanced its viscoelasticity modulus at concentration. They found that N increased as the surfactant
bulk and liquid film scales. Therefore, incorporating clay NPs with concentration increased, and it was highest for AMS. Core-flood
AOS foam can be useful in chemical EOR applications. Rezaei experiments demonstrated that the micelle size of AMS increased
et al. (2020a) investigated the performance of AOS with and from 100 nm to 1000 nm as crude oil exists in the produced
without hydrophilic silica NPs. It was shown that AOS causes a solution.
significant reduction in oil–water IFT. Further reduction in the Suleimanov et al. (2011) studied the performance of sulfanole
IFT was achieved with the addition of silica NPs. The optimum (alkyl aryl sodium sulfonate) in the presence and absence of
concentration of the NPs was determined to be 0.1 wt%. It was nonferrous metal nanoparticles (NPs) of sizes ranging from 70 nm
also shown that an aqueous solution of AOS and silica NPs had to 150 nm. The NPs were used at a concentration of 0.001 wt%
the ability to alter the wettability and increase the oil recovery in surfactant solution. It was found that the IFT was reduced
by 2.5–8.6% of OOIP in carbonate rocks. Despite that, Rezaei et al. as the surfactant concentration increased. The presence of the
(2020a) did not conduct wettability and EOR experiment using an NPs in surfactant solution caused further reduction in the IFT.
AOS solution in the absence of silica NPs. EOR experiments showed that the surfactant solution was capable
of increasing oil recovery by 1.7–2.1% of OOIP, which was then
5.1.4. Alkyl aryl sulfonates increased by 6.9–8.5% of OOIP using the NPs.
Alkyl aryl sulfonates (AAS) are anionic surfactants, which are
the most common type of the sulfonate surfactants (Cn H(2n+1) 5.2. Cationic surfactants
SO− +
3 X ). In sulfonate surfactants, the sulfur atom is connected
directly to the carbon atom in the alkyl group. Therefore, the The chemical structure of surfactants has a significant effect
molecules of sulfonate surfactants tend to have high stability on their ability to alter the wettability of carbonates such as
against the hydrolysis reaction (Tadros, 2014). In 1930s, AAS limestone rock (Hosseini et al., 2020; Pan et al., 2020). In general,
surfactants were introduced to the market as detergents. By 1945, cationic surfactants have significantly larger potential to change
these surfactants became the main type of synthetic detergents the wettability of carbonates compared with anionic surfactants.
used for laundry applications. The early preparation of AAS was The positively charged carbonate surfaces repulse the polar parts
3163
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

of the cationic surfactants. Accordingly, the nonpolar parts will formation brine in the presence of 0.5 wt% DTAB achieved higher
be oriented toward the rock surface. This in turn leaves the polar levels of oil recovery. According to Karimi et al. (2016), three
parts in contact with water at the rock–fluid interface, which factors could influence limestone wettability to different extents.
results in a more water-wet conditions (Castro Dantas et al., These include brine dilution, the use of the cationic surfactant
2014). In the following parts, we present several types of cationic DTAB and the presence of sulfate and magnesium ions.
surfactants that are mostly used in carbonates.
5.2.3. Alkyltrimethylammonium chlorides
5.2.1. Cetyltrimethylammonium bromide (CTAB) This group of surfactants can be expressed as Cn H(2n+1) N+
CTAB is a cationic surfactant that is also known as cetrimo- (CH3 )3 Cl− . In a study conducted by Chen et al. (2014), the ad-
nium bromide (Shahbazi et al., 2020) or hexadecyltrimethylam- sorption performance of three cationic surfactant onto limestone
monium bromide (Anas et al., 2020). It has the molecular formula powder was investigated. The studied surfactants include dode-
of C19 H42 BrN (Qian et al., 2020). CTAB can be used effectively cyltrimethylammonium chloride (DTAC), tetradecyltrimethylam-
to improve oil recovery in carbonate reservoirs (Daghlian Sofla monium chloride (TTAC) and octadecyltrimethylammonium chlo-
et al., 2016; Souraki et al., 2019). Mechanisms responsible for this ride (OTAC). A broad range of salinities reaching up to 223802.8
improvement of oil recovery include the ability of CTAB to (1) ppm was used during the adsorption measurement. Results
reduce the oil–water IFT (Kumari et al., 2019); (2) generate foam showed that surfactant adsorption increased due to the increase
in porous media (Telmadarreie and Trivedi, 2016); and (3) modify in salinity. Adsorption of DTAC was the lowest and it was rela-
rock wettability toward more water-wet conditions (Derikvand tively close to that of TTAC. However, adsorption of OTAC was
et al., 2020). substantially higher than that of the two surfactants. Chen et al.
Nandwani et al. (2017) studied the performance of CTAB under (2014) attributed the high adsorption of OTAC to its stronger
high salinity conditions. Their experiments showed that the sur- hydrophobic force resulted from its long hydrophobic chain.
face tension of aqueous solutions decreased due to the addition Therefore, it is easier for OTAC to leave the water phase and
of CTAB. However, the effect of CTAB was higher in brine (3 wt% adsorb onto limestone.
NaCl) compared to distilled water. In addition, the IFT between
brine (3 wt% NaCl) and oil decreased as the CTAB concentra- 5.3. Nonionic surfactants
tion increased. Core-flood experiments demonstrated that the
addition of CTAB resulted in an EOR of 32.1–54.8% of OOIP. Nonionic surfactants show hydrophilic behavior due to the
Kumari et al. (2019) investigated the effect of CTAB and SDS presence of polymerized glycol ether or glucose units in their
on the IFT and oil recovery. A brine solution that had a salinity structure (Nollet, 2005). These surfactants typically contain a
of 3 wt% was used in the experiments. They reported that the polar head, such as glucose in alkyl polyglucosides, and a linear or
oil–water IFT was substantially reduced with the addition of the branched alkyl tail. To synthesize nonionic surfactants, ethylene
surfactants. The impact of CTAB on the IFT was larger compared oxide or propylene oxide are usually added to compounds such
to SDS. However, a mixture of CTAB and SDS caused the largest as fatty acids, alkylphenols, fatty acid amides, fatty amines or
reduction in the IFT. Kumari et al. (2019) also showed that oil fatty alcohols (Castro et al., 2014). Nonionic surfactants are em-
recovery could be increased using CTAB, SDS and their mixtures ployed in a wide range of applications such pesticides, medicine,
by 18.3%, 14.6% and 24.0–39.2% of OOIP, respectively. food industry (Xia et al., 2020), detergents (Cheng et al., 2020)
and EOR processes (Bustamante-Rendón et al., 2020; Moham-
5.2.2. Dodecyltrimethylammonium bromide (DTAB) madshahi et al., 2020). There are different types of nonionic
DTAB, also denoted as C12 TAB, is a cationic surfactant that has surfactants; however, they are commonly classified as shown in
the molecular formula of C15 H34 BrN (Hosseini et al., 2020). In a Fig. 11.
study conducted by Sharma and Mohanty (2013), spontaneous Triton X-100 (TX-100) is a nonionic surfactant that belongs
imbibition tests were used to investigate the ability of DTAB to the alkylphenol ethoxylates (APEOs) (Farsang et al., 2019). On
to improve oil recovery from limestone at a high temperature the other hand, Tergitol 1-s-7 (Tg7) and Tergitol 15-s-9 (Tg9)
(92 ◦ C). Using a formation brine of high salinity, oil recovery belong to a class of nonionic surfactants called secondary alco-
reached up to 14.6% of OOIP. After brine imbibition, a solution hol ethoxylates (Yıldız and Demir, 2019). These surfactants have
of 1 wt% DTAB could achieve oil recovery of 72% of OOIP. Such been widely used in EOR applications. In a study conducted by
increase in oil recovery indicated that DTAB possesses the capa- Sarmah et al. (2020), the performance of Tg7, Tg9 and TX-100
bility to improve the wettability of carbonates, as demonstrated was investigated in terms of surface tension, IFT and oil recovery.
by spontaneous imbibition tests. The three surfactants were able to reduce the surface tension
Mohsenatabar Firozjaii et al. (2018) performed a study to of the solution. However, Tg7 and Tg9 induced higher impact
identify the effect of DTAB on wettability and oil recovery in on the surface tensions compared to TX-100. At their CMC, the
carbonates. First, they found the CMC to be 3 wt%, 4.4 wt% surface tension was 27.8, 29.0 and 41.0 mN/m in the presence of
and 4.8 wt% using the IFT, conductivity and pH techniques, re- Tg7, Tg9 and TX-100, respectively. When different blends of the
spectively. Next, the change in rock wettability was determined surfactants were used (at the CMC), the resulted surface tension
using the static sessile drop technique at DTAB concentrations was in the range of 27.5 to 29.2 mN/m. Regarding the crude
of 1 wt% to 10 wt%. The measurements indicated that DTAB had oil–water IFT, a blend of Tg7 and TX-100 achieved the lowest
the ability to alter rock wettability to more water-wet conditions. IFT. Sarmah et al. (2020) also performed core flood experiments
Furthermore, EOR experiments showed that surfactant flooding using Tg7, TX-100 and their blend. They found that additional oil
(4 wt% DTAB) achieved additional 19% oil recovery above that of recovery of 21.0% to 25.7% of OOIP was achieved.
waterflooding. Bera et al. (2017) investigated the EOR performance of three
In a study conducted by Karimi et al. (2016), the effect of DTAB Tergitol surfactants including Tg7, Tg9 and Tergitol 1-S-12 (Tg12),
on oil recovery from limestone was investigated by conducting which are ethoxylated nonionic surfactants. First, they found
spontaneous imbibition tests. Using a formation brine with a that Tg12 provides the highest foamability and foam stability
salinity of 196,014 ppm, oil recovery was 8% of OOIP. After that, in solutions of different salinities (distilled water, 2 wt% and
a formation brine containing 0.5 wt% DTAB raised oil recovery 4 wt% NaCl and synthetic seawater). Moreover, Tg12 exhibited
to about 55.9% of OOIP. Moreover, the use of 100 times diluted the largest micelle sizes, which was attributed to its longest
3164
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 11. Basic classification of nonionic surfactants (Nollet, 2005).

chains of ethylene oxide (EO). Bera et al. (2017) also reported packs. Wei et al. (2020) also reported that APG substantially influ-
that additional oil recovery from sand packs was 22.1%, 24.6% enced the contact angle and oil–water IFT even at concentrations
and 26.9% of OOIP, which was achieved by Tg7, Tg9 and Tg12, close to 0.01 wt%. Therefore, APG surfactant can be possibly used
respectively. in EOR projects.
Hosseini (2019) investigated the effect of TX-100 on rock APG (lauryl glucoside) was also considered by do Vale et al.
wettability in the presence of asphaltene deposition. Sandstone (2020) for chemical EOR. They reported that the surface tension
wettability was determined by measuring the contact angle (θ ) of the surfactant solution decreased as the APG concentration
of an oil drop using the sessile drop technique. First, the rock increased. They also reported that APG flooding at a concentration
exhibited water-wet properties (θ = 21◦ ). However, asphaltene of 2% could effectively increase oil recovery by 52.1% of OOIP
precipitation changed rock wettability into oil-wet state (θ = after waterflooding. Experiments conducted by Li et al. (2019)
112◦ ). TX-100 was used to prevent this unfavorable change in demonstrated that APG has outstanding emulsification and in-
wettability. As a result, the contact angle (θ ) of the oil drop was terfacial activity properties. Moreover, core-flood experiments
reduced from 112◦ to 34◦ , indicating that the water-wet state showed that APG could increase heavy oil recovery by 10.1% of
was restored. Hou et al. (2018) compared the effects of TX-100, OOIP at a salinity of 30 g/L under 90 ◦ C.
CTAB (cationic surfactant) and their mixture on surface tension
and quartz wettability. The results showed that TX-100 and the 5.4. Zwitterionic surfactants
surfactant mixture achieved higher reduction in surface tension
compared to CTAB. It was also shown that both surfactants could Zwitterionic surfactants are surface active compounds that
change quartz wettability toward water-wet conditions to differ- comprise both positive and negative charge centers located at
ent extents. However, the surfactant mixture presented a higher the headgroup and separated by a short methylene segment
impact on wettability. (Gerola et al., 2017b,a). The structure of a zwitterionic surfactant
Alkyl polyglucosides (APGs), shown previously in Fig. 11, are molecule is shown in Fig. 12. Zwitterionic surfactants contain long
nonionic biosurfactants (do Vale et al., 2020; Li et al., 2020b) hydrophobic chains that are usually connected to the positive
that can be produced from natural sources such as vegetable charge center. The adsorption and micellization characteristics of
oil, sugar and starch. APGs readily biodegrade in aerobic envi- zwitterionic surfactants rely on specifics of their structure. These
ronments. Therefore, they tend to impose a low environmental include, for instance, the nature of the headgroup, the length of
burden compared to other surfactants (Ríos et al., 2016). Wei et al. the inter-charge spacer (n) and the length of the hydrophobic
(2020) studied the potential for APG surfactant to be used for EOR alkyl tail (m). For example, as the value of m increases, micel-
applications in sandstone at high salinity and temperature. The lization becomes more favored. However, it reaches a maximum
used APG was basically synthesized from glucose (C6 H12 O6 ) and at n equals 3 or 4 denoting the number of methylene groups in
lauryl alcohol (C12 H26 O). The results showed that APG could sig- the inter-charge spacer (Gerola et al., 2017a).
nificantly reduce the surface tension at 90 ◦ C. It was also shown The positive charge in zwitterionic surfactants results from
that static APG adsorption decreased slightly as the temperature cationic moieties such as amines or quaternary ammonium
increased from 20 to 50, 80 and 115 ◦ C. In terms of the salinity groups. On the other hand, the negative charge occurs due to the
effect, static APG adsorption increased due to the addition of salts presence of anionic moieties such as sulfonic acid, carboxylic acid
at different concentrations. However, the impact of salinity was and esters of phosphoric and sulfuric acids (Kwaśniewska et al.,
lower when investigated using dynamic adsorption tests in sand 2020; Zhang et al., 2019b; Zhao et al., 2015). Recently, research
3165
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 12. Structure of a zwitterionic surfactant molecule.


Source: Modified from Gerola et al. (2017b).

has shown that zwitterionic surfactants can be incorporated in behavior of C16DmCB and its effects on rock wettability and
chemical EOR. These surfactants have high water solubility, no- spontaneous imbibition. The results showed that C16DmCB ad-
table interfacial effects, high foam stability and good resistance to sorption was higher in carbonates compared with sandstones.
elevated temperature and salinity (Kamal et al., 2019; Zhao et al., In addition, C16DmCB adsorption increased due to the increase
2015). in salinity and alkalinity in both types of rocks. In terms of
There are several types of zwitterionic surfactants that were rock wettability, it was altered by C16DmCB toward more water
shown to be promising for chemical EOR applications. For ex- wet. However, wettability alteration occurred faster in sandstones
ample, Da et al. (2018) reported that a surfactant called cetyl than in carbonates. Finally, spontaneous imbibition experiments
betaine could significantly reduce the IFT between CO2 and brine demonstrated that the addition of C16DmCB increased oil recov-
of high salinity. In addition, they reported that cetyl betaine could ery by 3.1% and 18.9% of OOIP in carbonates and sandstones,
stabilize CO2 -in-water (C/W) foam in carbonates. The C/W foam respectively.
was found to be stable at temperatures reaching up to 150 ◦ C
and under high salinity conditions. Their results indicated that 5.5. Surfactant screening criteria
cetyl betaine can be a suitable zwitterionic surfactant to be used
for CO2 mobility control under elevated temperature and salinity. Surfactant screening studies aim to identify efficient surfac-
Chen et al. (2019b) synthesized a zwitterionic surfactant called tant formulations that are stable under normal and harsh reser-
SPODP through modification of oleic acids. By investigating its voir conditions. Furthermore, these studies may predict technical
effect on the crude oil–water IFT, SPODP was found to have problems and provide solutions to these problems before they are
remarkable interfacial properties at high temperature and salinity encountered during field application. In order to be a good candi-
conditions without alkali addition. date for chemical EOR, a surfactant should exhibit the following
properties: (1) tolerance to reservoir salinity/hardness; (2) ther-
Kamal et al. (2019) investigated thermal and foaming char-
mal stability at reservoir temperature; (3) capability of reducing
acteristics of betaine-based polyoxyethylene surfactants. The in-
the IFT to ∼0.01–0.001 dynes/cm under reservoir conditions; (4)
vestigated surfactants were denoted as OPAC, OPAH and OPAS
effectiveness at low concentrations (0.1–0.3%); (5) low adsorp-
and they contained carboxybetaine-, hydroxysulfobetaine- and
tion onto reservoir rock (< 1 mg/g rock); (6) compatibility with
sulfobetaine-headgroups, respectively. It was found that the three
other additives (alkalis, polymers, NPs, etc.) to be incorporated in
zwitterionic surfactants were thermally stable over 3 months of
EOR; and (7) commercial availability at reasonable cost (Al-Amodi
exposure to a temperature of 90 ◦ C. Regarding the foamability,
et al., 2016; Foo et al., 2016; Han et al., 2013; Khalilinezhad et al.,
the three surfactants exhibited a comparable foaming behavior in
2018).
synthetic seawater (SW) and in deionized water (DW), regardless
Based on the previous criteria, surfactant screening can be
of the headgroup type. However, foam height was lower using
performed. First, the compatibility of the candidate surfactants
DW compared with SW. Lastly, Kamal et al. (2019) recommended
with polymers and other additives, if any, should be examined
that the effect of these surfactants be investigated using core- at reservoir conditions. Next, phase behavior tests are conducted
flood experiments. Hussain et al. (2019) reported that OPAC, under reservoir temperature and salinity to evaluate the long-
OPAH and OPAS have high solubility in DW, SW and formation term stability of the compatible surfactants. Common examples
water (FW). They also reported that the surfactants could tolerate of these tests include aqueous stability and salinity scans. At this
high temperature (90 ◦ C) in DW, SW and FW without observing stage, surfactant formulations exhibiting the characteristics of a
phase separation or precipitation. Winsor Type III microemulsion are selected. Following that, the
C16DmCB is a zwitterionic surfactant that is based on car- effect of the selected surfactants on the IFT between crude oil
boxybetaine. Kumar et al. (2019) studied the performance of and brine is evaluated. Next, surfactant adsorption onto reservoir
C16DmCB using wettability measurement and core flooding tests. rock is measured. Finally, core-flood experiments are conducted
Different brines were prepared in the presence of C16DmCB (0 under reservoir conditions using the surfactant formulations that
to 0.020 wt%) and alkali (0 to 2 wt%). The results demonstrated showed acceptable thermal stability, salt tolerance, ultra-low IFT
that C16DmCB could alter the wettability of a quartz surface and low adsorption. Based on oil recovery and other laboratory
from oil wet to water wet. Furthermore, the addition of alkali results and by considering the cost of the surfactants, a decision
to the surfactant solution induced larger effect on the wetta- can be made regarding pilot tests and field application.
bility. Regarding core-flood tests, surfactant flooding increased
the pressure difference across sandstone cores, and it also in- 5.6. Field-scale application of surfactant-assisted chemical EOR
creased oil recovery by 20.0% of OOIP. However, incorporating
alkali with surfactant flooding increased oil recovery by 26.8% of The number of studies reporting on the EOR performance of
OOIP. This occurred due to the increased viscosity of the injected surfactants in the field is very limited compared to the labora-
solution. Kumar and Mandal (2019) investigated the adsorption tory studies. As shown in Table 6, there are different types of
3166
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 6
Field cases of surfactant-assisted chemical EOR.
Field name Location Formation Formation properties Oil properties Surfactants Other chemicals Performance References
type
Daqing China Sandstone k = 670 mD ρo = 0.864 g/cm 3
ABS (0.1–0.3%) HPAM polymer An incremental oil Sun et al. (2020)
Oilfield P = 858.6 psi µo = 8.2 − 9.3 cp (1300–2000 mg/L) recovery of 30% was
(B-1-D T = 42.4 ◦ C NaOH (0.8–1.2%) achieved using
Block) S = 5611 ppm alkaline-surfactant–
polymer
(ASP).
East Texas-USA Carbonate k = 3 mD Oil gravity = 31 ◦ API Linear ethoxylated CO2 Surfactant solution Alcorn et al. (2020)
Seminole P = 3400 psig alcohol (C12−14 EO22 ) reduced CO2
Field (San T = 40 ◦ C (0.5 wt%) injectivity by 70%
Andres S = 70,000 ppm indicating an
Formation) increased control of
CO2 mobility.
– Colombia Sandstone P = 1350.3 psi Oil gravity = 19 ◦ API Anionic surfactants SiO2 NPs (100 mg/L) An incremental oil Franco et al. (2020)
T = 50 ◦ C (350–1000 mg/L) Salts (27,137 ppm) recovery of 30,035
bbls was achieved
from two pilot wells
after around 5
months.
West West Sandstone k = 10–250 md ρo = 0.87 g/cm3 IOS 24–28 (0.5 wt%) FLOPAAM 3230 An incremental oil Volokitin et al. (2018)
Salym Siberia– T = 83 ◦ C µo = 2 cp IOS 15–18 (0.2 wt%) polymer (0.25 wt%) recovery of 16% was
Field Russia S = 14,000–16,000 Na2 CO3 (0.8–2.0 wt%) achieved from ASP
ppm C4 H10 O (2.0 wt%) pilot.
Dagang China Sandstone Average µo = 67 cp Petroleum sulfonate Associative polymer Field-pilot results Guo et al. (2018)
Oilfield kg = 161×10−3 pm2 surfactant PS-985 AP-P7 (2000 mg/L) showed that oil
(Guan-109 T = 78 ◦ C (0.3 wt%) production increased
Reservoir) S = 26,983 ppm from 27.7 to 51.7
m3 /d using
surfactant–polymer
(SP) flooding.
Tanjung Indonesia Sandstone k = 10–100 md Oil gravity = 40 ◦ API Anionic surfactant – Surfactant flooding Aslam et al. (2017)
Field (Zone T = 60–63 ◦ C µo = 1.14 cp A127 (2.05%) caused a noticeable
A) S = 3800 ppm increase in oil
recovery.
Reservoir Oklahoma- Sandstone k = 0.8–300 md Oil gravity = 40 ◦ API SATa (0.084%) as a CPXG biopolymer SP flooding achieved Hsu et al. (2012)
SF USA Average k = 74 md µo = 4 cp primary surfactant (1800 ppm) an incremental oil
T = 42 ◦ C SST (0.105%) and SC01 recovery of 19%, as
S = 165,000 ppm (0.042%) as shown by Single-well
co-surfactants field test.
Semoga Indonesia Carbonate k = 3.79–14.9 md µo = 0.98 − 1.28 cp Nonionic surfactants – Water cut decreased Rilian et al. (2010)
Field P = 845–895 psi by 8% and the and Syaifudin (2002)
(Baturaja 3-month oil recovery
Formation) increased by 5800
bbls.
Daqing China Sandstone k = 18.5×10−3 µm2 – Nonionic alkanol acid Auxiliary agent The oil recovery Yin and Pu (2008)
Oilfield amide increased by 30% for
(Chao 522 Surfactant (1.0%) single well.
Block)
Cotton- Wyoming- Carbonate kg = 0.2 − 130 md Oil gravity = 30 ◦ API Polyoxyethylene – Surfactant-soak Weiss et al. (2006)
wood USA (based on cores) µo = 24.3 cp alcohol (1500 ppm) treatments of 23 and Xie et al. (2005)
Creek Field T = 60 ◦ C wells failed in 70% of
S = 30,755 ppm the treated wells.
Northern China Sandstone k = 577 md ρo = 0.86 g/cm3 ABS (0.1–0.15%) as a HPAM polymer The total ASP pilot Daoshan et al. (2004)
part of T = 45 ◦ C µo = 9–10 cp primary surfactant (800–1800 mg/L) recovery was 16.6% of
Daqing S = 4456 ppm RH biosurfactant NaOH (1.2%) OOIP, with a
Oilfield (0.2%) as a sacrificial significantly reduced
agent cost.
Cambridge Wyoming- Eolian k = 845 md Oil gravity = 20 ◦ API Petrostep B-100 Na2 CO3 (1.25 wt%) The incremental oil Vargo et al. (1999)
Minnelusa USA sandstone P = 1792 psi µo = 31 cp surfactant (0.1 wt%) Alcoflood 1275A recovery achieved
Reservoir with T = 55.6 ◦ C polymer (1450 mg/L) using ASP flooding
dolomite was estimated by
and 26.8% of OOIP.
anhydrite
cement

a
SAT, SST and SC01 are sodium sulfosuccinate, sodium alkyl ethoxylate sulfate and branched sodium diphenyl oxide disulfonate, respectively. RH is a rhamnolipid-fermentation-liquor biosurfactant.

surfactants that have been employed in chemical EOR projects. recovery at laboratory scale was obtained using surfactants only,
Most of these surfactants were applied in sandstone formations unlike that obtained in the field. Therefore, it can be concluded
with a limited application in carbonates. Unlike laboratory-scale that the laboratory studies (i.e., core-flood experiments) do not
studies, field-scale studies show that surfactants are mainly used accurately predict the performance of surfactants in the field.
in the presence of other additives. The majority of these additives This may occur because the conditions for core-flood experi-
are alkalis (e.g., NaOH and Na2 CO3 ) and polymers (e.g., HPAM ments are more controlled compared to the conditions of the
and AP-P7). Alkalis are usually incorporated to increase the pH, field tests. Furthermore, these experiments are usually conducted
hence reducing surfactant adsorption on reservoir rock. They also using carefully selected core samples, which are frequently ob-
contribute to additional IFT reduction by in situ generation of tained from outcrops that vary from reservoir rock. Therefore,
surfactants. On the other hand, polymers are added to provide the 1-D displacement inside these cores is different from the
further mobility control. displacement in oil reservoirs. In many cases, these reservoirs are
Fig. 13 compares the incremental oil recovery obtained using highly heterogeneous, and contain complex systems of fractures,
surfactants at field and laboratory scales, it also provides details vugs, faults, etc., which may adversely affect the sweep efficiency.
of incremental oil recovery provided by each type of surfactant at Despite its ability to increase oil recovery, surfactant flooding
laboratory scale. According to Fig. 13, surfactants could achieve has some limitations and challenges that require further research.
an incremental oil recovery of ∼16%–30% of OOIP in the field, First, surfactant screening is a high resource-consuming process.
which incorporated additives such as alkalis and polymers. On the It is usually necessary to examine several types of surfactants
other hand, incremental oil recovery obtained at laboratory scale to identify the optimal formulations for specific reservoir con-
extended over a wide range of ∼1.7% to 54.0% of OOIP, with most ditions. Second, surfactant adsorption onto reservoir rock can
values ranging between 7.2% and 26.0%. However, incremental oil be only minimized, which imposes additional material costs to
3167
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Fig. 13. Incremental oil recovery obtained by employing surfactants at field and laboratory scales.

Fig. 14. Possible stabilization mechanisms of a liquid film by: (a) monolayer of bridging particles, (b) bilayer of closed-packed particles, and (c) network of particle
aggregates inside the film.
Source: Reproduced from Horozov (2008).

EOR projects. Third, the produced oil can be part of an emulsion, Gbadamosi et al., 2019a; Rezk and Allam, 2019; Sharma et al.,
thereby increasing the difficulty of oil–water separation. Fourth, 2016; Zargar et al., 2020). Research showed that NPs improve the
field experience with surfactant flooding shows that they need to performance of EOR through multiple mechanisms, which mainly
be incorporated with alkalis and/or polymers in order to obtain include: (1) wettability alteration; (2) application of disjoining
significant oil recovery. This means that more costly chemicals pressure between oil droplets and rock surfaces; (3) pore channel
are required to be transported to remote locations (e.g., off- plugging; (4) IFT reduction between crude oil and the injected
shore production facilities) where the storage capacity is usually fluids; and (5) inhibition of asphaltene precipitation (Cheraghian
limited. In addition, the presence of such chemicals in the pro- and Hendraningrat, 2016; Hemmat Esfe and Esfandeh, 2020; Rez-
duced fluids requires more processing and disposal operations. vani et al., 2018; Rosestolato et al., 2019; Yousefvand and Jafari,
Finally, several types of surfactants do not tolerate high salinity; 2018; Zhang et al., 2014). In addition, NPs can be utilized to
therefore, seawater cannot be used with these surfactants. reduce the viscosity of heavy and semi-heavy crude oil depending
on the type, size and concentration of the NPs (Cheraghian et al.,
5.7. Application of nanoparticles in surfactant flooding 2020) as well as the temperature of the bulk fluid containing
crude oil, brine, NPs, etc. (Patel et al., 2018).
Nanoparticles (NPs) are usually defined as particulate sub- Beside the previous effects, NPs provide some additional ad-
stances with sizes or dimensions ranging between 1 nm and vantages when incorporated in surfactant-assisted EOR. In gen-
100 nm (Auffan et al., 2009; Boverhof et al., 2015). Recently, NPs eral, experiments demonstrated that nanoparticle–surfactant
have gained prominence in petroleum research. Several studies flooding improves oil recovery by ∼4%–11% of OOIP above that
have reported the use of clay, silica/fumed silica, aluminum ox- obtained by surfactant flooding (Betancur et al., 2020; Li et al.,
ide, zinc oxide and titanium oxide NPs to change the interfacial 2020c; Liu et al., 2020b; Wu et al., 2017; Zhong et al., 2020). In
properties of fluids and rocks aiming for improved oil recov- a study conducted by Cheraghian (2016a), it was found that the
ery (Cheraghian, 2017, 2016b; Cheraghian and Tardasti, 2012; addition of TiO2 NPs (2.4 wt%) to a surfactant solution (SDS, 1800
3168
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Table 7
Current research trends on the use of surfactants in chemical EOR.
Reference Surfactants used Focus of study Remarks
Ge et al. (2021) POAPMB (zwitterionic) Development of a proper An SP system was developed to
surfactant–polymer (SP) system for increase oil recovery by ∼15% of
EOR applications without alkali to OOIP under high-salinity conditions
alleviate scale formation
Han et al. (2020) AOS (anionic) Development of a proper SP system Different SP systems increased oil
APS (anionic) for EOR applications without alkali to recovery by 24.5–34.8% of OOIP
ABS (anionic) alleviate scale formation
CAHS (zwitterionic)
Kurnia et al. (2020) APS (anionic) Identification of proper mixtures of Anionic–zwitterionic surfactant
ABS (anionic) anionic and zwitterionic surfactants solutions reduced the IFT below
AAS (anionic) for EOR applications without alkali to 0.001 dynes/cm unlike singular
AES (anionic) alleviate scale formation surfactant solutions
ADAO (zwitterionic)
CAHS (zwitterionic)
Franco et al. (2020) Commercial anionic Incorporating nanoparticles in Silica NPs (100 mg/L) improved the
surfactants surfactant flooding for improved EOR EOR performance of surfactant
performance flooding by 5% OOIP
Xu et al. (2020) SDS (anionic) Incorporating nanoparticles in Silica NPs (1 wt%) improved the EOR
Aerosol-OT (anionic) surfactant flooding for improved EOR performance of methane-surfactant
LAPE (nonionic) performance foam by 30% OOIP
Betaine-based surfactant
(zwitterionic)
Omidi et al. (2020) CTAB (cationic) Incorporating nanoparticles in CTAB solution (1000 ppm) and
TR-880 (zwitterionic) surfactant flooding for improved EOR nanofluid prepared from
performance Fe3 O4 /eggshell (500 ppm) increased
oil recovery by 8.16% OOIP
Chaturvedi and Sharma (2021) SDS (anionic) Stabilization of nanofluids using Certain SDS concentrations can
surfactant for EOR applications prevent salt-induced NP
agglomeration depending on solution
salinity
Alsmaeil et al. (2020) CTAB (cationic) Development of nano-capsules for Only 46% of CTAB was released after
controlled surfactant release during 12 days of exposure to a saline brine
EOR (56,000 mg/L salts)
Ojo et al. (2020) Alkyl betaine surfactant Encapsulation of surfactant with Surfactant encapsulation increased oil
(zwitterionic) halloysite nanotubes for controlled recovery of surfactant flooding by
surfactant release during EOR 17%–26% of OOIP
Romero-Zerón et al. (2020) SDS (anionic) Molecular encapsulation of surfactant Surfactant encapsulation reduced
with β -cyclodextrin (β -CD) for dynamic surfactant adsorption onto
controlled surfactant release during sand–kaolin surfaces by 43.3%
EOR
Chen and Schechter (2020) ADAC (cationic) Investigation of the effect of Compared to ionic surfactants,
ATAC (cationic) surfactant molecular structure on nonionic surfactants provide higher
APS (anionic) rock wettability wettability alternation depending on
IOS (anionic) the tail length and EO number
AES (nonionic)
CAPB (zwitterionic)
Dashtaki et al. (2020) Vitagnus plant extract Development of natural surfactant for Vitagnus surfactant solution (3000
EOR applications ppm) achieved a tertiary recovery of
10.6% OOIP
do Vale et al. (2020) APG (nonionic) Utilization of biosurfactant in EOR APG solution (2%) achieved a tertiary
applications recovery of 52.1% OOIP
Khayati et al. (2020) Saponin (nonionic) Utilization of natural surfactant in Saponin solution (5 g/L) achieved a
EOR applications tertiary recovery of 6.2–8.4% OOIP
Eslahati et al. (2020) Saponin (nonionic) Utilization of natural surfactant in Saponin solution (4 wt%) increased
EOR applications the total recovery using spontaneous
imbibition by 19.2%

ppm) improves oil recovery by 4.85% above that obtained using surfactants. This mechanism involves the arrangement of parti-
solutions containing SDS alone. Flooding experiments in a glass cles during foam film drainage. As depicted in Fig. 14, stabilization
micromodel showed that TiO2 NPs provides more uniform flow of a liquid film occurs according to three cases: (1) formation of
patterns as well as a more stable front when compared to flooding a monolayer of bridging particles; (2) formation of a bilayer of
in the absence of the NPs. Such results indicate that TiO2 NPs have close-packed particles; and (3) formation of a network of particle
aggregates inside the film, which all hinder foam drainage and
the ability to improve the rheological properties of the injected
coalescence (Horozov, 2008). In addition to foam stabilization,
fluids during EOR processes. some types of NPs have the ability to reduce surfactant adsorp-
NPs can be used to stabilize foams such as those containing tion onto reservoir rock, thus increasing the availability of the
CO2 and N2 , which are generated using surfactants for EOR pur- surfactant for the process of oil recovery. Several types of NPs
poses. Horozov (2008) reported a mechanism through which NPs were identified to induce such a positive effect. These include,
stabilize emulsion droplets and foam bubbles in the presence of for example, clay, silica and aluminum oxide NPs (Cheraghian and
3169
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Nezhad, 2016; Wu et al., 2017; Yekeen et al., 2017b). Regarding Declaration of competing interest
the mechanism of adsorption reduction, a frequently reported
mechanism is the competitive adsorption. In this mechanism, the The authors declare that they have no known competing finan-
surfactant molecules and the NPs compete on the adsorption sites cial interests or personal relationships that could have appeared
existing on the rock surface, thereby decreasing the adsorption to influence the work reported in this paper.
area available for the surfactant molecules (Venancio et al., 2020;
Wu et al., 2017; Zargartalebi et al., 2014). Acknowledgments

6. Summary and future prospects We acknowledge Qatar National Library (QNL) for funding
the open access publication of this article. We also acknowledge
In this paper, we provided a review on the use of surfactants the support for this publication provided by the National Prior-
in chemical EOR. First, we presented a background on surfactants, ities Research Program grant NPRP11S-1210-170079 from Qatar
which mainly discussed surfactant characterization and phase National Research Fund.
behavior. Second, we presented the role of surfactants in the pro-
cess of oil recovery. Third, we discussed the process of surfactant References
adsorption onto reservoir rock during surfactant flooding. Fourth,
we reviewed the different types of surfactants that have been Abalkhail, N., Liyanage, P.J., Upamali, K.A.N., Pope, G.A., Mohanty, K.K., 2020.
Alkaline-surfactant-polymer formulation development for a HTHS carbonate
used in chemical EOR in addition to the experimental and field
reservoir. J. Petrol. Sci. Eng. 191, 107236. http://dx.doi.org/10.1016/j.petrol.
application of surfactant flooding. 2020.107236.
In most oil reservoirs, more than 50% of OOIP remains trapped Abbas, A.H., Moslemizadeh, A., Wan Sulaiman, W.R., Jaafar, M.Z., Agi, A., 2020. An
in the reservoir rock after the primary and secondary phases insight into a di-chain surfactant adsorption onto sandstone minerals under
different salinity-temperature conditions: Chemical EOR applications. Chem.
of oil recovery. Researchers have investigated surfactant flood-
Eng. Res. Des. 153, 657–665. http://dx.doi.org/10.1016/j.cherd.2019.11.021.
ing as a method that has the potential to recover additional Abed, M.A., Saxena, A., Bohidar, H.B., 2004. Micellization of alpha-olefin sulfonate
quantities of oil in the tertiary phase. They showed that there in aqueous solutions studied by turbidity, dynamic light scattering and
are several mechanisms through which surfactants improve oil viscosity measurements. Colloids Surf. A 233, 181–187. http://dx.doi.org/10.
recovery. These mechanisms mainly include IFT reduction, wetta- 1016/j.colsurfa.2003.11.016.
Abushaikha, A.S., Gosselin, O.R., 2008. Matrix-fracture transfer function in dual-
bility alteration, foam generation and emulsification. Based on the media flow simulation: Review, comparison and validation. In: Eur. Conf.
reviewed experimental studies, incremental oil recovery obtained Exhib.. http://dx.doi.org/10.2118/113890-MS.
using surfactants usually ranges between 7% and 26% of OOIP, AfzaliTabar, M., Rashidi, A., Alaei, M., Koolivand, H., Pourhashem, S., Askari, S.,
with several cases exceeding 50% of OOIP. On the other hand, field 2020. Hybrid of quantum dots for interfacial tension reduction and reservoir
alteration wettability for enhanced oil recovery (EOR). J. Mol. Liq. 307,
application of surfactants showed that surfactants can achieve
112984. http://dx.doi.org/10.1016/j.molliq.2020.112984.
incremental oil recovery of ∼16%–30% of OOIP. However, this Agi, A., Junin, R., Gbonhinbor, J., Onyekonwu, M., 2018. Natural polymer flow
increase in oil recovery could be obtained only with the assistance behaviour in porous media for enhanced oil recovery applications: A review.
of substances such as alkalis and polymers. J. Pet. Explor. Prod. Technol. 8, 1349–1362. http://dx.doi.org/10.1007/s13202-
018-0434-7.
Surfactant flooding is usually challenged by problems that are
Ahmadi, M.-A., Ahmad, Z., Phung, L.T.K., Kashiwao, T., Bahadori, A., 2016.
aggravated by harsh reservoir conditions. These problems mainly Evaluation of the ability of the hydrophobic nanoparticles of SiO2 in the EOR
include surfactant adsorption and instability under high (or nor- process through carbonate rock samples. Pet. Sci. Technol. 34, 1048–1054.
mal) values of temperature and salinity. As a result, researchers http://dx.doi.org/10.1080/10916466.2016.1148052.
proposed criteria to be applied during surfactant screening for Ahmadi, M.A., Shadizadeh, S.R., 2018. Spotlight on the new natural surfactant
flooding in carbonate rock samples in low salinity condition. Sci. Rep. 8,
certain reservoir conditions and rock type. Based on that, it is 10985. http://dx.doi.org/10.1038/s41598-018-29321-w.
recommended that surfactant adsorption not to exceed 1 mg/g Ahmadi, M.A., Shadizadeh, S.R., Chen, Z., 2018. Thermodynamic analysis of
rock. Furthermore, the surfactant under evaluation should pos- adsorption of a naturally derived surfactant onto shale sandstone reservoirs.
sess the capability to reduce the oil–water IFT to ∼0.01–0.001 Eur. Phys. J. Plus 133, 420. http://dx.doi.org/10.1140/epjp/i2018-12264-x.
Aktar, S., Saha, M., Mahbub, S., Halim, M.A., Rub, M.A., Hoque, M.A., Islam, D.M.S.,
dynes/cm. In addition, the surfactant should be effective at low
Kumar, D., Alghamdi, Y.G., Asiri, A.M., 2020. Influence of polyethylene
concentration ranging between 0.1% and 0.3%. Otherwise, exceed- glycol on the aggregation/clouding phenomena of cationic and non-ionic
ing the concentration limit may affect the economic viability of surfactants in attendance of electrolytes (NaCl & Na2SO4): An experimental
surfactant flooding. and theoretical analysis. J. Mol. Liq. 306, 112880. http://dx.doi.org/10.1016/
j.molliq.2020.112880.
Table 7 provides current research trends according to recently
Al-Amodi, A.O., Al-Mubaiyedh, U.A., Sultan, A.S., Kamal, M.S., Hussein, I.A.,
published studies in the field of surfactant-based EOR. According 2016. Novel fluorinated surfactants for enhanced oil recovery in carbonate
to Table 7, future research regarding the application of surfactants reservoirs. Can. J. Chem. Eng. 94, 454–460. http://dx.doi.org/10.1002/cjce.
for chemical EOR can be grouped around five research areas: (1) 22406.
development of surfactant or surfactant–polymer mixtures that Alcorn, Z.P., Føyen, T., Zhang, L., Karakas, M., Biswal, S.L., Hirasaki, G., Graue, A.,
2020. CO foam field pilot design and initial results. In: SPE Improv. Oil
can be employed in chemical EOR without using alkalis; there- Recover. Conf.. http://dx.doi.org/10.2118/200450-MS.
fore, alleviating the problem of scale build-up; (2) incorporating Ali, H., Soleimani, H., Yahya, N., Khodapanah, L., Sabet, M., Demiral, B.M.R., Hus-
nanoparticles that can improve the performance of surfactant sain, T., Adebayo, L.L., 2020. Enhanced oil recovery by using electromagnetic-
flooding; (3) development of surfactant encapsulation techniques assisted nanofluids: A review. J. Mol. Liq. 309, 113095. http://dx.doi.org/10.
1016/j.molliq.2020.113095.
to control surfactant release during the process of EOR; (4) in-
Aljuboori, F.A., Lee, J.H., Elraies, K.A., Stephen, K.D., 2019. Gravity drainage mech-
vestigation of the effect of surfactant molecular structure on anism in naturally fractured carbonate reservoirs; Review and application.
rock wettability and oil recovery; and (5) development of nat- Energies http://dx.doi.org/10.3390/en12193699.
ural surfactants/biosurfactants to reduce the environmental costs Almahfood, M., Bai, B., 2018. The synergistic effects of nanoparticle-surfactant
associated with EOR processes. Generally, experimental research nanofluids in EOR applications. J. Petrol. Sci. Eng. 171, 196–210. http://dx.
doi.org/10.1016/j.petrol.2018.07.030.
on the use of surfactants to improve the efficiency of EOR in
Alsmaeil, A.W., Enotiadis, A., Hammami, M.A., Giannelis, E.P., 2020. Slow release
forthcoming years is expected to revolve around the previously of surfactant using silica nanosized capsules. SPE J. 9, http://dx.doi.org/10.
mentioned fields of study. 2118/202479-PA, Preprint.

3170
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Álvarez, M.S., Longo, M.A., Deive, F.J., Rodríguez, A., 2019. Non-ionic surfactants Bourrel, M., Salager, J.L., Schechter, R.S., Wade, W.H., 1980. A correlation for
and ionic liquids are a suitable combination for aqueous two-phase sys- phase behavior of nonionic surfactants. J. Colloid Interface Sci. 75, 451–461.
tems. Fluid Phase Equilib. 502, 112302. http://dx.doi.org/10.1016/j.fluid.2019. http://dx.doi.org/10.1016/0021-9797(80)90470-1.
112302. Bouton, F., Durand, M., Nardello-Rataj, V., Serry, M., Aubry, J.-M., 2009. Clas-
Amirmoshiri, M., Zhang, L., Puerto, M.C., Tewari, R.D., Bahrim, R.Z.B.K., Fara- sification of terpene oils using the fish diagrams and the equivalent alkane
jzadeh, R., Hirasaki, G.J., Biswal, S.L., 2020. Role of wettability on the carbon (EACN) scale. Colloids Surf. A 338, 142–147. http://dx.doi.org/10.1016/
adsorption of an anionic surfactant on sandstone cores. Langmuir 36, j.colsurfa.2008.05.027.
10725–10738. http://dx.doi.org/10.1021/acs.langmuir.0c01521. Boverhof, D.R., Bramante, C.M., Butala, J.H., Clancy, S.F., Lafranconi, M., West, J.,
Anas, N.A., Fen, Y.W., Yusof, N.A., Omar, N.A., Ramdzan, N.S., Daniyal, W.M., Gordon, S.C., 2015. Comparative assessment of nanomaterial definitions and
2020. Investigating the properties of cetyltrimethylammonium bro- safety evaluation considerations. RTP 73, 137–150. http://dx.doi.org/10.1016/
mide/hydroxylated graphene quantum dots thin film for potential op- j.yrtph.2015.06.001.
tical detection of heavy metal ions. Materials http://dx.doi.org/10.3390/ Brantson, E.T., Ju, B., Opoku Appau, P., Akwensi, P.H., Peprah, G.A., Liu, N.,
ma13112591. Aphu, E.S., Annan Boah, E., Aidoo Borsah, A., 2020. Development of hybrid
Andersen, P.Ø., Evje, S., Kleppe, H., 2014. A model for spontaneous imbibition as low salinity water polymer flooding numerical reservoir simulator and smart
a mechanism for oil recovery in fractured reservoirs. Transp. Porous Media proxy model for chemical enhanced oil recovery (CEOR). J. Petrol. Sci. Eng.
101, 299–331. http://dx.doi.org/10.1007/s11242-013-0246-7. 187, 106751. http://dx.doi.org/10.1016/j.petrol.2019.106751.
Antón, R.E., Garcés, N., Yajure, A., 1997. A correlation for three-phase behavior of Bratovcic, A., Nazdrajic, S., 2020. Viscoelastic behavior of synthesized liquid
cation1c surfactant-oil-water systems. J. Dispers. Sci. Technol. 18, 539–555. soaps and surface activity properties of surfactants. J. Surfactants Deterg.
http://dx.doi.org/10.1080/01932699708943755. http://dx.doi.org/10.1002/jsde.12444, n/a.
Aslam, B.M., Ulitha, D., Swadesi, B., Fauzi, I., Marhaendrajana, T., Purba, F.I., Brown, K.E., 1982. Overview of artificial lift systems. J. Pet. Technol. 34,
Wardhana, A.I., Buhari, A., Hakim, R., Hasibuan, R., 2017. History match 2384–2396. http://dx.doi.org/10.2118/9979-PA.
to support interpretation of surfactant flooding pilot test in tanjung field. Bryan, J., Kantzas, A., 2009. Potential for alkali-surfactant flooding in heavy oil
In: SPE/IATMI Asia Pacific Oil Gas Conf. Exhib.. http://dx.doi.org/10.2118/ reservoirs through oil-in-water emulsification. J. Can. Pet. Technol. 48, 37–46.
186325-MS. http://dx.doi.org/10.2118/09-02-37.
Auffan, M., Rose, J., Bottero, J.-Y., Lowry, G.V., Jolivet, J.-P., Wiesner, M.R., 2009. Budhathoki, M., Barnee, S.H.R., Shiau, B.-J., Harwell, J.H., 2016a. Improved oil
Towards a definition of inorganic nanoparticles from an environmental, recovery by reducing surfactant adsorption with polyelectrolyte in high
health and safety perspective. Nature Nanotechnology 4, 634–641. http: saline brine. Colloids Surf. A 498, 66–73. http://dx.doi.org/10.1016/j.colsurfa.
//dx.doi.org/10.1038/nnano.2009.242. 2016.03.012.
Awolayo, A.N., Sarma, H.K., Nghiem, L.X., 2018. Brine-dependent recovery Budhathoki, M., Hsu, T.-P., Lohateeraparp, P., Roberts, B.L., Shiau, B.-J., Har-
processes in carbonate and sandstone petroleum reservoirs: Review of well, J.H., 2016b. Design of an optimal middle phase microemulsion for
laboratory-field studies, interfacial mechanisms and modeling attempts. ultra high saline brine using hydrophilic lipophilic deviation (HLD) method.
Energies http://dx.doi.org/10.3390/en11113020. Colloids Surf. A 488, 36–45. http://dx.doi.org/10.1016/j.colsurfa.2015.09.066.
Azam, M.R., Tan, I.M., Ismail, L., Mushtaq, M., Nadeem, M., Sagir, M., 2013. Static Burnham, P., Dollahon, N., Li, C.H., Viescas, A.J., Papaefthymiou, G.C., 2013.
adsorption of anionic surfactant onto crushed Berea Sandstone. J. Pet. Explor. Magnetization and specific absorption rate studies of ball-milled iron oxide
Prod. Technol. 3, 195–201. http://dx.doi.org/10.1007/s13202-013-0057-y. nanoparticles for biomedicine. J. Nanoparticles 2013, 1–13. http://dx.doi.org/
Azam, M.R., Tan, I.M., Ismail, L., Mushtaq, M., Nadeem, M., Sagir, M., 2014. Kinet- 10.1155/2013/181820.
ics and equilibria of synthesized anionic surfactant onto Berea Sandstone. J. Bustamante-Rendón, R.A., Pérez, E., Gama Goicochea, A., 2020. Comparing the
Dispers. Sci. Technol. 35, 223–230. http://dx.doi.org/10.1080/01932691.2013. efficiency of pure and mixed cationic and nonionic surfactants used in
783491. enhanced oil recovery by mesoscopic simulations. Fuel 277, 118287. http:
Azodi, M., Nazar, A.R.S., 2013. Experimental design approach to investigate the //dx.doi.org/10.1016/j.fuel.2020.118287.
effects of operating factors on the surface tension, viscosity, and stability Castor, T.P., Somerton, W.H., Kelly, J.F., 1981. In: Shah, D.O. (Ed.), Recovery
of heavy crude oil-in-water emulsions. J. Dispers. Sci. Technol. 34, 273–282. Mechanisms of Alkaline Flooding BT - Surface Phenomena in Enhanced Oil
http://dx.doi.org/10.1080/01932691.2011.646611. Recovery. Springer US, Boston, MA, pp. 249–291. http://dx.doi.org/10.1007/
Bai, S., Kubelka, J., Piri, M., 2020. Relationship between molecular charge 978-1-4757-0337-5_14.
distribution and wettability reversal efficiency of cationic surfactants on Castro, M.J.L., Ojeda, C., Cirelli, A.F., 2014. Advances in surfactants for agrochem-
calcite surfaces. J. Mol. Liq. 318, 114009. http://dx.doi.org/10.1016/j.molliq. icals. Environ. Chem. Lett. 12, 85–95. http://dx.doi.org/10.1007/s10311-013-
2020.114009. 0432-4.
Bales, B.L., Benrraou, M., Zana, R., 2002. Krafft temperature and micelle ionization Castro Dantas, T.N., Soares, A.P.J., Wanderley Neto, A.O., Dantas Neto, A.A.,
of aqueous solutions of cesium dodecyl sulfate. J. Phys. Chem. B 106, Barros Neto, E.L., 2014. Implementing new microemulsion systems in wet-
9033–9035. http://dx.doi.org/10.1021/jp021297l. tability inversion and oil recovery from carbonate reservoirs. Energy Fuels
Bansal, V.K., Shah, D.O., 1978. The effect of divalent cations (Ca++ and Mg++ ) 28, 6749–6759. http://dx.doi.org/10.1021/ef501697.
on the optimal salinity and salt tolerance of petroleum sulfonate and Chaturvedi, K.R., Sharma, T., 2021. Rheological analysis and EOR potential of sur-
ethoxylated sulfonate mixtures in relation to improved oil recovery. J. Am. factant treated single-step silica nanofluid at high temperature and salinity. J.
Oil Chem. Soc. 55, 367–370. http://dx.doi.org/10.1007/BF02669932. Petrol. Sci. Eng. 196, 107704. http://dx.doi.org/10.1016/j.petrol.2020.107704.
Belhaj, A.F., Elraies, K.A., Mahmood, S.M., Zulkifli, N.N., Akbari, S., Hussien, O.S., Chen, Z.-Z., Gang, H.-Z., Liu, J.-F., Mu, B.-Z., Yang, S.-Z., 2019b. A thermal-stable
2020. The effect of surfactant concentration, salinity, temperature, and pH and salt-tolerant biobased zwitterionic surfactant with ultralow interfacial
on surfactant adsorption for chemical enhanced oil recovery: A review. J. tension between crude oil and formation brine. J. Petrol. Sci. Eng. 181,
Pet. Explor. Prod. Technol. 10, 125–137. http://dx.doi.org/10.1007/s13202- 106181. http://dx.doi.org/10.1016/j.petrol.2019.06.045.
019-0685-y. Chen, S., Liu, H., Yang, J., Zhou, Y., Zhang, J., 2019a. Bulk foam stability and
Bera, A., Mandal, A., Belhaj, H., Kumar, T., 2017. Enhanced oil recovery by non- rheological behavior of aqueous foams prepared by clay particles and alpha
ionic surfactants considering micellization, surface, and foaming properties. olefin sulfonate. J. Mol. Liq. 291, 111250. http://dx.doi.org/10.1016/j.molliq.
Pet. Sci. 14, 362–371. http://dx.doi.org/10.1007/s12182-017-0156-3. 2019.111250.
Bera, A., Ojha, K., Mandal, A., Kumar, T., 2011. Interfacial tension and phase Chen, W., Schechter, D.S., 2020. Surfactant selection for enhanced oil recovery
behavior of surfactant-brine–oil system. Colloids Surf. A 383, 114–119. http: based on surfactant molecular structure in unconventional liquid reservoirs.
//dx.doi.org/10.1016/j.colsurfa.2011.03.035. J. Petrol. Sci. Eng. 107702. http://dx.doi.org/10.1016/j.petrol.2020.107702.
Betancur, S., Olmos, C.M., Pérez, M., Lerner, B., Franco, C.A., Riazi, M., Gallego, J., Chen, L., Zhang, G., Wang, L., Wu, W., Ge, J., 2014. Zeta potential of limestone in
Carrasco-Marín, F., Cortés, F.B., 2020. A microfluidic study to investigate a large range of salinity. Colloids Surf. A 450, 1–8. http://dx.doi.org/10.1016/
the effect of magnetic iron core-carbon shell nanoparticles on displacement j.colsurfa.2014.03.006.
mechanisms of crude oil for chemical enhanced oil recovery. J. Petrol. Sci. Cheng, K.C., Khoo, Z.S., Lo, N.W., Tan, W.J., Chemmangattuvalappil, N.G., 2020.
Eng. 184, 106589. http://dx.doi.org/10.1016/j.petrol.2019.106589. Design and performance optimisation of detergent product containing binary
Bhosle, M.R., Joshi, S.A., Bondle, G.M., 2020. An efficient contemporary multi- mixture of anionic-nonionic surfactants. Heliyon 6, e03861. http://dx.doi.org/
component synthesis for the facile access to coumarin-fused new thiazolyl 10.1016/j.heliyon.2020.e03861.
chromeno[4,3-b]quinolones in aqueous micellar medium. J. Heterocycl. Cheraghian, G., 2016a. Effects of titanium dioxide nanoparticles on the efficiency
Chem. 57, 456–468. http://dx.doi.org/10.1002/jhet.3802. of surfactant flooding of heavy oil in a glass micromodel. Pet. Sci. Technol.
Blunt, M., Fayers, F.J., Orr, F.M., 1993. Carbon dioxide in enhanced oil recov- 34, 260–267. http://dx.doi.org/10.1080/10916466.2015.1132233.
ery. Energy Convers. Manag. 34, 1197–1204. http://dx.doi.org/10.1016/0196- Cheraghian, G., 2016b. Improved heavy oil recovery by nanofluid surfactant
8904(93)90069-M. flooding-an experimental study. In: 78th EAGE Conference and Exhibition
Bondi, C.A.M., Marks, J.L., Wroblewski, L.B., Raatikainen, H.S., Lenox, S.R., Geb- 2016. European Association of Geoscientists & Engineers, pp. 1–5.
hardt, K.E., 2015. Human and environmental toxicity of sodium lauryl sulfate Cheraghian, G., 2017. Evaluation of clay and fumed silica nanoparticles on
(SLS): Evidence for safe use in household cleaning products. Environ. Health adsorption of surfactant polymer during enhanced oil recovery. J. Jpn. Pet.
Insights 9, http://dx.doi.org/10.4137/EHI.S31765, EHI.S31765. Inst. 60, 85–94. http://dx.doi.org/10.1627/jpi.60.85.

3171
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Cheraghian, G., Hendraningrat, L., 2016. A review on applications of nanotech- Eslahati, M., Mehrabianfar, P., Isari, A.A., Bahraminejad, H., Manshad, A.K.,
nology in the enhanced oil recovery part A: Effects of nanoparticles on Keshavarz, A., 2020. Experimental investigation of alfalfa natural surfactant
interfacial tension. Int. Nano Lett. 6, 129–138. http://dx.doi.org/10.1007/ and synergistic effects of Ca2+ , Mg2+ , and SO4 2− ions for EOR applications:
s40089-015-0173-4. Interfacial tension optimization, wettability alteration and imbibition studies.
Cheraghian, G., Kiani, S., Nassar, N.N., Alexander, S., Barron, A.R., 2017. Silica J. Mol. Liq. 310, 113123. http://dx.doi.org/10.1016/j.molliq.2020.113123.
nanoparticle enhancement in the efficiency of surfactant flooding of heavy Fakher, S., Abdelaal, H., Elgahawy, Y., Imqam, A., 2019. A characterization of
oil in a glass micromodel. Ind. Eng. Chem. Res. 56, 8528–8534. http://dx.doi. different alkali chemical agents for alkaline flooding enhanced oil recovery
org/10.1021/acs.iecr.7b01675. operations: An experimental investigation. SN Appl. Sci. 1, 1622. http://dx.
Cheraghian, G., Nezhad, S.S.K., 2016. Improvement of heavy oil recovery and role doi.org/10.1007/s42452-019-1662-2.
of nanoparticles of clay in the surfactant flooding process. Pet. Sci. Technol. Farajzadeh, R., Andrianov, A., Zitha, P.L.J., 2010. Investigation of immiscible
34, 1397–1405. http://dx.doi.org/10.1080/10916466.2016.1198805. and miscible foam for enhancing oil recovery. Ind. Eng. Chem. Res. 49,
1910–1919. http://dx.doi.org/10.1021/ie901109d.
Cheraghian, G., Rostami, S., Afrand, M., 2020. Nanotechnology in enhanced oil
Farajzadeh, R., Bertin, H., Rossen, W.R., 2020. Editorial to the special issue: Foam
recovery. Process http://dx.doi.org/10.3390/pr8091073.
in porous media for petroleum and environmental engineering—Experience
Cheraghian, G., Tardasti, S., 2012. Improved Oil Recovery By the Effi-
sharing. Transp. Porous Media 131, 1–3. http://dx.doi.org/10.1007/s11242-
ciency of Nano-Particle in Imbibition Mechanism. European Association of
019-01329-4.
Geoscientists & Engineers, http://dx.doi.org/10.3997/2214-4609.20142967.
Farsang, E., Gaál, V., Horváth, O., Bárdos, E., Horváth, K., 2019. Analysis of
Cholpraves, C., Rattanaudom, P., Suriyapraphadilok, U., Charoensaeng, A., non-ionic surfactant triton X-100 using hydrophilic interaction liquid chro-
Gani, R., 2017. The systematic screening methodology for surfactant flooding matography and mass spectrometry. Molecules http://dx.doi.org/10.3390/
chemicals in enhanced oil recovery. In: Espuña, A., Graells, M., Puig- molecules24071223.
janer, L.B.T.-C.A.C.E. (Eds.), 27 European Symposium on Computer Aided Foo, Y.Y., Tewari, R.D., Kok, K.C., Elrufai, A., Elbaloula, H., Elkheir, L., 2016.
Process Engineering. Elsevier, pp. 991–996. http://dx.doi.org/10.1016/B978- Laboratory evaluation of chemical EOR process for viscous and high EACN
0-444-63965-3.50167-7. oil in East African oilfields. In: Int. Pet. Technol. Conf.. http://dx.doi.org/10.
Cong, Y., Zhang, W., Liu, C., Huang, F., 2020. Composition and oil-water interfacial 2523/IPTC-18601-MS.
tension studies in different vegetable oils. Food Biophys. 15, 229–239. http: Foroozesh, J., Kumar, S., 2020. Nanoparticles behaviors in porous media: Appli-
//dx.doi.org/10.1007/s11483-019-09617-8. cation to enhanced oil recovery. J. Mol. Liq. 316, 113876. http://dx.doi.org/
Cornwell, P.A., 2018. A review of shampoo surfactant technology: Consumer 10.1016/j.molliq.2020.113876.
benefits, raw materials and recent developments. Int. J. Cosmet. Sci. 40, Franco, Carlos A., Giraldo, L.J., Candela, C.H., Bernal, K.M., Villamil, F., Montes, D.,
16–30. http://dx.doi.org/10.1111/ics.12439. Lopera, S.H., Franco, Camilo A., Cortés, F.B., 2020. Design and tuning
Da, C., Alzobaidi, S., Jian, G., Zhang, L., Biswal, S.L., Hirasaki, G.J., Johnston, K.P., of nanofluids applied to chemical enhanced oil recovery based on the
2018. Carbon dioxide/water foams stabilized with a zwitterionic surfactant surfactant–nanoparticle–brine interaction: From laboratory experiments to
at temperatures up to 150, ◦ C in high salinity brine. J. Petrol. Sci. Eng. 166, oil field application. Nanomaterials http://dx.doi.org/10.3390/nano10081579.
880–890. http://dx.doi.org/10.1016/j.petrol.2018.03.071. Gbadamosi, A.O., Junin, R., Manan, M.A., Agi, A., Oseh, J.O., Usman, J., 2019a. Effect
Daghlian Sofla, S.J., Sharifi, M., Hemmati Sarapardeh, A., 2016. Toward mecha- of aluminium oxide nanoparticles on oilfield polyacrylamide: Rheology,
nistic understanding of natural surfactant flooding in enhanced oil recovery interfacial tension, wettability and oil displacement studies. J. Mol. Liq. 296,
processes: The role of salinity, surfactant concentration and rock type. J. Mol. 111863. http://dx.doi.org/10.1016/j.molliq.2019.111863.
Liq. 222, 632–639. http://dx.doi.org/10.1016/j.molliq.2016.07.086. Gbadamosi, A.O., Junin, R., Manan, M.A., Agi, A., Yusuff, A.S., 2019b. An overview
of chemical enhanced oil recovery: Recent advances and prospects. Int. Nano
Daoshan, L., Shouliang, L., Yi, L., Demin, W., 2004. The effect of biosurfactant
Lett. 9, 171–202. http://dx.doi.org/10.1007/s40089-019-0272-8.
on the interfacial tension and adsorption loss of surfactant in ASP flooding.
Gbadamosi, A.O., Kiwalabye, J., Junin, R., Augustine, A., 2018. A review of gas
Colloids Surf. A 244, 53–60. http://dx.doi.org/10.1016/j.colsurfa.2004.06.017.
enhanced oil recovery schemes used in the North Sea. J. Pet. Explor. Prod.
Das, A., Nguyen, N., Nguyen, Q.P., 2020. Low tension gas flooding for secondary
Technol. 8, 1373–1387. http://dx.doi.org/10.1007/s13202-018-0451-6.
oil recovery in low-permeability, high-salinity reservoirs. Fuel 264, 116601. Ge, M.-R., Miao, S.-J., Liu, J.-F., Gang, H.-Z., Yang, S.-Z., Mu, B.-Z., 2021. Laboratory
http://dx.doi.org/10.1016/j.fuel.2019.116601. studies on a novel salt-tolerant and alkali-free flooding system composed of
Dashtaki, S.R.M., Ali, J.A., Manshad, A.K., Nowrouzi, I., Mohammadi, A.H., Ke- a biopolymer and a bio-based surfactant for oil recovery. J. Petrol. Sci. Eng.
shavarz, A., 2020. Experimental investigation of the effect of vitagnus 196, 107736. http://dx.doi.org/10.1016/j.petrol.2020.107736.
plant extract on enhanced oil recovery process using interfacial tension Ge, J., Wang, Y., 2015. Surfactant enhanced oil recovery in a high temperature
(IFT) reduction and wettability alteration mechanisms. J. Pet. Explor. Prod. and high salinity carbonate reservoir. J. Surfactants Deterg. 18, 1043–1050.
Technol. http://dx.doi.org/10.1007/s13202-020-00966-6. http://dx.doi.org/10.1007/s11743-015-1735-1.
Davarpanah, A., Mirshekari, B., 2018. Experimental study and field application Gerola, A.P., Costa, P.F.A., Nome, F., Quina, F., 2017a. Micellization and adsorption
of appropriate selective calculation methods in gas lift design. Pet. Res. 3, of zwitterionic surfactants at the air/water interface. Curr. Opin. Colloid
239–247. http://dx.doi.org/10.1016/j.ptlrs.2018.03.005. Interface Sci. 32, 48–56. http://dx.doi.org/10.1016/j.cocis.2017.09.005.
Davies, J.T., 1957. A quantitative kinetic theory of emulsion type. I. Physical Gerola, A.P., Costa, P.F.A., Quina, F.H., Fiedler, H.D., Nome, F., 2017b. Zwitterionic
chemistry of the emulsifying agent, in: Gas/Liquid and Liquid/Liquid In- surfactants in ion binding and catalysis. Curr. Opin. Colloid Interface Sci. 32,
terface. Proceedings of the International Congress of Surface Activity. pp. 39–47. http://dx.doi.org/10.1016/j.cocis.2017.10.002.
426–438. Glennie, A.R., Mohareb, M.M., Palepu, R.M., 2006. Thermodynamic and related
Derikvand, Z., Rezaei, A., Parsaei, R., Riazi, M., Torabi, F., 2020. A mechanistic ex- properties of alkyl cationic surfactants based on dimethyl and diethyl
perimental study on the combined effect of Mg2+ , Ca2+ , and SO4 2− ions and ethanol amines. J. Dispers. Sci. Technol. 27, 731–738. http://dx.doi.org/10.
a cationic surfactant in improving the surface properties of oil/water/rock 1080/01932690600660608.
system. Colloids Surf. A 587, 124327. http://dx.doi.org/10.1016/j.colsurfa. Grassia, P., Torres-Ulloa, C., Berres, S., Mas-Hernández, E., Shokri, N., 2016.
2019.124327. Foam front propagation in anisotropic oil reservoirs. Eur. Phys. J. E 39, 42.
Desnelli, D., Permana, Y., Radiman, C.L., 2015. Copolymerization of acrylamide http://dx.doi.org/10.1140/epje/i2016-16042-5.
with 9- and 10-acrylamidodecanoic acids. In: Macromol. Symp., vol. 353, pp. Griffin, W.C., 1949. Classification of surface-active agents by ‘‘HLB’’. J. Soc.
198–204. http://dx.doi.org/10.1002/masy.201550327. Cosmet. Chem. 1, 311–326.
Griffin, W.C., 1954. Calculation of HLB values of non-ionic surfactants. J. Soc.
Dicharry, C., Diaz, J., Torré, .J.-P., Ricaurte, M., 2016. Influence of the carbon
Cosmet. Chem. 5, 249–256.
chain length of a sulfate-based surfactant on the formation of CO2 , CH4 and
Grządka, E., Matusiak, J., 2020. Changes in the CMC/ZrO2 system properties in
CO2 –CH4 gas hydrates. Chem. Eng. Sci. 152, 736–745. http://dx.doi.org/10.
the presence of hydrocarbon, fluorocarbon and silicone surfactants. J. Mol.
1016/j.ces.2016.06.034.
Liq. 303, 112699. http://dx.doi.org/10.1016/j.molliq.2020.112699.
Dölle, S., Lechner, B.-D., Park, J.H., Schymura, S., Lagerwall, J.P.F., Scalia, G., 2012. Guo, F., He, J., Johnson, P.A., Aryana, S.A., 2017. Stabilization of CO2 foam
Utilizing the krafft phenomenon to generate ideal micelle-free surfactant- using by-product fly ash and recyclable iron oxide nanoparticles to improve
stabilized nanoparticle suspensions. Angew. Chem. Int. Ed. 51, 3254–3257. carbon utilization in EOR processes. Sustain. Energy Fuels 1, 814–822. http:
http://dx.doi.org/10.1002/anie.201106793. //dx.doi.org/10.1039/C7SE00098G.
Doranehgard, M.H., Siavashi, M., 2018. The effect of temperature dependent Guo, Y., Zhang, J., Zhang, X., Hu, J., Wang, W., Liang, Y., 2018. Investigation
relative permeability on heavy oil recovery during hot water injection and application of an associative polymer-surfactant binary system for a
process using streamline-based simulation. Appl. Therm. Eng. 129, 106–116. successful flooding pilot in a high-temperature, high-salinity, ordinary heavy
http://dx.doi.org/10.1016/j.applthermaleng.2017.10.002. oil reservoir. In: SPE EOR Conf. Oil Gas West Asia. http://dx.doi.org/10.2118/
El-hoshoudy, A.N., Desouky, S.E.M., Elkady, M.Y., Al-Sabagh, A.M., Betiha, M.A., 190411-MS.
Mahmoud, S., 2017. Hydrophobically associated polymers for wettability Gyan, P.S., Xie, C., Brantson, E.T., Atuahene, S., 2019. Computer modeling and sim-
alteration and enhanced oil recovery – Article review. Egypt. J. Pet. 26, ulation for undersaturated primary drive recovery mechanism. Adv. Mech.
757–762. http://dx.doi.org/10.1016/j.ejpe.2016.10.008. Eng. 11, http://dx.doi.org/10.1177/1687814019841948, 1687814019841948.

3172
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Hamidi, H., Mohammadian, E., Rafati, R., Azdarpour, A., Ing, J., 2015. The effect Hsu, T.-P., Lohateeraparp, P., Roberts, B.L., Wan, W., Lin, Z., Wang, X., Bud-
of ultrasonic waves on the phase behavior of a surfactant–brine–oil system. hathoki, M., Shiau, B., Harwell, J.H., 2012. Improved oil recovery by chemical
Colloids Surf. A 482, 27–33. http://dx.doi.org/10.1016/j.colsurfa.2015.04.009. flood from high salinity reservoirs-single-well surfactant injection test. In:
Hammond, P.S., Unsal, E., 2011. Spontaneous imbibition of surfactant so- SPE EOR Conf. Oil Gas West Asia. http://dx.doi.org/10.2118/154838-MS.
lution into an oil-wet capillary: Wettability restoration by surfactant- Hu, D., Mafi, A., Chou, K.C., 2016. Revisiting the thermodynamics of water
contaminant complexation. Langmuir 27, 4412–4429. http://dx.doi.org/10. surfaces and the effects of surfactant head group. J. Phys. Chem. B 120,
1021/la1048503. 2257–2261. http://dx.doi.org/10.1021/acs.jpcb.5b11717.
Han, M., AlSofi, A., Fuseni, A., Zhou, X., Hassan, S., 2013. Development of chem- Hussain, S.M.S., Kamal, M.S., Fogang, L.T., 2019. Synthesis and physicochemi-
ical EOR formulations for a high temperature and high salinity carbonate cal investigation of betaine type polyoxyethylene zwitterionic surfactants
reservoir. In: IPTC 2013: International Petroleum Technology Conference. containing different ionic headgroups. J. Mol. Struct. 1178, 83–88. http:
European Association of Geoscientists & Engineers, p. cp-350. //dx.doi.org/10.1016/j.molstruc.2018.09.094.
Han, X., Chen, Z., Zhang, G., Yu, J., 2020. Surfactant-polymer flooding formulated Illous, E., Ontiveros, J.F., Lemahieu, G., Lebeuf, R., Aubry, J.-M., 2020. Am-
with commercial surfactants and enhanced by negative salinity gradient. Fuel phiphilicity and salt-tolerance of ethoxylated and propoxylated anionic
274, 117874. http://dx.doi.org/10.1016/j.fuel.2020.117874. surfactants. Colloids Surf. A 601, 124786. http://dx.doi.org/10.1016/j.colsurfa.
Han, X., Kurnia, I., Chen, Z., Yu, J., Zhang, G., 2019. Effect of oil reactivity on 2020.124786.
salinity profile design during alkaline-surfactant-polymer flooding. Fuel 254, Inada, T., Koyama, T., Tomita, H., Fuse, T., Kuwabara, C., Arakawa, K., Fujikawa, S.,
115738. http://dx.doi.org/10.1016/j.fuel.2019.115738. 2017. Anti-ice nucleating activity of surfactants against silver iodide in
Haq, B., Al Shehri, D., Al Damegh, A., Al Muhawesh, A., Albusaad, M., Lardhi, M., water-in-oil emulsions. J. Phys. Chem. B 121, 6580–6587. http://dx.doi.org/
Barri, A., Iddris, A., Muhammed, N., Hossain, S.M.Z., Rahman, M.M., Aziz, A., 10.1021/acs.jpcb.7b02644.
2020a. The role of carbon nanotubes (CNTs) and carbon particles in green Ivanova, A.A., Phan, C., Barifcani, A., Iglauer, S., Cheremisin, A.N., 2020. Effect
enhanced oil recovery (GEOR) for Arabian crude oil in sandstone core. APPEA of nanoparticles on viscosity and interfacial tension of aqueous surfactant
J. 60, 133–142. solutions at high salinity and high temperature. J. Surfactants Deterg. 23,
Haq, B., Liu, J., Liu, K., Al Shehri, D., 2020b. The role of biodegradable surfactant 327–338. http://dx.doi.org/10.1002/jsde.12371.
in microbial enhanced oil recovery. J. Petrol. Sci. Eng. 189, 106688. http: Jha, N.K., Iglauer, S., Sangwai, J.S., 2018. Effect of monovalent and divalent salts
//dx.doi.org/10.1016/j.petrol.2019.106688. on the interfacial tension of n-heptane against aqueous anionic surfactant
Haruna, M.A., Gardy, J., Yao, G., Hu, Z., Hondow, N., Wen, D., 2020. Nanoparticle solutions. J. Chem. Eng. Data 63, 2341–2350. http://dx.doi.org/10.1021/acs.
modified polyacrylamide for enhanced oil recovery at harsh conditions. Fuel jced.7b00640.
268, 117186. http://dx.doi.org/10.1016/j.fuel.2020.117186. Jia, H., Lian, P., Leng, X., Han, Y., Wang, Q., Jia, K., Niu, X., Guo, M., Yan, H., Lv, K.,
Harutyunyan, L.R., Harutyunyan, R.S., 2019. Effect of amino acids on micellization 2019. Mechanism studies on the application of the mixed cationic/anionic
and micellar parameters of anionic surfactant alpha olefin sulfonate C14– surfactant systems to enhance oil recovery. Fuel 258, 116156. http://dx.doi.
C16 in aqueous solutions: Surface tension, conductometric, volumetric, and org/10.1016/j.fuel.2019.116156.
fluorescence studies. J. Chem. Eng. Data 64, 640–650. http://dx.doi.org/10. Jin, L., Budhathoki, M., Jamili, A., Li, Z., Luo, H., Ben Shiau, B.J., Delshad, M.,
1021/acs.jced.8b00886. Harwell, J.H., 2017. Predicting microemulsion phase behavior using physics
Hazarika, K., Gogoi, S.B., 2020. Effect of alkali on alkali–surfactant flooding in based HLD-NAC equation of state for surfactant flooding. J. Petrol. Sci. Eng.
an Upper Assam oil field. J. Pet. Explor. Prod. Technol. 10, 1591–1601. 151, 213–223. http://dx.doi.org/10.1016/j.petrol.2016.12.035.
http://dx.doi.org/10.1007/s13202-019-00794-3. Jin, L., Jamili, A., Li, Z., Lu, J., Luo, H., Ben Shiau, B.J., Delshad, M., Harwell, J.H.,
He, C., Fan, Z., Xu, A., Zhao, L., 2019. Foamy oil properties and horizontal well 2015. Physics based HLD–NAC phase behavior model for surfactant/crude
inflow performance relationship under solution gas drive. Geosystem Eng. oil/brine systems. J. Petrol. Sci. Eng. 136, 68–77. http://dx.doi.org/10.1016/j.
22, 151–160. http://dx.doi.org/10.1080/12269328.2018.1504698. petrol.2015.10.039.
Hemmat Esfe, M., Esfandeh, S., 2020. 3D numerical simulation of the enhanced Jin, F., Li, Q., He, Y., Luo, Q., Pu, W., 2020. Experimental study on enhanced oil
oil recovery process using nanoscale colloidal solution flooding. J. Mol. Liq. recovery method in tahe high-temperature and high-salinity channel sand
301, 112094. http://dx.doi.org/10.1016/j.molliq.2019.112094. reservoir: Combination of profile control and chemical flooding. ACS Omega
Höök, M., Davidsson, S., Johansson, S., Tang, X., 2014. Decline and depletion rates 5, 5657–5665. http://dx.doi.org/10.1021/acsomega.9b03306.
of oil production: A comprehensive investigation. Philos. Trans. R. Soc. Lond. Jing, W., Huiqing, L., Genbao, Q., Yongcan, P., Yang, G., 2019. Investigations on
Ser. A Math. Phys. Eng. Sci. 372, 20120448. http://dx.doi.org/10.1098/rsta. spontaneous imbibition and the influencing factors in tight oil reservoirs.
2012.0448. Fuel 236, 755–768. http://dx.doi.org/10.1016/j.fuel.2018.09.053.
Horozov, T.S., 2008. Foams and foam films stabilised by solid particles. Curr. Johnson, Jr., C.E., 1976. Status of caustic and emulsion methods. J. Pet. Technol.
Opin. Colloid Interface Sci. 13, 134–140. http://dx.doi.org/10.1016/j.cocis. 28, 85–92. http://dx.doi.org/10.2118/5561-PA.
2007.11.009. Joshi, S.J., Geetha, S.J., Desai, A.J., 2015. Characterization and application of bio-
Hosseini, E., 2019. Experimental investigation of effect of asphaltene deposition surfactant produced by bacillus licheniformis R2. Appl. Biochem. Biotechnol.
on oil relative permeability, rock wettability alteration, and recovery in 177, 346–361. http://dx.doi.org/10.1007/s12010-015-1746-4.
WAG process. Pet. Sci. Technol. 37, 2150–2159. http://dx.doi.org/10.1080/ Kachangoon, R., Vichapong, J., Santaladchaiyakit, Y., Srijaranai, S., 2020. Cloud-
10916466.2018.1482335. point extraction coupled to in-situ metathesis reaction of deep eutectic
Hosseini, S., Sabet, M., Zeinolabedini Hezave, A., A. Ayoub, M., Elraies, K.A., solvents for preconcentration and liquid chromatographic analysis of neoni-
2020. Effect of combination of cationic surfactant and salts on wettability cotinoid insecticide residues in water, soil and urine samples. Microchem. J.
alteration of carbonate rock. Energy Sources A 1–17. http://dx.doi.org/10. 152, 104377. http://dx.doi.org/10.1016/j.microc.2019.104377.
1080/15567036.2020.1778141. Kalantari Meybodi, M., Shokrollahi, A., Safari, H., Lee, M., Bahadori, A., 2015.
Hosseininoosheri, P., Lashgari, H.R., Sepehrnoori, K., 2016. A novel method A computational intelligence scheme for prediction of interfacial tension
to model and characterize in-situ bio-surfactant production in microbial between pure hydrocarbons and water. Chem. Eng. Res. Des. 95, 79–92.
enhanced oil recovery. Fuel 183, 501–511. http://dx.doi.org/10.1016/j.fuel. http://dx.doi.org/10.1016/j.cherd.2015.01.004.
2016.06.035. Kamal, M.S., Hussain, S.M., Fogang, L.T., 2019. Role of ionic headgroups on
Hosseinzade Khanamiri, H., Baltzersen Enge, I., Nourani, M., Stensen, J.Å., the thermal, rheological, and foaming properties of novel betaine-based
Torsæter, O., Hadia, N., 2016a. EOR by low salinity water and surfactant at polyoxyethylene zwitterionic surfactants for enhanced oil recovery. Process
low concentration: Impact of injection and in situ brine composition. Energy http://dx.doi.org/10.3390/pr7120908.
Fuels 30, 2705–2713. http://dx.doi.org/10.1021/acs.energyfuels.5b02899. Karatayev, M., Movkebayeva, G., Bimagambetova, Z., 2019. In: Mouraviev, N.,
Hosseinzade Khanamiri, H., Torsæter, O., Stensen, J.Å., 2016b. Effect of calcium Koulouri, A. (Eds.), Increasing Utilisation of Renewable Energy Sources:
in pore scale oil trapping by low-salinity water and surfactant enhanced Comparative Analysis of Scenarios Until 2050 BT - Energy Security: Policy
oil recovery at strongly water-wet conditions: In situ imaging by X-ray Challenges and Solutions for Resource Efficiency. Springer International
microtomography. Energy Fuels 30, 8114–8124. http://dx.doi.org/10.1021/ Publishing, Cham, pp. 37–68. http://dx.doi.org/10.1007/978-3-030-01033-
acs.energyfuels.6b01236. 1_3.
Hou, H., He, H., Wang, Y., 2020. Effects of SDS on the activity and conformation Karimi, M., Al-Maamari, R.S., Ayatollahi, S., Mehranbod, N., 2015. Mechanistic
of protein tyrosine phosphatase from thermus thermophilus HB27. Sci. Rep. study of wettability alteration of oil-wet calcite: The effect of magnesium
10, 3195. http://dx.doi.org/10.1038/s41598-020-60263-4. ions in the presence and absence of cationic surfactant. Colloids Surf. A 482,
Hou, B., Jia, R., Fu, M., Wang, Y., Bai, Y., Huang, Y., 2018. Wettability alteration 403–415. http://dx.doi.org/10.1016/j.colsurfa.2015.07.001.
of an oil-wet sandstone surface by synergistic adsorption/desorption of Karimi, M., Al-Maamari, R.S., Ayatollahi, S., Mehranbod, N., 2016. Wettability
cationic/nonionic surfactant mixtures. Energy Fuels 32, 12462–12468. http: alteration and oil recovery by spontaneous imbibition of low salinity brine
//dx.doi.org/10.1021/acs.energyfuels.8b03450. into carbonates: Impact of Mg2+ , SO4 2− and cationic surfactant. J. Petrol. Sci.
Hou, B., Wang, Y., Cao, X., Zhang, J., Song, X., Ding, M., Chen, W., 2016. Surfactant- Eng. 147, 560–569. http://dx.doi.org/10.1016/j.petrol.2016.09.015.
induced wettability alteration of oil-wet sandstone surface: Mechanisms Khaledialidusti, R., Kleppe, J., Enayatpour, S., 2017. Evaluation of surfactant
and its effect on oil recovery. J. Surfactants Deterg. 19, 315–324. http: flooding using interwell tracer analysis. J. Pet. Explor. Prod. Technol. 7,
//dx.doi.org/10.1007/s11743-015-1770-y. 853–872. http://dx.doi.org/10.1007/s13202-016-0288-9.

3173
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Khaleel, O., Teklu, T.W., Alameri, W., Abass, H., Kazemi, H., 2019. Wettability Li, Q., Huang, Y., Wen, D., Fu, R., Feng, L., 2020b. Application of alkyl polyglyco-
alteration of carbonate reservoir cores—laboratory evaluation using comple- sides for enhanced bioremediation of petroleum hydrocarbon-contaminated
mentary techniques. SPE Reserv. Eval. Eng. 22, 911–922. http://dx.doi.org/10. soil using Sphingomonas changbaiensis and Pseudomonas stutzeri. Sci. Total
2118/194483-PA. Environ. 719, 137456. http://dx.doi.org/10.1016/j.scitotenv.2020.137456.
Khalilinezhad, S.S., Mobaraki, S., Zakavi, M., Omidvar Sorkhabadi, M., Cher- Li, P., Ishiguro, M., 2016. Adsorption of anionic surfactant (sodium dodecyl
aghian, G., Jarrahian, K., 2018. Mechanistic modeling of nanoparticles- sulfate) on silica. Soil Sci. Plant Nutr. 62, 223–229. http://dx.doi.org/10.1080/
assisted surfactant flood. Arab. J. Sci. Eng. 43, 6609–6625. http://dx.doi.org/ 00380768.2016.1191969.
10.1007/s13369-018-3415-8. Li, P., Wang, Z., Ma, K., Chen, Y., Yan, Z., Penfold, J., Thomas, R.K., Campana, M.,
Khan, S., Khan, P., Hassan, M.I., Ahmad, F., Islam, A., 2019. Protein stability: Webster, J.R.P., Washington, A., 2020a. Multivalent electrolyte induced sur-
Determination of structure and stability of the transmembrane protein face ordering and solution self-assembly in anionic surfactant mixtures:
Mce4A from M. tuberculosis in membrane-like environment. Int. J. Biol. Sodium dodecyl sulfate and sodium diethylene glycol monododecyl sulfate.
Macromol. 126, 488–495. http://dx.doi.org/10.1016/j.ijbiomac.2018.12.183. J. Colloid Interface Sci. 565, 567–581. http://dx.doi.org/10.1016/j.jcis.2020.01.
Khattab, A., Mohamed, M., Basalious, E.B., 2020. Design of self-nanoemulsifying 032.
system to enhance absorption and bioavailability of poorly permeable Li, Z., Xu, D., Yuan, Y., Wu, H., Hou, J., Kang, W., Bai, B., 2020d. Advances of
Aliskiren hemi-fumarate. J. Drug Deliv. Sci. Technol. 57, 101646. http://dx. spontaneous emulsification and its important applications in enhanced oil
doi.org/10.1016/j.jddst.2020.101646. recovery process. Adv. Colloid Interface Sci. 277, 102119. http://dx.doi.org/
Khayati, H., Moslemizadeh, A., Shahbazi, K., Moraveji, M.K., Riazi, S.H., 2020. An 10.1016/j.cis.2020.102119.
experimental investigation on the use of saponin as a non-ionic surfactant Li, Y., Zhang, W., Kong, B., Puerto, M., Bao, X., Sha, O., Shen, Z., Yang, Y., Liu, Y.,
for chemical enhanced oil recovery (EOR) in sandstone and carbonate oil Gu, S., Miller, C., Hirasaki, G.J., 2016. Mixtures of anionic/cationic surfactants:
reservoirs: Ift, wettability alteration, and oil recovery. Chem. Eng. Res. Des. A new approach for enhanced oil recovery in low-salinity, high-temperature
160, 417–425. http://dx.doi.org/10.1016/j.cherd.2020.04.033. sandstone reservoir. SPE J. 21, 1164–1177. http://dx.doi.org/10.2118/169051-
Kok, M.V., 2012. Progress and recent utilization trends in petroleum recovery: PA.
Steam injection and in-situ combustion. Energy Sources A 34, 2253–2259. Liu, Z., Ghatkesar, M.K., Sudhölter, E.J.R., Singh, B., Kumar, N., 2019. Under-
http://dx.doi.org/10.1080/15567036.2010.504947. standing the cation-dependent surfactant adsorption on clay minerals in
Kondo, S., Shiohara, M., Maruyama, K., Fukaya, K., Yamada, K., Ogawa, S., Saito, S., oil recovery. Energy Fuels 33, 12319–12329. http://dx.doi.org/10.1021/acs.
2007. Effect of the hydrophilic-lipophilic balance (HLB) of surfactants in- energyfuels.9b03109.
cluded in the post-CMP cleaning chemicals on porous SiOC direct CMP. In: Liu, Z., Hedayati, P., Sudhölter, E.J.R., Haaring, R., Shaik, A.R., Kumar, N., 2020c.
2007 IEEE International Interconnect Technology Conferencee. pp. 172–174. Adsorption behavior of anionic surfactants to silica surfaces in the presence
http://dx.doi.org/10.1109/IITC.2007.382381. of calcium ion and polystyrene sulfonate. Colloids Surf. A 602, 125074.
Kumar, A., Mandal, A., 2017a. Synthesis and physiochemical characterization of http://dx.doi.org/10.1016/j.colsurfa.2020.125074.
zwitterionic surfactant for application in enhanced oil recovery. J. Mol. Liq. Liu, J., Liu, Z., Yuan, T., Wang, C., Gao, R., Hu, G., Xu, J., Zhao, J., 2020a. Synthesis
243, 61–71. http://dx.doi.org/10.1016/j.molliq.2017.08.032. and properties of zwitterionic gemini surfactants for enhancing oil recovery.
Kumar, S., Mandal, A., 2017b. A comprehensive review on chemically enhanced J. Mol. Liq. 311, 113179. http://dx.doi.org/10.1016/j.molliq.2020.113179.
water alternating gas/CO2 (CEWAG) injection for enhanced oil recovery. J. Liu, F., Wang, M., 2020. Review of low salinity waterflooding mechanisms:
Petrol. Sci. Eng. 157, 696–715. http://dx.doi.org/10.1016/j.petrol.2017.07.066. Wettability alteration and its impact on oil recovery. Fuel 267, 117112.
Kumar, A., Mandal, A., 2019. Critical investigation of zwitterionic surfactant http://dx.doi.org/10.1016/j.fuel.2020.117112.
for enhanced oil recovery from both sandstone and carbonate reservoirs: Liu, P., Yu, H., Niu, L., Ni, D., Zhao, Q., Li, X., Zhang, Z., 2020b. Utilization
Adsorption, wettability alteration and imbibition studies. Chem. Eng. Sci. 209, of Janus-silica/surfactant nanofluid without ultra-low interfacial tension for
115222. http://dx.doi.org/10.1016/j.ces.2019.115222. improving oil recovery. Chem. Eng. Sci. 228, 115964. http://dx.doi.org/10.
Kumar, N., Mandal, A., 2020. Wettability alteration of sandstone rock by sur- 1016/j.ces.2020.115964.
factant stabilized nanoemulsion for enhanced oil recovery—A mechanistic Liyanage, P.J., Lu, J., Arachchilage, G.W.P., Weerasooriya, U.P., Pope, G.A., 2015.
study. Colloids Surf. A 601, 125043. http://dx.doi.org/10.1016/j.colsurfa.2020. A novel class of large-hydrophobe tristyrylphenol (TSP) alkoxy sulfate sur-
125043. factants for chemical enhanced oil recovery. J. Petrol. Sci. Eng. 128, 73–85.
Kumar, A., Neale, G., Hornof, V., 1984. Effects of connate water composition on http://dx.doi.org/10.1016/j.petrol.2015.02.023.
interfacial tension behaviour of surfactant solutions. J. Can. Pet. Technol. 23, Lombardo, D., Kiselev, M.A., Magazù, S., Calandra, P., 2015. Amphiphiles
6. http://dx.doi.org/10.2118/84-01-03. self-assembly: Basic concepts and future perspectives of supramolecular
Kumar, A., Saw, R.K., Mandal, A., 2019. RSM optimization of oil-in-water approaches. Adv. Condens. Matter Phys. 2015, 1–22. http://dx.doi.org/10.
microemulsion stabilized by synthesized zwitterionic surfactant and its 1155/2015/151683.
properties evaluation for application in enhanced oil recovery. Chem. Eng. Luan, H., Gong, L., Yue, X., Nie, X., Chen, Q., Guan, D., Que, T., Liao, G.,
Res. Des. 147, 399–411. http://dx.doi.org/10.1016/j.cherd.2019.05.034. Su, X., Feng, Y., 2019. Micellar aggregation behavior of alkylaryl sulfonate
Kumari, R., Kakati, A., Nagarajan, R., Sangwai, J.S., 2019. Synergistic effect of surfactants for enhanced oil recovery. Molecules http://dx.doi.org/10.3390/
mixed anionic and cationic surfactant systems on the interfacial tension molecules24234325.
of crude oil-water and enhanced oil recovery. J. Dispers. Sci. Technol. 40, Machale, J., Majumder, S.K., Ghosh, P., Sen, T.K., 2019a. Role of chemical additives
969–981. http://dx.doi.org/10.1080/01932691.2018.1489280. and their rheological properties in enhanced oil recovery. Rev. Chem. Eng.
Kume, G., Gallotti, M., Nunes, G., 2008. Review on anionic/cationic surfactant 20180033. http://dx.doi.org/10.1515/revce-2018-0033.
mixtures. J. Surfactants Deterg. 11, 1–11. http://dx.doi.org/10.1007/s11743- Machale, J., Majumder, S.K., Ghosh, P., Sen, T.K., 2019b. Development of a novel
007-1047-1. biosurfactant for enhanced oil recovery and its influence on the rheological
Kunieda, H., Ishikawa, N., 1985. Evaluation of the hydrophile-lipophile balance properties of polymer. Fuel 257, 116067. http://dx.doi.org/10.1016/j.fuel.
(HLB) of nonionic surfactants. II. Commercial-surfactant systems. J. Colloid 2019.116067.
Interface Sci. 107, 122–128. http://dx.doi.org/10.1016/0021-9797(85)90155- Makavipour, F., Pashley, R.M., Rahman, A.F.M.M., 2019. Low-level arsenic removal
9. from drinking water. Glob. Chall. 3, 1700047. http://dx.doi.org/10.1002/gch2.
Kunieda, H., Shinoda, K., 1985. Evaluation of the hydrophile-lipophile balance 201700047.
(HLB) of nonionic surfactants. I. Multisurfactant systems. J. Colloid Interface Malik, N.A., Ali, A., 2016. Krafft temperature and thermodynamic study of
Sci. 107, 107–121. http://dx.doi.org/10.1016/0021-9797(85)90154-7. interaction of glycine, diglycine, and triglycine with hexadecylpyridinium
Kurnia, I., Zhang, G., Han, X., Yu, J., 2020. Zwitterionic-anionic surfactant mixture chloride and hexadecylpyridinium bromide: A conductometric approach. J.
for chemical enhanced oil recovery without alkali. Fuel 259, 116236. http: Mol. Liq. 213, 213–220. http://dx.doi.org/10.1016/j.molliq.2015.11.017.
//dx.doi.org/10.1016/j.fuel.2019.116236. Maneedaeng, A., Flood, A.E., 2017. Synergisms in binary mixtures of anionic and
Kwaśniewska, D., Chen, Y.-L., Wieczorek, D., 2020. Biological activity of quater- pH-insensitive zwitterionic surfactants and their precipitation behavior with
nary ammonium salts and their derivatives. Pathogens http://dx.doi.org/10. calcium ions. J. Surfactants Deterg. 20, 263–275. http://dx.doi.org/10.1007/
3390/pathogens9060459. s11743-016-1902-z.
Leslie Zhang, D., Liu, S., Puerto, M., Miller, C.A., Hirasaki, G.J., 2006. Wettability Marszall, L., 1976. Adsorption of nonionic surfactants at the oil-water interface
alteration and spontaneous imbibition in oil-wet carbonate formations. J. and emulsion inversion point. Colloid Polym. Sci. 254, 674–675. http://dx.
Petrol. Sci. Eng. 52, 213–226. http://dx.doi.org/10.1016/j.petrol.2006.03.009. doi.org/10.1007/BF01753698.
Li, G., Chen, L., Ruan, Y., Guo, Q., Liao, X., Zhang, B., 2019. Alkyl polyglycoside: Marszall, L., 1978. Relationship among emulsion type, emulsion stability and the
A green and efficient surfactant for enhancing heavy oil recovery at high- presence of additives. Fette Seifen Anstrichm. 80, 289–293. http://dx.doi.org/
temperature and high-salinity condition. J. Pet. Explor. Prod. Technol. 9, 10.1002/lipi.19780800708.
2671–2680. http://dx.doi.org/10.1007/s13202-019-0658-1. Maurya, N.K., Kushwaha, P., Mandal, A., 2017. Studies on interfacial and rheolog-
Li, Y., Di, Q., Hua, S., Jia, X., 2020c. The effect of foam system containing ical properties of water soluble polymer grafted nanoparticle for application
surfactant and silica nanoparticles on oil recovery of carbonate rocks. Energy in enhanced oil recovery. J. Taiwan Inst. Chem. Eng. 70, 319–330. http:
Sources A 1–12. http://dx.doi.org/10.1080/15567036.2020.1748766. //dx.doi.org/10.1016/j.jtice.2016.10.021.

3174
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

McGuire, P.L., Chatham, J.R., Paskvan, F.K., Sommer, D.M., Carini, F.H., 2005. Low Nordiyana, M.S.W., Khalil, M., Jan, B.M., Ali, B.S., Tong, C.W., 2016. Formation and
salinity oil recovery: An exciting new EOR opportunity for Alaska’s North phase behavior of winsor type III Jatropha curcas-based microemulsion sys-
slope. In: SPE West. Reg. Meet. http://dx.doi.org/10.2118/93903-MS. tems. J. Surfactants Deterg. 19, 701–712. http://dx.doi.org/10.1007/s11743-
Mehranfar, A., Ghazanfari, M.H., 2014. Investigation of the microscopic displace- 016-1814-y.
ment mechanisms and macroscopic behavior of alkaline flooding at different Nourafkan, E., Asachi, M., Hu, Z., Gao, H., Wen, D., 2018. Synthesis of stable
wettability conditions in shaly glass micromodels. J. Petrol. Sci. Eng. 122, nanoparticles at harsh environment using the synergistic effect of surfactants
595–615. http://dx.doi.org/10.1016/j.petrol.2014.08.027. blend. J. Ind. Eng. Chem. 64, 390–401. http://dx.doi.org/10.1016/j.jiec.2018.
Miller, C., Bageri, B.S., Zeng, T., Patil, S., Mohanty, K.K., 2020. Modified 04.002.
two-phase titration methods to quantify surfactant concentrations in Nowrouzi, I., Mohammadi, A.H., Manshad, A.K., 2020. Water-oil interfacial ten-
chemical-enhanced oil recovery applications. J. Surfactants Deterg. http: sion (IFT) reduction and wettability alteration in surfactant flooding process
//dx.doi.org/10.1002/jsde.12442, n/a. using extracted saponin from Anabasis setifera plant. J. Petrol. Sci. Eng. 189,
Miller, R.G., Sorrell, S.R., 2014. The future of oil supply. Philos. Trans. R. Soc. 106901. http://dx.doi.org/10.1016/j.petrol.2019.106901.
Lond. Ser. A Math. Phys. Eng. Sci. 372, 20130179. http://dx.doi.org/10.1098/ Ojo, O.F., Farinmade, A., John, V., Nguyen, D., 2020. A nanocomposite of hal-
rsta.2013.0179. loysite/surfactant/wax to inhibit surfactant adsorption onto reservoir rock
Miyake, M., Oyama, N., 2009. Effect of amidoalkyl group as spacer on aggre- surfaces for improved oil recovery. Energy Fuels 34, 8074–8084. http://dx.
gation properties of guanidine-type surfactants. J. Colloid Interface Sci. 330, doi.org/10.1021/acs.energyfuels.0c00853.
180–185. http://dx.doi.org/10.1016/j.jcis.2008.10.047. Olayiwola, S.O., Dejam, M., 2019. A comprehensive review on interaction of
Mohammadshahi, H., Shahverdi, H., Mohammadi, M., 2020. Optimization of nanoparticles with low salinity water and surfactant for enhanced oil
dynamic interfacial tension for crude oil–brine system in the presence of recovery in sandstone and carbonate reservoirs. Fuel 241, 1045–1057. http:
nonionic surfactants. J. Surfactants Deterg. 23, 445–456. http://dx.doi.org/10. //dx.doi.org/10.1016/j.fuel.2018.12.122.
1002/jsde.12372. Omidi, A., Manshad, A.K., Moradi, S., Ali, J.A., Sajadi, S.M., Keshavarz, A.,
Mohapatra, S.S., Jha, J.M., Srinath, K., Pal, S.K., Chakraborty, S., 2014. Enhance- 2020. Smart- and nano-hybrid chemical EOR flooding using Fe3 O4 /eggshell
ment of cooling rate for a hot steel plate using air-atomized spray with nanocomposites. J. Mol. Liq. 316, 113880. http://dx.doi.org/10.1016/j.molliq.
surfactant-added water. Exp. Heat Transfer 27, 72–90. http://dx.doi.org/10. 2020.113880.
1080/08916152.2012.719068. Pal, S., Mushtaq, M., Banat, F., Al Sumaiti, A.M., 2018. Review of surfactant-
Mohsenatabar Firozjaii, A., Derakhshan, A., Shadizadeh, S.R., 2018. An inves- assisted chemical enhanced oil recovery for carbonate reservoirs: Challenges
tigation into surfactant flooding and alkaline-surfactant-polymer flooding and future perspectives. Pet. Sci. 15, 77–102. http://dx.doi.org/10.1007/
for enhancing oil recovery from carbonate reservoirs: Experimental study s12182-017-0198-6.
and simulation. Energy Sources A 40, 2974–2985. http://dx.doi.org/10.1080/
Pan, T., Wang, R., Xiao, K., Ye, W., Dong, W., Xu, M., 2019. Continuous degradation
15567036.2018.1514439.
of phenanthrene in cloud point system by reuse of Sphingomonas polyaro-
Mokheimer, E.M.A., Hamdy, M., Abubakar, Z., Shakeel, M.R., Habib, M.A., Mah- maticivorans cells. AMB Express 9, 8. http://dx.doi.org/10.1186/s13568-019-
moud, M., 2018. A comprehensive review of thermal enhanced oil recovery: 0736-2.
Techniques evaluation. J. Energy Resour. Technol. 141. http://dx.doi.org/10.
Pan, F., Zhang, Z., Zhang, X., Davarpanah, A., 2020. Impact of anionic and cationic
1115/1.4041096.
surfactants interfacial tension on the oil recovery enhancement. Powder
Moon, T.L., Blanzat, M., Labadie, L., Perez, E., Rico-Lattes, I., 2001. Synthesis
Technol. 373, 93–98. http://dx.doi.org/10.1016/j.powtec.2020.06.033.
and physicochemical study of new surfactants derived from carboxylic acid
Patel, H., Shah, S., Ahmed, R., Ucan, S., 2018. Effects of nanoparticles and
sugars. J. Dispers. Sci. Technol. 22, 167–176. http://dx.doi.org/10.1081/DIS-
temperature on heavy oil viscosity. J. Petrol. Sci. Eng. 167, 819–828. http:
100105202.
//dx.doi.org/10.1016/j.petrol.2018.04.069.
Mukhopadhyay, C., Suba, M., Sivakumar, D., Dhamodharan, K., Rao, R.V.S., 2019.
Peris-Vicente, J., Albiol-Chiva, J., Roca-Genovés, P., Esteve-Romero, J., 2016.
Cloud point extractive spectrophotometric method for determination of
Advances on melamine determination by micellar liquid chromatography:
uranium in raffinate streams during spent nuclear fuel reprocessing. J.
A review. J. Liq. Chromatogr. Relat. Technol. 39, 325–338. http://dx.doi.org/
Radioanal. Nucl. Chem. 322, 743–750. http://dx.doi.org/10.1007/s10967-019-
10.1080/10826076.2016.1152482.
06704-5.
Nagarajan, R., 2002. Molecular packing parameter and surfactant self-assembly: Phukan, R., Gogoi, S.B., Tiwari, P., 2020. Effects of CO2 -foam stability, interfacial
The neglected role of the surfactant tail. Langmuir 18, 31–38. http://dx.doi. tension and surfactant adsorption on oil recovery by alkaline-surfactant-
org/10.1021/la010831y. alternated-gas/CO2 flooding. Colloids Surf. A 597, 124799. http://dx.doi.org/
10.1016/j.colsurfa.2020.124799.
Nandwani, S.K., Malek, N.I., Lad, V.N., Chakraborty, M., Gupta, S., 2017. Study
on interfacial properties of imidazolium ionic liquids as surfactant and Pogaku, R., Mohd Fuat, N.H., Sakar, S., Cha, Z.W., Musa, N., Awang Tajudin, D.N.A.,
their application in enhanced oil recovery. Colloids Surf. A 516, 383–393. Morris, L.O., 2018. Polymer flooding and its combinations with other chemi-
http://dx.doi.org/10.1016/j.colsurfa.2016.12.037. cal injection methods in enhanced oil recovery. Polym. Bull. 75, 1753–1774.
Naseri, N., Ajorlou, E., Asghari, F., Pilehvar-Soltanahmadi, Y., 2018. An up- http://dx.doi.org/10.1007/s00289-017-2106-z.
date on nanoparticle-based contrast agents in medical imaging. Artif. Cells Puerto, M., Hirasaki, G.J., Miller, C.A., Barnes, J.R., 2012. Surfactant systems
Nanomed. Biotechnol. 46, 1111–1121. http://dx.doi.org/10.1080/21691401. for EOR in high-temperature, high-salinity environments. SPE J. 17, 11–19.
2017.1379014. http://dx.doi.org/10.2118/129675-PA.
Nazrul Islam, M., Sharker, K.K., Sarker, K.C., 2015. Salt-induced modulation of Puntervold, T., Mamonov, A., Aghaeifar, Z., Frafjord, G.O., Moldestad, G.M.,
the krafft temperature and critical micelle concentration of benzyldimethyl- Strand, S., Austad, T., 2018. Role of kaolinite clay minerals in enhanced
hexadecylammonium chloride. J. Surfactants Deterg. 18, 651–659. http://dx. oil recovery by low salinity water injection. Energy Fuels 32, 7374–7382.
doi.org/10.1007/s11743-015-1696-4. http://dx.doi.org/10.1021/acs.energyfuels.8b00790.
Negin, C., Ali, S., Xie, Q., 2017. Most common surfactants employed in chemical Qian, G., Chen, J., Luo, L., Yu, T., Wang, Y., Jiang, W., Xu, Q., Feng, S., Yin, S.,
enhanced oil recovery. Petroleum 3, 197–211. http://dx.doi.org/10.1016/j. 2020. Industrially promising nanowire heterostructure catalyst for enhancing
petlm.2016.11.007. overall water splitting at large current density. ACS Sustain. Chem. Eng. 8,
Nesměrák, K., Němcová, I., 2006. Determination of critical micelle concentration 12063–12071. http://dx.doi.org/10.1021/acssuschemeng.0c03263.
by electrochemical means. Anal. Lett. 39, 1023–1040. http://dx.doi.org/10. Rattanaudom, P., Shiau, B.-J., Suriyapraphadilok, U., Charoensaeng, A., 2019.
1080/00032710600620302. Aqueous foam stabilized by hydrophobic SiO2 nanoparticles using mixed
Nguyen, T.T.L., Edelen, A., Neighbors, B., Sabatini, D.A., 2010. Biocompatible anionic surfactant systems under high-salinity brine condition. J. Surfactants
lecithin-based microemulsions with rhamnolipid and sophorolipid biosur- Deterg. 22, 1247–1263. http://dx.doi.org/10.1002/jsde.12320.
factants: Formulation and potential applications. J. Colloid Interface Sci. 348, Reham, S.S., Masjuki, H.H., Kalam, M.A., Shancita, I., Rizwanul Fattah, I.M.,
498–504. http://dx.doi.org/10.1016/j.jcis.2010.04.053. Ruhul, A.M., 2015. Study on stability, fuel properties, engine combustion, per-
Nguyen, T.T., Morgan, C., Poindexter, L., Fernandez, J., 2019. Application of the formance and emission characteristics of biofuel emulsion. Renew. Sustain.
hydrophilic–lipophilic deviation concept to surfactant characterization and Energy Rev. 52, 1566–1579. http://dx.doi.org/10.1016/j.rser.2015.08.013.
surfactant selection for enhanced oil recovery. J. Surfactants Deterg. 22, Rezaei, A., Abdollahi, H., Derikvand, Z., Hemmati-Sarapardeh, A., Mosavi, A.,
983–999. http://dx.doi.org/10.1002/jsde.12305. Nabipour, N., 2020a. Insights into the effects of pore size distribution on
Nikolova, C., Gutierrez, T., 2020. Use of microorganisms in the recovery of oil the flowing behavior of carbonate rocks: Linking a nano-based enhanced
from recalcitrant oil reservoirs: Current state of knowledge, technological oil recovery method to rock typing. Nanomaterials http://dx.doi.org/10.3390/
advances and future perspectives. Front. Microbiol.. nano10050972.
Niu, J., Liu, Q., Lv, J., Peng, B., 2020. Review on microbial enhanced oil recovery: Rezaei, A., Riazi, M., Escrochi, M., Elhaei, R., 2020b. Integrating surfactant,
Mechanisms, modeling and field trials. J. Petrol. Sci. Eng. 192, 107350. alkali and nano-fluid flooding for enhanced oil recovery: A mechanistic
http://dx.doi.org/10.1016/j.petrol.2020.107350. experimental study of novel chemical combinations. J. Mol. Liq. 308, 113106.
Nollet, L.M.L., 2005. Chromatographic Analysis of the Environment. CRC Press. http://dx.doi.org/10.1016/j.molliq.2020.113106.

3175
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Rezaeian, M.S., Mousavi, S.M., Saljoughi, E., Akhlaghi Amiri, H.A., 2020. Evaluation Salleh, I.K., Misra, S., Ibrahim, J.M.B.M., Panuganti, S.R., 2019. Micro-emulsion-
of thin film composite membrane in production of ionically modified water based dissolver for removal of mixed scale deposition. J. Pet. Explor. Prod.
applied for enhanced oil recovery. Desalination 474, 114194. http://dx.doi. Technol. 9, 2635–2641. http://dx.doi.org/10.1007/s13202-019-0643-8.
org/10.1016/j.desal.2019.114194. Samimi, F., Sakhaei, Z., Riazi, M., 2020. Impact of pertinent parameters on foam
Rezk, M.Y., Allam, N.K., 2019. Unveiling the synergistic effect of ZnO nanoparti- behavior in the entrance region of porous media: Mathematical modeling.
cles and surfactant colloids for enhanced oil recovery. Colloid Interface Sci. Pet. Sci. http://dx.doi.org/10.1007/s12182-020-00465-7.
Commun. 29, 33–39. http://dx.doi.org/10.1016/j.colcom.2019.01.004. Saravanan, S., Keerthana, S., 2017. Floquet instability of gravity-modulated salt
Rezvani, H., Khalilnezhad, A., Ganji, P., Kazemzadeh, Y., 2018. How ZrO2 nanopar- fingering in a porous medium. Ind. Eng. Chem. Res. 56, 2851–2864. http:
ticles improve the oil recovery by affecting the interfacial phenomena in the //dx.doi.org/10.1021/acs.iecr.6b03866.
reservoir conditions? J. Mol. Liq. 252, 158–168. http://dx.doi.org/10.1016/j. Sarmah, S., Gogoi, S.B., Xianfeng, F., Baruah, A.A., 2020. Characterization and
molliq.2017.12.138. identification of the most appropriate nonionic surfactant for enhanced oil
Riahinezhad, M., Romero-Zerón, L., McManus, N., Penlidis, A., 2017. Evaluating recovery. J. Pet. Explor. Prod. Technol. 10, 115–123. http://dx.doi.org/10.1007/
the performance of tailor-made water-soluble copolymers for enhanced oil s13202-019-0682-1.
recovery polymer flooding applications. Fuel 203, 269–278. http://dx.doi.org/ Saxena, N., Kumar, A., Mandal, A., 2019. Adsorption analysis of natural anionic
10.1016/j.fuel.2017.04.122. surfactant for enhanced oil recovery: The role of mineralogy, salinity,
Riazi, M., Golkari, A., 2016. The influence of spreading coefficient on carbonated alkalinity and nanoparticles. J. Petrol. Sci. Eng. 173, 1264–1283. http://dx.
water alternating gas injection in a heavy crude oil. Fuel 178, 1–9. http: doi.org/10.1016/j.petrol.2018.11.002.
//dx.doi.org/10.1016/j.fuel.2016.03.021.
Schön, J.H., 2015. Chapter 2 - pore space properties. In: Schön, J.H.B.T.-D. (Ed.),
Rilian, N.A., Sumestry, M., Wahyuningsih, W., 2010. Surfactant stimulation to
Physical Properties of Rocks. Elsevier, pp. 21–84. http://dx.doi.org/10.1016/
increase reserves in carbonate reservoir a case study in semoga field. In:
B978-0-08-100404-3.00002-0.
SPE Eur. Annu. Conf. Exhib.. http://dx.doi.org/10.2118/130060-MS.
Sengupta, B., Sharma, V.P., Udayabhanu, G., 2012. A study of the effect of
Ríos, F., Fernández-Arteaga, A., Lechuga, M., Jurado, E., Fernández-Serrano, M.,
the concentration of constituents on the characteristics of a cross-linked
2016. Kinetic study of the anaerobic biodegradation of alkyl polyglucosides
polyacrylamide gel. Pet. Sci. Technol. 30, 1865–1881. http://dx.doi.org/10.
and the influence of their structural parameters. Environ. Sci. Pollut. Res. 23,
1080/10916466.2010.493907.
8286–8293. http://dx.doi.org/10.1007/s11356-016-6129-z.
Shafiai, S.H., Gohari, A., 2020. Conventional and electrical EOR review: The
Riswati, S.S., Bae, W., Park, C., Permadi, A.K., Efriza, I., Min, B., 2019. Experimental
development trend of ultrasonic application in EOR. J. Pet. Explor. Prod.
analysis to design optimum phase type and salinity gradient of alkaline
Technol. http://dx.doi.org/10.1007/s13202-020-00929-x.
surfactant polymer flooding at low saline reservoir. J. Petrol. Sci. Eng. 173,
1005–1019. http://dx.doi.org/10.1016/j.petrol.2018.09.087. Shahbazi, S., Goodpaster, J.V., Smith, G.D., Becker, T., Lewis, S.W., 2020. Prepa-
Romero-Zerón, L., Wadhwani, D., Sethiya, R., 2020. Formulation of an encap- ration, characterization, and application of a lipophilic coated exfoliated
sulated surfactant system, ESS, via β -CD host–guest interactions to inhibit egyptian blue for near-infrared luminescent latent fingermark detection.
surfactant adsorption onto solid surfaces. Ind. Eng. Chem. Res. http://dx.doi. Forensic Chem. 18, 100208. http://dx.doi.org/10.1016/j.forc.2019.100208.
org/10.1021/acs.iecr.0c01891. ShamsiJazeyi, H., Verduzco, R., Hirasaki, G.J., 2014. Reducing adsorption of
Rosen, M.J., Wang, H., Shen, P., Zhu, Y., 2005. Ultralow interfacial tension for anionic surfactant for enhanced oil recovery: Part II. Applied aspects. Colloids
enhanced oil recovery at very low surfactant concentrations. Langmuir 21, Surf. A 453, 168–175. http://dx.doi.org/10.1016/j.colsurfa.2014.02.021.
3749–3756. http://dx.doi.org/10.1021/la0400959. Sharma, T., Iglauer, S., Sangwai, J.S., 2016. Silica nanofluids in an oilfield polymer
Rosestolato, J.C.S., Pérez-Gramatges, A., Lachter, E.R., Nascimento, R.S.V., 2019. polyacrylamide: Interfacial properties, wettability alteration, and applications
Lipid nanostructures as surfactant carriers for enhanced oil recovery. Fuel for chemical enhanced oil recovery. Ind. Eng. Chem. Res. 55, 12387–12397.
239, 403–412. http://dx.doi.org/10.1016/j.fuel.2018.11.027. http://dx.doi.org/10.1021/acs.iecr.6b03299.
Roy, J.C., Islam, M.N., Aktaruzzaman, G., 2014. The effect of NaCl on the krafft Sharma, G., Mohanty, K., 2013. Wettability alteration in high-temperature and
temperature and related behavior of cetyltrimethylammonium bromide in high-salinity carbonate reservoirs. SPE J. 18, 646–655. http://dx.doi.org/10.
aqueous solution. J. Surfactants Deterg. 17, 231–242. http://dx.doi.org/10. 2118/147306-PA.
1007/s11743-013-1510-0. Sheng, J.J., 2010. Modern Chemical Enhanced Oil Recovery: theory and Practice.
Royer, M., Nollet, M., Catté, M., Collinet, M., Pierlot, C., 2018. Towards a new Gulf Professional Publishing.
universal way to describe the required hydrophilic lipophilic balance of oils Sheng, J.J., 2015. Status of surfactant EOR technology. Petroleum 1, 97–105.
using the phase inversion temperature of C10E4/n-octane/water emulsions. http://dx.doi.org/10.1016/j.petlm.2015.07.003.
Colloids Surf. A 536, 165–171. http://dx.doi.org/10.1016/j.colsurfa.2017.07. Shi, S., Wang, Y., Wang, L., Jin, Y., Wang, T., Wang, J., 2015. Potential of
024. spontaneous emulsification flooding for enhancing oil recovery in high-
Rudyk, S., Al-Khamisi, S., Al-Wahaibi, Y., Afzal, N., 2019. Internal olefin sulfonate temperature and high-salinity oil reservoir. J. Dispers. Sci. Technol. 36,
foam coreflooding in low-permeable limestone at varying salinity. Energy 660–669. http://dx.doi.org/10.1080/01932691.2014.905954.
Fuels 33, 8374–8382. http://dx.doi.org/10.1021/acs.energyfuels.9b01762. Shinoda, K., Hutchinson, E., 1962. Pseudo-phase separation model for thermo-
Sabahi, N., Razfar, M.R., Hajian, M., 2017. Experimental investigation of dynamic calculations on micellar solutions1. J. Phys. Chem. 66, 577–582.
surfactant-mixed electrolyte into electrochemical discharge machining http://dx.doi.org/10.1021/j100810a001.
(ECDM) process. J. Mater. Process. Technol. 250, 190–202. http://dx.doi.org/
Shinoda, K., Sagitani, H., 1978. Emulsifier selection in water/oil type emulsions by
10.1016/j.jmatprotec.2017.07.017.
the hydrophile—lipophile balance—temperature system. J. Colloid Interface
Salager, J.L., Antón, R.E., Briceño, M.I., Choplin, L., Márquez, L., Pizzino, A., Ro- Sci. 64, 68–71. http://dx.doi.org/10.1016/0021-9797(78)90335-1.
driguez, M.P., 2003. The emergence of formulation engineering in emulsion
Showell, M., 1997. Powdered Detergents. CRC Press.
making—transferring know-how from research laboratory to plant. Polym.
Soleimani Zohr Shiri, M., Henderson, W., Mucalo, M.R., 2019. A review of
Int. 52, 471–478. http://dx.doi.org/10.1002/pi.1112.
the lesser-studied microemulsion-based synthesis methodologies used for
Salager, J.-L., Antón, R.E., Sabatini, D.A., Harwell, J.H., Acosta, E.J., Tolosa, L.I., 2005.
preparing nanoparticle systems of the noble metals, Os, Re, Ir and Rh.
Enhancing solubilization in microemulsions—State of the art and current
Materials http://dx.doi.org/10.3390/ma12121896.
trends. J. Surfactants Deterg. 8, 3–21. http://dx.doi.org/10.1007/s11743-005-
0328-4. Song, Z.-J., Li, M., Zhao, C., Yang, Y.-L., Hou, J.-R., 2020. Gas injection for
Salager, J.-L., Forgiarini, A.M., Bullón, J., 2013. How to attain ultralow interfacial enhanced oil recovery in two-dimensional geology-based physical model of
tension and three-phase behavior with surfactant formulation for enhanced Tahe fractured-vuggy carbonate reservoirs: Karst fault system. Pet. Sci. 17,
oil recovery: A review. Part 1. Optimum formulation for simple surfactant– 419–433. http://dx.doi.org/10.1007/s12182-020-00427-z.
oil–water ternary systems. J. Surfactants Deterg. 16, 449–472. http://dx.doi. Song, B., Zhao, J., Wang, B., Jiang, R., 2009. Synthesis and self-assembly of new
org/10.1007/s11743-013-1470-4. light-sensitive Gemini surfactants containing an azobenzene group. Colloids
Salager, J.-L., Marquez, N., Graciaa, A., Lachaise, J., 2000. Partitioning of ethoxy- Surf. A 352, 24–30. http://dx.doi.org/10.1016/j.colsurfa.2009.09.044.
lated octylphenol surfactants in microemulsion-oil-water systems: influence Souraki, Y., Hosseini, E., Yaghodous, A., 2019. Wettability alteration of carbonate
of temperature and relation between partitioning coefficient and physico- reservoir rock using amphoteric and cationic surfactants: Experimental
chemical formulation. Langmuir 16, 5534–5539. http://dx.doi.org/10.1021/ investigation. Energy Sources A 41, 349–359. http://dx.doi.org/10.1080/
la9905517. 15567036.2018.1518353.
Salager, J.L., Morgan, J.C., Schechter, R.S., Wade, W.H., Vasquez, E., 1979. Optimum Southwick, J.G., van den Pol, E., van Rijn, C.H.T., van Batenburg, D.W., Boersma, D.,
formulation of surfactant/water/oil systems for minimum interfacial tension Svec, Y., Anis Mastan, A., Shahin, G., Raney, K., 2016. Ammonia as alkali
or phase behavior. Soc. Pet. Eng. J. 19, 107–115. http://dx.doi.org/10.2118/ for alkaline/surfactant/polymer floods. SPE J. 21, 10–21. http://dx.doi.org/10.
7054-PA. 2118/169057-PA.
Salehi, M., Johnson, S.J., Liang, J.-T., 2008. Mechanistic study of wettability al- Standnes, D.C., Austad, T., 2000. Wettability alteration in chalk: 2. mechanism for
teration using surfactants with applications in naturally fractured reservoirs. wettability alteration from oil-wet to water-wet using surfactants. J. Petrol.
Langmuir 24, 14099–14107. http://dx.doi.org/10.1021/la802464u. Sci. Eng. 28, 123–143. http://dx.doi.org/10.1016/S0920-4105(00)00084-X.

3176
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Standnes, D.C., Austad, T., 2003. Wettability alteration in carbonates: Interaction Wang, J., Xiao, L., Liao, G., Zhang, Y., Guo, L., Arns, C.H., Sun, Z., 2018. Theoretical
between cationic surfactant and carboxylates as a key factor in wettabil- investigation of heterogeneous wettability in porous media using NMR. Sci.
ity alteration from oil-wet to water-wet conditions. Colloids Surf. A 216, Rep. 8, 13450. http://dx.doi.org/10.1038/s41598-018-31803-w.
243–259. http://dx.doi.org/10.1016/S0927-7757(02)00580-0. Wei, P., Li, J., Xie, Y., Huang, X., Sun, L., 2020. Alkyl polyglucosides for potential
Suleimanov, B.A., Ismailov, F.S., Veliyev, E.F., 2011. Nanofluid for enhanced oil application in oil recovery process: Adsorption behavior in sandstones under
recovery. J. Petrol. Sci. Eng. 78, 431–437. http://dx.doi.org/10.1016/j.petrol. high temperature and salinity. J. Petrol. Sci. Eng. 189, 107057. http://dx.doi.
2011.06.014. org/10.1016/j.petrol.2020.107057.
Suleymani, M., Ashoori, S., Ghotbi, C., Moghadasi, J., Kharrat, R., 2020. Static and Weiss, W.W., Xie, X., Weiss, J., Subramanian, V., Taylor, A.R., Edens, F.J., 2006.
dynamic behavior of foam stabilized by modified nanoparticles: Theoretical Artificial intelligence used to evaluate 23 single-well surfactant-soak treat-
and experimental aspects. Chem. Eng. Res. Des. 158, 114–128. http://dx.doi. ments. SPE Reserv. Eval. Eng. 9, 209–216. http://dx.doi.org/10.2118/89457-
org/10.1016/j.cherd.2020.04.003. PA.
Sun, C., Guo, H., Li, Y., Jiang, G., Ma, R., 2020. Alkali effect on alkali-surfactant- Winsor, P.A., 1948. Hydrotropy, solubilisation and related emulsification
polymer (ASP) flooding enhanced oil recovery performance: Two large-scale processes. Trans. Faraday Soc. 44, 376–398. http://dx.doi.org/10.1039/
field tests’ evidence. J. Chem. 2020, 2829565. http://dx.doi.org/10.1155/2020/ TF9484400376.
2829565. Winsor, P.A., 1954. Solvent Properties of Amphiphilic Compounds. Butterworths
Suzuki, H., 1970. Determination of critical micelle concentration of surfactant Scientific Publications.
Witthayapanyanon, A., Harwell, J.H., Sabatini, D.A., 2008. Hydrophilic–lipophilic
by ultraviolet absorption spectra. J. Am. Oil Chem. Soc. 47, 273–277. http:
deviation (HLD) method for characterizing conventional and extended sur-
//dx.doi.org/10.1007/BF02609491.
factants. J. Colloid Interface Sci. 325, 259–266. http://dx.doi.org/10.1016/j.jcis.
Syaifudin, A., 2002. Reservoir Simulation to Increase Recovery in Kaji Semoga
2008.05.061.
Field.
Wołowicz, A., Staszak, K., 2020. Study of surface properties of aqueous solutions
Tackie-Otoo, B.N., Ayoub Mohammed, M.A., Yekeen, N., Negash, B.M., 2020. Al-
of sodium dodecyl sulfate in the presence of hydrochloric acid and heavy
ternative chemical agents for alkalis, surfactants and polymers for enhanced
metal ions. J. Mol. Liq. 299, 112170. http://dx.doi.org/10.1016/j.molliq.2019.
oil recovery: Research trend and prospects. J. Petrol. Sci. Eng. 187, 106828.
112170.
http://dx.doi.org/10.1016/j.petrol.2019.106828.
Wu, Y., Chen, W., Dai, C., Huang, Y., Li, H., Zhao, M., He, L., Jiao, B., 2017.
Tadros, T.F., 2014. An Introduction to Surfactants. Walter de Gruyter. Reducing surfactant adsorption on rock by silica nanoparticles for enhanced
Tagavifar, M., Jang, S.H., Sharma, H., Wang, D., Chang, L.Y., Mohanty, K., oil recovery. J. Petrol. Sci. Eng. 153, 283–287. http://dx.doi.org/10.1016/j.
Pope, G.A., 2018. Effect of pH on adsorption of anionic surfactants on petrol.2017.04.015.
limestone: Experimental study and surface complexation modeling. Colloids Xia, Y., Zhou, J.-J., Gong, Y.-Y., Li, Z.-J., Zeng, E.Y., 2020. Strong influence of
Surf. A 538, 549–558. http://dx.doi.org/10.1016/j.colsurfa.2017.11.050. surfactants on virgin hydrophobic microplastics adsorbing ionic organic
Tayari, F., Blumsack, S., Johns, R.T., Tham, S., Ghosh, S., 2018. Techno-economic pollutants. Environ. Pollut. 265, 115061. http://dx.doi.org/10.1016/j.envpol.
assessment of reservoir heterogeneity and permeability variation on eco- 2020.115061.
nomic value of enhanced oil recovery by gas and foam flooding. J. Petrol. Xie, X., Weiss, W.W., Tong, Z.J., Morrow, N.R., 2005. Improved oil recovery
Sci. Eng. 166, 913–923. http://dx.doi.org/10.1016/j.petrol.2018.03.053. from carbonate reservoirs by chemical stimulation. SPE J. 10, 276–285.
Teklu, T.W., Li, X., Zhou, Z., Alharthy, N., Wang, L., Abass, H., 2018. Low-salinity http://dx.doi.org/10.2118/89424-PA.
water and surfactants for hydraulic fracturing and EOR of shales. J. Petrol. Xu, Z., Li, B., Zhao, H., He, L., Liu, Z., Chen, D., Yang, H., Li, Z., 2020. Investi-
Sci. Eng. 162, 367–377. http://dx.doi.org/10.1016/j.petrol.2017.12.057. gation of the effect of nanoparticle-stabilized foam on EOR: Nitrogen foam
Telmadarreie, A., Trivedi, J.J., 2016. New insight on carbonate-heavy-oil recovery: and methane foam. ACS Omega 5, 19092–19103. http://dx.doi.org/10.1021/
Pore-scale mechanisms of post-solvent carbon dioxide foam/polymer- acsomega.0c02434.
enhanced-foam flooding. SPE J. 21, 1655–1668. http://dx.doi.org/10.2118/ Xu, J., Zhang, Y., Chen, H., Wang, P., Xie, Z., Yao, Y., Yan, Y., Zhang, J., 2013.
174510-PA. Effect of surfactant headgroups on the oil/water interface: An interfacial
Thiruvengadam, S., Murphy, M., Tan, J.S., Miller, K., 2020. A generalized the- tension measurement and simulation study. J. Mol. Struct. 1052, 50–56.
oretical model for the relationship between critical micelle concentrations, http://dx.doi.org/10.1016/j.molstruc.2013.07.049.
pressure, and temperature for surfactants. J. Surfactants Deterg. 23, 273–303. Yan, L., Ding, W., 2019. Flooding performance evaluation of alkyl aryl sulfonate
http://dx.doi.org/10.1002/jsde.12360. in various alkaline environments. PLoS One 14, e0219627.
Tichelkamp, T., Vu, Y., Nourani, M., Øye, G., 2014. Interfacial tension between low Yang, J., Zhou, Y., 2020. An automatic in situ contact angle determination
salinity solutions of sulfonate surfactants and crude and model oils. Energy based on level set method. Water Resour. Res. 56, http://dx.doi.org/10.1029/
Fuels 28, 2408–2414. http://dx.doi.org/10.1021/ef4024959. 2020WR027107, e2020WR027107.
Trickett, K., Xing, D., Enick, R., Eastoe, J., Hollamby, M.J., Mutch, K.J., Rogers, S.E., Yekeen, N., Manan, M.A., Idris, A.K., Samin, A.M., 2017a. Influence of surfactant
Heenan, R.K., Steytler, D.C., 2010. Rod-like micelles thicken CO2 . Langmuir and electrolyte concentrations on surfactant adsorption and foaming charac-
26, 83–88. http://dx.doi.org/10.1021/la902128g. teristics. J. Petrol. Sci. Eng. 149, 612–622. http://dx.doi.org/10.1016/j.petrol.
Turan, K., Kaur, P., Manhas, D., Sharma, J., Verma, G., 2020. Novel insights into the 2016.11.018.
Yekeen, N., Manan, M.A., Idris, A.K., Samin, A.M., Risal, A.R., 2017b. Experimental
dispersed and acid-mediated surface modification of the carbon nanofibers.
investigation of minimization in surfactant adsorption and improvement in
Mater. Chem. Phys. 239, 121978. http://dx.doi.org/10.1016/j.matchemphys.
surfactant-foam stability in presence of silicon dioxide and aluminum oxide
2019.121978.
nanoparticles. J. Petrol. Sci. Eng. 159, 115–134. http://dx.doi.org/10.1016/j.
Türksoy, U., Bağci, S., 2000. Improved oil recovery using alkaline solutions in
petrol.2017.09.021.
limestone medium. J. Petrol. Sci. Eng. 26, 105–119. http://dx.doi.org/10.1016/
Yekeen, N., Padmanabhan, E., Idris, A.K., Ibad, S.M., 2019. Surfactant adsorption
S0920-4105(00)00025-5.
behaviors onto shale from malaysian formations: Influence of silicon dioxide
do Vale, T.O., de Magalhães, R.S., de Almeida, P.F., Matos, J.B.T.L., Chinalia, F.A.,
nanoparticles, surfactant type, temperature, salinity and shale lithology. J.
2020. The impact of alkyl polyglycoside surfactant on oil yields and its
Petrol. Sci. Eng. 179, 841–854. http://dx.doi.org/10.1016/j.petrol.2019.04.096.
potential effect on the biogenic souring during enhanced oil recovery (EOR).
Yıldız, D., Demir, M., 2019. Flame atomic absorption determination of copper in
Fuel 280, 118512. http://dx.doi.org/10.1016/j.fuel.2020.118512.
environmental water with cloud point extraction using isonitrosoacetophe-
Vargo, J., Turner, J., Vergnani, B., Pitts, M.J., Wyatt, K., Surkalo, H., Patterson, D., none 2-aminobenzoylhydrazone. J. Anal. Chem. 74, 437–443. http://dx.doi.
1999. Alkaline-surfactant-polymer flooding of the cambridge minnelusa field. org/10.1134/S1061934819050022.
In: SPE Rocky Mt. Reg. Meet. http://dx.doi.org/10.2118/55633-MS. Yin, D., Pu, H., 2008. A numerical simulation study on surfactant flooding and
Venancio, J.C.C., Nascimento, R.S.V., Pérez-Gramatges, A., 2020. Colloidal stability it’s field application in daqing oilfield. In: Eur. Conf. Exhib... http://dx.doi.
and dynamic adsorption behavior of nanofluids containing alkyl-modified org/10.2118/112424-MS.
silica nanoparticles and anionic surfactant. J. Mol. Liq. 308, 113079. http: You, Q., Wang, H., Zhang, Y., Liu, Y., Fang, J., Dai, C., 2018. Experimental study
//dx.doi.org/10.1016/j.molliq.2020.113079. on spontaneous imbibition of recycled fracturing flow-back fluid to enhance
Volokitin, Y., Shuster, M., Karpan, V., Koltsov, I., Mikhaylenko, E., Bondar, M., oil recovery in low permeability sandstone reservoirs. J. Petrol. Sci. Eng. 166,
Podberezhny, M., Rakitin, A., van Batenburg, D.W., Parker, A.R., de Kruijf, S., 375–380. http://dx.doi.org/10.1016/j.petrol.2018.03.058.
Southwick, J.G., de Reus, J., van den Pol, E., van der Heyden, F.H.J., Boels, L., Yousefvand, H.A., Jafari, A., 2018. Stability and flooding analysis of nanosil-
Wever, D.A.Z., Brewer, M., 2018. Results of alkaline-surfactant-polymer ica/NaC/HPAM/SDS solution for enhanced heavy oil recovery. J. Petrol. Sci.
flooding pilot at West salym field. In: SPE EOR Conf. Oil Gas West Asia. Eng. 162, 283–291. http://dx.doi.org/10.1016/j.petrol.2017.09.078.
http://dx.doi.org/10.2118/190382-MS. Yu, H., Lu, X., Fu, W., Wang, Y., Xu, H., Xie, Q., Qu, X., Lu, J., 2020. Determination
Wang, S., Chen, C., Yuan, N., Ma, Y., Ogbonnaya, O.I., Shiau, B., Harwell, J.H., of minimum near miscible pressure region during CO2 and associated gas
2019. Design of extended surfactant-only EOR formulations for an ultrahigh injection for tight oil reservoir in Ordos Basin. China Fuel 263, 116737.
salinity oil field by using hydrophilic lipophilic deviation (HLD) approach: http://dx.doi.org/10.1016/j.fuel.2019.116737.
From laboratory screening to simulation. Fuel 254, 115698. http://dx.doi.org/ van Zanten, R., 2011. Stabilizing viscoelastic surfactants in high-density brines.
10.1016/j.fuel.2019.115698. SPE Drill. Complet 26, 499–505. http://dx.doi.org/10.2118/141447-PA.

3177
O. Massarweh and A.S. Abushaikha Energy Reports 6 (2020) 3150–3178

Zapf, A., Beck, R., Platz, G., Hoffmann, H., 2003. Calcium surfactants: A re- Zhao, J., Dai, C., Ding, Q., Du, M., Feng, H., Wei, Z., Chen, A., Zhao, M., 2015.
view. Adv. Colloid Interface Sci. 100–102, 349–380. http://dx.doi.org/10.1016/ The structure effect on the surface and interfacial properties of zwitterionic
S0001-8686(02)00065-9. sulfobetaine surfactants for enhanced oil recovery. RSC Adv. 5, 13993–14001.
Zargar, G., Arabpour, T., Khaksar Manshad, A., Ali, J.A., Mohammad Sajadi, S., http://dx.doi.org/10.1039/C4RA16235H.
Keshavarz, A., Mohammadi, A.H., 2020. Experimental investigation of the Zhao, J., Torabi, F., 2020. Experimental investigation and modelling of CO2 -
effect of green TiO2 /quartz nanocomposite on interfacial tension reduction, foam flow in heavy oil systems. Can. J. Chem. Eng. 98, 147–157. http:
wettability alteration, and oil recovery improvement. Fuel 263, 116599. //dx.doi.org/10.1002/cjce.23573.
http://dx.doi.org/10.1016/j.fuel.2019.116599. Zhong, X., Li, C., Li, Y., Pu, H., Zhou, Y., Zhao, J.X., 2020. Enhanced oil recovery
Zargartalebi, M., Barati, N., Kharrat, R., 2014. Influences of hydrophilic and in high salinity and elevated temperature conditions with a zwitterionic
hydrophobic silica nanoparticles on anionic surfactant properties: Interfacial surfactant and silica nanoparticles acting in synergy. Energy Fuels 34,
and adsorption behaviors. J. Petrol. Sci. Eng. 119, 36–43. http://dx.doi.org/10. 2893–2902. http://dx.doi.org/10.1021/acs.energyfuels.9b04067.
1016/j.petrol.2014.04.010. Zhong, X., Pu, H., Zhou, Y., Zhao, J.X., 2019. Comparative study on the static
Zhang, J., Gao, H., Xue, Q., 2020. Potential applications of microbial enhanced oil adsorption behavior of zwitterionic surfactants on minerals in middle
recovery to heavy oil. Crit. Rev. Biotechnol. 40, 459–474. http://dx.doi.org/ Bakken formation. Energy Fuels 33, 1007–1015. http://dx.doi.org/10.1021/
10.1080/07388551.2020.1739618. acs.energyfuels.8b04013.
Zhang, L., Jian, G., Puerto, M., Southwick, E., Hirasaki, G., Biswal, S.L., 2019a. Zhu, D., Hou, J., Wang, J., Wu, X., Wang, P., Bai, B., 2018. Acid-alternating-
Static adsorption of a switchable diamine surfactant on natural and synthetic base (AAB) technology for blockage removal and enhanced oil recovery
minerals for high-salinity carbonate reservoirs. Colloids Surf. A 583, 123910. in sandstone reservoirs. Fuel 215, 619–630. http://dx.doi.org/10.1016/j.fuel.
http://dx.doi.org/10.1016/j.colsurfa.2019.123910. 2017.11.090.
Zhang, H., Nikolov, A., Wasan, D., 2014. Enhanced oil recovery (EOR) using Zulkifli, N.N., Mahmood, S.M., Akbari, S., Manap, A.A.A., Kechut, N.I., Elrais, K.A.,
nanoparticle dispersions: Underlying mechanism and imbibition experi- 2020. Evaluation of new surfactants for enhanced oil recovery applications
ments. Energy Fuels 28, 3002–3009. http://dx.doi.org/10.1021/ef500272r. in high-temperature reservoirs. J. Pet. Explor. Prod. Technol. 10, 283–296.
Zhang, R., Somasundaran, P., 2006. Advances in adsorption of surfactants http://dx.doi.org/10.1007/s13202-019-0713-y.
and their mixtures at solid/solution interfaces. Adv. Colloid Interface Sci.
123–126, 213–229. http://dx.doi.org/10.1016/j.cis.2006.07.004.
Zhang, Y., Sun, H.-Y., Bai, X., Li, Y., Zhang, J., Zhao, M., Huang, X., Feng, C.-
Y., Zhao, Y., 2019b. Exfoliation of layered double hydroxides by use of
zwitterionic surfactants in aqueous solution. J. Dispers. Sci. Technol. 40,
811–818. http://dx.doi.org/10.1080/01932691.2018.1484293.

3178

You might also like