You are on page 1of 37

Journal of Petroleum Science and Engineering 220 (2023) 111243

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Manifestations of surfactant-polymer flooding for successful field


applications in carbonates under harsh conditions: A comprehensive review
Anas M. Hassan a, *, Emad W. Al-Shalabi a, Waleed Alameri a, Muhammad Shahzad Kamal b,
Shirish Patil b, Syed Muhammad Shakil Hussain b
a
Petroleum Engineering Department, Khalifa University of Science and Technology, Abu Dhabi, United Arab Emirates
b
College of Petroleum Engineering and Geosciences, King Fahad University of Petroleum and Minerals, Dhahran, Saudi Arabia

A R T I C L E I N F O A B S T R A C T

Keywords: Most oil fields today are mature, and the majority of the reservoirs in the Middle East are carbonate rocks
Enhanced Oil Recovery (EOR) characterized by high temperature high salinity (HTHS), heterogeneous mineral composition, and natural
Chemical Enhanced Oil Recovery (cEOR) fractures. Enhanced oil recovery (EOR) methods are used for boosting oil recovery from the aged reservoirs
Surfactant-Polymer (SP) Flooding
beyond primary and secondary recovery stages. Nevertheless, it can be a challenging task to employ EOR
High-temperature and High -Salinity (HTHS)
techniques in these aged carbonate reservoirs. This is because carbonate reservoirs have mixed-to-oil-wet
carbonate reservoirs
wettability with temperatures exceeding 85 ◦ C and salinity of over 100,000 ppm, which renders secondary
EOR-methods such as waterflooding ineffective. Therefore, even though carbonate reservoirs contain 60–65% of
world’s remaining oil, with immense intrinsic economic prospects, the oil recovery process from carbonate
reservoirs remains a considerable challenge. Chemical-EOR (cEOR) techniques, like SP based cEOR, have shown
marked promise in improved oil recovery, mainly from clastic reservoirs with medium temperature and salinity,
unlike carbonate reservoirs. During SP-floodings, the surfactants get adsorbed due to the mineral composition of
the carbonate rocks, and polymer degradation occurs due to HTHS conditions. Consequently, new surfactants
and polymers have been structurally conformed and tested to improve their robustness and related recovery
efficacy. For instance, Guerbet alkoxy-carboxylate surfactants demonstrated good stability at temperatures over
100 ◦ C and salinities up-to 275,000 ppm, yielding tertiary recovery of 94.5% and low adsorption of 0.086 mg/g
of rock. The cationic Gemini surfactants, zwitterionic or amphoteric class of surfactants are also suitable for
HTHS carbonates. Besides, effective biosurfactants sourced from plant such as, soy, corn, etc., are non-toxic and
readily biodegradable. The hydrophobically associating polyacrylamide (HAPAM) and its modified nano­
composite derivative can act as replacement surfactants, due to their wettability altering and robust charac­
teristics. Novel polymers viz., NVP-based, novel smart thermoviscosifying polymers (TVP), soft microgel, and
sulfonated polymers, are also relevant to HTHS carbonate applications. Xanthan gum, scleroglucan, and schiz­
ophyllan biopolymers have been noted to resist HTHS and low permeability conditions, requiring lower con­
centration and having low adsorption. Recent surfactant-polymer (SP) formulations also can be applicable for
HTHS carbonates with excellent ternary recoveries (93.6%) and minimal retention (0.083 μg/g of rock). Such
low retention values suggest low surfactants cost with minimal environmental impact. Moreover, several suc­
cessful field applications in carbonates were conducted in preceding years; however, the performance of some
novel surfactants under HTHS carbonates is yet to be fully understood. Hence, this comprehensive review aims to
provide renewed perspectives on surfactant and polymer optimizations for field applications in HTHS carbonates.
A list of recommendations is presented as guidelines for efficient SP-flooding designs. This critical literature
appraisal furnishes an array of potential manifestations for successful field application of SP-flooding in HTHS
carbonates, which holds both economic and environmental feasibility.

* Corresponding author.
E-mail address: anas.hassan@ku.ac.ae (A.M. Hassan).

https://doi.org/10.1016/j.petrol.2022.111243
Received 16 March 2022; Received in revised form 17 October 2022; Accepted 10 November 2022
Available online 17 November 2022
0920-4105/© 2022 Elsevier B.V. All rights reserved.
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

1. Introduction The SP-flooding process has primarily been used in reservoirs with
mild salinity (60,000 ppm) and temperature under (80 ◦ C) (Green and
Over half of the oil reserves in the world are in carbonate reservoirs Willhite, 1998; Lake et al., 2014). However, as oilfields age, carbonate
(Akbar et al., 2000). Most of these carbonates present harsh conditions i. reservoirs are becoming increasingly common and significant. The
e., high temperature and high salinity (HTHS) along with heterogeneous SP-flooding is challenging in carbonate reservoirs because of the com­
mineralogy and vuggy structure. Fig. 1 demonstrates an example of a plex geological structure (i.e., heterogeneity) and harsh conditions (i.e.,
highly heterogeneous and vuggy carbonate reservoir core sample (Lu HTHS), that cause instability, degradation, and chemical adsorption of
et al., 2014a). A sizeable number of these carbonate reservoirs have the injected SP-fluids on the rock’s surface. Moreover, the imple­
natural fractures while being mixed-wet to oil-wet (Chilingar and Yen, mentation of SP-based cEOR is severely impacted by high adsorption of
1983; Roehl and Choquette, 2012). Previous studies have indicated that surfactants (e.g., anionic alkylbenzene sulfonates) on carbonates rocks
approximately 70% of global oil reserves in carbonates are not or clay surfaces due to the electrostatic interactions. This major issue of
extractable through conventional oil recovery methods, e.g., water­ surfactant adsorption on rocks or clay surfaces necessitates constant
flooding (Lake et al., 2014; Hartono et al., 2017; Moreno-Arciniegas and surfactant re-injections, downgrading the efficiency and economic
Babadagli, 2017). The conventional recovery methods (i.e., water­ viability (Adams and Schievelbein, 1987) of the SP-based cEOR process.
flooding) become ineffective in incremental oil recovery, owing to the Surfactant adsorption is mainly influenced by the surface chemistry of
oil-wet condition of the carbonate reservoir and the strong adsorption of most of the carbonate rocks. A complex dissolution behavior is exhibited
crude oil on the carbonate rocks. Consequently, enhanced oil recovery by some minerals contained in carbonate rocks, viz., dolomite (CaMg
(EOR) techniques are powerful tools to unlock a significant amount of (CO3)2), calcite (CaCO3), and magnesite (MgCO3) (Hiorth et al., 2010).
residual and remaining oils from the existing hydrocarbon carbonate To mitigate this surfactant adsorption problem, adding co-solvents
reservoirs through obtaining improved oil recovery beyond water cut together with mixtures of different surfactants structures can reduce
levels at an economical production rate. microemulsion viscosity and lower surfactant retention (Tagavifar et al.,
In chemical EOR (cEOR) techniques like surfactant and polymer- 2018). Another cause of retention is phase entrapment, which is mini­
based EOR (i.e., SP-flooding), the combined effects (i.e., synergistic ef­ mized through applying negative-salinity tong, i.e., when the polymer
fect) of surfactant and polymer can (i) change the wettability of the oil- drive of lower salinity (hybrid low salinity waterflooding) than the
wet carbonate rock surface toward more water-wet conditions, (ii) injected surfactant slug’s salinity is applied (Hassan et al., 2022c).
reduce the interfacial tension (IFT) between crude oil and water, and Therefore, as previously mentioned, to select the most appropriate
(iii) enhance the mobility control of the injected fluid (Rosen and cEOR method, proper screening of the chemicals (e.g., surfactants and
Kunjappu, 2012; Olajire, 2014; Ragab and Mansour, 2021). The average polymers) must be conducted for the reservoir conditions to increase
incremental oil recovery achieved has been over 15% OOIP (original oil chemical stability as well as reduce chemical adsorption and degrada­
in place) in all the known field tests so far; however, there is very little tion. Consequently, many studies have developed a number of potential
knowledge of HTHS carbonate field applications (Liao et al., 2017). surfactant and polymer recipes, which may lead to successful field ap­
The surfactants are typically amphiphilic organic molecules that plications of SP-flooding in carbonates in the near future. These studies
contain both hydrophobic (tails) and hydrophilic groups (heads). have focused on determining the best performing surfactant structure
Consequently, a surfactant molecule has both a water-soluble compo­ relationships with certain crude oils and brines at various temperatures
nent and an oil-soluble component, as shown in Fig. 2. Fig. 2a illustrates and salinity ranges (e.g., HTHS conditions) (Skauge and Palmgren,
the surfactants adsorbed between air and water or between oil and water 1989; Levitt et al., 2006; Yang et al., 2010; Solairaj et al., 2012; Adkins
interfaces i.e., the oil-soluble clusters protrude out of the water into the et al., 2012; Lu et al., 2014b, 2014c). During one of these studies on the
air or into the oil-phase, while Fig. 2b shows the water-soluble head conformation of surfactants and their effects, it was observed that a
clusters that remain in the water-phase (El-Hoshoudy et al., 2017a). longer hydrophobic tail increases the ability to solubilize oil, i.e., higher
During surfactant and polymer floodings in HTHS carbonates, the sur­ oil solubilization ratios are achieved along with lower optimum salin­
factants get adsorbed due to the mineral composition of the carbonate ities at certain temperatures (Lu et al., 2014b). However, excess length
rocks and gels may form in the micellar solution along with other forms of hydrophobe reduces the aqueous solubility. It was also found that the
of polymer degradation (Reed and Healy, 1977; Skauge and Palmgren, branching of bulky groups like benzene rings on the hydrophobe and
1989; Kovscek and Radke, 1993; Hirasaki et al., 2011; Lake et al., 2014). other heavy clusters in the structure impart polarity to the surfactant,
Therefore, chemical screening of surfactants and polymers for suitability enhancing the water solubility (Skauge and Palmgren, 1989).
and stability is a prerequisite for implementing surfactant and polymer Another highly desirable feature is creating low viscosity micro­
flooding projects. Accordingly, some studies have developed emulsions that equilibrate at a rapid rate while avoiding the formation
custom-designed i.e., an optimized surfactant for the best possible of gels or macroemulsions. To adjust the structure, propylene oxide (PO)
interactive relations with certain oils and brines at various and ethylene oxide (EO) chains are incorporated in the hydrophobe,
temperatures. which modifies the surfactant’s hydrophilic-to-lipophilic balance (HLB)

Fig. 1. A computerized tomography (CT) scan image of a highly heterogeneous and vuggy carbonate core sample (Lu et al., 2014a).

2
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Fig. 2. A schematic of a surfactant structure where (a) shows the hydrophilic head with water affinity, and (b) the hydrophobic tail, which is oil-liking or oil-wet
(modified after El-Hoshoudy et al., 2017a).

(Skauge and Palmgren, 1989). For instance, the new Alkyl carboxylates performance in carbonates is still obscure. A pilot study was conducted
(AEC) have been observed to have good resistance to hardness, aqueous using sulfonated polymer (ATBS >35 mol%) for possible use in a
solubility, and tolerance to pH alterations or changes in electrolyte gigantic carbonate field in the Middle East with 5% incremental oil re­
levels (Behler et al., 2001; Rosen and Kunjappu, 2012). These AEC covery outcome (Jabbar et al., 2018). Recently, two polymer-based EOR
surfactants can be utilized in SP-based EOR, which could reduce costs as pilot studies have been undertaken by ADNOC in two reservoirs, which
these AEC surfactants having large-hydrophobe were synthesized that are currently undergoing waterflooding (Masalmeh et al., 2019). Useful
may be used in lesser quantities to attain the required parameters. correlations were derived through analyzing large data sets, which
Guerbet alkoxy sulfates (GAS), Guerbet alkoxy carboxylates (GAC), predict the best surfactant structure by considering the impact of
Guerbet alkoxy betaine (GAB) and Tristyrylphenol alkoxy sulfates (TSP) changeable factors e.g., the propylene oxide (PO) and ethylene oxide
are different types of large-hydrophobe novel AEC surfactants. Several (EO) numbers, EACN of the oil, temperature, and the brine salinity (Lu
betaine surfactants were analyzed for SP-based cEOR due to their et al., 2014c). These factors had been neglected previously but have now
well-documented resistance to salt/divalent ion under HTHS reservoir been deemed necessary in perfecting these correlations. Therefore, the
conditions. The novel surfactants with large-hydrophobe such as Guer­ field applications and small-scale pilots conducted under HTHS car­
bet alkoxy-sulfates (GAS), -carboxylates (GAC), and -betaine (GAB), bonate conditions can be expected to have overall more accurate results
have exhibited stability at high temperatures (>100 ◦ C) and high sa­ using these models.
linities (up to 275,000 ppm). When corefloods were conducted with Hence, this work is a comprehensive literature review, on the po­
carbonate cores, the surfactant formulations yielded high tertiary oil tential scope of SP-flooding in carbonate reservoirs under harsh HTHS
recovery of 94.5% (GAC formulation), and low adsorption of 0.086 conditions. Additionally, environmental impacts and economic viability
mg/g rock (Lu et al., 2014a). The dynamic polymeric surfactants (sur­ are discussed. The recommendations can help enhance the perspectives
face-active monomers) are another viable surfactant type, which is of key parameters for designing successful SP based cEOR imple­
structurally amphiphilic i.e., they are hydrophobic and hydrophilic at mentation in harsh carbonate conditions. To the best of our knowledge,
the extremities (Gao et al., 2008). Along with this, their molecular for­ this contribution presents one of the few review articles existing in the
mations contain binary vinyl bonds, which can be polymerized (Reb field of surfactant polymer flooding in carbonate reservoirs under harsh
et al., 2000). This renders them more adaptive than conventional sur­ HTHS conditions. An overview of the existing review articles, which
factants and may be used for emulsification (Guyot, 2004). In addition, cover the topic of SP-flooding is shown in Appendix-A Table A-1. The
several biodegradable and non-toxic surfactants and polymers have latter table summarizes the chemicals (i.e., surfactant and polymer
been designed that will offer a greener approach to chemical EOR formulations) used in those studies, their HTHS tolerance, advantages,
(Solairaj et al., 2012). and limitations.
On the other hand, over the years several screening criteria for
polymers have been provided (Taber, 1983; Goodlett et al., 1986; Taber 2. Mechanisms of SP-flooding in HTHS carbonate reservoirs
et al., 1997; Wang and Lu, 2009; Gao, 2011; Kang et al., 2014). Novel
polymers such as ATBS, Scleroglucan, NVP-based polymers, hydropho­ According to previously published studies, the SP-floodings have
bic associative polymers, novel smart Thermoviscosifying polymers been found to possess vast potential in cEOR applications (Cockcroft
(TVP) (salt-induced viscosifiers), soft microgel polymers (SMG), and et al., 1988; Bera et al., 2011; Rosen and Kunjappu, 2012; Olajire, 2014;
sulfonated polymers have been synthesized and tested and exhibited Karović-Maričić et al., 2014; Sheng, 2010, 2015; Mandal, 2015; Liao
good tolerance and stability under HTHS carbonate conditions in the et al., 2017; Ragab and Mansour, 2021). The SP-flooding method was
core floods and field applications (Levitt et al., 2012; Hashmet et al., patented by Gogarty and Tosch for Marathon oil company as Mara-flood
2017; Haruna et al., 2019). For instance, a high-potential polymer for in the 1990s (Sheng, 2013a; Ragab and Mansour, 2021). The SP-flooding
HTHS carbonates is the hydrophobically associating polyacrylamide process involves the injection of a chemical slug into the reservoir
polymers (HAPAMs) and its nanocomposite derivative. They have consisting of water, polymer, surfactant, electrolyte (i.e., salt), a
shown excellent tolerance towards temperature, pH, and ion content in co-solvent (i.e., most often alcohol), which is called the micellar solu­
the periphery of the wells while functioning as viscosifiers or wettability tion, followed by a thick water polymer injection. At times, to improve
modifiers in HTHS and high-pressure reservoirs (Rodrigues et al., 2006; robustness of surfactants at certain temperatures and water salinities,
Gouveia et al., 2008; C. Zhong et al., 2008). Also, in this paper, HMAs co-surfactants may be needed in the formulation fluid. The role of the
(Hot-Melt Adhesive) polymers, sulfonated polyacrylamide (AN-125), inorganic salts (i.e., the water-soluble electrolytes) in the micellar slug is
and certain HPAM based polymers such as Flopaam 3630 and SAV10, to attain greater solution viscosity regulation. The injected polymer
have been detailed as some potential polymers for SP based cEOR in solution helps propagate the micellar solution slug to achieve mobility
HTHS carbonate rocks. Several successful field applications in HTHS control (Cockcroft et al., 1988). Also, decreased oil-water IFT that re­
carbonates were conducted in the past years; however, their sults from the SP-flooding helps regulate the mobility ratio by

3
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

generating a sizable oil bank and ultimately boosts the oil recovery, as forces, is a key indicator of enhanced displacement efficacy during the
shown in Fig. 3. cEOR processes. Thus, it is essential to either increase the displacing
In other words, simultaneous adsorption of secondary surfactant, fluid’s viscosity (i.e., water) or decrease the IFT between the oil and
ionized acid, along with unionized acid on rock and solution interface water in order to increase the capillary number value (Sheng, 2010,
leads to the creation of a compound layer of adsorption. Consequently, 2015). Typically, the oil-wet carbonate system has negative capillary
the behavior of dynamic IFT occurs (Cockcroft et al., 1988; Bera et al., pressure, and consequently, water is stopped from spontaneously
2011; Karović-Maričić et al., 2014; Mandal, 2015). Furthermore, the imbibing into the porous medium since the capillarity of the carbonate
wettability changes by surfactants have been attributed to mainly pH rock surface strongly attaches the oil. SP-flooding reduces the IFT by
and potential determining ions (PDIs) i.e., Ca2+, Mg2+, and SO2− 4 . The forming microemulsions by the oil-water interface, causing reduction of
involved reactions are controlled by the surface charges (zeta-poten­ capillary forces, which mobilizes the trapped oil.
tials), produced precisely at the brine-oil and brine-rock interfaces (i.e., The IFT values are primarily defined according to the factors of
crude oil-brine-rock interactions or COBR system). Likewise, the types surfactant concentration in aqueous solutions, molecular structure,
and properties of crude oil influence the optimal salinity parameter. The temperature, and salinity. In terms of ultralow IFT, high surfactant
extent of recovery of oil in carbonate reservoirs is also vastly affected by concentrations are more desirable than a low concentration. Greater
gravity, capillary, viscosity forces, and injection water composition. The quantities of oil are solubilized when there are high surfactant concen­
components of this micellar solution or slug, secure a phase-wise shift, trations present, and a type III Winsor solution is anticipated to form,
which involves water displacement to oil displacement that is devoid of which result in oil recovery enhancement (Hirasaki et al., 2011).
interfaces. These components of SP-flooding generally constitute the However, the critical micelle concentration (CMC), over which the IFT is
following amounts in the solution: (i) 10–15% of surfactant, (ii) 20–60% unaffected, must not be exceeded by the surfactant concentration. To
of water, (iii) 25–70% of oil, (iv) 3–4% of co-surfactant (Sheng, 2015). precisely design an effective SP-flooding, a proper understanding on the
Using alcohol as a co-solvent, increases the tendency of the micellar fluid basic mechanisms behind reduction in IFT dynamics between crude oil
to incorporate both water and oil (Cockcroft et al., 1988; Bera et al., and water is paramount (Sheng, 2015; Mandal, 2015). Fig. 4 illustrates
2011; Karović-Maričić et al., 2014; Mandal, 2015). the COBR system during SP-flooding. Hence, the capillary pressures
Thus, SP based cEOR is an effective technique that decreases oil- between the aqueous and oleic phases indicate the reduced oil-water
water IFT, aids in mobility ratio control by creating a large oil bank, IFT. As previously mentioned, the surfactant concentration, tempera­
causes wettability alterations to cumulatively improve recovery factor. ture change, and water-chemistry (i.e., salinity and composition) of the
A number of experiments has demonstrated the incremental oil recov­ aqueous solution have a significant impact on the extent at which this
ery, and the fluid-rock interactions (i.e., COBR system) that constitute reaction occurs at brine-oil and brine-rock interfaces (i.e., COBR
the underlying mechanisms such as IFT reduction, mobility control, and system).
wettability alteration of the carbonate rock surface. Investigators conducted imbibition tests on surfactants, which
created IFT reduction, and noted that, mechanism of oil recovery during
imbibition via IFT reduction, was most probably due to buoyancy and
2.1. Interfacial tension (IFT)
alteration of wettability. In case of certain anionic surfactants, extremely
low IFT values can be achieved so that the capillary pressure is
One of the main considerations in surfactant-polymer (SP) flooding
decreased to almost zero (Hirasaki and Zhang, 2003; Seethepalli et al.,
cEOR processes is interfacial tension (IFT). The relationship of three
2004b; Adibhatla and Mohanty, 2006; Lu et al., 2014a). As the capillary
types of forces i.e., the capillary, gravitational, and viscous forces, in the
pressure nears zero, the various forces (i.e., gravity and viscous) can lead
oil reservoirs, influences the amount and rate of oil recovery. Obtaining
to imbibition, which was seen in several experiments. In particular,
a high capillary number, which is the ratio of viscous forces to interfacial

Fig. 3. A simple illustration of SP based cEOR process.

4
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

latter figure shows the correlation between incremental recovery and


IFT at varying polymer viscosities, as the IFT is progressively increased
from ultra-low values (10− 3 mN/m), the incremental recovery increases
with maximum values ranging from 10− 2 and 10− 1 mN/m and then
starts to decrease, demonstrating that the highest recovery was attained
at low IFT and not at ultra-low IFT value.
In summary, mixing surfactants, oil, water, and electrolytes forms
microemulsions due to IFT reduction, which leads to creation of oil
banks. The IFT reduction is mostly attributed to buoyancy and wetta­
bility changes along with the combination of particular temperature,
water salinity, surfactants type, crude oil characteristics, and viscous
forces. Microemulsions are reviewed in the ensuing section, as their role
is fundamental in cEOR processes.

2.1.1. Phase behavior


Fig. 4. A schematic diagram representing the COBR system during SP-flooding.
In SP-flooding, to identify optimum surfactants and ascertain their
R-COO- denotes carboxylate ion (i.e., anionic surfactant) and Poly− denotes
polymer molecule in the aqueous solution.
stability, microemulsion phase behavior tests are conducted. These tests
help in deducing the oil and water solubilization ratios by measuring the
oil and water volumes in microemulsion phase. Microemulsions are
simulation results by Abbasi-Asl and co-workers illustrated that the
thermodynamically stable phases, formed under certain conditions
transverse pressure gradients between the fractures and matrix can push
(Bourrel and Schechter, 2010). The thermodynamically stable micro­
the surfactant deeper into the matrix in the dynamic imbibition process
emulsions are composed of homogeneous dispersions of two immiscible
(Abbasi-Asl et al., 2010).
fluids, which are commonly, hydrocarbons and water stabilized with
Moreover, the quality of crude oil, water salinity, and the type of
surfactant molecules, that can be either alone or blended with a
surfactants are important factors effecting IFT reduction. The IFT for
co-surfactant (Winsor, 1954). They are capable of ultra-low IFT reduc­
crude oil and brine system, measures around 7 mN/m at optimum
tion between oil and water and also cause wettability alteration of
salinity (6 wt% NaCl), and this is much lesser value than the IFT of crude
reservoir rocks. Rather than surfactants adsorbed in the bulk phase, the
oil-water system (48 mN/m) (Anderson, 1986; Crain, 2002; Mandal,
primary components of microemulsions are the surfactants adsorbed at
2015). In numerous researches on surfactants, static imbibition experi­
the interphase. The values of IFT are very low between microemulsion
ments were extensively used to examine different surfactants (Austad
and crude oil; and between microemulsion and water, usually ranging
et al., 1998; Xie et al., 2004; Seethepalli et al., 2004a; Zhang et al.,
from 10− 3 mN/m (Hirasaki et al., 2011). Examining the phase behavior
2006a; Sagi et al., 2013; Chen and Mohanty, 2013; Bourbiaux et al.,
of the surfactants/co-surfactant–brine–oil system is the best approach
2014; Wang and Mohanty, 2015). It was categorically observed that, in
for defining the IFT behavior of microemulsions. It is understood that
carbonate reservoirs the oil recovery is highly affected by gravity and
IFT behavior is the basic factor in estimating the oil recovery perfor­
capillary forces. However, viscous forces may also play an important
mance through the microemulsion flooding method (Winsor, 1954).
role (Mohammadi et al., 2009) Several observations indicated the
Winsor (1954), the first presented the elemental concept of the phase
Marangoni effect (Austad and Milter, 1997) and spontaneous emulsifi­
behavior, studied these phase behaviors and classified them as I, II, and
cation (Zhang et al., 2006b) to induce imbibition in some static imbi­
III types (Winsor, 1954). The shifts from one to another microemulsion
bition experiments. Goudarzi and co-workers (Goudarzi et al., 2012)
type depends upon temperature, brine composition (i.e., salinity and
have suggested via static imbibition experiments, that modifying the
hardness), oil composition, pressure, and the concentration and chem­
matrix block size affects the oil recovery. When surfactants, oil, water
ical structures of surfactants and co-solvents (Skauge and Fotland, 1990;
and electrolytes blend, thermodynamically stable phases called micro­
Bourrel and Schechter, 2010). In accordance with type of surfactant, the
emulsions are formed under certain conditions (Winsor, 1954; Bourrel
microemulsion phase behavior alters from Winsor I (lower phase) to
and Schechter, 2010). The dissolved micro-emulsions in oil or water, a
Winsor III (middle phase) to Winsor II (upper phase) through changes in
stable pressure gradient, and enhanced sweep efficiency via polymer
the ensuing conditions: (1) salinity increase, (2) alcohol (co-surfactant)
injection aid to significantly increase the oil recovery (Hill et al., 1973;
concentration increase, (3) surfactant molecular weight increase, (4) oil
Bera et al., 2011).
chain length (alkane carbon number, ACN) decrease, (5) temperature
Furthermore, the experimental study by Yu et al. (2010) demon­
change, (6) total surfactant concentration increase, (7) surfactant sol­
strated that there exists an optimum IFT, which leads to highest oil re­
ution/oil ratio increase, (8) surfactant hydrophile-lipophile balance
covery that surpasses the ultra-low IFT stated, as illustrated in Fig. 5. The
(HLB) decrease, (9) brine/oil ratio increase, as shown in Fig. 6 (Salager
et al., 2005).
Furthermore, Huh (1979) observed that the solubilization ratios
correspond to the IFT between oil and water (Huh, 1979). When IFT
values are about 10− 3 dyn/cm, it is regarded as sufficiently low to
mobilize residual oil from a reservoir. According to Healy et al. (1976),
the salinity value at which oil and water solubilization ratios are equal is
called optimum salinity (Healy et al., 1976). The lowest IFT is normally
seen at the optimum salinity, which is a prerequisite for designing a
cEOR flooding. Liyanage et al. (2012) conducted a microemulsion phase
behavior test on the alkoxylated variant of the Tristyrylphenol alkoxy
sulfate-based surfactant in an active oil (Liyanage et al., 2012). They
observed excellent water solubility, good stability in hard brine, excel­
lent oil recovery, and very low surfactant retention (Adkins et al., 2012;
Lu et al., 2014c). In another phase behavior test, the new Guerbet alkoxy
Fig. 5. Correlation between incremental recovery and IFT at varying polymer carboxylate (GAC) surfactants showed several advantages over other
viscosities. types of surfactants under a broad range of conditions (Iglauer et al.,

5
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Fig. 6. Microemulsion phase-behavior of surfactant-water-crude oil as a function of several variable parameters (i.e., temperature, salinity, surfactant concentration,
brine-oil ratio, etc.).

2009, 2010; Adkins et al., 2012; Lu et al., 2014b). interfacial energy between water and solid, σow is the IFT between oil
Therefore, microemulsion phase behavior tests help determine the and water, and θc is the contact angle, as shown in Fig. 7 (Lwisa, 2021).
oil and water solubilization ratios. The phase behaviors have been In the fractured carbonate reservoirs, when wettability changes from
classified as I, II, and III types (Winsor, 1954), and the dependency oil- or mixed-wet towards a more water-wetting state, it effectively
factors for phase behaviors have also been determined. One such factor enhances the oil recovery. Thus, a fundamental factor that considerably
is IFT between oil and water, and lowest IFT is observed at optimum enhances oil recovery in surfactant flooding is the mechanism of reser­
salinity, which is the salinity where oil and water solubilization ratios voir wettability alteration from oil-wet to intermediate-wet or greater
are equivalent. water-wet state (Yang and Zhou, 2020). The wettability alteration aids
in residual oil mobilization, which creates the spontaneous imbibition
process. In general, carbonate and fractured reservoirs are considered as
2.2. Wettability alteration oil-wet or mixed-wet; therefore, the process of water imbibition is either
deprived or unable to occur owing to the negative capillary pressure
Wettability has been unanimously established as a major criterion (Andersen et al., 2014). Consequently, for increasing the oil recovery
affecting improved and enhanced oil recovery (Zhou et al., 2000; from such reservoirs, the reservoir matrix’s wettability needs to be
Morrow and Mason, 2001; Hirasaki and Zhang, 2003). Wettability is the altered to more beneficent ranges so that it enables imbibition of water
feature of a fluid to move around and attach onto a solid surface when into the rock matrix and drive the oil out of the matrix into the fracture
another immiscible liquid is present (Crain, 2002). Anderson, proposed (Zhang et al., 2006a). The phenomena behind carbonate reservoir
three ways to determine wettability: I) Through contact angle mea­ wettability change via surfactants towards a more beneficial water-wet
surement, or by using the product of the contact angle and the IFT (the state are either through ion-pairing or surfactant adsorption mecha­
last term in Young’s equation); II) By measuring the relative quantities nisms. When using cationic surfactants, ion-pairing mechanism alters
of oil and water displaced in like situations; III) Observations of the wettability of carbonate rock, and this occurs due to the anionic
displacement or surface mechanisms in relation to water or oil phases components (primarily of the carboxylic group) adsorbed on the rock
(Anderson, 1986). This is illustrated by the following equation below surfaces from the crude oil reacting with the cationic surfactants’ posi­
(Anderson, 1986): tive heads. Consequently, ion-pairs soluble in the oleic phase shall be
created together with the surfactant monomers as a result of the elec­
σ os − σws + σ ow cos θc = 0 (1)
trostatic forces. Accordingly, the rock surface shall be altered to more
where σos is the interfacial energy between oil and solid, σws is the water-wet state as the oil shall be desorbed (Salehi et al., 2008).
Moreover, during numerous pervious investigations, it was observed
that dissolved salt has considerable effect on enhanced oil recovery;
certain makeup and the concentration of various wettability-altering
components in the injection water had significant impact on
improving oil recovery (Austad et al., 2012; Qiao et al., 2016). Other
studies showed that the crude oil-brine-rock (COBR) interaction and the
related wettability changes are mainly governed by pH and potential
determining ions (i.e., PDI) (Ca2+, Mg2+, and SO2−4 ), and that the extent
of oil recovery is based on the concentration of these ions (Zhang and
Austad, 2006; Lager et al., 2008; Puntervold, 2008; RezaeiDoust et al.,
2011; Austad et al., 2012; Zhang and Sarma, 2012; Mahani et al., 2015;
Hassan et al., 2022a). However, several researchers have contended that
it is challenging, to describe ionically modified brines in the crude
oil/brine/rock system with a single mechanism, due to the complex
geochemistry involved in this process (Omekeh et al., 2012; AlQuraishi
et al., 2015; Al-Shalabi and Sepehrnoori, 2017; AlSada and Mackay,
2017). Boampong et al. (2002) devised a triple layer surface complex­
ation model (TLM) in PHREEQC geochemical simulator (Parkhurst and
Appelo, 1999), which was then coupled with UTCHEM, which is a
reservoir simulator that implements multiple phases (Luo et al., 2016).
Fig. 7. Wettability measurement of crude oil-rock-brine system. The results showed that wettability is ultimately associated with

6
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

zeta-potentials at the oil-brine and rock-brine interfaces. Also, the au­ in the HTHS environment, large-scale commercial applications are yet to
thors reported that only at extreme magnitudes of zeta-potentials at the be conducted. Since the majority of carbonates come with challenging
interfaces, the carbonate rocks could change to highly oil-wet or conditions such as heterogeneous composition, mixed-wet to oil-wet
water-wet (Boampong et al., 2022). rock wettability as well as HTHS conditions, SP based cEOR (i.e., SP-
Furthermore, in a recent study, two main mechanisms, (i) the coating flooding) becomes ineffective in recovering incremental oil due to the
process, which is associated with the use of anionic surfactants, and (ii) instability of the surfactant and polymer solutions under these envi­
the cleaning mechanism for cationic surfactants were found to influence ronments (de Oliveira Silva et al., 2007; Algharaib et al., 2014;
the wettability alteration of carbonates through surfactant injection Al-Shalabi, 2018; Diab and Al-Shalabi, 2019). In heterogeneous car­
(Chowdhury et al., 2022). Chowdhury et al. (2022) observed that the bonates, the adsorption of the surfactant on minerals hinders the
hydrophobic surfactant tails interact with the adsorbed oil in the effective implementation of SP-flooding. Furthermore, degradations
monolayer of anionic surfactant molecules that coats the carbonate rock also occur at times, due to the crude oil types (i.e., polarity constituents
surface. As a result, the hydrophilic head groups of the surfactant cover a of crude oil) and the permeability of the porous media. In reservoirs
previously oil-wet rock surface, changing the rock’s wettability toward possessing high salinity (i.e., 270,000 ppm), high hardness brine (i.e.,
the condition of water-wet state. Also, the cationic surfactants are con­ 25,000 ppm), and high temperatures of 100–140 ◦ C, which are quin­
nected to the second process, which is referred to as the cleaning tessential in the Middle East oil reservoirs, the successful implementa­
mechanism that also affects wettability alteration in carbonates. tion of SP-based cEOR becomes highly challenging. This is owing to
Thus, wettability is recognized as a leading factor in improving oil solubility and incompatibility of chemicals in such hard-saline brines,
recovery. In carbonates, wettability shift from oil- or mixed-wet towards and thermal instability at such high temperatures. Divalent cations, such
more water-wet, leads to enhanced oil recovery. Dissolved salt with as calcium and magnesium, are present in significant proportions in
certain composition and concentration of wettability-modifying ele­ carbonate reservoirs (Ca2+ and Mg2+ ions >2000 ppm), which can
ments in the injection water affect EOR. It has been proposed that the reduce the efficacy of typically used (anionic) surfactants and polymers,
pH, potential PDI viz., Ca2+, Mg2+, and SO24, and zeta-potentials at the and aggravate adsorption (Levitt et al., 2011; Dupuis et al., 2017).
oil-brine and rock-brine interfaces affect wettability. Nonetheless, re­
searchers have concluded that it is debatable as to which particular 3.1. Factors affecting SP-flooding under harsh HTHS conditions
mechanism can fully define wettability alteration, due to many com­
plexities involved (Omekeh et al., 2012; AlQuraishi et al., 2015; AlSada Factors including, but not limited to, temperature, salinity, adsorp­
and Mackay, 2017). tion, and injected fluid concentrations, may influence the effectiveness
and performance of an SP-flooding process. These factors affect the IFT,
2.3. Mobility control micro-emulsion phase behavior, wettability, and mobility control,
which have an impact on the oil displacement mechanisms. In this sub-
In the SP-flooding process, the role of polymers is mobility control. section, the most essential factors namely, temperature, salinity, pH,
Polymers augment macroscopic displacement efficacy through mobility adsorption, and injected fluid concentrations are reviewed.
ratio reduction (Hamoud, 1995). Adding polymers to injection water
fulfils the purpose of increasing viscosity and keeping mobility ratio <1. 3.1.1. Effect of temperature
This causes greater volumetric sweep efficiency for EOR (Mandal, The reservoir temperature is an influential factor in the chemical
2015). Micro-emulsions are formed upon dissolving surfactants in water flooding process through several means. To be exact, the critical micelle
or oil, causing an oil bank to form (Bera et al., 2011). As polymers are concentration (CMC) of the solution, the IFT, and the injected chemical
injected, oil recovery increases considerably due to an enlarged oil bank (e.g., surfactant) adsorption are typically impacted by the reservoir
coupled with stabilizing of pressure and improvement in sweep effi­ temperature. In the case of SP-flooding, the hydrophobic tail’s solubility
ciency (Hill et al., 1973). In the SP-based cEOR slug and in the following in water mostly decreases as temperature increases, and the surfactant is
polymer flood injections, the amount of polymer must be sufficient so shifted to the oil-solution interface, lowering the IFT (Gao and Sharma,
that, desirable sweep efficiency is achieved without viscous fingering, 2013). Additionally, from the perspective of surfactant flooding,
which is common in heterogeneous reservoir like carbonates. Thus, adsorption of surfactant on the rock surface poses a challenge. Never­
appropriate mobility control is important in SP-based cEOR projects theless, since temperature has a substantial influence on surfactant
(Dean, 2011), and is related to reduced surfactant adsorption (Nelson, adsorption rate, when temperature increases, a considerable decrease in
1983). Hence, the polymers are used in SP-flooding as viscosifying the surfactant adsorption rate occurs. This is attributed to the kinetic
agents for maintaining mobility ratio between oil and water <1, which energy increase of the species (Fava and Eyring, 1956).
would lead to a sizeable oil bank formation and increased volumetric In a study by Azam et al. (2014), laboratory experiments were con­
sweep efficacy for cEOR. ducted to evaluate the effect of temperature on the anionic surfactant
adsorption into samples of Berea sandstone. It was concluded that when
3. Challenges associated with SP-flooding under harsh the temperature increased from 25 to 75 ◦ C the anionic surfactant
conditions adsorption slightly reduced from 0.9604 to 0.491 mg/g. Prior to this,
Ziegler and Handy (1981) carried out laboratory experiments to deter­
A harsh condition carbonate reservoir may be defined as, any mine the temperature effect on the nonionic surfactant adsorption into
reservoir, where two or more of the following conditions are met, and samples of sandstone rock. The results obtained from the static
which are the typical conditions in the Middle East oil reservoirs (Kli­ adsorption experiments proved that when surfactant concentrations are
menko et al., 2020): low, the adsorption rate is decreased along with the increase in tem­
perature ranging from 25 to 95 ◦ C. Because of this kinetic energy in­
• Reservoir temperature of over 85 ◦ C, crease at elevated temperatures, whereupon the interaction forces
• Formation water salinity above 100,000 ppm (i.e., Total Dissolved become are weakened. Whereas, at high surfactant concentrations, the
Solids or TDS), adsorption rate increased as the temperature increased.
• The concentration of divalent cations (hardness) above 1000 ppm, Furthermore, the surfactant adsorption process in general is classi­
and permeability lower than 100 mD. fied as exothermic process that is either one: (i) enthalpy driven, and (ii)
entropy driven; while a separate influence of temperature on surfactant
The cEOR methods have been successfully applied decades ago in adsorption occurs based on the adsorption density (Hirasaki and Zhang,
clastic reservoirs (i.e., sandstones). However, in carbonates, especially 2003; Jian et al., 2018; Belhaj et al., 2020). Consequently, when

7
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

surfactants with low adsorption density are present, and the tempera­ water interfaces, and the temperature has a minor effect on decrease in
ture is increased, the adsorption density is increased (i.e., through IFT.
enthalpy driven adsorption). Conversely, low adsorption density is Karnanda et al. (2013) discovered in their study that there was no
observed for high adsorption density surfactants (i.e., through effect of either temperature or pressure on the CMC of Triton X-405,
entropy-driven adsorption) in the event of temperature increase (Kamal which is a non-ionic surfactant; however, the CMC of Zonyyl FSE anionic
et al., 2017). Moreover, through their studies, Bera et al. (2011) estab­ surfactant decreased with pressure increase and evened out as the
lished that the adsorption capacity decrease (in case of entropy-driven temperature continued to increase until 80 ◦ C, exceeding which the CMC
adsorption) due to temperature increase, is based on two leading rea­ increased. Likewise, the study by Puerto et al. (2012) confirmed that the
sons; (i) there is decrease in solution viscosity corresponding to tem­ cloud point of non-ionic surfactants (such as alkoxylatedglycidyl ether
perature increase; thus, the diffusion rate decreases of the adsorbate sulfonate, or AGES) negatively influences the surfactant functioning in
amidst the external boundary layer and internal pores of reservoir rocks. the course of the cEOR process at high reservoir temperatures. The
(ii) The equilibrium adsorption capacity of the sand particles is affected adversarial consequence of separation of the aqueous saline solutions of
by temperature based on if the adsorption process is exothermic or AGES into two liquid phases occurs at the cloud point temperatures.
endothermic. Consequently, for the ionic and non-ionic surfactants, they should be
In contrast, where non-ionic surfactant is concerned, there is tested in advance for the Krafft and cloud points, respectively under
adsorption increase corresponding to temperature increase. With the reservoir conditions before selection for use (Schramm, 2000; Schott,
increase in temperature, the surfactants head groups are dehydrated, 2005).
rendering them loose hydrophobicity and turn more compressed, lead­ Moreover, temperature significantly influences the process of
ing to increased surface activity and enhanced adsorption amounts. It wettability alteration. With the increase in temperature within the car­
was also observed that at low concentrations, the non-ionic surfactant bonate reservoirs, the water-wetting process appears to take over. In the
adsorption onto the crushed Berea sandstone decreased as the temper­ investigations of Zhang et al. (2006b) and Austad et al. (2005), it was
ature increased; however, at high concentrations, the adsorption observed that spontaneous imbibition from chalk led to an increase in oil
increased as the temperature increased (Ziegler and Handy, 1981). recovery as the temperature increased. The cause of this improvement is
Furthermore, non-ionic surfactants can suffer phase separation (cloud due to the greater attraction of SO2−
4 ions to the carbonate surface when
point) at high temperatures, causing a reduction in surfactant concen­ temperatures are higher, which results in the negatively charged
tration (Hirasaki et al., 2011). carboxylate groups (R-COO-) of the sulfate ions becoming more attrac­
A zwitterionic surfactant was synthesized with extremely low CMC ted on the positively charged carbonate surface. This event releases the
denominations at high temperature of 95 ◦ C, in order that the adsorption oil from the rock surface, and alters the rock wettability to a more
rates may be decreased by reducing free monomers of the surfactant in water-wet condition. Moreover, non-ionic surfactants are able to
the solution (Bataweel and Nasr-El-Din, 2011). On the other hand, the favorably affect contact angle reduction existing between the non-ionic
bulk of the surfactants will degrade or precipitate when the tempera­ surfactant and carbonate surface at high-temperature situations (Gupta
tures go above 120 ◦ C (Negin et al., 2017). Likewise, it should be minded and Mohanty, 2010), whereas zwitterionic surfactants display a mixed
that each surfactant has an unlike CMC value, subject to change based on effect on wettability alteration when temperature increases (Adejare
the temperature (Emrani and Nasr-El-Din, 2015; Islam et al., 2015). et al., 2012).
When ionic surfactants are concerned, the Krafft point (Krafft temper­ Furthermore, the amphoteric and nonionic surfactants exhibit min­
ature), is the minimum temperature at which micelles may form in a imal carbonate adsorption and are compatible with HTHS conditions
solution (Dicharry et al., 2016). Hence, if the temperatures are lowered (Han et al., 2013; Wang et al., 2015; Jian et al., 2016; Ayirala et al.,
below the Krafft point, the solubility would decrease, rendering the 2019). The amphoteric surfactants can be particularly effective in
surfactant crystals to isolate from the solution. Accordingly, the ionic lowering the oil-water IFT, which is advantageous for mobilizing re­
surfactant solubility rises along with the temperature increase to reach sidual oil (Zhao et al., 2007a, 2007b; Guo et al., 2019; Ayirala et al.,
the Krafft point, at which the CMC is attained (Miyake and Oyama, 2021). Furthermore, the sulfates are prone to hydrolysis at high tem­
2009; Inada et al., 2017). Contrariwise, the non-ionic surfactants have peratures and alternatives like sulfonates have to be used instead (Lu,
their solubility categorized based on the cloud point, which is according 2014; Abalkhail et al., 2019). Also, the synergy of zwitterionic surfac­
to the surfactant structure, and the cloud point range is between 60 and tants, nanoparticles, and new polymers (HAPAM), have been shown to
120 ◦ C. If this temperature range is surpassed, the surfactant degrada­ be highly resilient and effective in inducing IFT and altering wettability
tion or precipitation begins (Negin et al., 2017; Aktar et al., 2020). At the of carbonate rocks in both ambient and elevated temperatures with high
cloud point, the precipitation and separation of non-ionic surfactants salinity (hardness) (Panthi et al., 2022; Afolabi et al., 2022; Larestani
occurs. This decreases the surfactant concentration, which in turn up­ et al., 2022; Meng et al., 2022; Joshi et al., 2022; Song and Hatzignatiou,
surges the IFT to eventually cause surfactant performance to decrease 2022). Thus, for SP cEOR application in HTHS carbonate reservoirs, a
(Hirasaki et al., 2011). number of newer surfactants and polymers can be utilized for formu­
In addition, as the temperature of a non-ionic surfactant solution lating robust SP solutions with desirable rheological properties, as dis­
meets the cloud point, the solution turns turbid, and species start to cussed in following sections.
isolate from the solution. This is attributed to the hydrating of the pol­
yoxyethylene groups in the surfactants when the cloud point tempera­ 3.1.2. Effects of salinity, hardness, and pH
ture is surpassed (Mukhopadhyay et al., 2019). Ye et al. (2008) Salinity strongly affects different microemulsion structures that are
conducted research to demonstrate that temperature is the principal influential on the wettability behavior of carbonate rock. The condition
cause that affects the IFT of a surfactant solution; as the temperature of saline environs considerably affects SP-flooding efficiency, due to
increases, the IFT existing amidst the crude oil and surfactant solution unfavorable impact on the surfactant’s solubility by the high-salinity
decreases. Furthermore, the dynamic IFT is also affected by tempera­ conditions. This is for the reason that surfactants incline towards
ture, given that the increasing temperature causes the time requisite for remaining in the oleic phase, resulting in deterrence of achieving IFT
the IFT to achieve the solution equilibrium to decrease. In spite of this, at reduction (El-Batanoney et al., 1999). Consequently, it is crucial that
the time surfactants are being mixed, the influence of temperature is optimum salinity must be reached at which equal amounts of oil and
overlooked by reason of the synergism effect. This is ascribed to the water is able to be solubilized so that microemulsions are formed (Hir­
surfactant aggregation of molecules at the oil and solution interface; asaki et al., 2011; Gao and Sharma, 2013). Chou and Shah (1981) per­
hence, the principal mechanism of IFT reduction is the compressed formed a laboratory study to show that highest oil recovery was attained
arrangement and improved capability to amend the properties of oil and at the point where connate water salinity and polymer solution were at

8
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

the optimum surfactant formulation salinity. number of hydroxyl groups on the surface of the rock is decreased due to
The anionic surfactants are vulnerable to high-salinity aqueous so­ the rock minerals reacting with the hydroxyl groups in the solution.
lutions, as the surfactants precipitation starts owing to the salt ions and Hence, cationic surfactant is adsorbed by the negatively charged hy­
surfactant molecules interacting (Al-Hashim et al., 1988). Furthermore, drated mineral oxides existing on the solid rock surface (Liu et al., 2010;
when there is an increase in salinity, the adsorption of anionic surfac­ Hanamertani et al., 2018; Azam et al., 2013; Belhaj et al., 2020). These
tants increases correspondingly on Berea sandstone because of the sur­ are described through the ensuing reactions regarding the mechanism of
face acidity increase and the electrical double layer (EDL) composition Berea sandstone attaining positive and negative charges due to low pH
that diminishes the electrostatic repulsion existing between the surfac­ and high pH, respectively. The same mechanism can be applied in the
tant monomers and the clay mineral surface (Kwok et al., 1993). In case of carbonate rock having positive and negative surface charges
addition, for the conventional anionic surfactants utilized in chemical based on the variation in pH values (Hassan et al., 2020).
cEOR processes, they adsorb more on limestone and sandstone as the As a result, for adsorption on Berea sandstone to be decreased when
divalent cations increase. This is because of the cations present and their anionic surfactant is present, alkalis like sodium hydroxide or sodium
adsorptions on the rock surface, resulting in diminished repulsion forces carbonate are utilized, which enhance the pH and the positive charge
existing between the anionic surfactant molecules and the rock surface’s density on the rock surfaces is decreased (Azam et al., 2013; Hirasaki
negative charges; which consecutively, lets the surfactant molecules to et al., 2011). This was substantiated in the study by Azam et al. (2013),
be adsorbed on the surface of rock (Azam et al., 2013). Also, when high in which they conducted laboratory investigations on the pH effect
Ca2+ concentration is present in limestone rocks, it causes increase in the causing static adsorption onto Berea sandstone by means of synthesized
surfactant precipitation (Al-Hashim et al., 1988). Conversely, when anionic surfactant. The outcomes showed that a positive charge was
divalent cations are present, they may lessen the cationic surfactants attained by the silica fraction in Berea sandstone at low pH, while at high
adsorption onto carbonate rocks due to the strong positively charged pH, a negative charge was attained. Thus, the anionic surfactants
cations (Ma et al., 2013). Moreover, anionic surfactants can alter car­ adsorption on the rock surface will be decreased at high pH (above 9.0),
bonate rock wettability towards greater water-wet states only if the caused by the repulsive forces existing amid the surfactant head group
brine salinity is low, predominantly in the presence of divalent ions such and the negative rock surface charges. In another study, Mushtaq et al.
as Mg2+ and Ca2+. The interaction of multivalent cations with anionic (2014) provided a description of the effect that pH changes in the near
surfactants leads to precipitation and the wettability altering effect in the PZC (pH of 7.98) value has on the anionic surfactant adsorption onto
carbonate rocks is diminished (Deng et al., 2021). It was experimentally the Malaysian sandstone surface. When there was no alkali, the surfac­
shown by Gupta and Mohanty (2011) that along with salinity increase, tant’s static adsorption was greater (3.66 and 4.49 mg/g) at pH values
there was decrease in the optimum surfactant concentration, that lead to lesser than 7.98, whereas, when there was alkali present at pH values
the optimal contact angle increase. higher than 7.98, substantial decrease in surfactant adsorption (0.43 and
Experimental investigations by Baviere et al. (1988) showed pro­ 0.86 mg/g) was observed. Likewise, studies were conducted by Bera
spective efficiency of alpha-olefin sulfonate (AOS), which is an anionic et al. (2011) using the cationic surfactant (cetyltrimethylammonium
surfactant, for cEOR applications under HTHS and high-hardness situ­ bromide) to determine its adsorption behavior on the clean sand parti­
ations. Their outcomes demonstrated the high salt tolerance of AOS cles for samples. Their study concluded that cationic surfactant
surfactant even though there were calcium ions present. Addition of an adsorption increases correspondingly to pH increases of the solution, for
alcohol or increasing the temperature, can achieve solubility for the AOS the reason that at high pH values, the cationic surfactant, which has
surfactant in high-salinity brines. Consequently, AOS surfactants have positively charged head groups, is highly attracted to
high solubility and IFT reduction at elevated temperatures throughout a negatively-charged sand rock surfaces.
vast salinities range. There was slight increase in surfactant adsorption On the other hand, it is apparent from the results of Tagavifar et al.
on kaolinite with salinity increase even though the surfactant concen­ (2018) that the adsorption of anionic surfactant on Indiana limestone
trations were high. Salinity affects the solubility, the surface activity and (ILS) declines virtually directly with pH values above 9. In high pH
the non-ionic surfactants’ adsorption. Triton X-100 (a non-ionic sur­ conditions (i.e., pH values above 9), the dominant mode of adsorption
factant) was found by Nevskaia et al. (1998) to display three distinctive on calcite is through significant hydrogen bonding. However, at low pH
categories of adsorption behavior when NaCl was present with the use of values, the foremost adsorption type on calcite and clay surfaces is
different quartz rocks types (QA, QB and QC). They observed that the charge-regulated monodentate (i.e., carboxylate or sulfate), irrespective
adsorption of Triton X-100 was reduced after NaCl was mixed to the of surface loading (Tagavifar et al., 2018). Figs. 8 and 9 illustrate the
solution on the QA sample and the adsorption increased after testing on progression of the surfactant’s adsorption on ILS core-samples in rela­
the QB rock; in addition, there were no changes noticeable from the QC tion to surfactant concentration, pH, ionic strength, and temperature.
rock tests. This distinct behavior is related to the salt cations and Initially, the adsorption isotherms is estimated for both systems A
different surface hydroxyl group’s interactions, on top of the surface (78 ◦ C) and B (24 ◦ C) when salinity, temperature, and pH were constant,
impurities being present instead of the interactions between salt and which is depicted in Fig. 8. Further, the pH was estimated to be 7.6 and
TX-100.
Moreover, another critical factor influencing the efficiency of SP-
flooding is pH, as the surfactant adsorption rate is impacted by surface
charges of the rock that are subject to variations by way of pH alter­
ations. The pH value at which the net charge on the surface (surface
charge density) is zero is called the point of zero charge (PZC). It is a
crucial attribute of the rock for the reason that at pH levels greater than
their PZC, the surfaces turn into negatively charged (Grigg et al., 2004).
By way of example, the anionic surfactants, when used in cEOR pro­
cesses provide a window of opportunity; because they do not undergo
significant losses through adsorption on account of the negative surface
charge on sandstone when pH levels are high (Azam et al., 2013). In a
pH environment lower than PZC, the rock’s hydroxyl minerals tend to
attain a positive charge, leads to an anionic surfactant adsorption in­ Fig. 8. Langmuir isotherm for systems A (78 ◦ C) and B (24 ◦ C) for surfactant
crease by reason of the negative charges of the surfactant’s group having adsorption on Indiana limestone (ILS). Dotes and Lines represent experiment
affinity to the rock surface. Although, when pH is greater than PZC, the and model calculations respectively (Tagavifar et al., 2018).

9
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

8.6 for A and B systems, respectively, which was not attuned; while the behaviors of adsorption in order to minimize surfactant adsorption.
rock–water interactions (primarily the calcite dissolution) controlled it. Researchers have examined the effects of solid surface interactions,
Symbols were employed to denote the experimental data, and the range surfactant properties, water chemistry, and functional groups on sur­
of experimental error was between 1 and 5% of the absolute measure­ factant adsorption (Chandar et al., 1987; Zhu and Gu, 1991; Hu and
ment value. Solid lines were used to signify the results of surface Bard, 1997; Somasundaran and Krishnakumar, 1997; Yu et al., 2003; Ma
complexation modeling using the Phreeqc software. Both the systems et al., 2013; Zhong and Claverie, 2013; Levitt and Bourrel, 2016). For
were similar to Langmuir monomer adsorption in their adsorption re­ some of the well-known adsorbents, studies have defined the adsorption
sults with comparable adsorption capabilities (i.e. approximately 3 sequence, the adsorbed layer structure (Bohmer and Koopal, 1992),
mg/g of rock). The earlier results acquired for anionic surfactant adsorption mechanisms (Zhang and Somasundaran, 2006), and
adsorption on calcites (Ahmadali et al. 1993; Ma et al. 2013; Shamsi­ adsorption kinetics (Tabor et al., 2009; Ahmadi and Shadizadeh, 2015).
Jazeyi et al., 2014a) coincide with the adsorption plateau of 3 mg/g of Although, for carbonate rock formations (Levitt and Bourrel, 2016) and
rock at low pH values that agrees to 5.3 and 4 μmol/m2 for systems A carbonate soil (Ishiguro and Koopal, 2016), which as we know have
and B, respectively, as shown in Fig. 9. In Fig. 9, the surfactant heterogeneous mineral composition and in rock surface interactivities,
adsorption plateau on ILS rock in relation to pH at the constant ionic the underlying mechanisms and their degree of impact on adsorption are
strength and temperature is shown for both A and B systems. One can hard to predict. In fact, there is yet to be conclusive numerical or
observe that both A and B systems demonstrate very comparable experimental evidence about the workings and the extent of the effect of
adsorption profiles and are primarily determined by pH for the surfac­ minerals on adsorption based on the variations in ionic strength and pH
tant adsorption on ILS rock. At pH values below 9, the same adsorption in the environment. This is firstly due to the reason that pure mineral
plateau of 3 mg/g of rock is obtained for both systems, that analogous to samples are used in adsorption tests, while the difference in mineral
the adsorption isotherm plateau, as shown in Fig. 8. In case the pH composition of carbonates (e.g., limestone and calcite) causes very
values exceed 9, the adsorption amount diminishes almost aligned with different adsorption behaviors in the actual reservoirs. Ma et al. (2013)
the pH values. The adsorption reduction alongside pH is partially studied this aspect and observed negligible cationic surfactant adsorp­
attributable to the positive charge decrease and the subsequent charge tion on artificial calcite, yet significant adsorption was observed on
reversal alongside the pH values. Because of this, the like charged samples of natural limestone which can be clearly linked to aluminum
anionic surfactant and calcite repel one another. Nevertheless, adsorp­ and silicon present in the rock (Ma et al., 2013). This also shows the
tion is observed as remaining high at around 1 mg/g of rock at a pH of existence of trace amounts of minerals viz., magnesite, Fe-dolomite,
11.9. anhydrite, clay, and oxides of metal in carbonate rocks (Folk, 1959;
Ma et al., 2013; Levitt et al., 2015). The second reason is that mineral
3.1.3. Effect of adsorption components have varying surface charges and dissolution behaviors.
When evaluating SP-flooding in cEOR processes, surfactant adsorp­ Generally, metal oxides have a point of zero charges greater (PZC) than
tion on rock surfaces in porous media is a crucial factor believed to be carbonate minerals components, which eventually causes the oxides and
the leading mechanism behind surfactant retention on rock. The sur­ carbonates to become polar opposites in sign at certain pH levels (Kos­
factant adsorption process into the reservoir rocks eventually creates a mulski, 2002; Stumm and Morgan, 2012). This means that there will be
decrease in surfactant concentration that consecutively depreciates the contiguous heterogeneous surface potential and this generates the
beneficial properties for the process of displacement and decreases the following (Na et al., 2007):
SP-flooding efficiency (Liu et al., 2020). Thus, surfactant adsorption is a
primary factor contributing to the economics of the SP-flooding project. • In carbonate dust, the zeta potential is unreliable in fully explaining
Several parameters are in control of the adsorption process such as the adsorption associated with charges (Moulin and Roques, 2003; Chen
surfactant type, rock morphology and mineralogy characteristics, et al., 2014; Kasha et al., 2015; Mahani et al., 2017).
charges on the rock surface, the electrolytes types present in the solu­ • Adsorption in non-uniform forms such as the charge-controlled
tion, and the co-surfactant and alcohols present (Liu et al., 2004). As forms against the hydrogen-bonded forms, and the form dynamics
surfactant retention is controlled, it will enable to secure low costs and effectively cause patch-wise adsorption sometimes (Nwidee et al.,
environmental preservation. For this purpose, additives (e.g., Alkali) are 2017).
incorporated with surfactants to cause pH increase and surfactant • Also, when the surfactant solution contacts the rock, the ionic
adsorption to be reduced (Nelson et al., 1984; Delshad et al., 2011; constitution will be different from one at the start owing to the
Solairaj et al., 2012; Sharma et al., 2015; Southwick et al., 2016, 2018; dissolution of mineral components (Cui et al., 2015).
Wang et al., 2018, 2019). • Furthermore, with a decrease in pH, usually the adsorption of
Moreover, it is essential to precisely measure the static and dynamic anionic surfactants decreases too (Somasundaran and Fuerstenau,
1966; Seethepalli et al., 2004b; Pham et al., 2015; Sharma et al.,
2015; Southwick et al., 2016) and there is a slight increase noticed
when ionic strength increases, and with the increase in the concen­
tration of divalent cations (ShamsiJazeyi et al., 2014b; Budhathoki
et al., 2016).

The third reason is that adsorption on clay platelets extremities at


locations that are like oxides, which has polymorphic tendency i.e., the
water chemistry might affect both the size and dynamics in a non-
uniform manner, as demonstrated in Fig. 10. In the latter figure, the
charge-regulated adsorption is the leading mechanism on both calcite
and platelets. There is presence of hydrogen bonding mechanism on
both minerals. However, calcite sites show higher adsorption capacity
than oxide-like sites on clay platelets edges. This was proven by Levitt
and Bourrel (2016), who observed that in cases of montmorillonite clay,
Fig. 9. Static surfactant adsorption plateau on Indiana limestone (ILS) versus and in an outcrop rock that had undergone oxidization, that was
pH for systems A (78 ◦ C) and B (24 ◦ C). Dotes and Lines represent experiment composed of sizeable amounts of clay and iron, using a simple sodium
and model calculations, respectively (Tagavifar et al., 2018). chloride rinse, the extent of the adsorption could be decreased. They

10
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

mineral. Due to the complication of carbonate surface charges, it is


difficult to conclude if it is the cationic or the anionic surfactants that
should be utilized in order to diminish the electrostatic attraction forces
between the surfactant molecules and the formation surfaces (Ahmadali
et al., 1993; Ma et al., 2013). Ahmadali et al. (1993) conducted exper­
imental valuations regarding the cationic surfactant’s adsorption on
carbonate minerals. They used two cationic surfactants (i.e., cetylpyr­
idinium chloride, or CPC, and dodecylpyridinium chloride, or DPC) with
synthetic calcite and natural dolomite as samples. They concluded that
the cationic surfactants proved to have considerably lesser adsorptions
in contrast to the anionic surfactant that had comparable hydrophobic
chain lengths. When salts were added with multivalent cations, cationic
surfactant adsorption decreased; however, the anionic surfactant
adsorption increased. This is as a result of the cations addition and the
Fig. 10. Adsorption degree and manner on a low surface-area calcite and high lattice ions contained in carbonate rocks considerably influenced the
surface-area clay platelets at a pH of approximately 8 (Tagavifar et al., 2018). surface charge of the minerals causing anionic surfactant adsorption to
increase. Additionally, their investigations concluded the likelihood of
associated the reduced adsorption to the exchange of divalent cations lessening the surfactant loss caused by adsorption into carbonate
from clays due to rinsing, as it decreased the surface charges (Levitt and reservoir rocks by way of using cationic surfactants mixed with the
Bourrel, 2016). Furthermore, the carbonate surface geochemistry and additive multivalent electrolyte in appropriate concentrations.
the surface charge of carbonate rock can influence the retention values In summary, surfactant retention can occur due to heterogeneity in
of the surfactants. For example, in a coreflood by Klimenko et al. (2020), mineral composition and surface interactivities of rock. The surface
it was observed that even a small density of calcium sites on the surface geochemistry and the surface charge of carbonate rock can influence the
is sufficient to retain the carboxylate surfactant used (Klimenko et al., extent of surfactants retention. For instance, natural limestone has
2020). aluminum and silicon present that causes retention, and even tiny cal­
Also, injection of microemulsions (i.e., composed of surfactants cium sites on rock surface led to carboxylate surfactant adsorption.
mixed with co-solvents, alkali, oil, and brine constituents) that are Adding alkali to reduce pH, and pre flush with low salinity water
highly viscous and have non-Newtonian behavior results in elevated flooding before SP-based EOR can reduce surfactant adsorption.
surfactant adsorption in the field, higher gradients of pressure that are
unsustainable, lower sweep efficiency, and immobile microemulsions 4. Novel surfactant formulations for SP-flooding
owing to the high viscosity at low shear rates. To alleviate micro­
emulsion viscosity, generally, surfactant formulation is optimized via a The most commonly used surfactants types are the Anionic, Cationic,
suitable co-solvent and/or via incorporating increased branching into Zwitterionic and Non-ionic surfactants. Anionic surfactants possess an
the hydrophobe of surfactant structure. The co-solvents of proper anionic effective cluster at their head, like the carboxylates, sulfate,
amounts are added to create microemulsions of low viscosity. Never­ phosphate, and sulfonate. Cationic head groups surfactants are pH-
theless, co-solvents can raise the price and intricacy of the process while dependent primary, secondary, or tertiary amines; i.e., the primary
causing increased IFT. In addition, studies on the microemulsions and secondary amines are positively charged when pH < 10. One of the
formed by novel surfactants mixed with polymers (and co-solvents and cationic surfactants is octenidine hydrochloride. Zwitterionic surfac­
alkali), have shown that the SP (co-solvent or alkali) formulations which tants have cationic and anionic heads together on the same molecule.
are weakened by shear will cause an increase in surfactant adsorption The cationic head is composed of primary, secondary, or tertiary amines
(Walker et al., 2012; Suniga et al., 2016). or quaternary ammonium cations, while the anionic head could be more
Furthermore, it is recommended to use anionic surfactants in sand­ flexible and have sulfonates included in it (Lwisa, 2021). Non-ionic
stone reservoirs for the reason that they adsorb relatively less in com­ surfactants are constituted by covalent bonds of oxygen-containing hy­
parison to other types of surfactants. Sandstone rocks comprise of drophilic groups, which are attached to hydrophobic parent structures.
substantial portion of silica minerals and tiny portions of carbonates and The hydrogen bonding imparts the oxygen groups’ water-solubility,
silicates. The negatively charged silica causes the electrostatic repulsion which is inversely proportional to temperature increase; i.e.,
forces existing amid the sandstone formation and the anionic surfactant increasing temperature causes the water solubility of non-ionic surfac­
to be high, which averts surfactant adsorption into the sandstone surface tants to decrease. Non-ionic surfactants are affected to a lesser degree by
(Ma et al., 2013). Nevertheless, there could be adsorption on account of water hardness than anionic surfactants, and their foamability is also
the clay minerals (mainly kaolinite and illite) present in sandstone, less strong. A pictorial chart which details the various classes of sur­
because the clay’s surface charges have originated from isomorphous factant and their common types are given below in Fig. 11.
substitution, lattice defects, and broken bonds. Thus, the anionic sur­ There are only minor differences among each type of non-ionic
factant adsorption on the clay minerals is regulated by the forces of surfactants, and their selection is mainly dependent on the costs based
electrostatic and van der Waals (Kwok et al., 1993; Iglauer et al., 2010). on distinct/required properties (e.g., effectiveness and efficiency,
In carbonate reservoirs, zwitterionic surfactants adsorption on car­ toxicity, dermatological compatibility, biodegradability) or based on
bonate rock surface grows with surfactant concentration increase by their approved usage as food ingredient. In the design stage, when it
means of electrostatic and hydrophobic interactions, along with micellar comes to selecting the surfactants, many times anionic surfactants are
systems. On the other hand, as temperature or salinity increased the the preferred choice as they tend to have less adsorption at pH ranging
adsorption is deemed as low (Nieto-Alvarez et al., 2014). In contrast, from neutral to high on sandstones and carbonates alike (Sheng, 2010,
cationic surfactants exhibit significantly low adsorption onto carbonate 2015). They may be modified to suit a wide array of environments and
rocks, on account of positive surface charge of calcite and dolomite are easily available at low cost in certain cases. The sulfates are suitable
minerals. For that reason, cationic surfactants are considered appro­ for implementation in low temperature while the sulfonates are useable
priate chemicals for carbonate reservoir applications (Lippmann, 2012). in harsh temperatures. A cationic scan may be applied as co-surfactants
On the other hand, it is shown in certain studies that hydrogen ions act since the performance of non-ionic surfactants for SP-based EOR is
as potential determining ions (PDIs) for carbonates, which connotes that poorer, in comparison to anionic surfactants. Furthermore, the sulfo­
the equilibrium pH of the solution regulates the surface charge on the nated hydrocarbons viz., alcohol propoxylate sulfate and alcohol

11
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Fig. 11. Flow diagram of surfactant categories based on head group nature.

propoxylate sulfonate is frequently applied for usage in Surfactant surfactants and polymers that are high-yielding in harsh conditioned
flooding. carbonates. Therefore, many novel surfactant and polymer families have
Requirements of an effective surfactant for SP-flooding can be sum­ been developed recently, which have been assessed carefully for appli­
med up as follows (El-Hoshoudy et al., 2017b): The role of a surfactant is cability. Since existing surfactants lack the required properties for
to reduce IFT between the trapped oil and injected water interfaces to extremely reduced IFT, water stability (i.e., wettability alteration to­
ultralow values (0.001 mN/m), which would increase the capillary wards more water-wet), robustness at elevated temperatures, minimal
number (Nc) by a factor of at least three orders of magnitude. This adsorption, resistance to extreme salinity and hardness. Formulating
further reduces the capillary force, leading to decreased oil contact angle new surfactants with these desirable properties is incumbent. Literature
thereby changing the wettability (Wu et al., 2010) which solubilizes in this area have introduced three categories of recent large-hydrophobe
trapped oil and increases mobilization. A surfactant is also selected or surfactants such as Guerbet alkoxy sulfates, Guerbet alkoxy carboxyl­
designed to alter the matrix rock wettability toward more water-wet ates, and Tristyrylphenol alkoxy sulfate surfactants. Other novel sur­
conditions, as this increases brine-imbibition rates. Another use of sur­ factants are Guerbet Alkoxy Sulfate surfactants (GAS), Guerbet Alkoxy
factants is the ability to modify the properties of polymeric systems that Carboxylate Surfactants (GAC), Guerbet alkoxy betaine surfactant
enables a range of applications (Holmberg, 2003). Further useful fea­ (GAB), and polymeric surfactant or surfmers. Their suitability for large
tures of surfactants in the process of enhanced oil recovery include, harsh conditioned carbonate reservoirs is described below. Additionally,
particle dispersion, emulsion stabilization, foam generation, reservoir latest surfactant structure correlation is detailed further in the paper,
wetting or wettability alteration, and many more (Schramm, 2000; Maia that may be used as benchmark for determining surfactant formulations
et al., 2009). Ultimately, a desirable surfactant has the characteristic of that are high-performing.
minimal adsorption on reservoir rocks and, long-term temperature sta­
bility, brine salinity, and hardness under reservoir conditions along with 4.1.1. Guerbet Alkoxy Sulfate surfactants (GAS)
easy compatibility with reservoir fluids, and most importantly have The Alkyl alkoxy sulfates having hydrophobes between 12 and 18,
tolerance to divalent cations such as Ca2+ and Mg2+ ions. and have been used widely for cEOR, (Levitt et al., 2006; Flaaten et al.,
Therefore, the most common surfactant types include the Anionic, 2008). However, they are susceptible to higher equivalent alkane carbon
Cationic, Zwitterionic and Non-ionic surfactants, which differ in their number (EACN) of crude oil and high temperatures. Also, a majority of
structure and properties. The surfactant is chosen according to its these alkyl alkoxy sulfates possess restricted branching. But when a
robustness and long-term thermal and brine salinity, and its capacity to significantly high carbon number is used for branching Guerbet alcohols
shift wettability and cause IFT reduction. (GA), it turns them into viable surfactant hydrophobes, which can be
very costly when in very-pure form. Even then, Guerbet alcohols may be
produced in cost effective manner if lower than quantitative conversion
4.1. Potential surfactants for SP-flooding under HTHS conditions is done during the procedure of alcohol dimerization. In addition, this
version of surfactants can exhibit below par hydrolytic stability when
As cEOR is extended to extreme reservoir environments (i.e., HTHS), temperatures exceed 60 ◦ C. The half-life of ether sulfate surfactants, i.e.,
it becomes evidence that traditional surfactants and polymers, the time for decay is approximately 4–6 years, at temperature of 85 ◦ C
commonly used in previous decades, are intolerant under harsh HTHS (Adkins et al., 2010). Thus, they are recommended to be mixed with one
conditions. The HTHS conditions necessitate the need for formulating

12
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

or more of other classes of surfactants (i.e., as co-surfactants) that have (Behler et al., 2001; Rosen and Kunjappu, 2012). Therefore, to employ
varying structures, to improve the efficacy of SP-flooding in HTHS car­ AEC surfactants for SP-based EOR with minimal expenses incurred,
bonates. Co-surfactants may also potentially improve aqueous stability large-hydrophobe AEC surfactants have been designed. The structure of
or limit microemulsion viscosity (Adkins et al., 2010). Likewise, by Guerbet alkoxy carboxylate (GAC) surfactant can be modified according
introducing propylene oxide (PO) and ethylene oxide (EO) to the GAC to the particular cEOR requirements. This design modifications are done
surfactant molecules, alkyl alkoxy carboxylates surfactants are formed, by changing the carbon numbers in the Guerbet alcohol and though
which are better suited for SP-flooding under harsh HTHS carbonates change of the propylene oxide (PO) and ethylene oxide (EO) groups’
conditions. numbers (Adkins et al., 2012). Fig. 12 shows an example of the poly­
meric non-ionic surfactants structures namely, the Propylene oxide (PO)
4.1.1.1. Laboratory experiments of novel GAS. A formulation containing and Ethylene oxide (EO). These design modifications are conducted by
0.25 wt% C32–7PO-14EO-sulfate, 0.25 wt% C20–24-IOS, and 0.5 wt% changing the carbon numbers in the Guerbet alcohol and the PO and EO
TEGBE cosolvent was examined in a phase behavior test at 100 ◦ C (Levitt groups’ numbers, which ere formulated by Nakagawa et al. (1998).
et al., 2006). A light oil (API of 29, μo of 4.0 cP) with a limited acid Additionally, these novel surfactants when used in combination with
number signifying an inactive oil, was employed for this formulation. cosurfactants and cosolvents have proven to form effective formulations
Upon observing the microemulsion phase behavior, the solubilization with desirable properties shown via microemulsion phase behavior and
ratio at optimal salinity was seen to be around 17, which was correlated coreflooding experiments (Lu et al., 2014c). These experiments were
to an exceedingly reduced IFT of 1.0 × 10− 3 dyn/cm. In the water-based conducted successfully on different types of oils with varying composi­
solution, 500 ppm Flopaam 3630 polymer was added, which was clear tions and were performed under harsh HTHS reservoir conditions (Lu
and stable at temperature of 100 ◦ C in the presence of 35,000 ppm of et al., 2014a, 2014c).
sodium carbonate (Levitt et al., 2006). The authors also conducted a
coreflood to deduce the efficacy of the surfactant formulation. The 4.1.2.1. Laboratory experiments of novel GAC. A new formulation con­
core-samples possessed a permeability of 2500 mD and a porosity of taining 0.5 wt% C28–45PO-60EO-carboxylate and 0.5 wt% C15–18-IOS
0.217. A 0.5 pore volume (PV) of surfactant mixture having a 0.5 wt% was tested, which displayed excellent phase behavior at 120 ◦ C without
formulation density, was injected and afterwards a polymer drive was a co-solvent when a medium gravity oil (API = 22, μo = 1.1 cP) was
introduced. There was 94.7% oil retrieval of the total residual oil satu­ employed. At an optimal salinity, the solubilization ratio was 13 (Lu
ration following the waterflood process and the eventual oil saturation et al., 2014c).
was approximately 0.02. On the rock, the retained surfactant was noted The formulation of 1.0 wt% C32–7PO-32EO-carboxylate and 1.0 wt
as 0.11 mg/g of rock, which is a favorable reduction for limiting the % C19–23-IOS was tested for microemulsion phase behavior at 100 ◦ C
expenses of the projects. Through the chemical flood, pressure declined after 545 days for a light oil (API = 34, μo = 0.5 cP) with a low TAN. The
to 0.4 psi/ft at a 1 ft/day frontal velocity. This minor pressure decline is optimal salinity was 36,000 ppm and the optimal solubilization ratio of
greatly advantageous as the pressure gradient inside reservoirs, gener­ 11 showed corresponding IFT reduction to be satisfactory at about
ally tend to hover around 1 psi/ft (Levitt et al., 2006). Thus, the in­ 0.003 dyn/cm, which is feasible for majority of cEOR schemes. To
vestigations show that GAS surfactants with co-surfactants and develop a substitute and to create cheaper co-surfactants, a C11-linear
cosolvent are resilient and effective under high temperature, which alkyl benzene sulfonate (ABS) was formulated and examined in order
could be potentially applicable for SP flooding in harsh HTHS carbonate to replace the IOS with the same oil (API = 34, μo = 0.5 cP). Another
reservoirs. formulation with 0.7 wt% C28–25PO-55EO-carboxylate and 0.3 wt%
C11-ABS was observed to have excellent microemulsion phase behavior
4.1.2. Guerbet Alkoxy Carboxylate Surfactants (GAC) in hard brine and its maximum solubilization ratio was 11 at optimal
Since the Guerbet alkoxy sulfates cannot tolerate temperatures over salinity of 30,000 ppm. It is well known that ABS cannot be used in hard
60 ◦ C, except when the pH is between 10 and 11 (Adkins et al., 2010). brine but, this experiment showed a harmonious combination of the
This elevated pH is maintained by means of traditional alkali additives. GAC and the ABS that enhanced the efficacy and the stability in water
However, this is not possible in cases where there is exposure to gypsum for both these surfactants (Lu et al., 2014c).
or anhydrite in the reservoir rock or when softening of water injection is Founded on the results of a preceding microemulsion phase behavior
not available. experiment, a core flood was performed (Lu et al., 2014c). A Silurian
Ether sulfonates can be an alternative as they can tolerate extreme dolomite core was taken, having highly vuggy form and high hetero­
temperature and extreme hardness (Puerto et al., 2012). However, they geneity (through visual inspections and CT scans), the brine perme­
are difficult to produce; rendering them expensive and their commercial ability was of 478 md and the porosity was 0.168. As GAC was being
procurement is limited. In this regard, Alkyl alkoxy carboxylates are a investigated for the first time in a dolomite core, a larger surfactant slug
potential substitute as they are of an anionic surfactant family that can was injected into the core of (PV × C of 50) %, after which a polymer
be applied in SP-flooding (Shaw, 1984; Li et al., 2000, 2005). Adding drive was injected. The pore volumes of the surfactant slug multiplied
propylene oxide (PO) and ethylene oxide (EO) to the molecule, can with the surfactant concentration (PV × C) provides the total mass of
create alkyl polyoxy propylene/ethylene (alkoxy) carboxylates (AEC) injected surfactant and is an important indicator when comparing cor­
which are multifunctional. When compared with alkyl carboxylates eflood efficiencies. Here PV stands for pore volume and C denotes sur­
(soap), these AECs have been shown to have excellent resistance to factant concentration. Having a low surfactant retention allows for
hardness, together with a favorable aqueous solubility and tolerance to
pH alterations or changes in electrolyte levels (Behler et al., 2001; Rosen
and Kunjappu, 2012). Therefore, the AEC are great substitutes for alkoxy
sulfates in certain types of reservoir environments. Other attractive
features of the AEC are that, they can be paired with versions from all the
varieties of surfactant categories and are also quite convenient to use
(Behler et al., 2001). Above all, the AEC can maintain stable state at very
high thermal conditions, can remain chemically constant and can be
used in acidic and alkaline pH alike (Mayer et al., 1983; Thomas, 2008;
Ramsey, 2019; Abalkhail et al., 2020), with no decay in carboxylate
functional group, which makes them especially suitable for cEOR Fig. 12. Structure of Propylene oxide (PO) and Ethylene oxide (EO) polymeric
non-ionic surfactants (modified after Nakagawa et al., 1998).

13
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

successful corefloods when (PV × C) % is low. The results showed the oil betaine surfactants that have been well assessed for SP flooding that
recovery at 90.5% of residual oil after the waterflood having a Sorc of have high interfacial activity at oil-water interface, have superior ther­
0.044 (HPLC i.e., High-Performance Liquid Chromatography was used mal stability and resistance to salt (divalent ion) under HTHS reservoir
to gauge the surfactant concentration and retention on the effluent conditions. In order to be able to resist HTHS conditions and handle
specimens). The overall surfactant retention on the rock was observed to severe technical challenges, surfactant structure was needed to be
be 0.33 mg/gm rock, the pressure decline throughout the chemical flood enhanced (Han et al., 2013). Thus, to improve thermal tolerance, salt
was noted as 1.5 psi/ft at a 2 ft/day frontal velocity and this low resistance and emulsification enhancement, Guerbet alkoxy betaine
pressure-drop is beneficial as is near to the gradient of field pressure. surfactant (GAB) was formulated and evaluated for a HTHS low
In order to further evaluate the efficacy of the formulation, a lesser permeability reservoir. The branched hydrophobic group and EO in the
concentration of surfactant was used in the slug, for the very same oil, a novel GAB surfactant formulation significantly heightened the emulsi­
different coreflood was performed in relatively homogenous Estaillade fying ability. On the other hand, sulfonate hydrophilic group in the GAB
limestone core having permeability of 187 md and a porosity of 0.276, structure imparted exceptional thermal stability. It was tested effec­
and containing 0.15 wt% C28–25PO-45EO-carboxylate, 0.15 wt% tively at TDS of high formation brine (274,823 mg/L) with divalent ion
C15–18-IOS. This core was injected with a 1.0 PV surfactant/polymer content of approximately 27,000 mg/L and at reservoir temperature of
(SP) slug with (PV × C of 30) % and afterwards a polymer drive was 95 ◦ C and formation permeability within 10–30 mD (Cai et al., 2019).
injected. It was observed that the oil output was 94.5% with Sorc at 0.008 Hence, the GAB surfactants are high potential candidates for HTHS
and the overall retention of surfactant on rock surface was 0.34 mg/g carbonate application using SP-flooding technique. This is also evinced
rock. In case of C15–18-IOS, there was overall surfactant retention of from the experimental studies in ensuing sections.
0.16 mg/g rock and for C28–25PO-45EO-carboxylate the final value of
retention of surfactant on rock was 0.18 mg/g rock. Furthermore, 4.1.3.1. Laboratory experiments of novel GAB. A novel Guerbet alkoxy
chromatographic separation was not observed, neither was any prefer­ betaine surfactant (GAB) was designed for SP flooding and its charac­
ential retention of these chemicals noticed. Moreover, because there are teristics were investigated (Cai et al., 2019). The betaine surfactant
high divalent cations (hardness) concentration and high salinity in many molecule was incorporated with ethylene oxide (EO) functional group to
reservoirs, robust surfactant formulations are necessary for these enhance the emulsification ability and Guerbet alcohol was chosen as
extreme conditions (Lu et al., 2014c). the hydrophobic group. In the first step, glycidyl ether was derived by a
Furthermore, another SP formulation for high optimum salinity and reaction of alkoxylated Guerbet alcohol and epoxy chloropropane, fol­
extreme hardness for a light oil (API of 36.5, μo of 1.8 cP) is 1.0 wt% lowed by deriving tertiary amine through glycidyl ether and dimethyl
C28–25PO-55EO-carboxylate, 0.5 wt% C15–18-IOS, 0.5 wt% C19–23- amine. Finally, through quaternization reaction of tertiary amine with
IOS, and 1.0 wt% TEGBE, which was investigated at 83 ◦ C. In this 3-chloro-2-hydroxyl propane-sulfonic acid sodium salt, the GAB sur­
formulation, TEGBE stands for triethylene glycol monobutyl ether used factant was developed. Lab investigative methods such as IFT, long term
as co-solvent. Both the co-surfactant (IOS i.e., internal olefin sulfonate stability, contact angle, and phase behavior were conducted for this GAB
was used here) and the co-solvent (TEGBE) aid in averting the surfactant surfactant to assess its overall performance. An excellent tolerance and
molecules from packing tightly in the micelles which can lead to weak desirable conformity to high temperature, high salinity (HTHS) reser­
mobility or high retention on formation rock surfaces. This formulation voir conditions were observed. This new GAB surfactant could be
examined was on an Estaillade limestone core specimen having applied in salinity up to 275,000 mg/L and temperatures ranging to
permeability of 144 md and porosity of 0.24. At the beginning, the 120 ◦ C. The diluted GAB solutions when used for crude oil exhibited
salinity of (formation) brine was 230,800 ppm and there was a 21,700- ultralow IFT having the concentration of weight from 0.03% to 0.20%.
ppm divalent cation concentration. For the preparation of the slug, brine Other formulation containing 0.5% GAB and 0.5% amid-betaine,
with a salinity of 67,000 ppm TDS was used and with a divalent cation showed Winsor III middle phase microemulsion formation with dehy­
concentration of 3700 ppm present. It was discovered that this GAC drated light oil in HTHS carbonate reservoir. At reservoir temperature of
surfactant was very efficient in enhanced oil recovery and upon intro­ 95 ◦ C and optimal salinity of 50,000 mg/L, the solubilization ratio
duction of 0.2 PV SP slug, the oil recovery was observed at 92% whereas, reached a value of 16.
the ultimate saturation was noted at 0.016, and surfactant adsorption Thus, the improved betaine surfactant GAB showed better thermal
value was 0.3 mg/g rock (Lu et al., 2014c). stability, higher interfacial activity, and intensified emulsification fea­
In another experiment, a potential surfactant formulation based on tures under HTHS carbonate conditions, in comparison to Guerbet
novel large-hydrophobe Guerbet alkoxy carboxylate surfactant was used alkoxy sulfate surfactant and conventional sulfobetaine with similar
and an IOS co-surfactant was designed for HTHS carbonate reservoir (Lu structure.
et al., 2014a). Initial temperature was 100 ◦ C and salinity was 116,969
ppm. They performed both static and dynamic core experiments, and in 4.1.4. Tristyrylphenol alkoxy sulfate surfactants (TSP)
the dynamic coreflood experiment, the oil saturation decreased to 0.14 The TSP-based EOR surfactants were designed and tested for the
when surfactant was used in a small amount without polymer and the purpose of creating large quantities and variations of feed-stocks of
core was extremely vuggy and fractured. The surfactant retention was effective and reliable cEOR surfactants for use in prospective schemes. It
only 0.086 mg/g rock. The resultant oil recovery was excellent with the is possible to reproduce this petrochemical based hydrophobe in huge
dynamic coreflood delivering a higher yield than the static imbibition amounts so that it can be useful for large scale field schemes. TSP is a
experiment at similar conditions, even though, there were fractures in large hydrophobe that is widely marketed, with the ethoxylated variant
the core and it was especially vuggy. Moreover, it is to be noted that of TSP being well established as agricultural dispersants and in surfac­
mobility control was discarded, and a very small amount of the surfac­ tant formulations for pesticides (Derian et al., 1993; Gubelmann-Bon­
tant slug had been inoculated. Therefore, the results of dynamic core­ neau et al., 1995). When in the state of a phenol, it readily becomes
flood showed much greater oil retrieval as opposed to the results from alkoxylated with propylene oxide (PO) and ethylene oxide (EO) forming
similar static imbibition experiments. nonionic surfactant with massive molecular weight. Also, the degree of
These investigations indicate that GAC formulations are effective in the propylene oxide (PO) and ethylene oxide (EO) levels can be grafted
SP-based EOR techniques in harsh carbonate conditions, making them to suit the HLB according to the oil/brine mixture’s requirements and
apposite for field applications. when it is sulfated with sulfamic acid, it forms anionic surfactant of
substantial molecular weight which is a very inexpensive procedure,
4.1.3. Guerbet alkoxy betaine surfactant (GAB) thus, TSP alkoxy sulfates can be easily bulk produced and marketed
The recent years have seen significant development in various

14
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

(Liyanage et al., 2012). A pictorial diagram of synthesis sequence of a saturation of oil at 0.024 towards the end. When HPLC (High-Perfor­
TSP surfactant for large-scale production is shown in Fig. 13. mance Liquid Chromatography) quantifications were conducted, the
Hence, in the investigations by Lu and co-workers, the micro­ surfactant adsorption value was 0.10 mg/g of rock, and the two sur­
emulsion phase behavior experiments of the new TSP surfactants factant constituents, showed no visible chromatographic separation in
showed extremely reduced IFT and a transparent water-based solution, the core (Liyanage et al., 2012).
beyond optimum salinity, for a highly viscous oil (difficult oil), when For the corefloods by Lu and co-workers, a surfactant formulation
every other formulation for this case had failed to perform effectively with content of 0.20 wt% of the TSP-7PO12EO-sulfate and 0.05 wt% of
(Yang et al., 2010). Furthermore, these surfactants’ efficacies and sodium dihexyl sulfosuccinate was investigated in an unconsolidated
employability, were tested via corefloods using a surrogate oil and a live reservoir-sand (k ≈ 4.6 Darcy, φ ≈ 29%) with the aim of retrieving live
oil at high-pressure, and the observations signified exceptional perfor­ oil that had been reconstituted. This live oil had 31 mol% of methane
mance with increased oil output (92.8% of the residual oil post water­ and 9 mol % of carbon dioxide content in it, with the pressure at satu­
flood with a final oil saturation of 0.013) and negligible retention of ration being 1490 psi. Additionally, this coreflood experiment was
surfactant on rock surface (0.10 mg/g of rock) (Lu et al., 2014c). conducted in a pressure of 1700 psi that was a marginally higher.
Thus, a series of experimental laboratory studies using many TSP Eventually, this surfactant formulation was able to recover 92.8% of the
surfactants have demonstrated the capability of these surfactants in total post waterflood residual oil after leaving the saturation value
effective SP-based cEOR trails. The TSP surfactants are stable in HTHS 0.013 at the end, and on the rock the surfactant adsorption was at a
environment and effective in recovering highly-viscous oils. value of 0.10 mg/g (Lu et al., 2014c).
In summary, the TSP surfactants have proven to be highly stable
4.1.4.1. Laboratory experiments of novel TSP. In a micro-emulsion phase during the laboratory experiments and delivered the desired outcomes.
behavior test, Liyanage and co-workers demonstrated the capability of They have yet to be extensively studied for stability and efficacy in very
the alkoxylated variant of these TSP-based surfactants in an active oil of high temperature and salinity.
medium type (API = 28, μo = 14 cP) that had a complete acid number of
1.0 mg KOH/g oil at 62 ◦ C. To suit this surfactant and this experiment, 4.1.5. Gemini surfactants
finding an ASP formulation having existing surfactants for this partic­ To assuage the problems of hydrolysis and precipitation in HTHS
ular oil proved to be a challenging task. Nevertheless, superb solubility carbonate reservoirs, cationic gemini surfactants have been synthesized
in water was observed for TSP alkoxy sulfates even as it retained the and characterized recently. The cationic gemini surfactants with aro­
activity at the surface of this oil, resulting in phase behavior and core­ matic rings in varying numbers within the spacer group, along with
flood experiments with successful conclusions. These synthesized for­ diverse counter ions, were synthesized (Hussain et al., 2019). The out­
mulations were then examined in outcrops as well as reservoir comes of the thermogravimetric characterizations reveal that these
corefloods by taking a surrogate oil and also an artificial live-oil, in novel gemini surfactants decompose at approximately 300 ◦ C. This ex­
which excellent oil output was yielded and there was minimal surfactant ceeds the prevailing temperature of ≥100 ◦ C of current oilfields.
adsorption thus meeting the experimental objectives (Liyanage et al., Moreover, with the addition of ethoxy units amid the lipophilic tail and
2012). lipophobic head groups, a considerably improved solubility is obtained
An expansive set of TSP alkoxy sulfates, that had molecular weights in both highly saline brine and in ordinary water with no precipitate or
ranging from 1250 to 3600 g/mol were developed and tested. In the any murkiness. It was also observed that when there were two aromatic
selected oil, the surfactant with the highest potential had the formula­ phenyl rings in the spacer groups, the gemini surfactants present a
tion containing 0.18 wt% TSP-35PO-20EO-sulfate and 0.12 wt% reduced CMC than one phenyl ring gemini surfactants (Hussain and
C20–24-IOS. There were no cosolvents added in this formulation, even Kamal, 2017; Hussain et al., 2019; Mahboob et al., 2022). Also, previous
then, at the temperature of 62 ◦ C in the reservoir, it performed well for studies observed that with short spacer length, the gemini surfactants
the selected oil. Also, upon conducting a survey of salinities using the produced single-phase microemulsions, and with more extended spacer
surrogate oil, elevated optimum solubilization ratios were observed groups, gemini surfactants exhibited typical (conventional) surfactant
along with a large extremely low IFT sites for WORs (water-to-oil ratios) behavior (Mahboob et al., 2022).
ranging from 1 to 9. Also, in the representative activity diagram it was Besides, gemini surfactants were applied for field trials in Algyo field,
observed that, there was a small negative descent across the tested Hungary and exhibited high tolerance to hardness which is much higher
WORs, which was highly desirable. Furthermore, in the coreflood for than the prevailing temperatures of ≥100 ◦ C in current oilfields (Puskas
Bentheimer sandstone having permeability of k ≈ 2 Darcy, and porosity et al., 2017). In one of the field applications, at Cottonwood Creek,
of φ ≈ 0.24, this very same surfactant formulation performed very well Bighom Basin, Wyoming, USA, the nonionic-surfactant polyoxyethylene
by retrieving 92.1% of the post waterflood residual oil leaving the alcohol was used with satisfactory field application results.
To sum it up, gemini surfactants have exceptionally high heat and

Fig. 13. A schematic of TSP synthesis sequence with x (PO)-y (EO) sulfate surfactants (modified after Liyanage et al., 2015).

15
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

salinity tolerance with excellent solubility. With supplementary studies, attained by only a well-balanced combination of surfactant properties
they can be developed to perform exceedingly well for SP-based EOR in (Puskas et al., 2018). These formulations have been employed at Algyo
carbonates under harsh conditions. field; Hungary for SP based cEOR. The field is a sandstone reservoir with
depth of 1800–1900 m TVD, 191.3 bars original formation pressure,
4.1.5.1. Laboratory experiments of gemini surfactant. Through enduring 98 ◦ C reservoir temperature, 70 mD oil permeability, and 23.7% average
laboratory research of over 10 years, Puskas and co-authors (Puskas porosity. The formation oil is of low viscosity (0.64 cP at 98 ◦ C and 190
et al., 2017, 2018) developed optimal compositions of surfactant mix­ bars) (Puskas et al., 2017).
tures, which consist of highly effective gemini surfactants, and supple­ Thus, even though a finely balanced combination is required for
mentary surfactants (co-surfactants), which were produced from waste achieving the target properties, gemini surfactants have exceptionally
materials or inexpensive raw materials. The gemini surfactants are a high heat and salinity tolerance with excellent solubility. With supple­
type of novel surfactants with multiple hydrophilic head groups and mentary studies, they can be developed to perform exceedingly well for
hydrophobic tail groups connected by a spacer at or near the head SP-based EOR in carbonates under harsh conditions.
groups. The exceptional properties of gemini surfactants, is the low
critical micelle concentration (CMC), desirable aqueous solubility, and 4.1.6. Zwitterionic or amphoteric surfactant
uncommon micelle structures and aggregation behavior. Also, because Surfactants are typically categorized into four main types: anionic,
they are highly efficient in oil/water IFT reduction and have certain cationic, non-ionic and zwitterionic (or amphoteric) surfactants (Gupta
rheological properties, they have caught the interest of the researchers et al., 2020; Lu and Pope, 2017; Mejia et al., 2019; Pei et al., 2018;
dedicated to EOR applications in field projects (Páhi et al., 2008, 2009; Saxena et al., 2019). The zwitterionic (amphoteric) surfactants exhibit
Bergström et al., 2015; Kamal, 2016). The synthesized gemini surfac­ either of the positive or negative charge depending on the solution pH
tants molecules with improved properties in comparison to conventional (Lv et al., 2011). The presence of cationic moieties such as, amines or
surfactants based on laboratory studies of parameters such as critical quaternary ammonium groups create the positive charge whereas the
micelle concentration, surface tension, IFT, and wettability. The new negative charge is created due to the anionic moieties such as sulfonic
surfactant compositions contained both gemini surfactants and acid, carboxylic acid (Zhang et al., 2019). The amphoteric surfactants
co-surfactants, to function as adsorption reducing agents for gemini can shift the wettability state of both sandstone and carbonate rocks
surfactants on the reservoir rock surface and to provide greater tolerance effectively into more water-wet surfaces. The amphoteric surfactants
to both salinity and hardness. The past decade has seen the development alter the sandstone surface wettability by way of ion-pairing between
of over 300 non-ionic surfactants (and co-surfactants), which were the amphoteric hydrophilic head and the adsorbed basic crude oil con­
examined and evaluated based on several parameters such as acid stituents on sandstone rock. As the surfactant concentration increases
number that demonstrates the total content of acidic additives that are the crude oil components are increasingly desorbed causing more
water soluble and insoluble. Apart from this, the saponification number water-wet rock surfaces. For carbonate rock surfaces, the wettability is
was analyzed as a parameter which is expressed in milligrams of po­ altered due to the ammonium ions and the adsorbed acidic crude oil
tassium hydroxide, as it is essential for the saponification of all free and constituents forming ion pairs on carbonate surfaces changing it to
bonded acid in 1 g sample. Another important parameter was water water-wet (Kumar and Mandal, 2019). Moreover, increasing zwitter­
solubility of surfactants measured, which was measured by determining ionic surfactant concentration results in contact angle reduction in
the water number, which characterizes the relative solubility of sur­ carbonate rocks (Souraki et al., 2019; Amirpour et al., 2015). For the
factants in both water and oil (Bergström et al., 2015). Next, the HLB reason that of the zwitterionic surfactants heads have amphoteric
was evaluated due to its involvement in the hydrophilic/lipophilic na­ charges on them, the wettability alteration process in carbonate rocks
ture of the surfactants, and may allude to the surfactant’s efficacy, and swift (Kumar and Mandal, 2017). The zwitterionic surfactants that
indicate if the main groups in the surfactant mixtures are the water demonstrated favorable characteristics for cEOR process are carboxyl
soluble or the oil soluble functional groups (Puskas et al., 2018). betaine type, hydroxyl sulfonate betaine type, didodecylmethylcarboxyl
In addition, over 1000 surfactant mixtures were developed from betaine, and alkyl dimethylpropane surfactants (Kamal et al., 2017;
Geminis and co-surfactants, and were tested in the laboratory selection Massarweh and Abushaikha, 2020). Furthermore, the amphoteric sur­
phase via complex laboratory screening methods, to determine the factants, which have low Critical Micelle Concentration (CMC) are
surface-active properties and their potential efficacy for SP based cEOR. applicable for HTHS conditions. However, when compared to other
The screening methods of surfactant mixtures that were most essential surfactants, these zwitterionic surfactants are significantly costlier (Lv
are, the solubilization efficiency and determining the oil displacement et al., 2011).
effect by thin layer chromatography. In solubilization efficiency tests,
special laboratory equipment was used to shake together the mixture of 4.1.6.1. Laboratory experiments of zwitterionic surfactant. In an experi­
surfactant solution and crude oil at room temperature (ambient). After a ment, OCT-1 amphoteric surfactant was studied for IFT reduction and
duration of rest, separation of the phases had occurred and the formed phase behavior with different surfactants in Saudi Arabian seawater of
phase’s volumes were measured. Also, after storage for 24 h at reservoir salinity 57,000 ppm. The results showed desirable IFT reduction,
temperature (98 ◦ C), the phase volumes were repeatedly evaluated. The without significant microemulsion phase behavior. However, these
aqueous phase, emulsion and oil amounts were measured in comparison surfactants are expensive and used only in certain conditions (Butt et al.,
to the total liquid volume. The solubilization efficiency was studied with 2003; Fuseni et al., 2013; El-Hoshoudy et al., 2015a). Similarly, Chen
regards to the diverse effects of the surfactant mixture on the emulsions et al. (2019) showed that adding zwitterionic co-surfactants to
types that would typically form in the natural reservoir due to the oil-swollen micelles can significantly improve the high salinity colloidal
original phases. Then, using thin layer chromatography, the displace­ stability of alkylbenzene sulfonates. Through investigational and
ment effect was determined after the thin layer was prepared on a glass atomistic molecular dynamics simulations it was found that the less stiff
plate, from original core sand. Crude oil was introduced to the thin layer oil/water interfaces with mixed surfactants, which created fewer but
and allowed to dry. Following which, the glass plate was submerged into larger homogeneous micelles, significantly reduced the poly-dispersity
the glass cylinder of surfactant aqueous solution mixture. Several hours of surfactant mixed with oil-swollen micelles. In addition, it was
later, the distance between the drop and the edge of the oil spill was observed that sulfonates underwent significant dehydration as a result of
measured, which specifies the displacement efficiency of the surfactant shielding from more bulged zwitterionic co-surfactants at the oil/water
mixture. From the results of the single laboratory tests, it could be said interfaces. The co-surfactants successfully reduced the total exposures of
that the displacement efficiency of surfactant packages would be sulfonates to the aqueous phase, which could lead to divalent ion

16
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

bridging and formation of large aggregates. Thus, it was discovered that compared to conventional surfactants (Guyot, 2004). Polymeric sur­
the highest levels of sulfonate dehydration are found at flat oil/water factants are also a feasible substitute for traditional chemical processes
interfaces of extremely big oil-swollen micelles, which show that small as they have excellent tolerance to harsh conditions and have better
quantity of mineral oils could further increase the high salinity colloidal mobility with lower molecular weight than most polymers. They provide
stability (Chen et al., 2019). Thus, even with a higher price point, the multifunctional mechanisms including interfacial tension (IFT) reduc­
zwitterionic surfactants can provide robust performance in HTHS car­ tion and increased viscosity. They are considered to be more adaptable
bonates. By using hybrid SP-based cEOR and improving SP formulations in improving recovery as a result of mobility control and fluid redistri­
the surfactant retention and project cost may be reduced to optimize the bution caused by induced wettability alteration (Raffa et al., 2015,
implementation. 2016a; Raffa, 2021; Afolabi et al., 2022). It must be noted that there is
rarely any explicit performance calculation of polymeric surfactants to
4.1.7. Other surfactants formulations be found in literature regarding EOR (Raffa et al., 2016b). Nonetheless,
Other SP formulations were developed by Panthi et al. (2022) at a polymeric surfactants can be suitable alternatives for use in cEOR via
high temperature (100 ◦ C), high salinity (57,000 ppm), and particularly SP-based techniques.
high hardness (2760 ppm) carbonate reservoir, resulting in extremely
low interfacial tension. Acrylamido-tertiary-butyl sulfonate copolymer 5.1.1.1. Laboratory experiments of polymeric surfactants. The synthesis,
SAV10xv was discovered to be stable in these HTHS environments. A rheology, surface/interfacial tension, wettability, adsorption phenom­
formulation of three surfactants viz., carboxylate, an IOS, and a betaine ena, and recovery performance of polymeric surfactants have been
in combination with a pre-flush, according to coreflooding results, covered extensively in the literature. The majority of research on poly­
generated cumulative oil recovery in the range of 77–82% OOIP and a meric surfactants has focused on synthetic, high molecular weight (over
residual oil saturation of 8–11%. The surfactant retention was 0.133 1 million Daltons or g/mol) polymeric surfactants, particularly those
mg/g of rock without the use of NaPA; however, after pre-flushing with that are hydrophobically modified or associate with polyacrylamides. It
NaPA, the retention was reduced to 0.017–0.038 mg/g of rock. was observed that unlike traditional polymers like HPAM, they possess
SP-flooding projects in HTHS carbonate reservoirs can be improved by more remarkable rheological and emulsifying capabilities. The
using these surfactant formulations with a preflush of sodium poly­ hydrophobically-associating features of polymeric surfactants are the
acrylate and very low surfactant retention. basis of their greater tolerances to salinity, temperature, and shear.
In Appendix-A, Table A-2 summarizes the several novel surfactants However, because of their restricted surface activities, they are unable to
formulations, which can be suitable for HTHS conditions in carbonate lower IFT to the extremely low levels required to create microemulsions
reservoirs during SP-based cEOR. and successfully mobilize the oil trapped in micropores. Even so, an
ultra-low IFT is not always necessary for an effective SP-flood, as newer
5. Novel polymer formulations for SP-flooding studies suggest (Raffa, 2021; Afolabi et al., 2022). Also, the technical
problems present in chemical flooding techniques can be eliminated
Polymers are used for mobility control and for imparting viscoelastic with polymeric surfactants, which include unwanted fluid-fluid in­
characteristics to the injection fluid in SP-based cEOR systems. Most teractions and chromatographic separation caused by selective adsorp­
commercial polymers such as HPAM cannot be employed in HTHS tion and mechanical entrapment (Raffa et al., 2016a). The
carbonate reservoirs owing to thermal instability, adsorption on rock, multi-component injection slugs like the SP slugs, are vulnerable to
loss of viscosity, mechanical degradation, etc. Lately, researchers have failure in complicated and highly heterogeneous reservoirs due to the
developed novel polymers that are capable of withstanding HTHS and gravity segregation of fluid components. For tertiary recovery applica­
low permeability of carbonates with good viscous power and shear tions in these formations, amphiphilic polymers are expected to be
tolerance. These novel polymer formulations with their respective ex­ successful. In addition, polymeric surfactants can eliminate surface fa­
periments have been listed in the following sections. cility operational complications such as, scale buildup and corrosion
caused by alkali chemicals in ASP (Olajire, 2014).
5.1. Potential polymers for SP-flooding under HTHS conditions Concerning rheological characteristics, polymeric surfactants exhibit
comparatively mild surface-active properties (Raffa et al., 2016b; Guo
The novel polymers investigated for SP-flooding in HTHS carbonates et al., 2016; Negin et al., 2017), which is potentially due to molecular
include the hydrophobically associating polyacrylamide polymers size of the high-weighted polymeric surfactants. It has been generally
(HAPAMs) and its derivatives, surfmers, HMAs, NVP based polymers, observed that low-weighted amphiphilic polymers are some of the few
and certain biopolymers. These were evaluated for their physio- that can achieve extremely low IFT (Barakat et al., 1989; Cao and Li,
chemical properties, for instance, viscosity, rheology, salt tolerance, 2002; Wu et al., 2017). At their best, polymeric surfactants have been
and emulsification. Also, micro-models and coreflooding tests were observed to achieve IFTs of 0.5–1 mN/m (Cao and Li, 2002) and surface
conducted to determine the mechanism and the effectiveness of the tensions of 25–30 mN/m (Mehrabianfar et al., 2021). The emulsification
polymers during the SP based cEOR process. Furthermore, field appli­ capacity of polymeric surfactants depends on the length of the hydro­
cations were carried out using some novel polymers. Moreover, to study phobic branch chains (Maltesh and Somasundaran, 1992; Li et al.,
the chemical structure of these polymers (latexes), many methods were 2018). The surfactant concentration increases the properties of emulsi­
used such as, FTIR (Fourier-transform infrared spectroscopy), 13C-NMR fication and surface tension/IFT reduction until the CMC is reached.
(13C Nuсleаr Mаgnetiс Resоnаnсe), 1 H-NMR (Hydrogen 1 Nuсleаr Beyond CMC, the additional surfactant molecules mainly migrate from
Mаgnetiс Resоnаnсe), SEM (scanning electron microscope), X -ray surfaces into solutions, to form aggregates with various morphologies,
diffraction, and TEM (transmission electron microscope). The results and based on either oil or brine being the dwelling medium, the mor­
and potential implication for field applications in HTHS carbonates have phologies are cylindrical, spherical, or even club-shaped (Cao and Li,
been described in the ensuing sections. 1999; Hong et al., 2007). When these assemblies can capture oil globs,
emulsification follows. The non-ionic amphiphilic polymers are tolerant
5.1.1. Polymeric surfactant to divalent cations up to reasonably high levels (Sarsenbekuly et al.,
Polymeric surfactants are a type of functional surfactants, having an 2017). By adding functional groups to the scheme, such as through
amphiphilic structure i.e., a hydrophobic tail and hydrophilic head fluorination and sulfonation (Akbari et al., 2017; Li et al., 2018), the
group (Gao et al., 2008) but with an additional feature of polymerizable functionality of the hydrophobically modified polyacrylamide is
vinyl double bonds (Reb et al., 2000) in their molecular conformation, improved. Till date, the cEOR capacity of the low molecular weight
which imparts physicochemical properties that are more dynamic

17
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

(102–106 Da) polymeric surfactants have been investigated in a number surfmer created by adding 0.106 mol of 1-vinylimidazole drops into
of studies, although the findings are not conclusive, and requires further 0.106 mol of 4-dodecyl benzene sulfonic acid in presence of ethyl ace­
research. In addition, the EOR potentials have not been sufficiently tate (150 ml), which was kept in an ice immersion under N2 ambiance
substantiated by the coreflood or sand-pack tests, despite the fact that (El-Hoshoudy et al., 2017a). The first synthesis of vinyl monomers
the bulk rheology as well as surface and interfacial tension tests pro­ (Suresh KI and Bartsch E. 2013) was reported by Freedman et al. (1958),
duced positive findings. The lack of investigations on in-situ fluid-rock which were used as emulsifying agents (Summers and Eastoe, 2003;
interactions and displacement experiments, also made it difficult to Tsvetkov et al., 2014). Common polymerizable groups such as vinyl,
establish key tendencies in rheological and interfacial behaviors so far allyl, acrylate, methacrylate, styryl, and acrylamide have by employed
(Afolabi et al., 2022). for various uses. The position of the polymerizable group “H-type” on
Furthermore, Li et al. (2018) designed two new polymeric surfac­ the hydrophilic head group or the position of “T-type” polymerizable
tants, SP-1 and SP-2. These were synthesized by incorporating varying group in the hydrophobic tail have a major influence on the
hydrophobic groups on to polymer chains, and then the physio-chemical self-assembly and properties and property of the surfactant (Samakande
properties such as viscosity, rheology, salt tolerance and emulsification, et al., 2006, 2007).
were gauged. Micro-models and coreflooding tests were conducted to Therefore, surfmers are highly dynamic structural components that
identify the mechanism and the effectiveness of the surfmers during EOR are useful as polymerizable emulsifiers, as solid surface modifiers, and
application. Later, a field application was carried out in order to for synthesizing highly viscous water-soluble HAPAM polymers.
investigate the field applicability of polymeric surfactants for EOR. Re­ In addition, when surfmers are employed as hydrophobic monomers
sults showed that the viscosity of both the surfmers was considerable in and copolymerized with acrylamide (AM) it forms hydrophobically
comparison to a polymer with moderate molecular weight. The surfmer associating polyacrylamide (i.e., HAPAM), an extensively used chemical
SP-1 (with alkyl sulfonate functional groups) showed greater viscosity in enhanced oil recovery techniques, due to improved viscosifying
and salt tolerance in comparison to the SP-2 (with benzene sulfonic acid power and the ability of salt tolerance (Xue et al., 2004). These polymers
functional groups), whereas SP-2 exhibited superior emulsification. are developed by either micellar copolymerization of hydrophilic and
Based on the results of micro-models and coreflooding experiments, hydrophobic monomers (Camail et al., 2000) or via grafting
both SP-1 and SP-2 were able to achieve over 10% extra oil recovery cross-linking segments of hydrophobic chain onto the main chain of
factor following the tertiary polymer flooding, with SP-1 having a better HPAM. A large quantity of surfactants with small molecules are required
performance due to its higher viscosity than SP-2. Moreover, in the pilot so that the hydrophobic monomer is solubilized into micelles, even
test it was observed that, if a proper injection strategy was employed, though, these small molecule surfactants additives impart many un­
there could be considerable improvement in oil recovery using poly­ wanted effects (Guyot, 2004). In the process of HAPAM polymerization,
meric surfactant flooding (Li et al., 2018). the hydrophilic surfmers dissolve in an aqueous phase leading to a ho­
Hence, the polymeric surfactants are potential candidate for SP- mogeneous phase copolymerization of hydrophilic surfmers and acryl­
based cEOR techniques on account of their tolerance under HTHS, and amide (Gao et al., 2007), which does not require any surfactant additive.
shear, which can be further researched for low-cost, biodegradable al­ Moreover, when CMC of surfmer is surpassed, a micro block copoly­
ternatives to conventional polymers and surfactants that are much more merization mechanism is performed involving surfmer that is inserted
expensive and non-biodegradable. onto backbone of acrylamide main chain structure. This imparts greater
hydrophobic properties, viscosifying property of HAPAM, and enhanced
5.1.2. Surfmer and hydrophobically associating polyacrylamide polymers tolerance of HAPAM towards salinity (Gao et al., 2004; Yasin et al.,
(HAPAMs) 2015). The surfmer cannot separate from the polymer chain due to the
Contrary to conventional surfactants, the surface activity of surfmers surfmer being copolymerized with monomer and attached in its main
is akin to vinyl monomers in its capability of being initiated and poly­ chain (Xu and Chen, 2004). These enhanced stability properties of these
merized. Surfmers are of high importance in cEOR (Wang and Wu, 2015) polymers (Texter, 2001) imply greater mechanical stability (Borzenkov
owing to their amphipathic property and polymerizable characteristics. et al., 2015), electrolyte stability of the latex (Atta et al., 2014), reduced
Moreover, surfmers can be employed for several processes, such as, surfactant migration (Xu and Chen, 2004), and surface charge density
in synthesizing inorganic/organic nanocomposite, surfmers can be used control (Borzenkov and Hevus, 2014).
as polymerizable emulsifiers, as surface modification of solid substances, HAPAMs have become the subject of many research and industrial
and more importantly in generating novel water-soluble hydrophobi­ laboratories for SP-based cEOR (Bastiat et al., 2004; Feng et al., 2005).
cally associating polymers (HAPAM) that have exceptional viscous They have potentially advantageous characteristics such as the capa­
power (Summers and Eastoe, 2003). Fig. 14 shows the structure of a bility of the hydrophobic groups to form intermolecular hydrophobic
associations in nanodomains in water-based solutions, when a critical
association concentration (C*) is surpassed (Niu et al., 2001). This
causes the formation of a 3D-transient network structure (Candau and
Selb, 1999; Niu et al., 2001) of heightened ionic strength medium,
lending exceptional viscosifying ability (Shashkina et al., 2003; Gao
et al., 2005), better rheological characteristics, and enhanced tolerance
towards salts compared to the unmodified HPAM (Maia et al., 2009). It
can cause IFT reduction at the rock/fluid confluences because of hy­
drophobic moieties that form aggregates or micelles. These polymers
have exhibited a singular kind of adsorption isotherm (Hu et al., 2015)
which is seen as a wettability altering capacity.
These polymers resist mechanical degradation due to high shear
stress which commonly occur in pumps and the surrounding of the well
bore area, the reason being that the physical links between chains are
disrupted prior to any irreversible degradation and these polymers re­
Fig. 14. A structural illustration of the synthesis of a Surfmer by combining form and can maintain their viscosity when shear decreases (Evani and
0.106 mol of 1-vinylimidazole drops with 0.106 mol of 4-dodecyl benzene Rose, 1987). They have shown very good tolerance towards tempera­
sulfonic acid in presence of ethyl acetate (i.e., 150 ml), under N2 bath (modified ture, pH, and ion content in the periphery of the wells.
after El-Hoshoudy et al., 2017a). Thus, HAPAM polymers are potential candidates for SP based EOR,

18
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

by functioning as viscosifiers or wettability modifiers in HTHS and high- exceptional non-Newtonian behaviors (i.e., shear thinning behavior).
pressure reservoirs (Rodrigues et al., 2006; Gouveia et al., 2008; Zhong These correspond well to in situ reservoir stimuli such as temperature,
et al., 2008). They may also be used for reservoir simulation models and ionic strength, pH, and shear stress and also, have thermal, rheological,
for tertiary oil recovery processes (Ranjbar and Schaffie, 2000). and salt resistant capability in reservoir conditions, apart from
enhancing the sweeping efficiency as well. Their behavior is very similar
5.1.2.1. Laboratory experiments of surfmer and HAPAMs. El-Hoshoudy to IFT agents such as surfactants, as these also reduced the IFT to ul­
and co-workers (El-Hoshoudy et al., 2017b) authored works on hydro­ tralow values, which mobilizes residual crude oil. Upon evaluating the
phobically associating polyacrylamide (HAPAM) and its modified wettability using static sessile drop method, it was observed that
nanocomposite derivative, developed through free radical emulsion HAPAM copolymer and HAPAM-SiO2 nanocomposite were capable of
polymerization. The chemical structure of these latexes was determined altering rock wettability from oil-wet to water-wet, which significantly
using various methods viz., FTIR (Fourier-transform infrared spectros­ increases recovery factor based on the unanimous agreement of petro­
copy), 13C-NMR (13C Nuclear Magnetic Resonance), 1 H-NMR leum engineers that more oil is recovered from water-wet reservoirs
(Hydrogen 1 Nuclear Magnetic Resonance), SEM (scanning electron than oil-wet reservoirs. According to the team, this was the first ever
microscope), X-ray diffraction and TEM (transmission electron micro­ observation of polymers that altered sandstone rock wettability and the
scope). The DLS or dynamic light scattering was used to establish the recovered oil amount reached 26% from OOIP.
particle sizes and distribution of particle sizes, whereas to determine the Thus, this novel copolymer and its nanocomposite may be used as
thermal characteristics, TGA (thermal gravimetric analysis), and wettability-modifying agents in SP based cEOR in HTHS oil fields unlike
through the aid of DSC (differential scanning calorimetry) methods the commercial polyacrylamides, which renders them very high-
(El-Hoshoudy et al., 2015a, 2016). potential candidates for industrial scale energy supply via EOR appli­
Fig. 15 illustrates the structure and the synthesis of a HAPAM-SiO2 cations. For applicability in HTHS carbonates, more conclusive research
nanocomposite, with “n” number of the copolymer chain unit may be needed.
(El-Hoshoudy et al., 2017a). The synthesis of this polymer was done via
polymerization method, in which 3-aminopropyl triethoxy silane and 5.1.3. Hot-Melt Adhesive (HMA) polymers
KPS-initiator were independently incorporated at the identical time at Hot-melt Adhesive (HMA) polymers are categorized according to the
60 ◦ C reaction temperature (El-Hoshoudy et al., 2017b). Then rheo­ hydrophobic moieties into four groups, i.e., i) the associative polymers
logical and solution properties studies were conducted under virtual with a single associating block, ii) telechelic, iii) multisticker, and iv)
reservoir conditions, based on concentration of the polymer, salinity and combined HMA polymers. When HMA polymers contact surfactants,
temperature of reservoir, and the shear rate. Results showed decent viscosity of the HMA polymers is either enhanced or reduced, based on
robustness towards temperature and salinity, good IFT reduction and the concentration of the added surfactant and the competition between
better viscosity features (El-Hoshoudy et al., 2017a). intermolecular and intramolecular relations. The adsorption of HMA
The evaluation of the rheological properties included tolerance to­ polymers initially decreases with increasing salt concentration, but
wards salinity, temperature, polymer concentration, and shear robust­ above a certain concentration the adsorption increases with increasing
ness. Corefloods were conducted using linear pressurized packed model salinity. Thus, due to desirable rheological properties and low adsorp­
(El-Hoshoudy et al., 2015b, 2016, and 2017a). The team concluded the tion of HMA polymers with random monomer distribution rather than
following from the investigations: That the HAPAM-SiO2 nanocomposite HMA polymers with blocky microstructure which form hydrophobic
was developed by incorporating silica nanoparticles through a single associations (Tanaka et al., 1992; Huang and Santore, 2002), researchers
shot synthesis using the Aza-Michael addition reaction, to alleviate the are attempting to extend applications of HMA polymers to HTHS car­
shortages that occur due to agglomeration and coagulation of modified bonate reservoirs.
silica particles as the emulsion polymerization reactions are undergoing.
This HAPAM copolymer and HAPAM-SiO2 nanocomposite had the 5.1.3.1. Laboratory experiments of HMA polymer. HMA polymer type
exact desired characteristics such as retaining the original viscosity and was created by Zhong et al. (2009) by modifying the AM hydrophilic

Fig. 15. Illustration of HAPAM-SiO2 nanocomposite copolymer structure, synthesized by polymerization of 3-aminopropyl integrated with triethoxy silane and KPS-
initiator at the same period when reaction temperature was at 60 ◦ C (modified after El-Hoshoudy et al., 2017a).

19
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

backbone of a terpolymer with AMPS and hydrophobic monomer reservoirs.


2-vinylnaphtahlene was synthesized. This was intended for medium and Therefore, this section details and reviews the latest surfactant-based
low permeability oil reservoirs with HTHS. Compared to PAM, this field projects successfully conducted in carbonates. Moreover, we
terpolymer displayed better firmness and its viscosity increased with the discuss the essential lessons learned from these recent SP-based cEOR
increase in both salts concentration and temperature during laboratory field trails, which would ultimately improve the operative know-how for
testing. In another study, an HMA was designed by Gouveia et al. successful SP-flooding designs and field projects in HTHS carbonate
(2009), where an associating PAM using poly (propylene glycol) Mon­ reservoirs.
omethacrylate was synthesized as the hydrophobic oligomer through a
micellar copolymerization using an SDS as a surfactant. A
thermo-thickening behavior was exhibited by the HMA polymer in the 6.1. Field applications for SP-based cEOR
presence of SDS. It was suggested by the authors that low IFT could be
expected owing to the presence of the surfactant although experimental The following are examples of successful SP-based cEOR field ap­
evidence was not provided. The above-mentioned characteristics render plications in carbonate reservoirs, along with some relevant stand-alone
HMA polymers a potential candidate for possible SP-based EOR appli­ surfactant and polymer flooding field trails that have been effective in
cation in HTHS carbonates, although more dedicated studies and achieving their inception goals.
required.
6.1.1. Abu Al bukhoosh (ABK), Abu Dhabi, UAE (Al-Amrie et al., 2015)
5.1.4. Other polymers formulations A one-spot surfactant-polymer (SP) pilot in Middle East (i.e., Abu
In an investigation by Klimenko et al. (2020), Superpusher™ SAV10 Dhabi, UAE) was successfully conducted. The offshore target field,
(SNF Floerger), a terpolymer with functional groups of acrylamides known as Abu Al Bukhoosh (ABK), was a high-temperature, high-
(AM), Acrylamido-Tert-Butyl-Sulfonate (ATBS), and N-Vinyl-­ salinity, low-permeability carbonate reservoir. The oil-bearing carbon­
Pyrrolidone (NVP) was used for SP based EOR injections and all of them ate reservoir is of a Lower Cretaceous age, located 180 km offshore of
exhibited good thermal stability in the presence of divalent cations Abu Dhabi, and since 1974, it is operated by Total (Al-Amrie et al.,
(Masalmeh et al., 2019). These were all of medium molecular weight. 2015). The SP-formulation was synthesized by Total Research and
The polymer solutions were pre-sheared with a mixer, in order to mimic Development (R&D) labs, according to a new, hydrophilicity-enhanced
degradation at surface facilities. The pre-shear before injecting also molecule that is tolerant to reservoirs with high-temperature, high-­
helps to achieve better polymer injectivity by approximately 30% salinity. Moreover, a new technique was fostered for optimization of
decrease in viscosity (at 10 sec− 1) (Jouenne et al., 2019). The concen­ polymer molecular weight distribution, with the aim of in-situ viscosity
tration of polymer used was 2350 ppm in SP slug and 5300–5400 ppm in decoupling from near-wellbore injectivity. The final synthesized
Polymer slug. The number of viscosity measurements is difficult to formulation contained 0.4 pore volume (PV) SP slug with 1.35% of
reduce at such high temperatures, the extended model of Jouenne et al. active surfactant, plus 1% clarifier, and SAV-225 polymer (SNF
(2019) aided in predicting the polymer concentration essential to attain Floerger) in an 80 g/L brine equivalent to a hypothetical softened
the target viscosity. Whereby, upon measuring the differential pressure mixture of seawater and water from local aquifer. Subsequently, a
value on the capillary installed at the core inlet, there was noticed polymer drive of softened seawater consisting of AN-125 polymer (SNF
excellent concurrence with the calculated value. Viscosity measure­ Floerger) was injected, for enforcing a negative salinity gradient amidst
ments helped in evaluating the stability. The recorded pressure profiles the 230 g/L formation brine, 80 g/L SP slug, and 42 g/L seawater. There
taken during the cEOR flooding, showed no evidence of polymer accu­ was no requirement of a pre-flush owing to the SP-formulation design
mulation and plugging, even if the injection time could have been too and its application. In the analog limestone cores, the measured value
short for this phenomenon to ensue. Fairly low polymer adsorption is for the residual oil saturation to chemicals (Sorc) was 5% ± 2%, which
expected, based on the observation that aqueous phase of the effluent was consistent with a recovery factor (RF) of 90% ± 4%.
turned viscous instantaneously after the polymer bank arrived. The
temperature for experiments was between 105 and 120◦ and the for­ 6.1.1.1. Key implications. Despite, the one-spot pilot test recovery
mation water salinity was 297,000 TDS ppm (Klimenko et al., 2020). agreed strongly with the recovery in analog cores with Sorc of 4%, and
Thus, most re HPAM, AM, ATBS, and NVP based polymers, proved to be recovery factor = 90% (Al-Amrie et al., 2015). However, it was noted
tolerant towards HTHS conditions during laboratory trails. that, due to the substantial heterogeneity on the core-scale of the
In Appendix-A, Table A-3 summarizes several novel polymers for­ reservoir limestone, the oil bank formation was probably thwarted,
mulations, which can be suitable for HTHS conditions in carbonate which resulted in smaller recoveries of Sorc: 13% ± 2%, RF: 84% ± 4%.
reservoirs during SP-based cEOR. The cause for the lesser recovery in reservoir cores could be mainly due
to the short length of the core relating to the mixing zone (Levitt et al.,
6. Field applications for SP-based cEOR and lessons learned 2012). Nonetheless, the injectivity of water injection preceding the
Single Well Tracer Tests (SWTT) was considerably lower than the ex­
Large field trials of SP-based cEORs have been implemented world­ pected rate derived from the design process due to the low formation
wide since 1980 (Ragab and Mansour, 2021). Approximately 32 permeability. Also, the surfactant-slug had to be diluted in order to
SP-based cEOR field applications had been conducted worldwide; mitigate this injectivity issue. In addition, controlled shear degradation
however, that number has risen since, with more recent reported works would optimize the distribution of polymer molecular weight for
of SP-based cEOR field pilots. These field applications show that on maximized injectivity in tight reservoirs (Jouenne et al. 2019). The
average, SP-flooding can achieve over 15–25% OOIP (Liao et al., 2017). ultra-high molecular weight polymer fraction can be eradicated by
Nevertheless, most of the target oilfields have been sandstone forma­ employing a minimum of 20–25% (by viscosity) of shear-degradation,
tions, while the field trails for carbonate oilfields are far less in number, which is executed on a moderate molecular weight polymer, instead
especially of carbonates under harsh conditions (HTHS). Among these of a low-molecular weight (MW) polymer. Notwithstanding, in this pilot
few SP field applications in carbonate reservoirs are Abu Al Bukhoosh case, surfactant alone was adequate to desaturate oil in the
(ABK), Abu Dhabi, UAE (Al-Amrie et al., 2015) and Raudhatain Lower near-wellbore area; granting that mobility control could still be required
Burgan Reservoir, Kuwait (Al-Murayri et al., 2018, 2019a, and 2017). for attaining satisfactory sweep efficiency when reservoir scale appli­
Thus, meticulous efforts have been put into improving SP-formulations cations were conducted. The two thermally stable polymers selected
by synthesizing novel surfactants and polymers especially designed to showed poor injectivity in the in low-permeability carbonates, thereby
have higher temperature and salinity tolerance in HTHS carbonate could not by completely assessed owing to the tested well having low

20
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

injectivity. Even when MW distribution modification was done for 6.1.3. Offshore, south China sea (Harris et al., 2020)
SAV-225, it did not suffice for the 5-mD formation, still, further observed Shell directed a single well chemical tracer (SWCT) test using a low
justifications of injectivity loss could be possible. On the other hand, the complexity SP-formulation for offshore environments in the South China
optimized AN-125 polymer, was injected effectively in a 2 mD core, Sea (Harris et al., 2020), which is mainly composed of organic reef li­
indicating its utility in low-permeability carbonate reservoirs. These thology. The SP formulation was seawater-based to prevent the water
polymer injectivity trails were conducted concurrent to the surfactant treatment complexities from occurring when under commercial appli­
slug injection operations in the1-spot pilot area. Thus, regardless of cation circumstances (Southwick et al. 2019). Following the
having no mobility control due to lack of polymer, the developed sur­ SP-injection, the residual oil saturation decreased from 20-27% to 7–9%
factant slug efficaciously mobilized the oil in the near-wellbore area. range. The SP-formulation was composed of an alkyl propoxy sulfate
However, surfactant retention was undesirably high with 0.5–0.7 mg/g (APS) and an alkyl ethoxy sulfate (AES) mixture devoid of any cosol­
of rock. A cost-effective project may be designed by achieving surfactant vents. In the case of application with seawater only (no salinity gradient)
reduction via decreasing the phase trapping via cosolvents or additives the APS and AES combined formulation yielded exceptionally greater oil
(Tagavifar et al. 2016). Surfactant retention can also be restricted recovery in comparison to the APS and internal olefin sulfonate (IOS)
through diminishing the electrostatic interactions with calcite minerals mixture in the selected outcrop. The SP-formulation utilized seawater
via swapping the cationic surfactants and polymers with other appro­ without any supplementary water treatment other than those that are
priate surfactants and polymers (Levitt et al., 2016). It is also worth typically employed for water flooding such as, filtration,
mentioning that specifically designed topsides were dedicated and de-oxygenation, and such (Harris et al., 2020).
installed for this SP cEOR pilot plan. Moreover, an analysis of the results
obtained with the analog core, reservoir core, and one-spot pilot, shows 6.1.3.1. Key implications. From the field test, it was realized that the
that whenever the heterogeneity of the core-scale thwarts oil bank for­ seawater-based formulations can be efficiently combined on a barge and
mation, the functioning of homogeneous analog cores could be better injected to a production well installed on platforms. It was observed that
predictors of the pilot performance in the field. Thus, this pilot project the highest oil recovery was obtained through using surfactant mixtures
proved the efficacy of SP-based cEOR in HTHS carbonate reservoirs, that produce formulations, which are below optimum (i.e., Winsor Type-
through appropriate SP-formulations. 1 phase behavior) with reservoir crude oil. Correspondingly, these
below optimum formulations evade high injection pressures that are
6.1.2. Raudhatain Lower Burgan Reservoir, Kuwait (Al-Murayri et al. observed for optimum formulations in low permeability outcrop rocks.
2019a) The recovery from these SP-formulations is an identical amount of oil in
SP-based cEOR was implemented in Raudhatain Lower Burgan the swept zone of the reservoir rock. The total recovery in the reservoir
(RALB) reservoir in Kuwait that has 82 ◦ C temperature and in-situ rock is lower than that in the selected outcrop because of more hetero­
salinity of over >200,000 ppm with light-oil. A Single Well Chemical geneity that leads to crude oil getting bypassed (Harris et al., 2020;
Tracer Test (SWCTT) was performed with the objective of determining Southwick et al. 2019).
the efficacy of lab-optimized surfactant-polymer formulation (Al-Mur­
ayri et al. 2017 and Al-Murayri et al., 2018) for the target reservoir. The 6.1.4. Block X in tahe oilfield, China (Yin et al., 2021)
SWCTT design work would validate the optimal injection/production In the channel reservoir of in Tahe oilfield, China, which has high-
sequence identified during the core scale floods regarding minimal temperature of 100–110 ◦ C, high-salinity of over >2.1 × 105 mg/L,
volumes, rates, and duration. There were several sensitivity scenarios and is part of thick carbonate strata. In order to conform, regulate and
tested in order to establish the ambiguities. Subsequently, the field enhance the oil recovery, the combined flooding was conducted by
application of SWCTT took place and the results were evaluated and employing dispersed particle gel (DPG) and surfactant. The pilot test
matched with the formerly attained results of laboratory as well as comprised of an introduction to the geological backgrounds and devel­
simulation tests. The SWCTT was conducted with the purpose of vali­ oping history of the block X. In combined flooding, the DPG is useful as
dating the polymer and surfactant solutions efficiencies in relation to the an in-depth conformance control agent in order to enhance the sweep
reduction in residual oil saturation and injectivity of water. For this, the efficiency, and a subsequent surfactant solution slug is utilized for
before and after the chemical flooding values of residual oil saturation displacement efficiency improvement. The long-term stability of DPG
were appraised during which the injection rates and corresponding lasts 15 days, which warrants the efficacy of in-depth conformance
wellhead pressures were continuously monitored. The injection control. Wherein it increases from its original 0.543 μm–35.5 μm
sequence of the SWCTT included the preliminary water-flooding, then following the 7 days of aging in the 2.17 × 105 mg/L reservoir water at
tracer injection followed, with subsequent phases of soaking and pro­ 110 ◦ C reservoir temperature. The optimization process revealed that
duction to measure the oil saturation after water flooding. The pre-flush 0.35% NAC-1+ 0.25% NAC-2 surfactant solution having IFT of 3.2 ×
was subsequently trailed by a main-slug containing surfactant concen­ 10− 2 mN/m can easily create a moderately stable emulsion with the
tration of 5000 ppm and polymer concentration of 500 ppm, which was dehydrated crude oil. During the double core flooding, the conformance
followed by a post-flush with the polymer only. Next, with in order to control performance is validated through the fluid diversion following
quantify the oil saturation after chemical flooding, sea-water push, fol­ the combined flooding leading to incremental EOR of 21.3% conse­
lowed by tracer injection, soaking and production were conducted quently. Following the 14 years of the utilization of Block X, the for­
(Al-Murayri et al., 2019a). mation energy declined rapidly owing to deficiency of large bottom
water. While, the water fingering also resulted in a reduced production
6.1.2.1. Key implications. The result of simulation work before the rate and water cut increase. Next, following the combined flooding in Y
SWCTT test had been conducted, had resulted in positive oil desatura­ well group comprised of 1 injector and 3 producers, there was marked
tion levels following chemical flooding within 10 ft from the well. Then increased in the average dynamic liquid level, daily production, and the
again, the pilot results showed that a difference existed between the real breakthrough time of the tracing agent. While there was noted a
SWCTT results and the lab and simulation results attained beforehand. decrease in the water cut and injectivity index. Also, the injected fluid’s
The probable causes for the noticeable difference amid the real field data distribution rate and real-time monitoring verified the performance of
and lab/simulation results are under investigation to optimize the SP- conformance control and the oil recovery from this group of wells was
based cEOR for improved oil recovery from the aged reservoirs (Al augmented by an excess of 3000 tons of oil (Yin et al., 2021).
Murayri et al., 2019a).
6.1.4.1. Key implications. The examination of this combined flooding

21
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

from laboratory experiments to field case analysis, showed that the in April 2019, with continuous mode surfactant injection though 2
heterogeneity was adjusted via DPG used in conjunction with EOR corresponding long-horizontal production wells at 4000 ppm of active
surfactant to increase displacement efficiency, and occurs synergisti­ concentration in sea water. The injection continued for 7 months with
cally in this HTHS carbonates. The selection and performance standards the entire injection volume of 1000 MT. The surfactant injection ended
of the combined flooding also provided some hands-on experiences and in November 2019, and subsequently seawater injection was carried out.
principles for combined flooding (Yin et al., 2021). The objective of the project was mainly to estimate the field scale effi­
cacy of the developed solution (Rohilla et al. 2021).
6.1.5. Baturaja formation, semoga field, Indonesia (Rilian and Sumestry,
2010; Sheng, 2013b) 6.1.6.1. Key implications. The outcomes of the surfactant flooding pilot
The Semoga field reservoir rock located in Indonesia is highly oil-wet demonstrated that there was no injectivity loss or viscous emulsions
with initial water contact angle measured as 78◦ . The Baturaja Forma­ formation in the reservoir from the surfactant injection, indicating the
tion, a carbonate reservoir in the Semoga field had a confirmed volume developed surfactant formulations efficacy and stability under the car­
of approximately 317,856 acre-ft (77 ft net pay thickness), and was in bonate reservoir conditions. One of the injectors (Injector-01), in which
low production due to depletion. The formation has 127 wells, including surfactant had been injected, achieved 8% yearly water injectivity on
17 shut-in wells, 28 water injection wells, and 82 producing wells. In account of the surfactant injection. Furthermore, both the producer
this project surfactant works and wettability changes acquired in the wells for the pattern exhibited incremental oil response. So far, no sur­
laboratory were mirrored in the field pilot, and increased oil mobility factant breakthrough from produced water analysis have been observed.
and water-wet wettability state to a certain degree were observed. A Further details of the project, the chemicals used and acquired experi­
non-ionic surfactant was used for the field application for the reason that ences from the field project may be found elsewhere (Rohilla et al.
the homogeneity of reservoir could not accommodate a cationic sur­ 2021).
factant. The wettability changes in the laboratory tests coincided with
those observed in the field application, i.e., the wettability changed by a 6.1.7. Bashkirian reservoirs, bobryk, Ukraine (Varfolomeev et al., 2020)
small degree from oil-wet towards water-wet, thereby improving The Bashkirian dolostone reservoirs of Bobryk oil and gas condensate
mobility of oil. Further studies are ongoing to analyze the efficacy of the field in Ukraine is a fractured carbonate reservoir with heavy oil. The
surfactant work in the field project, to secure the water cut reduction reservoir has low temperature of 23 ◦ C and an ultra-high salinity of
using the surfactant, and to prevent reverse conning and other unde­ 217.5 g/L, and low injectivity restricts the reservoir’s effective devel­
sirable effects. Surfactant stimulation resulted in a reduction of water opment. Thus, the efficiency of surfactant injection (developed in prior
cut of around 8% and an increase in total oil production for this project laboratory experiments) was investigated through a pilot test. A
of roughly 5800 bbls over the course of three months. Further investi­ nonionic surfactant was selected for the pilot test based on its demon­
gation is underway to determine the effectiveness of the surfactant work strated capability to cause IFT reduction and wettability alteration, as
in this project and to confirm if the reduction in water cut was caused by well as its oil displacement efficacy, ability to reduce injectivity, pro­
the surfactant and no other effects, such as reverse conning etc. To duction performance, cost-effectiveness, and availability (Varfolomeev
reduce “subjectivity” in pressure build-up (PBU) analysis, the results et al., 2020). During the pilot test the nonionic surfactant exhibited
have to be verified using the techniques such as, coring analysis, time- excellent influence on water injectivity improvement. Following the
lapse saturation logs, and single well tracers, which are used to eval­ surfactant injection, the decrease in the wellhead pressure and down­
uate the influence of surfactant stimulation on changes in reservoir hole pressure ranged from 40 at and 171.5 at to 24 at and 153.4 at,
characteristics. In conclusion, the increased oil output was around 5800 respectively. The incremental oil production due to the surfactant in­
bbls over the course of three months. The net present value (NPV) at jection was approximately 1094 tones.
three months was $24,013; at one year, it was 58,189; and in 11 days the
pay-out time (POT) of the project was $15,180. Hence, the outcomes 6.1.7.1. Key implications. The successful effect of the selected nonionic
were optimistic, and the oil recovery was expected to be more sub­ surfactant projects its superb potential as a candidate for SP applications
stantial if the technique i.e., surfactant huff and puff method was to be in high salinity carbonates. Each surfactant was measured at three
applied in high production wells instead of high water cut wells (Rilian different concentrations while screening so that the optimum surfactant
and Sumestry, 2010). formulation was acquired. Whereas, it is expected that using the sur­
factant will result in well-injectivity enhancement along with increased
6.1.5.1. Key implications. The surfactant stimulation resulted in inten­ oil displacement efficacy for composite heavy oil carbonate deposits
sification in skin damage in the X-2 well, which was observed through (Varfolomeev et al., 2020). Therefore, this benchmark study strengthens
the pressure build-up (PBU) test analysis (Rilian and Sumestry, 2010). the technological basis for furthering the field application plans of
Nevertheless, the decrease in water cut in both the wells was attributed surfactant-based flooding towards harsh conditioned carbonate
to the surfactant stimulation rather than any other cause (such as reverse reservoirs.
conning). In conclusion, the surfactant works and the wettability alter­
ations in the laboratory corroborated with the field project, in which the 6.1.7.2. Lower Shaunavon oilfield, Canada (Stephenson et al., 2019). In a
oil mobility had increased and the wettability conditions had been separate field application case, microemulsion additives were employed
transformed from oil-wet towards slightly water-wet state (Rilian and in Lower Shaunavon oilfield located in SW Saskatchewan, Canada. The
Sumestry, 2010). Shaunavon oilfield is classified as unconventional oilfield based on
permeability, temperature, calcite mineralogy, and oil with higher
6.1.6. Block 5, Al-shaheen field, Qatar (Rohilla et al. 2021) content of asphaltene, along with high-salinity formation water. The
A chemical based EOR application in Block 5, Al-Shaheen Field, ability of microemulsion additive to function as delivery platform for
Qatar, which is a giant offshore carbonate reservoir was performed using solvent (terpene) and surfactant mixtures across a given rock volume
a surfactant formulation with wettability altering capability. The car­ was used as a basis for the analysis There are numerous formulations
bonate reservoir has low permeability (i.e., 5 mD), and is extended with available, each of which is customized for a specific set of reservoir and
elongated horizontal wells with injection well and production wells in fluid properties (Stephenson et al., 2019). Microemulsion additives
an alternate sequence. Surfactant injection method for cEOR was sys­ lessen the impairment and transform the energetics of surfaces, and can
tematically conducted in stages to reduce the risks involved in field-wide improve oil mobilization. The optimized microemulsion additives were
cEOR applications (Rohilla et al., 2020; 2021). The project was initiated created according to the fluid-fluid and fluid-rock interactions for the

22
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

target reservoir through laboratory experiments. The increases in oil summarized in Table A-5 in Appendix-A.
production from the field trail were attributed to changes in injectivity
of water, which was the key markers of success for both core floods and 6.2. Essential lessons learned (i.e., key points summary)
pilot tests (Stephenson et al., 2019).
• Evaluation of Chemical SP-Formulation Synthesis. It was Low
6.1.7.3. Key implications. The core flood response in unconventional complexity chemical combinations for SP-formulations are prefer­
reservoirs was insignificant with <5% incremental oil recovery able, such as the surfactant formulations that used seawater without
observed, partly attributed to the asphaltenes coating formed on the any other water treatment. Besides, the usually performed water
surface of the core during wettability restoration. Nevertheless, the pilot treatment viz., filtration, and de-oxygenation methods are preferable
response was optimistic with high recovery. Through evaluating this too. Novel hydrophilicity-enhanced molecule are stable under HTHS
inconsistency, it was inferred that using dependable screening tools for carbonate reservoirs. Moreover, no pre-flush is required for SP-
unconventional reservoirs was necessary prior to core flood tests. formulation designs that use appropriate designs such as, softened
Moreover, it was clear that the microemulsions had positive influence in seawater mixtures. Novel polymers proven to be stable under harsh
conventional as well as unconventional reservoirs, which could be carbonate environments include hydrophilicity-enhanced surfac­
further improved with optimization according to fluid-fluid and fluid- tants; SAV-225 (AM-AMPS-NVP terpolymer), AN-125 SH (AM-AMPS
rock interactions within the crude oil-rock-brine system. More details copolymer) and a polyacrylamide-based polymer with high 2-acryl­
of the chemicals used, adopted mythology and field operation data may amido-tertiary-butyl sulfonic acid (ATBS) content called SAV 10.
be found in the given publications (Stephenson et al., 2019). Also, a new technique cultivated for optimization of polymer mo­
lecular weight distribution can achieve in-situ viscosity decoupling
6.1.8. Abu Dhabi, UAE (Baloch et al. 2021) from near-wellbore injectivity.
The Abu Dhabi National Oil Company (ADNOC) planned for a • Injectivity of SP-Based cEOR. It was observed that nonionic sur­
Polymer-based cEOR, with the intention of enhanced displacement and factant shows excellent effect on improving water injectivity, which
sweep efficiencies in carbonate reservoirs. These polymer-based EOR can achieve enhanced well-injectivity accompanied by increased oil
methods included polymer injections (polymer injection test/PIT), displacement efficiency for HTHS carbonate reservoirs. For low
simultaneous injection of miscible gas and polymer (SIMGAP), simul­ permeability formations, the injectivity of water injection preceding
taneous injection of water and polymer (SIWAP), low salinity polymer, the Single Well Tracer Tests (SWTT). In certain scenarios of low
and others. The target reservoirs in Abu Dhabi are carbonate reservoir permeability outcrop rocks, below optimum formulations. Can avoid
formation with high temperature (~100–130 ◦ C), high salinity forma­ the high injection pressures that are observed for optimum formu­
tion brine (~200,000 ppm), along with high concentrations of divalent lations and are more preferred in such conditions was considerably
ions brine in the formation brine (~18,000 ppm of Ca++ and Mg++). lower, where in diluting the surfactant slug may mitigate this
Since, ADNOC planned on widening the polymer flooding scope towards injectivity problem. Additionally, for oil-wet carbonates reservoirs
harsher oilfield conditions, which had higher content of H2S. Conse­ with low permeability, water injectivity can be improved via sur­
quently, a PIT was designed to prepare for a multi-well pilot test, in factant injection by wettability alteration towards water-wetness.
which polymer flooding is applied to dissolve H2S concentrations of Furthermore, the use of controlled shear degradation optimizes the
20–40 ppm and to substantiate the injectivity at typical field conditions, polymer molecular weight distribution, which leads to maximized
and in-situ polymer functioning. As stable polymers in these harsh en­ injectivity in tight reservoirs such as, carbonate reservoirs with low-
vironments, experimental studies identified a polyacrylamide-based permeability formations.
polymer with high 2-acrylamido-tertiary-butyl sulfonic acid (ATBS) • SP-Based cEOR Injection Equipment. The optimal chemical in­
content called SAV 10. The SAV 10 polymer was stable in the presence of jection facilities for SP offshore oilfields that use the seawater-based
H2S (500 ppm) and/or O2 (up to 150 ppb) as well. Also, this ATBS formulations are the portable injection skid-mounted units on the
polymer had favorable injectivity behavior that could be easily modified platform applicable in the case of shallow water fields. In the case of
for injection according to the target reservoirs settings (Masalmeh, S. deepwater fields, the floating production storage and offloading
et al. 2019). In 2019, a primary polymer injectivity test (PIT) was per­ (FPSO) system is preferable.
formed at 121.1 ◦ C and salinity of over 200,000 ppm, as well as at low • involve combining it on a barge and then injecting it to a production
H2S content (Rachapudi et al., 2020; Leon et al., 2021). The PIT was well installed on platforms. In other words, these are the portable
completed over a duration of five months, i.e., February 2021 to July injection skid-mounted unit on the platform applicable in the case of
2021, and then was followed by chase water flood till December 2021. shallow water fields (Han et al. 2006). Another facility is the polymer
During the PIT, the overall polymer solution injection volume was 108, preparation in maturation tank on FPSO and the polymer injection
392 barrels (Baloch et al. 2021). pump installed the platform and injection facilities on the FPSO for
deepwater fields.
6.1.8.1. Key implications. Initial outcomes from the PIT indicated the • Timing to Start a SP-Based cEOR Project. The appropriate sched­
key performance indicators had been attained, including an expected uling for the commencement of a SP-based cEOR field trial requires
viscosity yield and beneficial injectivity at target rates that matched further research. In general, at the earliest stage of expansion of the
with the laboratory data. The near-wellbore polymer behavior was SP field operations the anticipated favorable results are observed,
ascertained using a down-hole shut-in tool (DHSIT) to measure the nonetheless the base case reference in water flooding is absent.
pressure fall-off (PFO) data, in order to help in the PIT program opti­
mization. It is anticipated to enable the polymer cEOR assessments of 7. Overcoming challenges in SP-based cEOR: future perspectives
many other massive heterogeneous carbonate reservoirs in the middle
east, which possess harsh conditions. Hence, based on the properties of As stated previously, carbonate reservoir is characterized by high
novel polymer established over the course of the PIT, the optimization of temperature of over 85 ◦ C, high formation water salinity above 100,000
the models for field and sector can be done in the future (Baloch et al. ppm (TDS), divalent cations (hardness) above 1000 ppm, and perme­
2021). ability lower than 100 mD, and make up nearly 60% the world’s oil
Finally, a list of published field studies on SP-based cEOR for car­ assets (Akbar et al., 2000). Critical problems are produced due to these
bonate reservoirs has been summarized in Table A-4 in Appendix-A. harsh conditions when applying oil recovery techniques including ter­
Likewise, some of the chemical-based cEOR projects worldwide are tiary recovery methods of cEOR. In the last decade, the SP-based cEOR
technologies have been further developed to overcome many of the

23
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

challenges of harsh carbonates reservoirs, which are evidenced through zone’’ that cause early water/gas breakthrough and loss of oil produc­
a number of successful experimental and field investigations. However, tion. To surmount this issue, mobility control agents, such as foam, may
on account of the complex carbonate reservoir characteristics, which be taken as a practicable measure (Talebian et al., 2014, 2015). Foam is
also include fractures, heterogeneity and oil-wet wettability, the SP in­ suggested for countering the channeling, viscous fingering, and gravity
jection cEOR applications, may not always be successful. override problems during Gas (e.g., CO2 and N2) flooding, which occurs
Therefore, in this review, we enumerate the primary challenges of due to carbonate matrix heterogeneity (Taber et al., 1997; Masalmeh
SP-based cEOR application in carbonate reservoirs, and discuss the new et al., 2011; Lake et al., 2014; Yuncong et al., 2014). The Foam Assisted
and upcoming technologies and materials (i.e., chemicals) that mitigate Water Alternating Gas (FAWAG) injections are proposed as measures to
these issues, to pave the way for more efficient SP-flooding imple­ improve volumetric sweep efficiency and lower the gas-oil ratio, leading
mentations in HTHS carbonate rocks. These advanced approaches can to effective oil recovery (Aarra et al., 2002; Blaker et al., 2002; Masal­
tackle chemical instability, adsorption on carbonate surface, as well as meh et al., 2011, 2012; Tunio et al., 2012). If the foam films can
environmental and economic concerns. Also, the novel SP formulations maintain stability during propagation, a successful FAWAG injection can
are able to withstand HTHS conditions of carbonates and create very low be performed. However, foam stability in the presence of oil is still a
retention on carbonates, while inducing IFT reduction and wettability challenge during foam injection (i.e., FAWAG), which requires consid­
alteration of carbonate rock from oil-wet to water-wet state. Moreover, erable attention. Based on this concept, a novel hybrid EOR technology,
the synergic effect created through combining low salinity water, smart- which merges smart water (i.e., ionically modified composition brine)
water, and foam flooding with SP-flooding, by way of hybrid SP-based for stabilization of the foam and water films and foam flooding for IFT
techniques, can further augment the residual oil displacement effect, reduction and wettability alteration called smart water assisted foam
resulting in high oil recoveries. Thus, the review enlightens concisely on (SWAF) flooding was introduced by Hassan et al. (2020, 2022b),
the imminent perspectives of SP-based cEOR in HTHS carbonate reser­ particularly for harsh carbonate reservoirs. Fig. 16 illustrates the injec­
voirs via the most recent and effective approaches for successful field ted scheme scenarios of the SWAF technology, which start with smart
applications. water injection followed by surfactant aqueous solution (SAS) and gas
(Hassan et al., 2022a).
7.1. Hybrid low salinity water (LSW) injection with SP-flooding (i.e., The numerical and laboratory investigations demonstrated the effi­
LSW-SP) ciency of the hybrid SWAF method, with highly encouraging results
pertaining to foamability and foam stability; while the dual-
The polymers and surfactants are prone to destabilization under high improvement effects (i.e., water and foam films stability) of the
salinity and high temperature during SP-based cEOR process. The method, produced exceptional cumulative oil recovery of 92%, as shown
typical problems encountered when polymers are used for cEOR include in Fig. 17 (Hassan et al., 2020, 2021; 2022a, and 2022b). The latter
low injectivity as viscosity is lost or complete plugging of injection wells, figure shows smart water-assisted foam (SWAF) injection process for the
degradation of polymers, incomplete polymer dissolution, and pump best-case scenario (i.e., MgCl2). The method works as an improve­
failures. The low injectivity creates reduced sweep efficiency and d/enhanced (IOR/EOR) recovery method for the ionically-modified
diminished residual oil displacement. Furthermore, polymer retention brine, which goes up to 39% RF for the primary stage (i.e., formation
on rock surface can permanently impair the formation and prevent water), 11% RF (i.e., 50% cumulative oil original in core, OIIC) for the
further successful cEOR applications. Hence, a very promising new secondary stage (i.e., smart water injection), and 92% cumulative OIIC
approach to polymer plugging and polymer degradation is the hybrid for the tertiary stage (SAG injection i.e., SAS, gas, SAS, and gas).
low salinity water polymer (LSW-P) flooding, which combines low
salinity water (LSW) injection (Gbadamosi et al., 2022) and polymer 7.3. Biosurfactants, biopolymers, nanoparticles (NPs), and ionic liquids
flooding (Al-Murayri et al., 2019b; Hassan et al., 2022d) techniques. (ILs)
The LSP process is initiated by treating the source brine so that its
salinity is much lower than the formation brine salinity, after which the Recently, applying modern techniques like biotechnology, nano­
polymer is mixed with the lower salinity brine. This creates synergies in technology (NPs), and ionic liquids (ILs) technology in cEOR is
the LSWP method, which is capable of improving both displacement becoming more prevalent, due to the perceived environmentally, eco­
efficiency and sweep efficiency (Al-Shalabi et al., 2022). The polymers nomic, and productivity enhancement potentials.
play the role of mobility controller of the injected fluid that leads to
improved sweep efficiency (Sorbie, 1991, 2013), whereas the hybrid 7.3.1. Biosurfactants and biopolymers formulations
synergy of polymers employed together with low salinity aqueous so­ The scope of biosurfactants and biopolymers in cEOR and their ca­
lutions, modifies the wettability state of the formation and results in pacity for IFT reduction, wettability alteration, and mobility control,
reduced water-cut and water/oil ratio (Karimov et al., 2020). Also, due which are advantageous in SP-based cEOR, have been the subject of
to reduced salinity of the injection water, the polymer concentration numerous studies (Al-Wahaibi et al., 2014; Gao, 2018; Geetha et al.,
required for the LSWP process is low (Vermolen et al., 2014; Shiran and 2018; Seo et al., 2018; Hassan et al., 2019; Musa et al., 2021; Tack­
Skauge, 2014), which is likely to reduce project costs, and as smaller ie-Otoo et al., 2022).
injection facilities are enough, back-produced polymer in high concen­ The bio-cEOR floodings can be a potentially effective approach to
trations is mitigated. Thus, combined LSW injection with polymer (and improving sweep efficiency in fractured carbonate reservoirs, according
SP) floodings is a comparatively new hybrid EOR technology, which to Gomari (2022), who researched the impact of polymer, biopolymer,
provides a synergic effect to enhance the displacement and sweep effi­ and biosurfactant elements of chemical EOR and bio-cEOR on oil re­
cacies (Al-Shalabi et al., 2022). Hence, with the breakthrough in LSWP covery performance in fractured carbonate reservoirs. The findings
and LSWSP techniques, greater oil recovery is expected in a showed that the suggested method can enhance oil recovery, with bio­
cost-effective process from HTHS carbonate reservoirs. surfactant/polymer/water flooding showing the greatest rates up to
16%. It was discovered that the inclusion of the biosurfactant into the
7.2. Hybrid smart water assisted foam (SWAF) injection with SP-flooding system changed the rock matrix’s wettability from oil-wet to
(i.e., SWAF-SP) neutral-wet, facilitating the oil movement toward fractures (Gomari
et al. 2021; Gomari, 2022). Furthermore, several more recent laboratory
The SP-flooding cEOR for an oil-wet HTHS carbonate reservoir may investigations have successfully demonstrated the efficacy of bio­
not be successful at all times due to these harsh conditions. In fractured surfactants with polymers (or biopolymers) and/or alkali, which can
carbonate formations, high-permeability fractures can act like ‘‘thief affect IFT reduction and wettability alteration of carbonates, and can

24
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Fig. 16. Schematic diagram of SWAF process. (a) Injection of SWAF solution in presence of crude oil, (b) SWAF solution interacts with crude oil, and (c) oil detaches
from rock and moves toward production well (Hassan et al., 2022b).

Fig. 17. Smart water-assisted foam (SWAF) injection process for the best-case scenario (i.e., MgCl2). (Hassan et al., 2022b).

prove to be a cheaper low-risk alternative to conventional surfactants Moreover, Kalpakci et al. (1990) stated that scleroglucan was stable
(Al-Ghailani et al., 2018; Bassir and Shadizadeh, 2020; Nasiri and Biria, thermally for 720 days at 100 ◦ C. The main issue with application of
2020; Bera et al., 2021; Udoh and Vinogradov, 2021; Janaghi et al., scleroglucan for cEOR is its very low filterability (Kulawardana et al.,
2022; Gomari, 2022; Wu et al., 2022). 2012). Xanthan gum is another biopolymer which can be mixed with a
Moreover, Biosurfactants’ hydrophilic head and hydrophobic tail range fluid additives and surfactants for application in tertiary oil re­
interact with the molecules of water and crude oil, respectively. In covery formulations at temperatures up to 90 ◦ C (Olajire, 2014). Also,
recent studies, Pseudomonas aeruginosa HAK01, was discovered by according to some studies, Schizophyllan, which is also a biopolymer,
Khademolhosseini et al. (2019a), who produced a rhamnolipid bio­ can be a potential polymer for SP based EOR under HTHS reservoirs with
surfactant that demonstrated remarkable capacity in IFT reduction and low permeability (Huang and Sorbie, 1993; Leonhardt et al., 2014).
wettability alteration from oil-wet to the water-wet state of the sample Newly developed modifications of polyacrylamides (e.g., NVP-AM) have
glass surface. Moreover, good performance in EOR can be achieved as, been designed to be functional and stable at higher temperature appli­
biosurfactants showed high stability in harsh temperature, pH, and cations (Kulawardana et al., 2012). Hence, some biopolymers and bio­
salinity conditions. surfactants can withstand HTHS carbonate reservoir conditions, and can
Furthermore, Scleroglucan biopolymer has a rigid structure, imparts be modified and improved to function aptly for SP flooding in harsh
high viscosity and shear resistance in aqueous solution with its triple environments.
helix nature that enhances its thermal stability (Kulawardana et al.,
2012). However, the reported thermal stability data studies vary. Ac­ 7.3.2. Nanoparticles (NPs) formulations
cording to Davison (Davison and Mentzer, 1982), scleroglucan remains Many studies showed that nanoparticles (NPs) may change the
intact in original state for 500 days at 90 ◦ C whereas, Rivenq et al. wettability to more water-wet conditions and successfully recover oil
(1992) reported that the thermal stability limit to be 100 ◦ C for 60 days. (Moghaddam et al., 2015; Al-Anssari et al., 2016; Ali et al., 2018;

25
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Moghadam and Salehi, 2019; Patel et al., 2022) while reducing IFT recovery application. To ascertain their surface and interfacial behavior
(El-Diasty and Aly, 2015; Zhang et al., 2016; Youssif et al., 2018; Kuang under reservoir conditions, SAILs were examined in aqueous solution.
et al., 2018). In addition, various studies suggested that the use of NPs The synthesized SAILs had good surface-active characteristics, high ef­
and surfactants can enhance oil recovery through synergistic effects ficiency on the interface adsorption, and a decreased critical micellar
(Khademolhosseini et al., 2015; Mohajeri et al., 2015; Eshraghi et al., concentration of 200 ppm (C16mimBr). The IFT was decreased to
2017; Emadi et al., 2017). For instance, Zargartalebi et al. (2015) extremely low values (between 10− 4 and 10− 3) across a variety of
demonstrated that the use of hydrophilic silica NPs and sodium dodecyl salinity, concentration, and temperature ranges, providing important
sulfate (SDS) decreased IFT and improved oil recovery from sandstone support for the use of surfactants in EOR. The development of the
rocks. The outcomes were in agreement with Ahmadi and Shadizadeh’s Winsor-III micro-emulsion was also proven by the phase behavior ex­
reports (Ahmadi and Shadizadeh, 2013). Furthermore, a combination of periments, rendering it a robust performance formulation (Pillai and
NPs and surfactants changed the system’s wettability to the water-wet Mandal, 2022).
condition, as demonstrated by Nwidee et al. (2017). Zhao et al. (2018) Moreover, combinatory flooding approaches were explored through
examined the synergistic effects of silica NPs and TX-100 (a non-ionic gauging the potential of an alkali-surfactant (AS) formulation for
surfactant) on oil recovery, and the outcomes demonstrated that silica improved oil recovery that combines monoethanolamine (ETA) and 1-
nanoparticles combined with this surfactant were more capable to affect hexadecyl-3-methyl imidazolium bromide (C16mimBr). The study
wettability than a pure surfactant solution. compared the usual formulation of sodium metaborate and cetyl­
Joshi et al. (2022) conducted evaluations via visual settling, dynamic trimethylammonium bromide (CTAB) (NaBO2). The research proved
light scattering (DLS) technique, high Zeta potential, and turbiscan that CTAB and C16mimBr exhibited comparable surface activity and
analysis. They found that the SiO2 nanoparticles (NPs) nano-fluids were aggregation patterns. The ETA-C16mimBr system was shown to work
very stable for use in SP based EOR. When nanoparticles were spread in well with brine that has large divalent cations content. With an ultra-low
the polymer solution or the zwitterionic surfactant CPAB (Cetylpir­ IFT of 7.6 × 10− 3 mN/m, studies on interfacial characteristics revealed
idinium bromide), SiO2 nanofluid stability was significantly improved. that the ETA-C16mimBr system outperformed the NaBO2-CTAB system
For flooding investigations, a SiO2 NPs concentration of 800 ppm was in terms of IFT reduction capabilities. In the presence of salt, the ETA-
determined to be adequate. However, in low permeability carbonates C16mimBr system’s IFT reduction performance improved, resulting in
the SP flooding was slightly more effective in oil recovery, which was an ultra-low IFT of 2.3 × 10− 3 mN/m. Even under highly salinized cir­
attributed to nanoparticles clogging pores. Also, high permeability cumstances with a 15 % wt. NaCl solution, the system maintained an
surfactants were considered the most suitable for these NPs nanofluid ultra-low IFT. Although the wetting power of CTAB was not increased by
approach. Using 10–60 nm oil swollen micelles dispersed by cocami­ the addition of NaBO2, synergism was visible for the ETA-C16mimBr
dopropyl hydroxysultaine zwitterionic cosurfactant at high salinity (56, system in modifying the carbonate rock surface. Even at high tempera­
000 mg/L) and temperature (100 ◦ C), a highly stable “Nano-Surfactant tures, the ETA-C16mimBr and NaBO2-CTAB systems both demonstrated
(NS)" fluid was created via encapsulating Petroleum sulfonate (PS) salt the ability to create stable emulsions (Tackie-Otoo et al., 2022).
surfactants, which are insoluble in high-salinity water. The Finally, several other studies have shown the potential of ionic fluids
nano-surfactant fluid can be created swiftly in the field through a as an alternative cEOR approach by being an alternative to conventional
straightforward process at room temperature requiring minimal energy. chemicals, with the capacity to reduce IFT and alter wettability partic­
The NS fluid preparation and evaluations show their enduring chemical ularly in carbonates at HTHS conditions, affording a low-cost high
and colloidal stability at 100 ◦ C, along with a IFT reduction of 3 orders of performing alternative (Sakthivel, 2021; Atilhan and Santiago, 2021;
magnitude i.e., from 10 to 0.008 mN/m, between crude oil and Esfandiarian et al., 2021; Sanati et al., 2021; Barari et al., 2021; Somoza
high-salinity water, and enriched mobilization of the trapped crude oil et al., 2021, 2022).
from the carbonate rock, as described by Gizzatov et al. (2019). Their
work illustrated the use of NS formulations to improve oil mobilization 7.4. Environmental impact
for a range of reservoir conditions during Surfactant-based cEOR. The PS
salt surfactants, which are among the most widely available and inex­ In the process of SP-based cEOR, majority of the chemicals used i.e.,
pensive industrial surfactants, can be delivered with precision using the surfactants, polymers, alkali, co-solvents, etc., are toxic and non-
nano-surfactant (NS) fluid, which is lucrative, effective, and ecologically biodegradable, which leads to environmental concerns such as,
benign. contamination of ground water and formation damage. The environ­
Finally, according to the investigation of Khademolhosseini et al. mental concerns become significant if these chemicals are retained in
(2019b), high uniformity of spherical NPs led to better distribution and the reservoir at high levels. Moreover, because carbonate reservoirs are
more efficient interactions with crude oil components, which increased naturally fractured and possess heterogeneity, the oil recovery is greatly
recovery. Furthermore, in the coreflooding test, they noticed that the decreased due to high adsorption of surfactants. In addition, reservoir
biosurfactant-spherical NP solution produced 26.1% more oil recovery damage caused by surfactant adsorption hinders any further EOR-
after brine flooding than the biosurfactant-sponge-like NP solution process implementations, which makes the extraction of oil recovery
(25.1%). Finally, biosurfactant-NP solutions outperformed biosurfactant more energy intensive and thus, leading to high CO2 footprint. There­
solutions in terms of improving the wettability alteration of the oil-wet fore, to maintain reasonable expenditure for SP-based EOR and for
carbonate rock. preservation of pristine reservoir condition, surfactant retention must be
abated. One of the suggested approaches of reducing surfactant reten­
7.3.3. Ionic liquid (ILs) formulations tion is using greener alternatives to existing surfactants (i.e., such as new
Recently, a number of research works have suggested surface-active plant sourced surfactants or biosurfactants) since they exhibited
ionic liquids (SAILs), which are functional ionic liquids with amalgam­ remarkable stability in HTHS carbonates environments. Laboratory
ated properties of surfactant and ionic liquid as an alternative to con­ studies have successfully tested the effectiveness of biosurfactants in
ventional surfactant, polymer, and alkaline-based cEOR technologies. reducing IFT and altering carbonate rock wettability towards more
These novel functional ionic liquids represent an exciting alternative, water-wet, which increases recovery (Al-Ghailani et al., 2018; Bassir and
particularly for EOR processes in harsh reservoir conditions of Shadizadeh, 2020; Nasiri and Biria, 2020; Bera et al., 2021; Udoh and
carbonates. Vinogradov, 2021; Janaghi et al., 2022; Gomari, 2022; Wu et al., 2022).
A study on the physicochemical properties of synthesized SAILs using Similarly, widely used polymers such as HPAM, could have potential
FT-IR, 1H NMR, and TGA validated the chemical structure and thermal groundwater impacts and formation damage due to HPAM having very
stability of the synthesized C12mimBr and C16mimBr, for increased oil low biodegradation (Millemann et al., 1982). Additionally, polymer

26
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

adsorption may lead to successive cEOR processes becoming ineffective. most recent extrapolations described in this analysis will progress our
Therefore, biopolymers such as Xanthan gum sourced from bio-waste understanding to a great extent for enhanced designs of systemized
can be employed, which is easily biodegradable and compatible with cEOR projects meant for harsh HTHS carbonate reservoirs.
various fluid additives and surfactants (and biosurfactants) for appli­ Hence, this paper presents a comprehensive literature review on the
cation in SP-based flooding at temperatures up to 90 ◦ C. Additionally, potential of SP-based cEOR applications in carbonates under harsh
Scleroglucan is stable for long durations at temperatures up to 100 ◦ C conditions of HTHS from the perspective of successful laboratory, core-
(Davison and Mentzer, 1982; Kalpakci et al., 1990; Rivenq et al., 1992; flood, and field applications. The main conclusions of this work can be
Kulawardana et al., 2012) and Schizophyllan is also applicable in HTHS summarized as follows:
conditions, which are both biodegradable biopolymers (Huang and
Sorbie, 1993; Leonhardt et al., 2014). • Studies showed the predominant factors influencing oil recovery in
In summary, eco-friendly chemicals such as biosurfactants and bio­ carbonates to be the capillary forces, concentration of potential
polymers could be good alternative candidates for SP-flooding under determining ions, gravity and viscous forces, and pH.
HTHS carbonates conditions in case of high-levels of surfactant and • The new Guerbet alkoxy sulfates (GAS), Guerbet alkoxy carboxylates
polymer adsorptions. (GAC), and Guerbet alkoxy betaine surfactant (GAB), demonstrated
good stability at temperatures over 100 ◦ C and salinities up to
8. Summary, conclusions, and recommendations 275,000 ppm with very low adsorption on carbonates and high ter­
tiary oil recovery.
Significant amounts of remaining and residual oil saturations exist in • This study also includes novel cationic Gemini surfactant formula­
the fractured low-permeability HTHS carbonate reservoirs. In order to tions with their temperature tolerance being approximately 300 ◦ C
achieve maximum total oil production, SP-based cEOR technologies are and salinity tolerance of over 220,000 ppm.
adapted. At the moment, a growing number of SP-based cEOR pilot • The hydrophobically associating polymers (HAPAM), that are highly
projects are in progress worldwide. The highly challenging temperature viscosifying chemicals, and its modified nanocomposite derivative
and salinity (i.e., HTHS) environments render it difficult to procure the have good wettability altering capabilities as well as tolerance to
appropriate surfactants and polymers required. The harsh HTHS high temperature and high heterogeneity.
conditioned carbonate reservoirs hold over two thirds of the world’s oil • Hot-melt Adhesive (HMA) has desirable rheological properties and
reserves, and are characterized by fractured formations, heterogeneous low adsorption in HTHS carbonates. E.g., SAV10, an HPAM based
mineral composition, and vuggy structure. In carbonate reservoirs, oil polymer, and Sulfonated Polyacrylamide (AN-125) showing excel­
recovery by SP-flooding is severely impacted as surfactants and poly­ lent thermal and salinity tolerance in SP based EOR under HTHS
mers degrade significantly in extreme temperature, and salinity (HTHS). carbonates.
Moreover, surfactant retention can occur due to heterogeneity in min­ • Furthermore, Polymeric surfactants, biopolymers like Xanthan gum,
eral composition and surface interactivities of rock. The surface Scleroglucan and Schizophyllan, can be applied in tertiary oil re­
geochemistry and the surface charge of carbonate rock can influence the covery formulations for SP-based cEOR, in HTHS and low perme­
extent of surfactants retention. For instance, natural limestone has ability reservoir conditions effectively.
aluminum and silicon present that causes retention, and even tiny cal­ • Surface-active ionic liquids (SAILs) reduce IFT to ultra-low values
cium sites on rock surface led to carboxylate surfactant adsorption. and induce wettability alteration towards a more water-wetting state
Adding alkali to reduce pH, and pre-flush with low salinity water in harsh conditioned carbonates.
flooding before SP-based cEOR can reduce surfactant adsorption. • Nano-technology innovations include the nanoparticles (NPs) as
Therefore, in order to meet the prerequisites for HTHS, heterogeneity potential surfactant careers for SP-based EOR in HTHS carbonates,
in mineral composition, and surface interactivities of carbonate reser­ causing wettability alteration and high oil recovery, especially, the
voirs, there is still a lack of availability of suitable surfactants and biosurfactant-NP solutions that are highly effective in carbonate
polymers for SP-based cEOR implementations, except the surfactants rocks.
that exhibited potential in the laboratory scale experiments such as • Novel polymers such as, ATBS, NVP-based, hydrophobic associative,
those described in this paper. Fortunately, the laboratory and field trials smart thermoviscosifying polymers (TVP), soft microgel polymers,
of these surfactants demonstrate high potential. and sulfonated polymers have shown to be tolerant to harsh
Furthermore, in SP systems (i.e., COBR-system), polymers are the carbonates.
major mobility control employed, hence, both surfactants and polymers • SWAF-SP flooding techniques can mitigate retention and improve oil
that are appropriate for carbonate reservoirs must be regarded for recovery by combining smart-water assisted foam flooding with SP-
advanced cEOR applications. For this reason, recently, many novel flooding (i.e., up to 92% cumulative OIIC for the tertiary stage).
surfactant and polymer formulations have been designed in a way that • LSW-SP flooding can create as negative salinity gradient and it can
can withstand the harsh environment of HTHS carbonate rocks. In reduce retention while also augmenting the efficiency of surfactant
addition, the laboratory and field trials demonstrate high potential of flood, and is used for preconditioning the reservoir.
hybrid SP-based cEOR methods that are combined with low salinity • The progressive engineering of novel surfactants and polymers that
water (LSW) injection, and foam flooding via low salinity water polymer can withstand HTHS carbonate conditions is aimed at exploitation of
(LSWP) flooding and smart-water assisted foam (SWAF) flooding pro­ previously inaccessible oil reservoirs. Thus, this paper reiterates the
cesses, respectively. These superior imminent technologies will be able fact that surfactants, SP formulations, and other technologies have
to decrease quite a few operational and logistic problems (i.e., increase evolved to be highly stable in harsh carbonate environments. Newer
productivity, and reduce cost and environmental impact) of SP-based works have focused on explaining the fluid-rock interactions during
EOR for carbonate rocks. These technologies are anticipated to be SP flooding process with respect to these chemical agents.
adopted in the life-cycle methodology of SP-based EOR, which involves
reservoir stimulations that are carefully prepared, starting from the field Finally, the analysis of the latest progress suggests that surfactant-
planning and development phase. Thus, for prospective projects, the polymer formulations can be optimized for improved chemical stabil­
framework and operative know-hows learned from experimental out­ ity, as well as lower surfactant/polymer adsorption and degradation
comes can be utilized for the considerable growth of efficient oil pro­ (polymers) in HTHS carbonate reservoirs. This would create environ­
duction globally. Therefore, advancing EOR to new heights requires mentally acceptable, cost-effective SP-based flooding designs. However,
imperative development of cEOR methods that have been improved on there are more conclusive works to anticipate since there are very few
cost-effectiveness, HTHS tolerance, and eco-conscious parameters. The field applications so far.

27
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Declaration of competing interest Acknowledgments

The authors declare that they have no known competing financial This work was supported by the KFUPM-KU Joint Research Program.
interests or personal relationships that could have appeared to influence Author(s) at Khalifa University acknowledge the support received under
the work reported in this paper. award # KFUPM-KU-2020-14. Author(s) at KFUPM would like to
acknowledge the support received under grant # KU-201-001.
Data availability

No data was used for the research described in the article.

Appendix-A
Table A1
An overview of the existing review articles, which cover the topic of SP-based cEOR including the materials used in these studies, temperature and salinity ranges,
advantages, and limitations

Reference Chemicals Used Experimental Conditions Remarks

Temp Salinity (ppm) Advantages (OIIP%) Limitations


(◦ C)

Gao et al. Petroleum sulfonate, HPAM, Comb shaped 71–80 3904–12,404 The work discusses the projects that Ranges of temperature and salinity do not
(2014) polymer successfully used SP-flooding. The meet HTHS carbonates conditions. The
recovery reached 5.5% OIIP. study is concentrated on oil fields in
China.
Raffa et al. A tapered copolymer of vinyl acetate, acrylic – – According to the study, using polymeric The application of polymeric surfactants
(2016a) acid, Alternate co-polymers of styrene and surfactants may improve EOR is the only focus of the research. HTHS
secondary amines, HPAM modified with performance. It provides an OIIP carbonates conditions are not highlighted
hydrophobic monomer recovery between 17 and 41%. in the study. The temperature and salinity
ranges are excluded.
Pal et al. Anionic, nonionic, and cationic surfactants, 25–100 Up to 11,000 Application of surfactant-based cEOR in Instead of SP-flooding, the study
(2018) Co-surfactant alkanols, biosurfactant, sugar- carbonate formations with a 12–36% concentrated more on surfactant-assisted
based surfactant OIIP oil recovery range. flooding. Salinity range does not meet the
criteria for high salinity.
Jain et al. Anionic surfactant, co-surfactant, polymer, 93.3 – The significance of surfactant and The study provides an overview of
(2022) alkali polymer floodings is emphasized in the surfactant and polymer flooding (i.e., not
paper. Depending on the reservoir emphasizing SP-flooding). The
characteristics, it can recover between formulations of surfactant and polymers
7% and 26% OIIP. used are not specified. No salinity range is
given.
Bashir Anionic: Alkyl-aryl sulfonic acid. Anionic: 25–120 2800–117,000 The review study classifies the The study focuses on surfactant flooding
et al. Internal olefin sulfonate. Cationic: surfactants-based flooding’s oil recovery without taking the co-surfactant (i.e.,
(2022) Hexadecyltrimethylam. Non-ionic: synthetic techniques. It can recover between 13% polymer) effect into account. The study is
biosurfactant and 90% OIIP. focused on sandstone and carbonate
reservoirs.
Han et al. Polyacrylamide-type polymers, Surfactant 50–80 3000–150,000 The study reviews field applications and The study’s primary focus is on surfactant
(2013) type learnings from offshore chemical flooding potential, assessment of the co-
flooding. It provides a recovery of 5%– surfactant (i.e., polymer) effect. The two
15% OIIP. main areas of research are sandstone and
carbonate reservoirs.

Table A2
Description of different novel surfactants viable for SP-based cEOR in HTHS carbonate reservoirs

Surfactant Formulation Temperature Salinity (ppm) Function


(◦ C)

Guerbet Alkoxy Sulfate 0.25 wt% C32–7PO-14EO-sulfate, 0.25 wt% C20–24- 100 35,000 ppm of Solubilization ratio at optimal salinity was around 17,
Surfactants (GAS) IOS, and 0.5 wt% TEGBE cosolvent sodium with exceptionally low IFT, surfactant retention of
carbonate 0.11 mg/g of rock with 94.7% oil retrieval
Guerbet Alkoxy 0.5 wt% C28–45PO-60EO-carboxylate and 0.5 wt% 120 30,000 Excellent phase behavior at 120 ◦ C, oil recovery at
Carboxylate C15–18-IOS 90.5% of residual oil and surfactant retention of 0.33
Surfactants (GAC) mg/g of rock
Guerbet Alkoxy 1.0 wt% C28–25PO-55EO-carboxylate, 0.5 wt% 83 230,800 Oil recovery was observed at 92% and Surfactant
Carboxylate C15–18-IOS, 0.5 wt% C19–23-IOS, and 1.0 wt% TEGBE adsorption value was 0.3 mg/g of rock
Surfactants (GAC)
Guerbet Alkoxy 0.5 wt% C28–25PO-25EO-carboxylate and 0.5 wt% 100 116,969 Excellent oil recovery of 65.9% with surfactant
Carboxylate C15–18-IOS retention was 0.086 mg/g of rock
Surfactants (GAC)
Tristyrylphenol Alkoxy 0.18 wt% TSP-35PO-20EO-sulfate and 0.12 wt% 62 55,000 92.1% oil retrieval with 0.10 mg/g of rock adsorption
Sulfate Surfactants C20–24-IOS (No cosolvents added) and no chromatographic separation visible in the core
(TSP)
0.50% GAB + 0.50% ADB + 1.0% n-butanol (for phase 120 275,000
behavior test)
(continued on next page)

28
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Table A2 (continued )
Surfactant Formulation Temperature Salinity (ppm) Function
(◦ C)

Novel Guerbet Alkoxy Excellent tolerance to (HTHS) reservoir conditions


Betaine Surfactant and exhibited ultralow IFT for crude oil and
(GAB) intensified emulsification features
Gemini Surfactants – 98 – Field application at 190 bars reservoir pressure in
Hungary
Alkaline-Surfactant- 0.50 wt% C28–35PO-65EO-COONa, 0.25 wt%, C15-18 100 60,000 Exceptional tertiary oil output and minimal surfactant
Polymer Formulation IOS, 0.25 wt% C19–28IOS, 1.0 wt% ethoxylated adsorption
diisopropylamine (DIPA-10EO) and 0.75 wt% NH3

Table A3
Description of different novel polymers viable for SP-based cEOR in HTHS carbonate reservoirs

Polymer Monomer Group Use Main Function

Polymeric Surfactant Polymerizable groups viz., vinyl, allyl, Surface modification of solid substances, generating Altering properties of the surfactant for EOR
(surfmer) acrylate, methacrylate, styryl, and novel water-soluble hydrophobically associating
acrylamide polymers with high viscous power
Hydrophobically Micellar copolymerization of hydrophilic Resistant towards temperature, pH, and ion content Function as viscosifiers or wettability
Associating and hydrophobic monomers in the periphery of the wells modifiers in HTHS and high-pressure
Polyacrylamide Polymers reservoirs
(HAPAMs)
Flopaam 3630 HPAM based polymer Good tolerance of high temperature (100 ◦ C) at Mobility control with less shear sensitive
35,000 ppm salinity behavior under HTHS conditions
HAPAM Copolymer and HAPAM Retention of original viscosity and excellent non- Substitute for current surfactants, through
HAPAM-SiO2 Newtonian behaviors (i.e., correspond well to in- wettability altering nature and outstanding
Nanocomposite situ shear thinning behavior) tolerance to high temperature and high
heterogeneity oil fields
Hot-melt Adhesive (HMA) AM hydrophilic backbone of a terpolymer Surfactants could either enhance or reduce the Low adsorption in HTHS carbonates, better
Polymers modified with AMPS and hydrophobic viscosity of the HMA polymers depending on the rigidity, increased viscosity with salts and
monomer 2-vinylnaphtahlene (and other concentration added and the intermolecular and temperature, possible IFT reduction
variations) intramolecular associations
SAV10 ATBS-based polymer Tested at 297 g/L TDS and 105–120 ◦ C Thermally stable polymer for SP based EOR
in HTHS carbonates
Sulfonated Polyacrylamide Copolymer of AM and ATBS (25% Tested at formation brine with TDS of 229,870 ppm Low retention on carbonate rocks with
(AN-125) sulfonation) at 100 ◦ C HTHS

Table A4
Summary of reported literature on SP-based cEOR field applications in carbonate reservoirs

Reference Oilfield Temp Chemicals Used & Salinity (ppm) Remarks


(◦ C)
Recovery (%) Flooding Scheme

Al-Amrie et al. ABK oilfield, 83 Novel stable classes of hydrophilicity- Oil recovery of OOIP 90% SP slug and the polymer drive, although
(2015); Levitt Offshore, Abu Dhabi, enhanced surfactants; SAV-225 (AM- OOIP; Effective Sor reduction during actual operation the polymer
et al., (2016) SP UAE. AMPS-NVP terpolymer) from SNF from 40% to 4% in a tight drive was not required in this case.
Floerger; AN-125 SH (AM-AMPS carbonate with HTHS and low
copolymer) from SNF Floerger; Salinity permeability
is 230,000 (ppm)
Al-Murayri et al. Raudhatain Lower 82 Surfactant concentration of 5000 ppm; Positive oil desaturation Pre-flush was injected before the main-
(2019a, 2018, and Burgan (RALB) Polymer concentration of 500 ppm; In- levels following chemical slug injection, then followed by a post-
2017) SP Reservoir, Kuwait situ salinity of over >200,000 ppm flooding within 10 ft from the flush with the polymer only.
well Afterwards, sea-water push conducted,
then tracer injection, soaking and
production was conducted
Harris et al. (2020) Offshore the South – Alkyl propoxy sulfate (APS) and an alkyl The residual oil saturation Single Well Chemical Tracer Test
SP China Sea Field, ethoxy sulfate (AES) mixture with no co- decreased from 20-27% to (SWCTT)
China solvents; The salinity formulation was 7–9% range.
seawater-based
Yin et al., (2021) SP Block X, Tahe 100–110 Dispersed particle gel (DPG) and Incremental cEOR of 21.3% Combined flooding of DPG for in-depth
oilfield, Xinjiang, surfactant of the formulation 0.35% OIIP; The oil recovery from conformance control to enhance the
China NAC-1+ 0.25% NAC-2. In-situ salinity given well group was sweep efficiency, followed by surfactant
around 220,500 (ppm) increased by 3000 tons solution slug for improving
displacement efficiency
Rilian and Sumestry Semoga field, 83 Nonionic surfactants; Oil with 38 API◦ Oil recovery volume of Surfactant well stimulation via Huff-n-
(2010); Sheng Baturaja Formation, and 0.84 cP viscosity 58,000 bbl during the course Puff processes; Operation implemented
(2013b) Indonesia of 3 months in 3 steps: (i) pre-flush, (ii) main flush
and (iii) over-flush
Rohilla et al. (2021) Offshore Al-Shaheen 60 Surfactant concentration of 200,000 Incremental cEOR of 30–35% Long-term surfactant injection through
Surfactant Field, Block 5, Qatar (ppm); Salinity is 79,700 (ppm) OIIP due to wettability continuous mode at 4000 ppm of active
(continued on next page)

29
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Table A4 (continued )
Reference Oilfield Temp Chemicals Used & Salinity (ppm) Remarks
(◦ C)
Recovery (%) Flooding Scheme

alteration; No surfactant concentration in sea water, after which


breakthrough sea water was injected
Varfolomeet al., Bashkirian 23 Nonionic surfactant; In-situ salinity Incremental cEOR of 1094 After surfactant injection, the nonionic
(2020) Surfactant Reservoirs, Ukraine around 217,500 (ppm) tones, via IFT reduction and surfactant showed good influence on
wettability alteration water injectivity enhancement and in
reducing wellhead and downhole
pressure
Stephenso et al., Lower Shaunavon 82 Emulsions additives; In-situ salinity of Pilot test showed optimistic Microemulsions additives act as a
(2019) Additives Saskatchewa, Canada over >200,000 ppm results with high cEOR delivery platform for surfactants,
solvents, and co-solvents into the
reservoir for enhanced waterflooding
Baloch et al. (2021) Abu Dhabi, UAE, The 121.1 SAV 10, which is a polyacrylamide- Good injectivity recover Polymer injection test/(PIT).; The
Polymer Abu Dhabi National based polymer with high 2-acrylamido- through achieving favorable overall polymer solution injection
Oil Company tertiary-butyl sulfonic acid (ATBS) indicators (PIT) volume was 108,392 barrels
(ADNOC) content; In-situ salinity of over
>200,000 ppm

Table A5
Surfactant-polymer based cEOR field projects worldwide (information was taken from Llano et al., 2013)

Project Name Location Operator Type of SP-flooding

ANGSI CEOR Asia PETRONAS SP, and ASP


St. Joseph Asia Petronas ASP
Daqing Asia Daqing Oilfield Polymer, ASP trials
Shengli Oilfield Asia Shengli Oilfield Polymer, SP, ASP
Dagang Oilfield Asia Petrochina cEOR
Mangala Field Asia Cair Energy India Pty Ltd ASP-flooding
Unnamed Middle East TOTAL SP-flooding
Oman Middle East Shell Polymer, SP, and ASP
Al Shaheen Middle East Maersk Oil Polymer and ASP
Battrum North America Hyak Energy ASP-flooding
Trembley Oilfield North America – ASP-flooding
ASP flooding pilot North America OXY ASP-flooding
Lawrence North America Rex Energy ASP-flooding
Saskatchewan North America Talisman ASP-flooding
ASP Colombia South America Williams Energy ASP-flooding
HTHS EOR Project – Total SP-flooding

References Ahmadi, M.A., Shadizadeh, S.R., 2013. Induced effect of adding nano silica on adsorption
of a natural surfactant onto sandstone rock: experimental and theoretical study.
J. Petrol. Sci. Eng. 112, 239–247.
Aarra, M.G., Skauge, A., Martinsen, H.A., 2002. FAWAG: a breakthrough for EOR in the
Ahmadi, M.A., Shadizadeh, S.R., 2015. Experimental investigation of a natural surfactant
north sea. In: SPE Annual Technical Conference and Exhibition.
adsorption on shale-sandstone reservoir rocks: static and dynamic conditions. Fuel
Abalkhail, N., Liyanage, P.J., Upamali, K.A.N., Pope, G.A., Mohanty, K.K., 2020.
159, 15–26.
Alkaline-surfactant-polymer formulation development for a HTHS carbonate
Akbar, M., Vissapragada, B., Alghamdi, A.H., Allen, D., Herron, M., Carnegie, A.,
reservoir. J. Petrol. Sci. Eng. 191, 107236.
Dutta, D., Olesen, J.-R., Chourasiya, R.D., Logan, D., 2000. A snapshot of carbonate
Abbasi-Asl, Y., Pope, G.A., Delshad, M., 2010. Mechanistic modeling of chemical
reservoir evaluation. Oilfield Rev. 12 (4), 20–21.
transport in naturally fractured oil reservoirs. In: SPE Improved Oil Recovery
Akbari, S., Mahmood, S.M., Tan, I.M., Ghaedi, H., Ling, O.L., 2017. Assessment of
Symposium.
polyacrylamide-based co-polymers enhanced by functional group modifications with
Adams, W.T., Schievelbein, V.H., 1987. Surfactant flooding carbonate reservoirs. SPE
regards to salinity and hardness. Polymers 9 (12), 647.
Reservoir Eng. 2 (4), 619–626.
Aktar, S., Saha, M., Mahbub, S., Halim, M.A., Rub, M.A., Hoque, M.A., Islam, D.S.,
Adejare, O.O., Nasralla, R.A., Nasr-El-Din, H.A., 2012. The effect of viscoelastic
Kumar, D., Alghamdi, Y.G., Asiri, A.M., 2020. Influence of polyethylene glycol on the
surfactants and a mutual solvent on the wettability of a carbonate rock. In: SPE
aggregation/clouding phenomena of cationic and non-ionic surfactants in
International Production and Operations Conference & Exhibition. OnePetro.
attendance of electrolytes (NaCl & Na2SO4): an experimental and theoretical
Adibhatla, B., Mohanty, K.K., 2006. Oil Recovery from Fractured Carbonates by
analysis. J. Mol. Liq. 306, 112880.
Surfactant-Aided Gravity Drainage: Laboratory Experiments and Mechanistic
Al-Amrie, O., Peltier, S., Pearce, A., Al-Yafei, A., Morel, D., Bourrel, M., Bursaux, R.,
Simulations. SPE/DOE Symposium on Improved Oil Recovery.
Cordelier, P., Jouenne, S., Juilla, H., Klimenko, A., 2015. The first successful
Adkins, S., Arachchilage, G.P., Solairaj, S., Lu, J., Weerasooriya, U., Pope, G., 2012.
chemical EOR pilot in the UAE: one spot pilot in high temperature, high salinity
Development of thermally and chemically stable large-hydrophobe alkoxy
carbonate reservoir. In: Abu Dhabi International Petroleum Exhibition and
carboxylate surfactants. In: SPE Improved Oil Recovery Symposium.
Conference. OnePetro.
Adkins, S., Liyanage, P.J., Pinnawala Arachchilage, G.W., Mudiyanselage, T.,
Al-Anssari, S., Barifcani, A., Wang, S., Maxim, L., Iglauer, S., 2016. Wettability alteration
Weerasooriya, U., Pope, G.A., 2010. A new process for manufacturing and stabilizing
of oil-wet carbonate by silica nanofluid. J. Colloid Interface Sci. 461, 435–442.
high-performance EOR surfactants at low cost for high-temperature, high-salinity oil
Al-Ghailani, T., Al-Wahaibi, Y.M., Joshi, S.J., Al-Bahry, S.N., Elshafie, A.E., Al-Bemani, A.
reservoirs. In: SPE Improved Oil Recovery Symposium.
S., 2018. Alkaline-biosurfactant-biopolymer process and its potential for enhancing
Afolabi, F., Mahmood, S.M., Yekeen, N., Akbari, S., Sharifigaliuk, H., 2022. Polymeric
oil recovery in Omani oil field. In: SPE EOR Conference at Oil and Gas West Asia.
surfactants for enhanced oil recovery: a review of recent progress. J. Petrol. Sci. Eng.
Al-Murayri, M.T., Kamal, D.S., Suniga, P., Fortenberry, R., Britton, C., Pope, G.A.,
208, 109358.
Liyanage, P.J., Jang, S.H., Upamali, K.A., 2017. October. Improving ASP
Ahmadali, T., Gonzalez, M.V., Harwell, J.H., Scamehorn, J.F., 1993. Reducing surfactant
performance in carbonate reservoir rocks using hybrid-alkali. In: SPE Annual
adsorption in carbonate reservoirs. SPE Reservoir Eng. 8 (2), 117–122.
Technical Conference and Exhibition. OnePetro.
Algharaib, M., Alajmi, A., Gharbi, R., 2014. Improving polymer flood performance in
high salinity reservoirs. J. Petrol. Sci. Eng. 115, 17–23.

30
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Al-Hashim, H.S., Celik, M.S., Oskay, M.M., Al-Yousef, H.Y., 1988. Adsorption and Bassir, S.M., Shadizadeh, S.R., 2020. Static adsorption of a new cationic biosurfactant on
precipitation behaviour of petroleum sulfonates from Saudi Arabian limestone. carbonate minerals: application to EOR. Petrol. Sci. Technol. 38 (5), 462–471.
J. Petrol. Sci. Eng. 1 (4), 335–344. Bastiat, G., Grassl, B., Khoukh, A., François, J., 2004. Study of sodium dodecyl sulfate−
Ali, J.A., Kolo, K., Manshad, A.K., Mohammadi, A.H., 2018. Recent advances in poly (propylene oxide) methacrylate mixed micelles. Langmuir 20 (14), 5759–5769.
application of nanotechnology in chemical enhanced oil recovery: effects of Bataweel, M.A., Nasr-El-Din, H.A., 2011. Alternatives to minimize scale precipitation in
nanoparticles on wettability alteration, interfacial tension reduction, and flooding. carbonate cores caused by alkalis in asp flooding in high salinity/high temperature
Egyptian journal of petroleum 27 (4), 1371–1383. applications. In: SPE European Formation Damage Conference.
Al-Murayri, M.T., Hassan, A.A., Abdullah, M.B., Abdulrahim, A.M., Marlière, C., Baviere, M., Bazin, B., Noik, C., 1988. Surfactants for EOR: olefin sulfonate behavior at
Hocine, S., Tabary, R., Suzanne, G.P., 2019a. Surfactant/polymer flooding: chemical- high temperature and hardness. SPE Reservoir Eng. 3 (2), 597–603.
formulation design and evaluation for Raudhatain lower burgan reservoir, Kuwait. Behler, A., Biermann, M., Hill, K., Raths, H.-C., Victor, M.E.S., Uphues, G., Texter, J.,
SPE Reservoir Eval. Eng. 22 (3), 923–940. 2001. Industrial surfactant syntheses. Surfactant Sci. Ser. 1–44.
Al-Murayri, M.T., Kamal, D.S., Al-Sabah, H.M., AbdulSalam, T., Al-Shamali, A., Belhaj, A.F., Elraies, K.A., Mahmood, S.M., Zulkifli, N.N., Akbari, S., Hussien, O.S., 2020.
Quttainah, R., Glushko, D., Britton, C., Delshad, M., Liyanage, J., 2019b. Low- The effect of surfactant concentration, salinity, temperature, and pH on surfactant
salinity polymer flooding in a high-temperature low-permeability carbonate adsorption for chemical enhanced oil recovery: a review. J. Pet. Explor. Prod.
reservoir in west Kuwait. In: SPE Kuwait Oil and Gas Show and Conference. Technol. 10, 125–137.
Al-Murayri, M.T., Al-Qenae, A., AlRukaibi, D., Chatterjee, M., Hewitt, P., 2017. Design of Bera, A., Kumar, S., Foroozesh, J., 2021. Biosurfactant role in microbial enhanced oil
a partitioning inter-well tracer test for a chemical EOR pilot targeting the Sabriyah recovery. Green Sustainable Process for Chemical and Environmental Engineering
Mauddud carbonate reservoir in Kuwait. In: SPE Kuwait Oil & Gas Show and and Science 1–33.
Conference. OnePetro. Bera, A., Ojha, K., Mandal, A., Kumar, T., 2011. Interfacial tension and phase behavior of
AlQuraishi, A.A., AlHussinan, S.N., AlYami, H.Q., 2015. Efficiency and recovery surfactant-brine–oil system. Colloids Surf., A 383 (1–3), 114–119.
mechanisms of low salinity water flooding in sandstone and carbonate reservoirs. Bergström, L.M., Tehrani-Bagha, A., Nagy, G., 2015. Growth behavior, geometrical
Offshore Mediterranean Conference and Exhibition. shape, and second CMC of micelles formed by cationic gemini esterquat surfactants.
AlSada, A., Mackay, E., 2017. Low salinity water flooding possible mechanisms and the Langmuir 31 (16), 4644–4653.
impact of injected sulphate variation on oil recovery in carbonate reservoirs: Blaker, T., Aarra, M.G., Skauge, A., Rasmussen, L., Celius, H.K., Martinsen, H.A.,
compositional modeling approach. In: SPE Kingdom of Saudi Arabia Annual Vassenden, F., 2002. Foam for gas mobility control in the Snorre field: the FAWAG
Technical Symposium and Exhibition. project. SPE Reservoir Eval. Eng. 5 (4), 317–323.
Al-Shalabi, E.W., 2018. Numerical modeling of biopolymer flooding in high-temperature Boampong, L.O., Rafati, R., Haddad, A.S., 2022. A calibrated surface complexation model
high-salinity carbonate cores. In: Offshore Technology Conference Asia. for carbonate-oil-brine interactions coupled with reservoir simulation-Application to
Al-Shalabi, E.W., Alameri, W., Hassan, A.M., 2022. Mechanistic modeling of hybrid low controlled salinity water flooding. J. Petrol. Sci. Eng. 208, 109314.
salinity polymer flooding: role of geochemistry. J. Petrol. Sci. Eng. 210, 110013. Bohmer, M.R., Koopal, L.K., 1992. Adsorption of ionic surfactants on variable-charge
Al-Shalabi, E.W., Sepehrnoori, K., 2017. Low Salinity and Engineered Water Injection for surfaces. 1. Charge effects and structure of the adsorbed layer. Langmuir 8 (11),
Sandstone and Carbonate Reservoirs. Gulf Professional Publishing. 2649–2659.
Al-Wahaibi, Y., Joshi, S., Al-Bahry, S., Elshafie, A., Al-Bemani, A., Shibulal, B., 2014. Borzenkov, M., Hevus, O., 2014. Surface Active Monomers: Synthesis, Properties, and
Biosurfactant production by Bacillus subtilis B30 and its application in enhancing oil Application. Springer.
recovery. Colloids Surf. B Biointerfaces 114, 324–333. Borzenkov, M., Mitina, N., Lobaz, V., Hevus, O., 2015. Synthesis and properties of novel
Amirpour, M., Shadizadeh, S.R., Esfandyari, H., Ahmadi, S., 2015. Experimental surface-active monomers based on derivatives of 4-hydroxybutyric acid and 6-
investigation of wettability alteration on residual oil saturation using nonionic hydroxyhexanoic acid. J. Surfactants Deterg. 18 (1), 133–144.
surfactants: capillary pressure measurement. Petroleum 1 (4), 289–299. Bourbiaux, B., Fourno, A., Nguyen, Q.-L., Norrant, F., Robin, M., Rosenberg, E.,
Andersen, P., Evje, S., Kleppe, H., 2014. A model for spontaneous imbibition as a Argillier, J.-F., 2014. Experimental and numerical assessment of chemical EOR in oil-
mechanism for oil recovery in fractured reservoirs. Transport Porous Media 101 (2), wet naturally-fractured reservoirs. In: SPE Improved Oil Recovery Symposium.
299–331. Bourrel, M., Schechter, R.S., 2010. Microemulsions and Related Systems: Formulation,
Anderson, W., 1986. Wettability literature survey-part 2: wettability measurement. Solvency, and Physical Properties. Editions Technip.
J. Petrol. Technol. 38 (11), 1246–1262. Budhathoki, M., Barnee, S.H.R., Shiau, B.-J., Harwell, J.H., 2016. Improved oil recovery
Atilhan, M., Santiago, A., 2021. Review on chemical enhanced oil recovery: utilization of by reducing surfactant adsorption with polyelectrolyte in high saline brine. Colloids
ionic liquids and deep eutectic solvents. J. Petrol. Sci. Eng. 205, 108746. Surf. A Physicochem. Eng. Asp. 498, 66–73.
Atta, A.M., Dyab, A.K.F., Al-Lohedan, H.A., AlJenady, K.A., 2014. Novel reactive Butt, H., Graf, K., Kappl, M., 2003. Physics and Chemistry of Interfaces. Wiley-VCH
polymerizable nonyl phenol ethoxylate surfactants as emulsifier in non-aqueous Verlag GmbH and Co. KGaA, Weinheim, FRG, FRG.
emulsion polymerization. Polym. Sci. B 56 (6), 770–787. Cai, H., Wang, Q., Luo, W., Wang, H., Zhou, X., Li, J., Zheng, Y., 2019. Guerbet alkoxy
Austad, T., Matre, B., Milter, J., Saevareid, A., Øyno, L., 1998. Chemical flooding of oil betaine surfactant for surfactant-polymer flooding in high temperature, high salinity
reservoirs 8. Spontaneous oil expulsion from oil-and water-wet low permeable chalk reservoirs. SPE International Conference on Oilfield Chemistry.
material by imbibition of aqueous surfactant solutions. Colloids Surf. A Camail, M., Margaillan, A., Martin, I., Papailhou, A.L., Vernet, J.L., 2000. Synthesis of N-
Physicochem. Eng. Asp. 137 (1–3), 117–129. alkyl-and N-arylalkylacrylamides and micellar copolymerization with acrylamide.
Austad, T., Milter, J., 1997. Spontaneous imbibition of water into low permeable chalk at Eur. Polym. J. 36 (9), 1853–1863.
different wettabilities using surfactants. International Symposium on Oilfield Candau, F., Selb, J., 1999. Hydrophobically-modified polyacrylamides prepared by
Chemistry. micellar polymerization. Adv. Colloid Interface Sci. 79 (2–3), 149–172.
Austad, T., Shariatpanahi, S.F., Strand, S., Black, C.J.J., Webb, K.J., 2012. Conditions for Cao, Y., Li, H., 1999. Synthesis of a novel family of polymeric surfactants with low
a low-salinity enhanced oil recovery (EOR) effect in carbonate oil reservoirs. Energy interfacial tension by ultrasonic method. Polym. J. 31 (11), 920–923.
Fuel. 26 (1), 569–575. Cao, Y., Li, H., 2002. Interfacial activity of a novel family of polymeric surfactants. Eur.
Austad, T., Strand, S., Høgnesen, E.J., Zhang, P., 2005. Seawater as IOR fluid in fractured Polym. J. 38 (7), 1457–1463.
chalk. In: SPE International Symposium on Oilfield Chemistry. Chandar, P., Somasundaran, P., Turro, N.J., 1987. Fluorescence probe studies on the
Ayirala, S.C., Boqmi, A., Alghamdi, A., AlSofi, A., 2019. Dilute surfactants for wettability structure of the adsorbed layer of dodecyl sulfate at the alumina—water interface.
alteration and enhanced oil recovery in carbonates. IOR 2019–20th European J. Colloid Interface Sci. 117 (1), 31–46.
Symposium on Improved Oil Recovery 2019 (1). Chen, H., Gizzatov, A., Abdel-Fattah, A.I., 2019. Molecular assembly of surfactant
Ayirala, S., Sofi, A., Li, Z., Xu, Z., 2021. Surfactant and surfactant-polymer effects on mixtures in oil-swollen micelles: implications for high salinity colloidal stability.
wettability and crude oil liberation in carbonates. J. Petrol. Sci. Eng. 207, 109117. J. Phys. Chem. B 124 (3), 568–576.
Azam, M.R., Tan, I.M., Ismail, L., Mushtaq, M., Nadeem, M., Sagir, M., 2013. Static Chen, L., Zhang, G., Wang, L., Wu, W., Ge, J., 2014. Zeta potential of limestone in a large
adsorption of anionic surfactant onto crushed Berea sandstone. J. Pet. Explor. Prod. range of salinity. Colloids Surf. A Physicochem. Eng. Asp. 450, 1–8.
Technol. 3 (3), 195–201. Chen, P., Mohanty, K.K., 2013. Surfactant-mediated spontaneous imbibition in carbonate
Azam, S., Darlington, A., Gibbs-Davis, J.M., 2014. The influence of concentration on rocks at harsh reservoir conditions. SPE J. 18 (1), 124–133.
specific ion effects at the silica/water interface. J. Phys. Condens. Matter 26 (24), Chilingar, G. v, Yen, T.F., 1983. Some notes on wettability and relative permeability of
244107. carbonate rocks. II. Energy Sources 7 (1), 67–75.
Baloch, S.A., Leon, J.M., Masalmeh, S.K., Chappell, D., Brodie, J., Romero, C., Al Chou, S.I., Shah, D.O., 1981. The effect of H2O and D2O on colloidal properties of
Mazrouei, S.H., Al Tenaiji, A., Al Balooshi, M., Igogo, A., Azam, M.U.F., 2021. surfactant solutions and microemulsions. J. Colloid Interface Sci. 80 (1), 49–57.
Expanding polymer injectivity tests on a second giant carbonate UAE oil reservoir at Chowdhury, S., Shrivastava, S., Kakati, A., Sangwai, J.S., 2022. Comprehensive review
high salinity & high temperature conditions. In: Abu Dhabi International Petroleum on the role of surfactants in the chemical enhanced oil recovery process. Ind. Eng.
Exhibition & Conference. OnePetro. Chem. Res. 61 (1), 21–64.
Barakat, Y., Gendy, T.S., Mohamad, A.I., Youssef, A.F.M., 1989. Polymeric surfactants for Cockcroft, P., Anli, J., Duignan, J., 1988. EOR Potential of Indonesian Reservoirs.
enhanced oil recovery. Part I—critical micelle concentration of some ethoxylated Crain, E.R., 2002. Crain’s Petrophysical Handbook. Spectrum 2000 Mindware Limited.
alkylphenol—formaldehyde nonionics. Br. Polym. J. 21 (5), 383–389. Cui, L., Ma, K., Abdala, A.A., Lu, L.J., Tanakov, I., Biswal, S.L., Hirasaki, G.J., 2015.
Barari, M., Lashkarbolooki, M., Abedini, R., 2021. Interfacial properties of crude oil/ Adsorption of a switchable cationic surfactant on natural carbonate minerals. SPE J.
imidazolium based ionic liquids in the presence of NaCl and Na2SO4 during EOR 20 (1), 70–78.
process. J. Mol. Liq. 327, 114845. Davison, P., Mentzer, E., 1982. Polymer flooding in north sea reservoirs. Soc. Petrol. Eng.
Bashir, A., Haddad, A.S., Rafati, R., 2022. A review of fluid displacement mechanisms in J. 22 (3), 353–362.
surfactant-based chemical enhanced oil recovery processes: analyses of key De Oliveira Silva, H.G., Pires, A.J.V., da Cunha Neto, P.A., de Carvalho, G.G.P., Veloso, C.
influencing factors. Petrol. Sci. M., Ferreira daSilva, F., 2007. Digestibilidade de dietas contendo silagem de capim-

31
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

elefante amonizado e farelo de cacau ou torta de dendê em ovinos. Rev. Bras. Zootec. macromolecular chains of a water-soluble copolymer. J. Appl. Polym. Sci. 96 (3),
36 (2). 714–722.
Dean, R.M., 2011. Selection and Evaluation of Surfactants for Field Pilots. The University Gao, B., Yu, Y., Jiang, L., 2007. Studies on micellar behavior of anionic and surface-
of Texas at Austin. M.Sc. thesis. active monomers with acrylamide type in aqueous solutions. Colloids Surf. A
Delshad, M., Han, C., Veedu, F.K., Pope, G.A., 2011. A simplified model for simulations Physicochem. Eng. Asp. 293 (1–3), 210–216.
of alkaline-surfactant-polymer floods. In: SPE Reservoir Simulation Symposium. Gao, C.H., 2011. Scientific research and field applications of polymer flooding in heavy
Deng, X., Kamal, M.S., Patil, S., Hussain, S.M.S., Zhou, X., Mahmoud, M., 2021. oil recovery. J. Pet. Explor. Prod. Technol. 1 (2–4), 65–70.
Wettability alteration of locally synthesized cationic gemini surfactants on carbonate Gao, C., 2018. Experiences of microbial enhanced oil recovery in Chinese oil fields.
rock. J. Mol. Liq. 344, 117817. J. Petrol. Sci. Eng. 166, 55–62.
Derian, P.-J., Guerin, G., Fiard, J.-F., 1993. Microemulsions of pyrethroids: phase Gao, C., Shi, J., Zhao, F., 2014. Successful polymer flooding and surfactant-polymer
diagrams and effectiveness of tristyrylphenol based surfactants. In: Pesticide flooding projects at Shengli Oilfield from 1992 to 2012. J. Pet. Explor. Prod.
Formulations and Application Systems: Twelfth Volume. ASTM International. Technol. 4 (1), 1–8.
Diab, W.N., Al-Shalabi, E.W., 2019. Recent developments in polymer flooding for Gbadamosi, A., Patil, S., Al Shehri, D., Kamal, M.S., Hussain, S.S., Al-Shalabi, E.W.,
carbonate reservoirs under harsh conditions. Offshore Technology Conference Brasil. Hassan, A.M., 2022. Recent advances on the application of low salinity
Dicharry, C., Diaz, J., Torré, J.P., Ricaurte, M., 2016. Influence of the carbon chain waterflooding and chemical enhanced oil recovery. Energy Rep. 8, 9969–9996.
length of a sulfate-based surfactant on the formation of CO2, CH4 and CO2–CH4 gas Geetha, S.J., Banat, I.M., Joshi, S.J., 2018. Biosurfactants: production and potential
hydrates. Chem. Eng. Sci. 152, 736–745. applications in microbial enhanced oil recovery (MEOR). Biocatal. Agric. Biotechnol.
Dupuis, G., Antignard, S., Giovannetti, B., Gaillard, N., Jouenne, S., Bourdarot, G., 14, 23–32.
Zaitoun, A., 2017. A new thermally stable synthetic polymer for harsh conditions of Gizzatov, A., Mashat, A., Kosynkin, D., Alhazza, N., Kmetz, A., Eichmann, S.L., Abdel-
Middle East reservoirs. In: Part I. Thermal Stability and Injection in Carbonate Cores. Fattah, A.I., 2019. Nanofluid of petroleum sulfonate nanocapsules for enhanced oil
Abu Dhabi International Petroleum Exhibition and Conference. recovery in high-temperature and high-salinity reservoirs. Energy Fuel. 33 (11),
El-Batanoney, M., Abdel-Moghny, T., Ramzi, M., 1999. The effect of mixed surfactants on 11567–11573.
enhancing oil recovery. J. Surfactants Deterg. 2 (2), 201–205. Gomari, K.E., 2022. Hybrid Polymer/biopolymer/biosurfactant Flooding Influence on
El-Diasty, A.I., Aly, A.M., 2015. Understanding the mechanism of nanoparticles the Displacement Efficiency in Fractured Carbonate Reservoirs. PhD Dissertation.
applications in enhanced oil recovery. In: SPE North Africa Technical Conference Teesside University.
and Exhibition. Gomari, K.E., Hughes, D., Islam, M., Gomari, S.R., 2021. Application of water-soluble
El-Hoshoudy, A.N., Desouky, S.E.M., Al-Sabagh, A.M., Betiha, M.A., My, E., polymer/biopolymer combined with a biosurfactant in oil-wet fractured carbonate
Mahmoud, S., 2017b. Evaluation of solution and rheological properties for reservoirs. ACS Omega 6 (24), 15674–15685.
hydrophobically associated polyacrylamide copolymer as a promised enhanced oil Goodlett, G.O., Honarpour, M.M., Chung, F.T., Sarathi, P.S., 1986. The role of screening
recovery candidate. Egyptian Journal of Petroleum 26 (3), 779–785. and laboratory flow studies in EOR process evaluation. In: SPE Rocky Mountain
El Hoshoudy, A., Desouky, S., Al-sabagh, A., El-kady, M., Betiha, M., Mahmoud, S., Regional Meeting.
2015a. Synthesis and characterization of polyacrylamide crosslinked copolymer for Goudarzi, A., Delshad, M., Mohanty, K.K., Sepehrnoori, K., 2012. Impact of matrix block
enhanced oil recovery and rock wettability alteration. Int. J. Oil Gas Coal Eng. 3 (4), size on oil recovery response using surfactants in fractured carbonates. In: SPE
43–55. Annual Technical Conference and Exhibition.
El-Hoshoudy, A.N., Desouky, S.E.M., Betiha, M.A., Alsabagh, A.M., 2016. Use of 1-vinyl Gouveia, L.M., Grassl, B., Müller, A.J., 2009. Synthesis and rheological properties of
imidazole based surfmers for preparation of polyacrylamide–SiO2 nanocomposite hydrophobically modified polyacrylamides with lateral chains of poly (propylene
through aza-Michael addition copolymerization reaction for rock wettability oxide) oligomers. J. Colloid Interface Sci. 333 (1), 152–163.
alteration. Fuel 170, 161–175. Gouveia, L.M., Paillet, S., Khoukh, A., Grassl, B., Müller, A.J., 2008. The effect of the
El-Hoshoudy, A.N., Desouky, S.M., Betiha, M.H., Alsabagh, A.M., 2017a. Hydrophobic ionic strength on the rheological behavior of hydrophobically modified
polymers flooding. Application and Characterization of Surfactants. InTechopen polyacrylamide aqueous solutions mixed with sodium dodecyl sulfate (SDS) or
75–95. cetyltrimethylammonium p-toluenesulfonate (CTAT). Colloids Surf. A Physicochem.
El-Hoshoudy, A.N., Desouky, S.E.M., Elkady, M.Y., Alsabagh, A.M., Betiha, M.A., Eng. Asp. 322 (1–3), 211–218.
Mahmoud, S., 2015b. Investigation of optimum polymerization conditions for Green, D.W., Willhite, G.P., 1998. Enhanced Oil Recovery, vol. 6. Henry L. Doherty
synthesis of cross-linked polyacrylamide-amphoteric surfmer nanocomposites for Memorial Fund of AIME, Society of Petroleum Engineers.
polymer flooding in sandstone reservoirs. International Journal of Polymer Science, Grigg, R.B., Baojun, B., Yi, L., 2004. Competitive adsorption of a hybrid surfactant system
2015. onto five minerals, berea sandstone, and limestone. In: SPE Annual Technical
Emadi, S., Shadizadeh, S.R., Manshad, A.K., Rahimi, A.M., Mohammadi, A.H., 2017. Conference and Exhibition. OnePetro.
Effect of nano silica particles on Interfacial Tension (IFT) and mobility control of Gubelmann-Bonneau, I.V., Mailhe, P.A., Perrin, M.A., 1995. Tristyrylphenol surfactants
natural surfactant (Cedr Extraction) solution in enhanced oil recovery process by in agricultural formulations: properties and challenges in applications. In: Pesticide
nano-surfactant flooding. J. Mol. Liq. 248, 163–167. Formulations and Application Systems: Fourteenth Volume.
Emrani, A.S., Nasr-El-Din, H.A., 2015. Stabilizing CO2-foam using nanoparticles. In: SPE Guo, H., Li, Y., Kong, D., Ma, R., Li, B., Wang, F., 2019. Lessons learned from alkali/
European Formation Damage Conference And Exhibition. OnePetro. surfactant/polymer-flooding field tests in China. SPE Reservoir Eval. Eng. 22 (1),
Esfandiarian, A., Maghsoudian, A., Shirazi, M., Mohammadi, M., Kord, S., Tamsilian, Y., 78–99.
2021. Experimental investigation of using ionic-liquids as alternatives of surfactants Guo, L., Han, M., Fuseni, A., AlSofi, A., 2016. Laboratory investigation of polymeric
in enhanced-oil-recovery processes for harsh carbonate reservoirs. 82nd EAGE surfactants for EOR in high salinity and high temperature reservoir. In: SPE EOR
Annual Conference and Exhibition 2021. No. 1. Conference at Oil and Gas West Asia.
Eshraghi, S., Kazemzadeh, Y., Qahramanpour, M., Kazemi, A., 2017. Investigating effect Gupta, R., Mohanty, K.K., 2010. Temperature effects on surfactant-aided imbibition into
of SiO2 nanoparticle and sodium-dodecyl-sulfate surfactant on surface properties: fractured carbonates. SPE J. 15 (3), 588–597.
wettability alteration and IFT reduction. J. Petrol Environ. Biotechnol. 8, 349–353. Gupta, R.A., Mohanty, K.K., 2011. Wettability alteration mechanism for oil recovery
Evani, S., Rose, G.D., 1987. Water soluble hydrophobe association polymers. Polym. from fractured carbonate rocks. Transport Porous Media 87 (2), 635–652.
Mater. Sci. Eng. 57, 477–481. Gupta, S.K., Sharma, A., Tapadia, K., 2020. A convenient nanodrop spectrophotometric
Fava, A., Eyring, H., 1956. Equilibrium and kinetics of detergent adsorption–a approach to determine cationic surfactant Tetraoctylammonium bromide (TOAB) in
generalized equilibration theory. J. Phys. Chem. 60 (7), 890–898. environmental and biological samples. Chem. Pap. 74 (6), 1791–1798.
Feng, Y., Billon, L., Grassl, B., Bastiat, G., Borisov, O., François, J., 2005. Guyot, A., 2004. Advances in reactive surfactants. Adv. Colloid Interface Sci. 108, 3–22.
Hydrophobically associating polyacrylamides and their partially hydrolyzed Hamoud, A.S., 1995. Enhanced Oil Recovery in Carbonate Reservoirs. Al Baath
derivatives prepared by post-modification. 2. Properties of non-hydrolyzed polymers University, Hims.
in pure water and brine. Polymer 46 (22), 9283–9295. Hanamertani, A.S., Pilus, R.M., Idris, A.K., Irawan, S., Tan, I.M., 2018. Ionic liquids as a
Flaaten, A., Nguyen, Q.P., Zhang, J., Mohammadi, H., Pope, G.A., 2008. ASP chemical potential additive for reducing surfactant adsorption onto crushed Berea sandstone.
flooding without the need for soft water. In: SPE Annual Technical Conference and J. Petrol. Sci. Eng. 162, 480–490.
Exhibition. Han, M., AlSofi, A., Fuseni, A., Zhou, X., Hassan, S., 2013. Development of chemical EOR
Folk, R.L., 1959. Practical petrographic classification of limestones. AAPG (Am. Assoc. formulations for a high temperature and high salinity carbonate reservoir. In: IPTC
Pet. Geol.) Bull. 43 (1), 1–38. 2013: International Petroleum Technology Conference, Cp-350.
Freedman, H.H., Manson, J.P., Medalia, A.I., 1958. Polysoaps. II. The preparation of Harris, R.R., Paramanathan, S., Dickson, S.J., Bouwmeester, R.C., Southwick, J.G.,
vinyl soaps1. J. Org. Chem. 23 (1), 76–82. Keijzer, E., van Rijn, C.H., Batenburg, D.W., 2020 August. Evaluation of seawater
Fuseni, A., Han, M., Al-Mobith, A., 2013. Phase Behavior and Interfacial Tension based surfactant polymer formulation in a single well test offshore. In: SPE Improved
Properties of an Amphoteric Surfactant for EOR Application, pp. 1–14. Oil Recovery Conference. OnePetro.
Gao, B., Sharma, M.M., 2013. A new family of anionic surfactants for enhanced-oil- Hartono, A.D., Hakiki, F., Syihab, Z., Ambia, F., Yasutra, A., Sutopo, S., Efendi, M.,
recovery applications. SPE J. 18 (5), 829–840. Sitompul, V., Primasari, I., Apriandi, R., 2017. Revisiting EOR projects in Indonesia
Gao, B., Guo, H., Wang, J., Zhang, Y., 2008. Preparation of hydrophobic association through integrated study: EOR screening, predictive model, and optimization. In:
polyacrylamide in a new micellar copolymerization system and its hydrophobically SPE/IATMI Asia Pacific Oil and Gas Conference and Exhibition.
associative property. Macromolecules 41 (8), 2890–2897. Haruna, M.A., Nourafkan, E., Hu, Z., Wen, D., 2019. Improved polymer flooding in harsh
Gao, B., Wu, N., Li, Y., 2004. Study on tercopolymer of acrylamide containing strong environments by free-radical polymerization and the use of nanomaterials. Energy
anions and hydrophobic association blocks. Acta Polym. Sin. 5, 736–742. Fuel. 33 (2), 1637–1648.
Gao, B., Wu, N., Li, Y., 2005. Interaction between the strong anionic character of strong Hashmet, M.R., AlSumaiti, A.M., Qaiser, Y., AlAmeri, W., 2017. Laboratory investigation
anions and the hydrophobic association property of hydrophobic blocks in and simulation modeling of polymer flooding in high-temperature, high-salinity
carbonate reservoirs. Energy Fuel. 31 (12), 13454–13465.

32
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Hassan, A.M., Ayoub, M., Eissa, M., Al-Shalabi, E.W., Almansour, A., Alquraishi, A., Jian, G., Puerto, M., Wehowsky, A., Miller, C., Hirasaki, G.J., Biswal, S.L., 2018.
2022a. Foamability and foam stability screening for smart water assisted foam Characterizing adsorption of associating surfactants on carbonates surfaces.
flooding: a new hybrid EOR method. In: International Petroleum Technology J. Colloid Interface Sci. 513, 684–692.
Conference. Joshi, D., Maurya, N.K., Kumar, N., Mandal, A., 2022. Experimental investigation of
Hassan, A.M., Ayoub, M., Eissa, M., Al-Shalabi, E.W., Al-Mansour, A., Al-Quraishi, A., silica nanoparticle assisted Surfactant and polymer systems for enhanced oil
2022b. Increasing reservoir recovery efficiency through laboratory-proven hybrid recovery. J. Petrol. Sci. Eng., 110791
smart water-assisted foam (SWAF) flooding in carbonate reservoirs. Energies 15 (9), Jouenne, S., Levache, B., Joly, M., Hourcq, C., Questel, M., Heurteux, G., 2019. Universal
3058. viscosifying behavior of acrylamide-based polymers used in EOR-application for QA/
Hassan, A.M., Al-Shalabi, E.W., AlAmeri, W., Kamal, M.S., Patil, S., Shakil Hussain, S.M., QC, viscosity predictions and field characterization. IOR 2019–20th European
2022c. New insights into hybrid low salinity polymer (LSP) flooding through a Symposium on Improved Oil Recovery 1, 1–23.
coupled geochemical-based modeling approach. In: SPE Annual Technical Kalpakci, B., Jeans, Y.T., Magri, N.F., Padolewski, J.P., 1990. Thermal stability of
Conference and Exhibition. OnePetro. scleroglucan at realistic reservoir conditions. In: SPE/DOE Enhanced Oil Recovery
Hassan, A.M., Al-Shalabi, E.W., Ayoub, M.A., 2022d. Updated perceptions on polymer- Symposium.
based enhanced oil recovery toward high-temperature high-salinity tolerance for Kamal, M.S., 2016. A review of gemini surfactants: potential application in enhanced oil
successful field applications in carbonate reservoirs. Polymers 14 (10), 2001. recovery. J. Surfactants Deterg. 19 (2), 223–236.
Hassan, A., Ayoub, M., Eissa, M., Bruining, H., Al-Mansour, A., Al-Quraishi, A., 2021. Kamal, M.S., Hussein, I.A., Sultan, A.S., 2017. Review on surfactant flooding: phase
A new hybrid improved and enhanced oil recovery IOR/EOR process using smart behavior, retention, IFT, and field applications. Energy Fuel. 31 (8), 7701–7720.
water assisted foam SWAF flooding in carbonate rocks; A laboratory study approach. Karnanda, W., Benzagouta, M.S., AlQuraishi, A., Amro, M.M., 2013. Effect of
In: International Petroleum Technology Conference. temperature, pressure, salinity, and surfactant concentration on IFT for surfactant
Hassan, A.M., Ayoub, M., Eissa, M., Bruining, H., Zitha, P., 2020. Study of surface flooding optimization. Arabian J. Geosci. 6 (9), 3535–3544.
complexation modeling on a novel hybrid enhanced oil recovery (EOR) method; Kang, P.-S., Lim, J.-S., Huh, C., 2014. Screening criteria for application of EOR processes
smart-water assisted foam-flooding. J. Petrol. Sci. Eng. 195, 107563. in offshore fields. In: The Twenty-Fourth International Ocean and Polar Engineering
Hassan, A.M., Ayoub, M., Eissa, M., Musa, T., Bruining, H., Farajzadeh, R., 2019. Exergy Conference.
return on exergy investment analysis of natural-polymer (Guar-Arabic gum) Karimov, D., Hashmet, M.R., Pourafshary, P., 2020. A laboratory study to optimize ion
enhanced oil recovery process. Energy 181, 162–172. composition for the hybrid low salinity water/polymer flooding. In: Offshore
Healy, R.N., Reed, R.L., Stenmark, D.G., 1976. Multiphase microemulsion systems. Soc. Technology Conference Asia.
Petrol. Eng. J. 16 (3), 147–160. Karović-Maričić, V., Leković, B., Danilović, D., 2014. Factors influencing successful
Hill, H.J., Reisberg, J., Stegemeier, G.L., 1973. Aqueous surfactant systems for oil implementation of enhanced oil recovery projects. Podzemni Radovi 22 (25), 41–50.
recovery. J. Petrol. Technol. 25 (2), 186–194. Kasha, A., Al-Hashim, H., Abdallah, W., Taherian, R., Sauerer, B., 2015. Effect of Ca2+,
Hiorth, A., Cathles, L.M., Madland, M.V., 2010. The impact of pore water chemistry on Mg2+ and SO42− ions on the zeta potential of calcite and dolomite particles aged with
carbonate surface charge and oil wettability. Transport Porous Media 85 (1), 1–21. stearic acid. Colloids Surf. A Physicochem. Eng. Asp. 482, 290–299.
Hirasaki, G.J., Miller, C.A., Puerto, M., 2011. Recent advances in surfactant EOR. SPE J. Khademolhosseini, R., Jafari, A., Mousavi, S.M., Hajfarajollah, H., Noghabi, K.A.,
16 (4), 889–907. Manteghian, M., 2019a. Physicochemical characterization and optimization of
Hirasaki, G., Zhang, D.L., 2003. Surface chemistry of oil recovery from fractured, oil-wet, glycolipid biosurfactant production by a native strain of Pseudomonas aeruginosa
carbonate formation. In: International Symposium on Oilfield Chemistry. HAK01 and its performance evaluation for the MEOR process. RSC Adv. 9 (14),
Holmberg, K., 2003. Novel Surfactants: Preparation Applications and Biodegradability, 7932–7947.
Revised and Expanded, vol. 114. Crc Press. Khademolhosseini, R., Jafari, A., Mousavi, S.M., Manteghian, M., 2019b. Investigation of
Hong, P., Fa, C., Wei, Y., Sen, Z., 2007. Surface properties and synthesis of the cellulose- synergistic effects between silica nanoparticles, biosurfactant and salinity in
based amphoteric polymeric surfactant. Carbohydr. Polym. 69 (4), 625–630. simultaneous flooding for enhanced oil recovery. RSC Adv. 9 (35), 20281–20294.
Hu, K., Bard, A.J., 1997. Characterization of adsorption of sodium dodecyl sulfate on Khademolhosseini, R., Jafari, A., Shabani, M.H., 2015. Micro scale investigation of
charge-regulated substrates by atomic force microscopy force measurements. enhanced oil recovery using nano/bio materials. Procedia Materials Science 11,
Langmuir 13 (20), 5418–5425. 171–175.
Hu, S.-S., Zhang, L., Cao, X.-L., Guo, L.-L., Zhu, Y.-W., Zhang, L., Zhao, S., 2015. Influence Klimenko, A., Molinier, V., Dubos, F., Joly, M., Saint-Loubert, M., Jouenne, S., Passade-
of crude oil components on interfacial dilational properties of hydrophobically Boupat, N., Bourrel, M., 2020. Surfactant–polymer flooding at high temperature and
modified polyacrylamide. Energy Fuel. 29 (3), 1564–1573. high salinity: promising lab scale experiments in challenging conditions. In: Abu
Huang, Y., Santore, M.M., 2002. Dynamics in adsorbed layers of associative polymers in Dhabi International Petroleum Exhibition and Conference.
the limit of strong backbone− surface attractions. Langmuir 18 (6), 2158–2165. Kosmulski, M., 2002. The pH-dependent surface charging and the points of zero charge.
Huang, Y., Sorbie, K.S., 1993. Scleroglucan behavior in flow through porous media: J. Colloid Interface Sci. 253 (1), 77–87.
comparison of adsorption and in-situ rheology with xanthan. SPE International Kovscek, A.R., Radke, C.J., 1993. Fundamentals of Foam Transport in Porous Media.
Symposium on Oilfield Chemistry. Kuang, W., Saraji, S., Piri, M., 2018. A systematic experimental investigation on the
Huh, C., 1979. Interfacial tensions and solubilizing ability of a microemulsion phase that synergistic effects of aqueous nanofluids on interfacial properties and their
coexists with oil and brine. J. Colloid Interface Sci. 71 (2), 408–426. implications for enhanced oil recovery. Fuel 220, 849–870.
Hussain, S.M.S., Kamal, M.S., 2017. Effect of large spacer on surface activity, thermal, Kulawardana, E.U., Koh, H., Kim, D.H., Liyanage, P.J., Upamali, K.A., Huh, C.,
and rheological properties of novel amido-amine cationic gemini surfactants. J. Mol. Weerasooriya, U., Pope, G.A., 2012. Rheology and transport of improved EOR
Liq. 242, 1131–1137. polymers under harsh reservoir conditions. In: SPE Improved Oil Recovery
Hussain, S.M.S., Kamal, M.S., Murtaza, M., 2019. Effect of aromatic spacer groups and Symposium.
counterions on aqueous micellar and thermal properties of the synthesized Kumar, A., Mandal, A., 2017. Synthesis and physiochemical characterization of
quaternary ammonium gemini surfactants. J. Mol. Liq. 296, 111837. zwitterionic surfactant for application in enhanced oil recovery. J. Mol. Liq. 243,
Iglauer, S., Wu, Y., Shuler, P., Tang, Y., Goddard III, W.A., 2009. Alkyl polyglycoside 61–71.
surfactant–alcohol cosolvent formulations for improved oil recovery. Colloids Surf. A Kumar, A., Mandal, A., 2019. Critical investigation of zwitterionic surfactant for
Physicochem. Eng. Asp. 339 (1–3), 48–59. enhanced oil recovery from both sandstone and carbonate reservoirs: adsorption,
Iglauer, S., Wu, Y., Shuler, P., Tang, Y., Goddard III, W.A., 2010. New surfactant classes wettability alteration and imbibition studies. Chem. Eng. Sci. 209, 115222.
for enhanced oil recovery and their tertiary oil recovery potential. J. Petrol. Sci. Eng. Kwok, W., Nasr-El-Din, H.A., Hayes, R.E., Sethi, D., 1993. Static and dynamic adsorption
71 (1–2), 23–29. of a non-ionic surfactant on Berea sandstone. Colloids Surf. A Physicochem. Eng.
Inada, T., Koyama, T., Tomita, H., Fuse, T., Kuwabara, C., Arakawa, K., Fujikawa, S., Asp. 78, 193–209.
2017. Anti-ice nucleating activity of surfactants against silver iodide in water-in-oil Lager, A., Webb, K.J., Black, C.J.J., Singleton, M., Sorbie, K.S., 2008. Low salinity oil
emulsions. J. Phys. Chem. B 121 (27), 6580–6587. recovery-an experimental investigation1. Petrophysics-The SPWLA Journal of
Ishiguro, M., Koopal, L.K., 2016. Surfactant adsorption to soil components and soils. Adv. Formation Evaluation and Reservoir Description 49 (1).
Colloid Interface Sci. 231, 59–102. Lake, L.W., Johns, R., Rossen, B., Pope, G.A., 2014. Fundamentals of Enhanced Oil
Islam, N., Sharker, K.K., Sarker, K.C., 2015. Salt-induced modulation of the Krafft Recovery, vol. 1. Society of Petroleum Engineers Richardson, TX.
temperature and critical micelle concentration of Larestani, A., Mousavi, S.P., Hadavimoghaddam, F., Ostadhassan, M., Sarapardeh, A.H.,
benzyldimethylhexadecylammonium chloride. J. Surfactants Deterg. 18 (4), 2022. Predicting the surfactant-polymer flooding performance in chemical enhanced
651–659. oil recovery: cascade neural network and gradient boosting decision tree. Alex. Eng.
Jabbar, M., Xiao, R., Teletzke, G.F., Willingham, T., Al Obeidli, A., Al Sowaidi, A., J. 61 (10), 7715–7731.
Britton, C., Delshad, M., Li, Z., 2018. Polymer EOR assessment through integrated Leon, J.M., Masalmeh, S.K., Xu, S., AlSumaiti, A.M., BinAmro, A.A., Baslaib, M.A., 2021.
laboratory and simulation evaluation for an offshore Middle East carbonate December. Analysis of the World’s First Polymer Injectivity Test in a Carbonate
reservoir. In: Abu Dhabi International Petroleum Exhibition and Conference. Reservoir Under Extreme Harsh Conditions in ADNOC’s Reservoirs. In: Abu Dhabi
Jain, S., Pachisia, H., Sharma, A., Patel, S., Patel, S., Ragunathan, B., 2022. A systematic International Petroleum Exhibition & Conference. OnePetro.
review–Chemical EOR using surfactants and polymers. Mater. Today Proc. Leonhardt, B., Ernst, B., Reimann, S., Steigerwald, A., Lehr, F., 2014. Field testing the
Janaghi, P., Amani, H., Naseri, A., Kariminezhad, H., 2022. Accurate prediction of polysaccharide Schizophyllan: results of the first year. In: SPE Improved Oil
enhanced oil recovery from carbonate reservoir through smart injection of Recovery Symposium.
nanoparticle and biosurfactant. J. Petrol. Sci. Eng. 216, 110772. Levitt, D., Bourrel, M., 2016. Adsorption of EOR chemicals under laboratory and
Jian, G., Puerto, M., Wehowsky, A., Dong, P., Johnston, K.P., Hirasaki, G.J., Biswal, S.L., reservoir conditions, part III: chemical treatment methods. In: SPE Improved Oil
2016. Static adsorption of an ethoxylated nonionic surfactant on carbonate minerals. Recovery Conference.
Langmuir 32 (40), 10244–10252.

33
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Levitt, D., Dufour, S., Pope, G.A., Morel, D.C., Gauer, P.R., 2011. Design of an ASP flood Masalmeh, S., AlSumaiti, A., Gaillard, N., Daguerre, F., Skauge, T., Skuage, A., 2019.
in a high-temperature, high-salinity, low-permeability carbonate. Int. Petrol. Extending polymer flooding towards high-temperature and high-salinity carbonate
Technol. Conf. reservoirs. In: Abu Dhabi International Petroleum Exhibition and Conference.
Levitt, D., Jackson, A., Heinson, C., Britton, L.N., Malik, T., Dwarakanath, V., Pope, G.A., Masalmeh, S.K., Wei, L., Blom, C.P., 2011. Mobility control for gas injection in
2006. Identification and evaluation of high-performance EOR surfactants. In: SPE/ heterogeneous carbonate reservoirs: comparison of foams versus polymers. In: SPE
DOE Symposium on Improved Oil Recovery. Middle East Oil and Gas Show and Conference.
Levitt, D., Klimenko, A., Jouenne, S., Chamerois, M., Bourrel, M., 2012. Design Masalmeh, S.K., Wei, L., Hillgartner, H., Al-Mjeni, R., Blom, C., 2012. Developing high
challenges of chemical EOR in high-temperature, high salinity carbonates. In: Abu resolution static and dynamic models for waterflood history matching and EOR
Dhabi International Petroleum Conference and Exhibition. evaluation of a Middle Eastern carbonate reservoir. In: Abu Dhabi International
Levitt, D., Klimenko, A., Jouenne, S., Passade-Boupat, N., Cordelier, P., Morel, D., Petroleum Conference and Exhibition.
Bourrel, M., 2016. April. Designing and injecting a chemical formulation for a Massarweh, O., Abushaikha, A.S., 2020. The use of surfactants in enhanced oil recovery:
successful off-shore chemical EOR pilot in a high-temperature, high-salinity, low- a review of recent advances. Energy Rep. 6, 3150–3178.
permeability carbonate field. In: SPE Improved Oil Recovery Conference. OnePetro. Mayer, E.H., Berg, R.L., Carmichael, J.D., Weinbrandt, R.M., 1983. Alkaline injection for
Levitt, D.B., Weatherl, R.K., Harris, H.W., McNeil, R.I., Didier, M., Loriau, M., Gaucher, E. enhanced oil recovery-A status report. J. Petrol. Technol. 35 (1), 209–221.
C., Bourrel, M., 2015. Adsorption of EOR chemicals under laboratory and reservoir Mehrabianfar, P., Bahraminejad, H., Manshad, A.K., 2021. An introductory investigation
conditions, Part 1-iron abundance and oxidation state. In: IOR 2015-18th European of a polymeric surfactant from a new natural source in chemical enhanced oil
Symposium on Improved Oil Recovery cp-445. recovery (CEOR). J. Petrol. Sci. Eng. 198, 108172.
Li, B., Liu, Z., Fei, C., Lv, J., Chen, X., Li, Y., 2018. Polymeric surfactant for enhanced oil Mejia, L., Tagavifar, M., Xu, K., Mejia, M., Du, Y., Balhoff, M., 2019. Surfactant flooding
recovery-microvisual, core-flood experiments and field application. In: SPE EOR in oil-wet micromodels with high permeability fractures. Fuel 241, 1117–1128.
Conference at Oil and Gas West Asia. Meng, R., Wang, C., Jin, J., Wang, R., Deng, L., 2022. Self-assembly of hydrophobically
Li, G., Xu, J., Mu, J., Zhai, L., Shui, L., Chen, W., Jiang, J., Chen, F., Guo, D., Lin, W., associating amphiphilic polymer with surfactant and its effect on nanoemulsion.
2005. Design and application of an alkaline-surfactant-polymer flooding system in Colloids Surf. A Physicochem. Eng. Asp. 642, 128599.
field pilot test. J. Dispersion Sci. Technol. 26 (6), 709–717. Millemann, R.E., Haynes, R.J., Boggs, T.A., Hildebrand, S.G., 1982. Enhanced oil
Li, G.-Z., Mu, J.-H., Li, Y., Xiao, H.-D., Gu, Q., 2000. What is the criterion for selecting recovery: environmental issues and state regulatory programs. Environ. Int. 7 (3),
alkaline/surfactant/polymer flooding formulation: phase behavior or interfacial 165–177.
tension. J. Dispersion Sci. Technol. 21 (3), 305–314. Miyake, M., Oyama, N., 2009. Effect of amido alkyl group as spacer on aggregation
Liao, G., Wang, Q., Wang, H., Liu, W., Wang, Z., 2017. Chemical flooding development properties of guanidine-type surfactants. J. Colloid Interface Sci. 330 (1), 180–185.
status and prospect. Acta Pet. Sin. 38 (2), 196–207. Moghadam, A.M., Salehi, M.B., 2019. Enhancing hydrocarbon productivity via
Lippmann, F., 2012. Sedimentary Carbonate Minerals, vol. 6. Springer Science & wettability alteration: a review on the application of nanoparticles. Rev. Chem. Eng.
Business Media. 35 (4), 531–563.
Liu, S., Li, R.F., Miller, C.A., Hirasaki, G.J., 2010. Alkaline/surfactant/polymer processes: Moghaddam, R.N., Bahramian, A., Fakhroueian, Z., Karimi, A., Arya, S., 2015.
wide range of conditions for good recovery. SPE J. 15 (2), 282–293. Comparative study of using nanoparticles for enhanced oil recovery: wettability
Liu, Y., Lee, J.Y., Hong, L., 2004. In situ preparation of poly (ethylene oxide)–SiO2 alteration of carbonate rocks. Energy Fuel. 29 (4), 2111–2119.
composite polymer electrolytes. J. Power Sources 129 (2), 303–311. Mohajeri, M., Hemmati, M., Shekarabi, A.S., 2015. An experimental study on using a
Liu, Z., Zhou, G., Li, S., Wang, C., Liu, R., Jiang, W., 2020. Molecular dynamics nanosurfactant in an EOR process of heavy oil in a fractured micromodel. J. Petrol.
simulation and experimental characterization of anionic surfactant: influence on Sci. Eng. 126, 162–173.
wettability of low-rank coal. Fuel 279, 118323. Mohammadi, H., Delshad, M., Pope, G.A., 2009. Mechanistic modeling of alkaline/
Liyanage, P.J., Lu, J., Arachchilage, G.W.P., Weerasooriya, U.P., Pope, G.A., 2015. surfactant/polymer floods. SPE Reservoir Eval. Eng. 12 (4), 518–527.
A novel class of large-hydrophobe tristyrylphenol (TSP) alkoxy sulfate surfactants for Moreno-Arciniegas, L., Babadagli, T., 2017. Experimental analysis of optimal
chemical enhanced oil recovery. J. Petrol. Sci. Eng. 128, 73–85. thermodynamic conditions for heavy-oil/bitumen recovery considering effective
Liyanage, P.J., Solairaj, S., Arachchilage, G.W., Linnemeyer, H.C., Kim, D.H., solvent retrieval. SPE Reservoir Eval. Eng. 20 (1), 149–160.
Weerasooriya, U., Pope, G.A., 2012. Alkaline surfactant polymer flooding using a Morrow, N.R., Mason, G., 2001. Recovery of oil by spontaneous imbibition. Curr. Opin.
novel class of large hydrophobe surfactants. In: SPE Improved Oil Recovery Colloid Interface Sci. 6 (4), 321–337.
Symposium. Moulin, P., Roques, H., 2003. Zeta potential measurement of calcium carbonate.
Llano, V., Henthorne, L., Walsh, J., 2013. Water management for EOR applications- J. Colloid Interface Sci. 261 (1), 115–126.
sourcing, treating, reuse and recycle. In: Offshore Technology Conference. Musa, T.A., Ibrahim, A.F., Nasr-El-Din, H.A., Hassan, A., 2021. New insights into guar
Lu, J., 2014. Development of Novel Surfactants and Surfactant Methods for Chemical gum as environmentally friendly polymer for enhanced oil recovery in high-salinity
Enhanced Oil Recovery. The University of Texas at Austin. PhD Dissertation. and high-temperature sandstone reservoirs. Journal of Petroleum Exploration and
Lu, J., Britton, C., Solairaj, S., Liyanage, P.J., Kim, D.H., Adkins, S., Arachchilage, G.W., Production 11 (4), 1905–1913.
Weerasooriya, U., Pope, G.A., 2014b. Novel large-hydrophobe alkoxy carboxylate Mukhopadhyay, C., Suba, M., Sivakumar, D., Dhamodharan, K., Rao, R.V., 2019. Cloud
surfactants for enhanced oil recovery. SPE J. 19 (6), 1024–1034. points extractive spectrophotometric method for determination of uranium in
Lu, J., Goudarzi, A., Chen, P., Kim, D.H., Delshad, M., Mohanty, K.K., Sepehrnoori, K., raffinate streams during spent nuclear fuel reprocessing. J. Radioanal. Nucl. Chem.
Weerasooriya, U.P., Pope, G.A., 2014a. Enhanced oil recovery from high- 322 (2), 743–750.
temperature, high-salinity naturally fractured carbonate reservoirs by surfactant Mushtaq, M., Tan, I.M., Ismail, L., Lee, S.Y., Nadeem, M., Sagir, M., 2014. Oleate ester-
flood. J. Petrol. Sci. Eng. 124, 122–131. derived nonionic surfactants: synthesis and cloud point behavior studies.
Lu, J., Liyanage, P.J., Solairaj, S., Adkins, S., Arachchilage, G.P., Kim, D.H., Britton, C., J. Dispersion Sci. Technol. 35 (3), 322–328.
Weerasooriya, U., Pope, G.A., 2014c. New surfactant developments for chemical Na, C., Kendall, T.A., Martin, S.T., 2007. Surface-potential heterogeneity of reacted
enhanced oil recovery. J. Petrol. Sci. Eng. 120, 94–101. calcite and rhodochrosite. Environ. Sci. Technol. 41 (18), 6491–6497.
Lu, J., Pope, G.A., 2017. Optimization of gravity-stable surfactant flooding. SPE J. 22 (2), Nakagawa, H., Watanabe, S., Fujita, T., Sakamoto, M., 1998. Characteristic properties of
480–493. cutting fluid additives made from the derivatives of some polymeric nonionic
Luo, H., Al-Shalabi, E.W., Delshad, M., Panthi, K., Sepehrnoori, K., 2016. A robust surface-active agents. J. Surfactants Deterg. 1 (2), 207–211.
geochemical simulator to model improved-oil-recovery methods. SPE J. 21 (1), Nasiri, M.A., Biria, D., 2020. Extraction of the indigenous crude oil dissolved
55–73. biosurfactants and their potential in enhanced oil recovery. Colloids Surf. A
Lv, W., Bazin, B., Ma, D., Liu, Q., Han, D., Wu, K., 2011. Static and dynamic adsorption of Physicochem. Eng. Asp. 603, 125216.
anionic and amphoteric surfactants with and without the presence of alkali. J. Petrol. Negin, C., Ali, S., Xie, Q., 2017. Most common surfactants employed in chemical
Sci. Eng. 77 (2), 209–218. enhanced oil recovery. Petroleum 3 (2), 197–211.
Lwisa, E., 2021. Chemical Enhanced Oil Recovery. Nelson, R.C., 1983. The effect of live crude on phase behavior and oil-recovery efficiency
Ma, K., Cui, L., Dong, Y., Wang, T., Da, C., Hirasaki, G.J., Biswal, S.L., 2013. Adsorption of surfactant flooding systems. SPE J. 23 (3), 501–510.
of cationic and anionic surfactants on natural and synthetic carbonate materials. Nelson, R.C., Lawson, J.B., Thigpen, D.R., Stegemeier, G.L., 1984. Cosurfactant-enhanced
J. Colloid Interface Sci. 408, 164–172. alkaline flooding. In: SPE Enhanced Oil Recovery Symposium.
Mahani, H., Keya, A.L., Berg, S., Bartels, W.-B., Nasralla, R., Rossen, W.R., 2015. Insights Nevskaia, D.M., Guerrero-Ruız, A., López-González, J.D.D., 1998. Adsorption of
into the mechanism of wettability alteration by low-salinity flooding (LSF) in polyoxyethylenic nonionic and anionic surfactants from aqueous solution: effects
carbonates. Energy Fuel. 29 (3), 1352–1367. induced by the addition of NaCl and CaCl2. J. Colloid Interface Sci. 205 (1), 97–105.
Mahani, H., Menezes, R., Berg, S., Fadili, A., Nasralla, R., Voskov, D., Joekar-Niasar, V., Nieto-Alvarez, D.A., Zamudio-Rivera, L.S., Luna-Rojero, E.E., Rodríguez-Otamendi, D.I.,
2017. Insights into the impact of temperature on the wettability alteration by low Marín-León, A., Hernández-Altamirano, R., Mena-Cervantes, V.Y., Chávez-
salinity in carbonate rocks. Energy Fuel. 31 (8), 7839–7853. Miyauchi, T.E., 2014. Adsorption of zwitterionic surfactant on limestone measured
Mahboob, A., Kalam, S., Kamal, M.S., Hussain, S.M.S., Solling, T., 2022. EOR Perspective with high-performance liquid chromatography: micelle–vesicle influence. Langmuir
of microemulsions: a review. J. Petrol. Sci. Eng. 208, 109312. 30 (41), 12243–12249.
Maia, A.M.S., Borsali, R., Balaban, R.C., 2009. Comparison between a polyacrylamide Niu, Y., Jian, O., Zhu, Z., Wang, G., Sun, G., 2001. Research on hydrophobically
and a hydrophobically modified polyacrylamide flood in a sandstone core. Mater. associating water-soluble polymer used for EOR. In: SPE International Symposium
Sci. Eng. C 29 (2), 505–509. on Oilfield Chemistry.
Maltesh, C., Somasundaran, P., 1992. Binding of sodium dodecyl sulfate to polyethylene Nwidee, L.N., Lebedev, M., Barifcani, A., Sarmadivaleh, M., Iglauer, S., 2017. Wettability
oxide at the silica—water interface. J. Colloid Interface Sci. 153 (1), 298–301. alteration of oil-wet limestone using surfactant-nanoparticle formulation. J. Colloid
Mandal, A., 2015. Chemical flood enhanced oil recovery: a review. Int. J. Oil Gas Coal Interface Sci. 504, 334–345.
Technol. 9 (3), 241–264.

34
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Olajire, A.A., 2014. Review of ASP EOR (alkaline surfactant polymer enhanced oil Development In Laboratory to the Implementation and Production Monitoring in an
recovery) technology in the petroleum industry: prospects and challenges. Energy Offshore Reservoir. In SPE Improved Oil Recovery Conference. OnePetro.
77, 963–982. Rohilla, N., Katiyar, A., Saxena, P., Pal, M., Rozowski, P., Gentilucci, A., 2021. Successful
Omekeh, A., Friis, H.A., Fjelde, I., Evje, S., 2012. Modeling of ion-exchange and solubility wettability alteration pilot in an offshore reservoir: laboratory analysis to support
in low salinity water flooding. In: SPE Improved Oil Recovery Symposium. planning, implementation and interpretation. In: IOR 2021, vol. 2021. European
Páhi, A.B., Király, Z., Mastalir, Á., Dudás, J., Puskás, S., Vágó, Á., 2008. Thermodynamics Association of Geoscientists & Engineers, pp. 1–14. No. 1.
of micelle formation of the counterion coupled gemini surfactant bis (4-(2-dodecyl) Rosen, M.J., Kunjappu, J.T., 2012. Surfactants and Interfacial Phenomena. John Wiley
benzenesulfonate)-Jeffamine salt and its dynamic adsorption on sandstone. J. Phys. and Sons.
Chem. B 112 (48), 15320–15326. Sagi, A.R., Puerto, M., Bian, Y., Miller, C.A., Hirasaki, G.J., Salehi, M., Thomas, C.P.,
Páhi, A.B., Király, Z., Puskás, S., 2009. Mass spectrometric characterization of the non- Kwan, J.T., 2013. Laboratory studies for surfactant flood in low-temperature, low-
ionic gemini surfactant Surfynol 465 and a microcalorimetric study of its micelle salinity fractured carbonate reservoir. In: SPE International Symposium on Oilfield
formation in water. Colloids Surf. A Physicochem. Eng. Asp. 345 (1–3), 13–17. Chemistry.
Pal, S., Mushtaq, M., Banat, F., Al Sumaiti, A.M., 2018. Review of surfactant-assisted Sakthivel, S., 2021. Wettability alteration of carbonate reservoirs using imidazolium-
chemical enhanced oil recovery for carbonate reservoirs: challenges and future based ionic liquids. ACS Omega 6 (45), 30315–30326.
perspectives. Petrol. Sci. 15 (1), 77–102. Salager, J.L., Antón, R.E., Sabatini, D.A., Harwell, J.H., Acosta, E.J., Tolosa, L.I., 2005.
Panthi, K., Sotomayor, M., Balhoff, M.T., Mohanty, K.K., 2022. Low retention surfactant- Enhancing solubilization in microemulsions—state of the art and current trends.
polymer process for a HTHS carbonate reservoir. J. Petrol. Sci. Eng. 214, 110516. J. Surfactants Deterg. 8 (1), 3–21.
Parkhurst, D.L., Appelo, C.A.J., 1999. User’s guide to PHREEQC (Version 2): a computer Salehi, M., Johnson, S.J., Liang, J.T., 2008. Mechanistic study of wettability alteration
program for speciation, batch-reaction, one-dimensional transport, and inverse using surfactants with applications in naturally fractured reservoirs. Langmuir 24
geochemical calculations. Water-Resources Investigations Report 99 (4259), 312. (24), 14099–14107.
Patel, M.C., Ayoub, M.A., Hassan, A.M., Idress, M.B., 2022. A novel ZnO nanoparticles Samakande, A., Hartmann, P.C., Cloete, V., Sanderson, R.D., 2007. Use of acrylic based
enhanced surfactant based viscoelastic fluid systems for fracturing under high surfmers for the preparation of exfoliated polystyrene–clay nanocomposites.
temperature and high shear rate conditions: synthesis, rheometric analysis, and fluid Polymer 48 (6), 1490–1499.
model derivation. Polymers 14 (19), 4023. Samakande, A., Hartmann, P.C., Sanderson, R.D., 2006. Synthesis and characterization of
Pei, H., Shu, Z., Zhang, G., Ge, J., Jiang, P., Qin, Y., Cao, X., 2018. Experimental study of new cationic quaternary ammonium polymerizable surfactants. J. Colloid Interface
nanoparticle and surfactant stabilized emulsion flooding to enhance heavy oil Sci. 296 (1), 316–323.
recovery. J. Petrol. Sci. Eng. 163, 476–483. Sanati, A., Rahmani, S., Nikoo, A.H., Malayeri, M.R., Busse, O., Weigand, J.J., 2021.
Pham, T.D., Kobayashi, M., Adachi, Y., 2015. Adsorption of anionic surfactant sodium Comparative study of an acidic deep eutectic solvent and an ionic liquid as chemical
dodecyl sulfate onto alpha alumina with small surface area. Colloid Polym. Sci. 293 agents for enhanced oil recovery. J. Mol. Liq. 329, 115527.
(1), 217–227. Sarsenbekuly, B., Kang, W., Fan, H., Yang, H., Dai, C., Zhao, B., Aidarova, S.B., 2017.
Pillai, P., Mandal, A., 2022. Synthesis and characterization of surface-active ionic liquids Study of salt tolerance and temperature resistance of a hydrophobically modified
for their potential application in enhanced oil recovery. J. Mol. Liq. 345, 117900. polyacrylamide based novel functional polymer for EOR. Colloids Surf. A
Puerto, M., Hirasaki, G.J., Miller, C.A., Barnes, J.R., 2012. Surfactant systems for EOR in Physicochem. Eng. Asp. 514, 91–97.
high-temperature, high-salinity environments. SPE J. 17 (1), 11–19. Saxena, N., Kumar, A., Mandal, A., 2019. Adsorption analysis of natural anionic
Puntervold, T., 2008. Waterflooding of Carbonate Reservoirs: EOR by Wettability surfactant for enhanced oil recovery: the role of mineralogy, salinity, alkalinity and
Alteration. nanoparticles. J. Petrol. Sci. Eng. 173, 1264–1283.
Puskas, S., Vago, A., Toro, M., Ordog, T., Kalman, G., Hanzelik, P., Bihari, Z., Blaho, J., Schott, H., 2005. Krafft points and cloud points of polyoxyethylated nonionic surfactants.
Tabajdi, R., Dekany, I., 2018. Surfactant-polymer EOR from laboratory to the pilot. Part 1 Krafft points in binary surfactant-water systems. Tenside Surfactants Deterg.
In: SPE EOR Conference at Oil and Gas West Asia. 42 (6), 356–367.
Puskas, S., Vago, A., Törő, M., Ördög, T., Kálmán, G.Y., Tabajd, R., Dekany, I., Dudas, J., Schramm, L.L., 2000. Surfactants: Fundamentals and Applications in the Petroleum
Nagy, R., Bartha, L., 2017. First surfactant-polymer EOR injectivity test in the Algyő Industry. Cambridge university press.
field, Hungary. IOR 2017-19th European Symposium on Improved Oil Recovery 1, Seethepalli, A., Adibhatla, B., Mohanty, K.K., 2004a. Physicochemical interactions
1–18. during surfactant flooding of fractured carbonate reservoirs. SPE J. 9 (4), 411–418.
Qiao, C., Johns, R., Li, L., 2016. Understanding the chemical mechanisms for low salinity Seethepalli, A., Adibhatla, B., Mohanty, K.K., 2004b. Wettability alteration during
waterflooding. 78th EAGE Conference and Exhibition 1, 1–13. surfactant flooding of carbonate reservoirs. In: SPE/DOE Symposium on Improved
Rachapudi, R.V., Alshehhi, S.S., BinAmro, A.A., Masalmeh, S.K., Dey, A., Al Nuimi, S.M., Oil Recovery.
Kenawy, M.M., Fabbri, C., Romero, C., Xu, S., Mabrook, M., 2020. November. World Seo, S., Mastiani, M., Mosavati, B., Peters, D.M., Mandin, P., Kim, M., 2018. Performance
First Polymer Injectivity Test in High Salinity and High Temperature Carbonate evaluation of environmentally benign nonionic biosurfactant for enhanced oil
Reservoir, Case Study from a Giant Reservoir in UAE. In: Abu Dhabi International recovery. Fuel 234, 48–55.
Petroleum Exhibition & Conference. OnePetro. ShamsiJazeyi, H., Verduzco, R., Hirasaki, G.J., 2014b. Reducing adsorption of anionic
Raffa, P., 2021. Design and synthesis of low molecular weight and polymeric surfactants surfactant for enhanced oil recovery: Part I. Competitive adsorption mechanism.
for enhanced oil recovery. In: Surfactants in Upstream E and P. Springer, Cham, Colloids Surf. A Physicochem. Eng. Asp. 453, 162–167.
pp. 3–37. ShamsiJazeyi, H., Verduzco, R., Hirasaki, G.J., 2014a. Reducing adsorption of anionic
Raffa, P., Broekhuis, A.A., Picchioni, F., 2016a. Polymeric surfactants for enhanced oil surfactant for enhanced oil recovery: Part II. Applied aspects. Colloids Surf. A
recovery: a review. J. Petrol. Sci. Eng. 145, 723–733. Physicochem. Eng. Asp. 453, 168–175.
Raffa, P., Broekhuis, A.A., Picchioni, F., 2016b. Amphiphilic copolymers based on PEG- Sharma, H., Dufour, S., Arachchilage, G.W.P.P., Weerasooriya, U., Pope, G.A.,
acrylate as surface active water viscosifiers: towards new potential systems for Mohanty, K., 2015. Alternative alkalis for ASP flooding in anhydrite containing oil
enhanced oil recovery. J. Appl. Polym. Sci. 133 (42). reservoirs. Fuel 140, 407–420.
Raffa, P., Wever, D.A.Z., Picchioni, F., Broekhuis, A.A., 2015. Polymeric surfactants: Shashkina, Y.A., Zaroslov, Y.D., Smirnov, V.A., Philippova, O.E., Khokhlov, A.R.,
synthesis, properties, and links to applications. Chem. Rev. 115 (16), 8504–8563. Pryakhina, T.A., Churochkina, N.A., 2003. Hydrophobic aggregation in aqueous
Ragab, A., Mansour, E.M., 2021. Enhanced oil recovery: chemical flooding. Geophysics solutions of hydrophobically modified polyacrylamide in the vicinity of overlap
and Ocean Waves Studies 51. concentration. Polymer 44 (8), 2289–2293.
Ramsey, M., 2019. Schlumberger Oilfield Glossary. Schlumberger. Shaw, J.E., 1984. Carboxylate surfactant systems exhibiting phase behavior suitable for
Ranjbar, M., Schaffie, M., 2000. Improved treatment of acrylamide co-and terpolymers enhanced oil recovery. JAOCS (J. Am. Oil Chem. Soc.) 61 (8), 1395–1399.
for water control in gas-producing and storage wells. J. Petrol. Sci. Eng. 26 (1–4), Sheng, J.J., 2010. Modern Chemical Enhanced Oil Recovery: Theory and Practice. Gulf
133–141. Professional Publishing.
Reb, P., Margarit-Puri, K., Klapper, M., Müllen, K., 2000. Polymerizable and Sheng, J.J., 2013a. A comprehensive review of alkaline-surfactant-polymer (ASP)
nonpolymerizable isophthalic acid derivatives as surfactants in emulsion flooding. In: SPE Western Regional and AAPG Pacific Section Meeting 2013 Joint
polymerization. Macromolecules 33 (21), 7718–7723. Technical Conference.
Reed, R.L., Healy, R.N., 1977. Some physicochemical aspects of microemulsion flooding: Sheng, J.J., 2013b. Enhanced Oil Recovery Field Case Studies. Gulf Professional
a review. Improved Oil Recovery by Surfactant and Polymer Flooding 383–437. Publishing.
RezaeiDoust, A., Puntervold, T., Austad, T., 2011. Chemical verification of the EOR Sheng, J.J., 2015. Status of surfactant EOR technology. Petroleum 1 (2), 97–105.
mechanism by using low saline/smart water in sandstone. Energy Fuel. 25 (5), Shiran, B.S., Skauge, A., 2014. Similarities and differences of low salinity polymer and
2151–2162. low salinity LPS (linked polymer solutions) for enhanced oil recovery. J. Dispersion
Rilian, N.A., Sumestry, M., 2010. Surfactant stimulation to increase reserves in carbonate Sci. Technol. 35 (12), 1656–1664.
reservoir: a case study in Semoga Field. In: SPE EUROPEC/EAGE Annual Conference Skauge, A., Fotland, P., 1990. Effect of pressure and temperature on the phase behavior
and Exhibition. of microemulsions. SPE Reservoir Eng. 5 (4), 601–608.
Rivenq, R.C., Donche, A., Nolk, C., 1992. Improved scleroglucan for polymer flooding Skauge, A., Palmgren, O., 1989. Phase behavior and solution properties of ethoxylated
under harsh reservoir conditions. SPE Reservoir Eng. 7 (1), 15–20. anionic surfactants. SPE International Symposium on Oilfield Chemistry.
Rodrigues, J. de A., Lachter, E.R., de Sá, C.H., de Mello, M., Nascimento, R.S.V., 2006. Solairaj, S., Britton, C., Lu, J., Kim, D.H., Weerasooriya, U., Pope, G.A., 2012. New
New multifunctional polymeric additives for water-based muds. In: SPE Annual correlation to predict the optimum surfactant structure for EOR. In: SPE Improved
Technical Conference and Exhibition. Oil Recovery Symposium.
Roehl, P.O., Choquette, P.W., 2012. Carbonate Petroleum Reservoirs. Springer Science Somasundaran, P., Fuerstenau, D.W., 1966. Mechanisms of alkyl sulfonate adsorption at
and Business Media. the alumina-water interface1. J. Phys. Chem. 70 (1), 90–96.
Rohilla, N., Katiyar, A., Rozowski, P.M., Gentilucci, A., Patil, P.D., Pal, M., Saxena, P., Somasundaran, P., Krishnakumar, S., 1997. Adsorption of surfactants and polymers at
2020, August. Field Trial for Wettability Alteration Using Surfactants: Formulation the solid-liquid interface. Colloids Surf. A Physicochem. Eng. Asp. 123, 491–513.

35
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Somoza, A., Arce, A., Soto, A., 2021. Oil recovery tests with ionic liquids: a review and Walker, D.L., Britton, C., Kim, D.H., Dufour, S., Weerasooriya, U., Pope, G.A., 2012. The
evaluation of 1-decyl-3-methylimidazolium triflate. Petrol. Sci. impact of microemulsion viscosity on oil recovery. In: SPE Improved Oil Recovery
Somoza, A., Tafur, N., Arce, A., Soto, A., 2022. Design and performance analysis of a Symposium.
formulation based on SDBS and ionic liquid for EOR in carbonate reservoirs. Wang, D., Maubert, M., Pope, G.A., Liyanage, P.J., Jang, S.H., Upamali, K.A., Chang, L.,
J. Petrol. Sci. Eng. 209, 109856. Tagavifar, M., Sharma, H., Ren, G., 2019. Reduction of surfactant retention in
Song, W., Hatzignatiou, D.G., 2022. On the reduction of the residual oil saturation limestones using sodium hydroxide. SPE J. 24 (1), 92–115.
through the injection of polymer and nanoparticle solutions. J. Petrol. Sci. Eng. 208, Wang, J., Han, M., Fuseni, A.B., Cao, D., 2015. Surfactant adsorption in surfactant
109430. polymer flooding in carbonate reservoirs. SPE 172700-MS. In: SPE Middle East Oil
Sorbie, K.S., 1991. Introduction to polymer flooding. In: Polymer-Improved Oil and Gas Show Conference.
Recovery. Springer, Dordrecht, pp. 1–5. Wang, J., Xiao, L., Liao, G., Zhang, Y., Guo, L., Arns, C.H., Sun, Z., 2018. Theoretical
Sorbie, K.S., 2013. Polymer-improved Oil Recovery. Springer Science and Business investigation of heterogeneous wettability in porous media using NMR. Sci. Rep. 8
Media. (1), 1–14.
Souraki, Y., Hosseini, E., Yaghodous, A., 2019. Wettability alteration of carbonate Wang, L., Mohanty, K., 2015. Enhanced oil recovery in gas flooded carbonate reservoirs
reservoir rock using amphoteric and cationic surfactants: experimental investigation. by wettability-altering surfactants. SPE J. 20 (1), 60–69.
Energy Sources, Part A Recovery, Util. Environ. Eff. 41 (3), 349–359. Wang, W., Lu, X., 2009. Property evaluation of EOR technology by means of expansive
Southwick, J., van den Pol, E., van Rijn, C.H.T., van Batenburg, D.W., a Manap, A.A., granular crosslinked polymer. In: SPE Production and Operations Symposium.
Mastan, A.K., Zulkifli, N.N., 2019. Surfactant flooding in offshore environments. In: Wang, Y., Wu, F., 2015. Amphiphilic acrylamide-based copolymer with porphyrin
IOR 2019–20th European Symposium on Improved Oil Recovery, vol. 2019. pendants for the highly selective detection of Hg2+ in aqueous solutions. Polymer 56,
European Association of Geoscientists & Engineers, pp. 1–21. No. 1. 223–228.
Southwick, J.G., Brewer, M.L., Pieterse, S.G., van Batenburg, D.W., 2018. Ammonia as Winsor, P.A., 1954. Solvent Properties of Amphiphilic Compounds. Butterworths
alkali for high acid number oils. In: SPE EOR Conference at Oil and Gas West Asia. Scientific Publications.
Southwick, J.G., van den Pol, E., van Rijn, C.H., van Batenburg, D.W., Boersma, D., Wu, B., Xiu, J., Yu, L., Huang, L., Yi, L., Ma, Y., 2022. Biosurfactant production by
Svec, Y., Mastan, A.A., Shahin, G., Raney, K., 2016. Ammonia as alkali for alkaline/ Bacillus subtilis SL and its potential for enhanced oil recovery in low permeability
surfactant/polymer floods. SPE J. 21 (1), 10–21. reservoirs. Sci. Rep. 12 (1), 1–10.
Stephenson, T., Oswald, D., Dwyer, P., Brown, D., Ndefo, E., Smith, J., Gaffney, B., Wu, G., Yuan, C., Ji, X., Wang, H., Sun, S., Hu, S., 2017. Effects of head type on the
Kiran, S., 2019. Core floods vs. Field pilot–effectiveness of microemulsions in stability of gemini surfactant foam by molecular dynamics simulation. Chem. Phys.
conventional and unconventional waterfloods. In: SPE Annual Technical Conference Lett. 682, 122–127.
and Exhibition. OnePetro. Wu, Y., Iglauer, S., Shuler, P., Tang, Y., Goddard III, W.A., 2010. Branched alkyl alcohol
Stumm, W., Morgan, J.J., 2012. Aquatic Chemistry: Chemical Equilibria and Rates in propoxylated sulfate surfactants for improved oil recovery. Tenside Surfactants
Natural Waters, vol. 126. John Wiley and Sons. Deterg. 47 (3), 152–161.
Summers, M., Eastoe, J., 2003. Applications of polymerizable surfactants. Adv. Colloid Xie, X., Weiss, W.W., Tong, Z., Morrow, N.R., 2004. Improved oil recovery from
Interface Sci. 100, 137–152. carbonate reservoirs by chemical stimulation. In: SPE/DOE Symposium on Improved
Suniga, P.T., Fortenberry, R., Delshad, M., 2016. Observations of microemulsion Oil Recovery.
viscosity for surfactant EOR processes. In: SPE Improved Oil Recovery Conference. Xu, X., Chen, F., 2004. Modified emulsifier-free emulsion polymerization of butyl
Taber, J.J., 1983. Technical screening guides for the enhanced recovery of oil. In: SPE methacrylate with ionic or/and nonionic comonomers. J. Appl. Polym. Sci. 92 (5),
Annual Technical Conference and Exhibition. 3080–3087.
Taber, J.J., Martin, F.D., Seright, R.S., 1997. EOR screening criteria revisited-Part 1: Xue, W., Hamley, I.W., Castelletto, V., Olmsted, P.D., 2004. Synthesis and
introduction to screening criteria and enhanced recovery field projects. SPE characterization of hydrophobically modified polyacrylamides and some
Reservoir Eng. 12 (3), 189–198. observations on rheological properties. Eur. Polym. J. 40 (1), 47–56.
Tabor, R.F., Eastoe, J., Dowding, P., 2009. Adsorption and desorption of nonionic Yang, H., Britton, C., Liyanage, P.J., Solairaj, S., Kim, D.H., Nguyen, Q.,
surfactants on silica from toluene studied by ATR-FTIR. Langmuir 25 (17), Weerasooriya, U., Pope, G.A., 2010. Low-cost, high-performance chemicals for
9785–9791. enhanced oil recovery. In: SPE Improved Oil Recovery Symposium.
Tackie-Otoo, B.N., Mohammed, M.A.A., Zalghani, H.A.B.M., Hassan, A.M., Murungi, P.I., Yang, J., Zhou, Y., 2020. An automatic in situ contact angle determination based on level
Tabaaza, G.A., 2022. Interfacial properties, wettability alteration and emulsification set method. Water Resour. Res. 56 (7).
properties of an organic alkali–surface active ionic liquid system: implications for Yasin, A., Zhou, W., Yang, H., Li, H., Chen, Y., Zhang, X., 2015. Shape memory hydrogel
enhanced oil recovery. Molecules 27 (7), 2265. based on a Hydrophobically-Modified polyacrylamide (HMPAM)/α-CD mixture via a
Tagavifar, M., Jang, S.H., Sharma, H., Wang, D., Chang, L.Y., Mohanty, K., Pope, G.A., host-guest approach. Macromol. Rapid Commun. 36 (9), 845–851.
2018. Effect of pH on adsorption of anionic surfactants on limestone: experimental Ye, F., Guo, H., Zhang, H., 2008. Biomimetic synthesis of oriented hydroxyapatite
study and surface complexation modeling. Colloids Surf. A Physicochem. Eng. Asp. mediated by nonionic surfactants. Nanotechnology 19 (24), 245605.
538, 549–558. Yin, X., Zhao, T., Yi, J., 2021, November. The Combined Flooding of Dispersed Particle
Tagavifar, M., Herath, S., Weerasooriya, U.P., Sepehrnoori, K., Pope, G., 2016. Gel and Surfactant for Conformance Control and EOR: From Experiment to Pilot
Measurement of microemulsion viscosity and its implications for chemical EOR. In: Test. In: SPE International Conference on Oilfield Chemistry. OnePetro.
SPE Improved Oil Recovery Conference. OnePetro. Youssif, M.I., El-Maghraby, R.M., Saleh, S.M., Elgibaly, A., 2018. Silica nanofluid
Talebian, S.H., Masoudi, R., Tan, I.M., Zitha, P.L.J., 2014. Foam assisted CO2-EOR: a flooding for enhanced oil recovery in sandstone rocks. Egyptian Journal of
review of concept, challenges, and future prospects. J. Petrol. Sci. Eng. 120, Petroleum 27 (1), 105–110.
202–215. Yu, K.M.K., Steele, A.M., Zhu, J., Fu, Q., Tsang, S.C., 2003. Synthesis of well-dispersed
Talebian, S.H., Tan, I.M., Sagir, M., Muhammad, M., 2015. Static and dynamic foam/oil nanoparticles within porous solid structures using surface-tethered surfactants in
interactions: potential of CO2-philic surfactants as mobility control agents. J. Petrol. supercritical CO2. J. Mater. Chem. 13 (1), 130–134.
Sci. Eng. 135, 118–126. Yu, Q., Jiang, H., Zhao, C., 2010. Study of interfacial tension between oil and surfactant
Tanaka, R., Williams, P.A., Meadows, J., Phillips, G.O., 1992. The adsorption of polymer flooding. Petrol. Sci. Technol. 28 (18), 1846–1854.
hydroxyethyl cellulose and hydrophobically modified hydroxyethyl cellulose onto Yuncong, G.A.O., Mifu, Z.H.A.O., Jianbo, W.A.N.G., Chang, Z.O.N.G., 2014. Performance
polystyrene latex. Colloid. Surface. 66 (1), 63–72. and gas breakthrough during CO2 immiscible flooding in ultra-low permeability
Texter, J., 2001. Reactions and synthesis in surfactant systems. Surfactant Science Series, reservoirs. Petrol. Explor. Dev. 41 (1), 88–95.
ACS. Zargartalebi, M., Kharrat, R., Barati, N., 2015. Enhancement of surfactant flooding
Thomas, S., 2008. Enhanced oil recovery-an overview. Oil and Gas Science and performance by the use of silica nanoparticles. Fuel 143, 21–27.
Technology-Revue de l’IFP 63 (1), 9–19. Zhang, D.L., Liu, S., Puerto, M., Miller, C.A., Hirasaki, G.J., 2006a. Wettability alteration
Tsvetkov, N.V., Lebedeva, E.V., Lezov, A.A., Podseval’nikova, A.N., Akhmadeeva, L.I., and spontaneous imbibition in oil-wet carbonate formations. J. Pet. Sci. Eng. 52
Zorin, I.M., Bilibin, A.Y., 2014. Macromolecules of poly-(12- (1− 4), 213–226.
acryloylaminododecanoic acid) in organic solvent: synthesis and molecular Zhang, D., Liu, S., Yan, W., Puerto, M., Hirasaki, G.J., Miller, C.A., 2006b. Favorable
characteristics. Polymer 55 (7), 1716–1723. attributes of alkali-surfactant-polymer flooding. In: SPE/DOE Symposium on
Tunio, S.Q., Chandio, T.A., Memon, M.K., 2012. Comparative study of FAWAG and Improved Oil Recovery.
SWAG as an effective EOR technique for a Malaysian field. Res. J. Appl. Sci. Eng. Zhang, H., Ramakrishnan, T.S., Nikolov, A., Wasan, D., 2016. Enhanced oil recovery
Technol. 4 (6), 645–648. driven by nanofilm structural disjoining pressure: flooding experiments and
Udoh, T., Vinogradov, J., 2021. Controlled salinity-biosurfactant enhanced oil recovery microvisualization. Energy Fuel. 30 (4), 2771–2779.
at ambient and reservoir temperatures-an experimental study. Energies 14 (4), 1077. Zhang, P., Austad, T., 2006. Wettability and oil recovery from carbonates: effects of
Varfolomeev, M.A., Ziniukov, R.A., Yuan, C., Khairtdinov, R.K., Sitnov, S.A., Sudakov, V. temperature and potential determining ions. Colloids Surf. A Physicochem. Eng. Asp.
A., Zhdanov, M.V., Mustafin, A.Z., Usmanov, S.A., Sattarov, A.I., Glukhov, M.S., 279 (1–3), 179–187.
2020 October. Optimization of carbonate heavy oil reservoir development using Zhang, R., Somasundaran, P., 2006. Advances in adsorption of surfactants and their
surfactant flooding: from laboratory screening to pilot test. In: SPE Russian mixtures at solid/solution interfaces. Adv. Colloid Interface Sci. 123, 213–229.
Petroleum Technology Conference. OnePetro. Zhang, W., Mao, J., Yang, X., Zhang, H., Zhao, J., Tian, J., Lin, C., Mao, J., 2019.
Vermolen, E., Almada, P.M., Wassing, B.M., Ligthelm, D.J., Masalmeh, S.K., 2014. Low- Development of a sulfonic gemini zwitterionic viscoelastic surfactant with high salt
salinity polymer flooding: improving polymer flooding technical feasibility and tolerance for seawater-based clean fracturing fluid. Chem. Eng. Sci. 207, 688–701.
economics by using low-salinity make-up brine. In: International Petroleum Zhang, Y., Sarma, H., 2012. Improving waterflood recovery efficiency in carbonate
Technology Conference. reservoirs through salinity variations and ionic exchanges: a promising low-cost
Smart-Waterflood approach. In: Abu Dhabi International Petroleum Conference and
Exhibition.

36
A.M. Hassan et al. Journal of Petroleum Science and Engineering 220 (2023) 111243

Zhao, M., Lv, W., Li, Y., Dai, C., Wang, X., Zhou, H., Zou, C., Gao, M., Zhang, Y., Wu, Y., Zhong, W., Claverie, J.P., 2013. Probing the carbon nanotube-surfactant interaction for
2018. Study on the synergy between silica nanoparticles and surfactants for the preparation of composites. Carbon 51, 72–84.
enhanced oil recovery during spontaneous imbibition. J. Mol. Liq. 261, 373–378. Zhou, X., Morrow, N.R., Ma, S., 2000. Interrelationship of wettability, initial water
Zhao, Z., Bi, C., Qiao, W., Li, Z., Cheng, L., 2007b. Dynamic interfacial tension behavior saturation, aging time, and oil recovery by spontaneous imbibition and
of the novel surfactant solutions and Daqing crude oil. Colloid. Surface. waterflooding. SPE J. 5 (2), 199–207.
Physicochem. Eng. Aspect. 294 (1–3), 191–202. Zhu, B.-Y., Gu, T., 1991. Surfactant adsorption at solid-liquid interfaces. Adv. Colloid
Zhao, Z., Li, Z., Qiao, W., Wang, G., 2007a. Surface and interfacial tension behavior of Interface Sci. 37 (1–2), 1–32.
novel alkyl methylnaphthalene sulfonate surfactant solutions. Chapter 1. In: Fong, P. Ziegler, V.M., Handy, L.L., 1981. Effect of temperature on surfactant adsorption in
A. (Ed.), Colloid and Surface Research Trends. Nova Science Publishers Inc., porous media. Soc. Petrol. Eng. J. 21 (2), 218–228.
pp. 1–76 Al-Murayri, M.T., Al-Kharji, A.A., Kamal, D.S., Al-Ajmi, M.F., Al-Ajmi, R.N., Al-
Zhong, C., Huang, R., Xu, J., 2008. Characterization, solution behavior, and Shammari, M.J., Al-Asfoor, T.H., Badham, S.J., Bouma, C., Brown, J. and Suniga, H.
microstructure of a hydrophobically associating nonionic copolymer. J. Solut. Chem. P., 2018, March. Successful Implementation of a One-Spot Alkaline-Surfactant-
37 (9), 1227–1243. Polymer ASP Pilot in a Giant Carbonate Reservoir. In SPE EOR Conference at Oil and
Zhong, C., Luo, P., Ye, Z., Chen, H., 2009. Characterization and solution properties of a Gas West Asia. OnePetro.
novel water-soluble terpolymer for enhanced oil recovery. Polym. Bull. 62 (1),
79–89.

37

You might also like